|
REVIEW article
Front. Immunol., 07 March 2024
Sec. Molecular Innate Immunity
Volume 15 - 2024 | https://doi.org/10.3389/fimmu.2024.1359600
This article is part of the Research TopicReactive Oxygen Species (ROS) Signaling and Immune Diseases.View all 9 articles
ROS signaling in innate immunity via oxidative protein modifications
The innate immune response represents the first-line of defense against invading pathogens. Reactive oxygen species (ROS) and reactive nitrogen species (RNS) have been implicated in various aspects of innate immune function, which involves respiratory bursts and inflammasome activation. These reactive species widely distributed within the cellular environment are short-lived intermediates that play a vital role in cellular signaling and proliferation and are likely to depend on their subcellular site of formation. NADPH oxidase complex of phagocytes is known to generate superoxide anion radical (O2•−) that functions as a precursor for antimicrobial hydrogen peroxide (H2O2) production, and H2O2 is utilized by myeloperoxidase (MPO) to generate hypochlorous acid (HOCl) that mediates pathogen killing. H2O2 modulates the expression of redox-responsive transcriptional factors, namely NF-kB, NRF2, and HIF-1, thereby mediating redox-based epigenetic modification. Survival and function of immune cells are under redox control and depend on intracellular and extracellular levels of ROS/RNS. The current review focuses on redox factors involved in the activation of immune response and the role of ROS in oxidative modification of proteins in macrophage polarization and neutrophil function.
선천성 면역 반응은
침입하는 병원체에 대한 첫 번째 방어선입니다.
반응성 산소 종(ROS)과 반응성 질소 종(RNS)은
호흡기 발작과 염증성 사슬 활성화와 관련된
선천성 면역 기능의 다양한 측면에 관여합니다.
세포 환경 내에 널리 분포되어 있는 이러한 반응성 종은
세포 신호 전달과 증식에 중요한 역할을 하는 수명이 짧은 중간체이며,
세포 내 형성 부위에 의존할 가능성이 높습니다.
식세포의 NADPH 산화효소 복합체는
항균성 과산화수소(H2O2) 생산의 전구체로 작용하는
슈퍼옥사이드 음이온 라디칼(O2•−)을 생성하는 것으로 알려져 있으며,
H2O2는 골수과산화효소(MPO)에 의해 병원균을 죽이는 데 관여하는
차아염소산(HOCl)을 생성하는 데 사용됩니다.
H2O2는
산화-환원 반응에 반응하는 전사 인자,
즉 NF-kB, NRF2, HIF-1의 발현을 조절함으로써
산화-환원 기반의 후성유전학적 변형을 매개합니다.
면역 세포의 생존과 기능은
산화-환원 조절을 받으며,
세포 내외의 ROS/RNS 수준에 따라 달라집니다.
이 리뷰에서는
면역 반응 활성화에 관여하는 산화-환원 인자와 대식세포 분화 및
호중구 기능에서 단백질의 산화적 변형에 있어서 ROS의 역할에 초점을 맞춥니다.
1 Introduction
In biology, reactive species formed by redox reaction or electronic excitation hold a growing interest due to their significant impact on the spectrum of pathological processes, including inflammation and aging (1). Based on the nature of reactive atoms, they are named reactive oxygen species (oxygen), reactive nitrogen species (nitrogen) and reactive sulfur species (sulfur), respectively. Produced by nearly all organisms and cells, ROS consists of two subclasses: highly reactive radical and non-radical species. Free oxygen radicals include superoxide anion (O2•-), hydroxyl (HO•), peroxyl (ROO•), and alkoxyl (RO•), while hydrogen peroxide (H2O2), singlet oxygen (1O2), hypochlorite anion and ozone represent the non-radical species (2) (Figure 1). ROS subcellular origin and its levels within the cellular environment defines its role. Its function as physiological secondary messengers in signal transduction becoming increasingly apparent; for example, ROS oxidize sulfhydryl (SH) groups of cysteine residues in protein kinases, including protein kinase A(PKA), protein kinase C (PKC), Calcium–calmodulin (CaM)-dependent protein kinase II (CaMKII) and receptor tyrosine kinase (RTK), which activate and phosphorylate their protein targets involved in signaling (3–5).
1 서론
생물학에서
산화 환원 반응이나 전자 여기(excitation)에 의해 형성된 반응성 종은
염증과 노화를 포함한 병리학적 과정의 스펙트럼에 상당한 영향을 미치기 때문에
점점 더 많은 관심을 받고 있습니다(1).
반응성 원자의 성질에 따라
각각 반응성 산소 종(oxygen), 반응성 질소 종(nitrogen), 반응성 황 종(sulfur)으로 명명됩니다.
거의 모든 유기체와 세포에서 생성되는 ROS는
반응성이 높은 라디칼과 비라디칼 종이라는 두 가지 하위 분류로 구성됩니다.
highly reactive radical and
non-radical species.
자유 산소 라디칼에는
슈퍼옥사이드 음이온(O2-), 하이드록실(HO-), 퍼옥실(ROO-), 알콕실(RO-)이 포함되며,
과산화수소(H2O2), 일중항 산소(1O2), 차아염소산 음이온, 오존은
ROS의 세포 내 기원과 세포 환경 내 수준에 따라
그 역할이 결정됩니다.
신호 전달에서
생리적 2차 메신저로서의 기능이 점점 더 분명해지고 있습니다.
예를 들어,
ROS는
단백질 키나아제 A(PKA),
단백질 키나아제 C(PKC),
칼슘-칼모둘린(CaM) 의존성 단백질 키나아제 II(CaMKII), 수용체 티로신 키나아제(RTK)를 포함한
단백질 키나아제에서
시스테인 잔기의 설프히드릴(SH) 그룹을 산화시킵니다.
ROS oxidize sulfhydryl (SH) groups of cysteine residues in protein kinases, including
protein kinase A(PKA),
protein kinase C (PKC),
Calcium–calmodulin (CaM)-dependent protein kinase II (CaMKII) and
receptor tyrosine kinase (RTK),
which activate and phosphorylate their protein targets involved in signaling.
이 그룹은
신호 전달에 관여하는 단백질 표적을
Figure 1
Figure 1 (A) Types of Reactive oxygen species and their (B) intracellular sources. Figure 1B was Created with BioRender.com.
Reported as the major subspecies within cells, O2•- and H2O2 highly differ in their chemical parameter and functions. An increase in cellular O2•- levels remains closely associated with oxidative stress and cellular damage, wherein the oxidation of biological macromolecules and irreversible protein inactivation disturb cellular signaling events (6). H2O2 is highly diffusible and relatively stable compared to O2•- and is considered a pleiotropic physiological signaling agent. In general, H2O2 at physiological pH mediates signaling via oxidation of cysteine residues, wherein the exposed thiol group (Cys-SH) are deprotonated to the thiolate group (Cys-S-) and susceptible to oxidation. H2O2-dependent signaling occurs with a lower intracellular concentration of approximately 1-100 nM, mediating reversible oxidation of the thiolate group to a sulfenic group (Cys-SOH) and covalent linkage of cysteine residues by disulfide bonds (Cys-S-S-Cys) altering its activity and localization (7). An antioxidant defense system can reverse such protein oxidation and thus serve as essential redox switches in various cellular processes. Nevertheless, excessive production of H2O2 mediates irreversible nonspecific oxidation of proteins, resulting in a state referred to as oxidative stress (8). This inherent duality of ROS serving as a beneficial secondary messenger in signaling and causing harmful effects through the accumulation of protein adducts signifies its antagonistic pleiotropy (6).
ROS regulate cellular homeostasis and signaling events within a cellular environment, thus serving as primary modulators of cellular dysfunction and contributing to disease pathology. In almost every subcellular organelle, including cytoplasm, endoplasmic reticulum (ER), mitochondria and peroxisomes, ROS are generated as byproducts as a part of its basal metabolic function (9, 10). An imbalance in ROS generation and scavenging by antioxidants results in pathological conditions, including cancer, neurodegenerative disorders and atherosclerosis. Studies carried out so far substantiate the role of ROS in various metabolic processes like mitochondrial ROS promoting monocyte migration, wherein NADPH oxidase (NOX)-mediated cytosolic ROS production aids in innate immunity (11). Depending on cell type, tissue environment and source, ROS participates in normal physiological processes or contribute to metabolic dysfunction and inflammatory signaling in sterile and infectious inflammation.
세포 내에서 주요 아종으로 보고된 O2•-와 H2O2는
화학 매개변수와 기능 면에서 매우 다릅니다.
세포 내 O2•- 수치의 증가는
산화 스트레스 및 세포 손상과 밀접한 관련이 있으며,
생물학적 거대 분자의 산화와 돌이킬 수 없는
단백질 불활성화가 세포 신호 전달 과정을 방해합니다(6).
H2O2는
O2•-에 비해 확산성이 높고 상대적으로 안정적이며,
다발성 생리 신호 전달 물질로 간주됩니다.
일반적으로,
생리학적 pH에서 H2O2는
시스테인 잔기의 산화를 통해 신호를 전달하는 매개체 역할을 합니다.
이 과정에서 노출된 티올기(Cys-SH)는
탈양성자화되어 티올레이트기(Cys-S-)가 되고
산화되기 쉽습니다.
H2O2 의존 신호 전달은
약 1-100 nM의 낮은 세포 내 농도에서 발생하며,
티올산염 그룹의 가역적 산화를 설펜산 그룹(Cys-SOH)으로, 그리고
시스테인 잔기의 공유 결합을 이황화 결합(Cys-S-S-Cys)으로 매개하여
그 활동과 국소화를 변화시킵니다(7).
항산화 방어 시스템은
이러한 단백질 산화를 역전시킬 수 있으며,
따라서 다양한 세포 과정에 필수적인 산화 환원 스위치 역할을 합니다.
그럼에도 불구하고,
H2O2의 과도한 생산은
돌이킬 수 없는 비특이적 단백질 산화를 유발하여
산화 스트레스(oxidative stress)라고 불리는 상태를 초래합니다(8).
신호 전달에 유익한 2차 메신저 역할을 하는 ROS의 이러한 고유한 이중성은
단백질 부가물의 축적을 통해 유해한 영향을 유발한다는 것을 의미하며,
이는 ROS의 길항성 다중성(antagonistic pleiotropy)을 나타냅니다(6).
ROS는
세포 내 환경 내에서 세포 항상성과 신호 전달 과정을 조절함으로써
세포 기능 장애의 주요 조절자 역할을 하며,
질병의 병리학에 기여합니다.
세포질, 소포체(ER), 미토콘드리아, 퍼옥시좀을 포함한
거의 모든 세포 소기관에서,
ROS는 기초 대사 기능의 일부로서 부산물로 생성됩니다(9, 10).
항산화제에 의한 ROS 생성 및 제거의 불균형은
암, 신경 퇴행성 장애, 동맥 경화증 등의 병리학적 상태를 초래합니다.
지금까지 수행된 연구들은
미토콘드리아 ROS가 단핵구 이동을 촉진하는 등
다양한 대사 과정에서 ROS의 역할을 입증하고 있으며,
NADPH 산화효소(NOX)에 의한 세포질 ROS 생성이
선천성 면역에 도움이 된다는 사실도 밝혀졌습니다(11).
세포 유형, 조직 환경, 출처에 따라
ROS는
정상적인 생리 과정에 관여하거나
무균 및 감염성 염증에서 대사 기능 장애 및 염증 신호 전달에 기여합니다.
2 ROS in sensing and reaction to damage
2.1 ROS in infectious inflammation
During the initial stage of an inflammatory process, immune cells activated in response to invading pathogens or agents release various inflammatory mediators, aiding in increased vascular permeability and leukocyte migration towards the tissue injury site (12). ROS regulates various intracellular adhesion molecules, including ICAM-1, VCAM-1 and selectin expression, ensuring their interaction with leukocytes and transendothelial migration. On the contrary, enhanced expression of superoxide dismutase reduces leukocyte binding to endothelium by decreasing the expression of adhesion molecules (13). Additionally, a gradient increase in H2O2 at the tissue injury site is an early factor in leukocyte recruitment (14). Cytokines, including tumor necrosis factor (TNF), vascular endothelial growth factor (VEGF), and NOX stimulate cell migration and adhesion (15).
Metabolic reprogramming, which includes changes in the activity of fatty acid oxidation (FAO), tricarboxylic acid (TCA) cycle, glycolysis levels, involvement of pentose phosphate pathway and mitochondrial respiration mediates the phenotypical and functional role of myeloid cells.
Over the past two decades, immune cell functions, namely proliferation and differentiation, have been linked to various metabolic pathways; thus, immunometabolism is inseparable from redox reactions (11). Thus, bioenergetic and biosynthetic demands for T and B cell response, macrophages and dendritic cells are regulated by metabolic pathway and their resulting breakdown products (16). In aerobic organisms, during oxidative phosphorylation in mitochondria, the electron transport chain shuttles electrons to molecular oxygen (O2), producing free oxygen radicals. ROS are well known for their role in inflammation through respiratory bursts, wherein macrophages and neutrophils phagocytize pathogens and cellular debris. The process activates the NOX assembly and activation, which triggers further ROS production (17). NOXs play a vital role in inflammatory response through respiratory burst and neutrophil extracellular trap formation. NOX-derived ROS are reportedly involved in angiogenesis, a process of the growth of new vessels and a key event in the proliferative phase of inflammation. It has been reported that NOX remains localized at the leading edge of migrating cells and is linked to actin and IQGAP1 protein (Ras GTPase-activating-like protein), and any disruption at this binding site results in impaired migration of endothelial cells (18).
Nevertheless, enhanced production of reactive species has to be regulated to avoid oxidation of biomolecules leading to cellular toxicity and cell death. Alterations in redox balance are often associated with chronic immune activation, such as autoimmune disorders and neurodegenerative diseases. Therefore, the redox state is an intrinsic cellular and systemic homeostasis indicator.
2 손상 감지와 반응에 관여하는 ROS
2.1 감염성 염증에 관여하는 ROS
염증 과정의 초기 단계에서 침입한
병원체나 물질에 반응하여
활성화된 면역 세포는
다양한 염증 매개체를 방출하여 혈관 투과성을 높이고
조직 손상 부위로 백혈구가 이동하도록 돕습니다(12).
ROS는
ICAM-1, VCAM-1, 셀렉틴 발현을 포함한
다양한 세포 내 부착 분자를 조절하여
백혈구와의 상호 작용과 내피 투과성 이동을 보장합니다.
반대로,
슈퍼옥사이드 디스뮤타제의 발현이 증가하면,
접착 분자의 발현이 감소하여 백혈구가 내피에 결합하는 것이 감소합니다(13).
또한,
조직 손상 부위에서 H2O2의 점진적 증가는
백혈구 모집의 초기 요인입니다(14).
종양 괴사 인자(TNF),
혈관 내피 성장 인자(VEGF),
NOX를 포함한 사이토카인은
세포 이동과 접착을 촉진합니다(15).
대사 재프로그래밍은
지방산 산화(FAO), 트리카르복실산(TCA) 사이클, 당분해 수준, 펜토스 인산 경로의 관여,
미토콘드리아 호흡의 변화를 포함하며
골수성 세포의 표현형 및 기능적 역할을 매개합니다.
지난 20년 동안 면역세포 기능,
즉 증식과 분화는
다양한 대사 경로와 관련되어 왔습니다.
따라서
면역 대사는
산화 환원 반응과 불가분의 관계에 있습니다(11).
따라서
T세포와 B세포 반응, 대식세포, 수지상 세포에 대한
생체 에너지와 생합성 요구는
대사 경로와 그 결과로 생성되는 분해 산물에 의해 조절됩니다(16).
호기성 유기체에서
미토콘드리아의 산화적 인산화 과정에서
전자 수송 사슬은 전자를 분자 산소(O2)로 이동시켜
자유 산소 라디칼을 생성합니다.
ROS는
호흡기 파열을 통해 염증에 관여하는 것으로 잘 알려져 있으며,
과정에서 대식세포와 호중구가 병원체와 세포 파편을 식균합니다.
이 과정은
NOX 조립과 활성화를 활성화하여
추가적인 ROS 생성을 유발합니다(17).
NOX는
호흡기 파열과 호중구 세포외 트랩 형성을 통한
염증 반응에 중요한 역할을 합니다.
NOX에서 유래된 ROS는
새로운 혈관 성장 과정인 혈관 신생과 염증의 증식 단계에서
핵심적인 사건에 관여하는 것으로 알려져 있습니다.
NOX는
이동하는 세포의 선단부에 국한되어 있으며,
액틴과 IQGAP1 단백질(Ras GTPase 활성화 단백질)과 연결되어 있으며,
이 결합 부위가 파괴되면 내피 세포의 이동이 손상된다는 보고가 있습니다(18).
그럼에도 불구하고,
반응성 종의 생산 증가는
세포 독성과 세포 사멸을 초래하는 생체 분자의 산화를 피하기 위해
규제되어야 합니다.
산화 환원 균형에 변화가 생기면
자가 면역 질환이나 신경 퇴행성 질환과 같은
만성 면역 활성화가 자주 발생합니다.
따라서,
산화 환원 상태는
본질적인 세포 및 전신 항상성 지표입니다.
2.2 ROS in sterile inflammation
The release of local damage-associated molecular patterns (DAMPs) in response to tissue damage elicits enhanced cytokine production via sterile inflammation. Cytokines belonging to the interleukin-1(IL-1) family have been proposed to be essential drivers of sterile inflammation, which recruits neutrophils and macrophages to the damaged site. During sterile inflammation, host-derived DAMPs released by stressed cells contribute to macrophage polarization that aids in resolving sterile inflammation and restores homeostasis (19). In endothelial cells, under hyperglycemic conditions, an excessive glucose load triggers ROS formation in mitochondria, impairing its functions and causing cellular damage through interaction with various cellular constituents, including DNA, proteins and lipids (20). These ROS trigger the activation of pro-inflammatory transcription factors, namely NFκB and activating protein-1(AP-1), resulting in enhanced inflammatory cytokines/chemokines expression. In addition, activated endothelial cells attract monocytes that promote inflammation and macrovascular and microvascular injury (21).
Macrophages, especially adipose tissue macrophages (ATM), drive diabetic pathology. In healthy individuals, adipocytes secrete adiponectin, which induces M2-like polarization of ATM and suppresses ROS and its related pathway genes (22). On the contrary, in obesity conditions, a reduction in adiponectin levels induces M1-like polarized macrophages, thereby enhancing glucose consumption through GLUT1 (glucose transporter 1). Thus, an interplay between obesity and hyperglycemia promotes ROS formation, glycolytic metabolism, and the release of pro-inflammatory cytokine mediators by macrophages (23). Hence, focusing more on metabolic reprogramming during macrophage polarization events is essential to identify potential targets for treating inflammation and metabolic disorders. A mitochondrial reactive oxidative stress increase was reported in mice subjected to ventilator-induced lung injury, activating NLRP3 to produce IL-1β and lung inflammation in combination with TLR4 signaling (24).
2.2 무균성 염증의 ROS
조직 손상에 대한 반응으로
국소 손상 관련 분자 패턴(DAMP)이 방출되면
무균성 염증을 통해 사이토카인 생산이 증가합니다.
인터루킨-1(IL-1) 계열에 속하는 사이토카인은
손상 부위에 호중구와 대식세포를 모집하는
무균성 염증의 필수적인 촉진제로 제안되었습니다.
무균성 염증 동안,
스트레스를 받은 세포에 의해 방출된 숙주 유래 DAMP는
무균성 염증의 해결을 돕고 항상성을 회복하는 데 도움이 되는
대식세포 분극화에 기여합니다(19).
내피 세포에서 고혈당 상태가 되면
과도한 포도당 부하가
미토콘드리아에서 ROS 형성을 유발하여
그 기능을 손상시키고
DNA, 단백질, 지질 등 다양한 세포 구성 요소와의 상호 작용을 통해
세포 손상을 유발합니다(20).
https://pmc.ncbi.nlm.nih.gov/articles/PMC3856020/
이러한 ROS는
전염증성 전사 인자,
즉 NFκB 및 활성화 단백질-1(AP-1)의 활성화를 유발하여
염증성 사이토카인/케모카인 발현을 강화합니다.
또한,
활성화된 내피세포는
염증과 거대혈관 및 미세혈관 손상을 촉진하는
단핵구(monocyte)를 유인합니다(21).
대식세포,
특히 지방조직 대식세포(ATM)는 당뇨병 병리를 유발합니다.
건강한 사람의 경우,
지방세포는 지방분해효소인 아디포넥틴(adiponectin)을 분비하는데,
이 물질은 ATM의 M2형 분화를 유도하고
ROS와 관련 경로 유전자를 억제합니다(22).
반대로,
비만 상태에서는 아디포넥틴 수치가 감소하면
M1-like 분화 대식세포가 유도되어,
GLUT1(포도당 수송체 1)을 통한 포도당 소비가 증가합니다.
따라서,
비만과 고혈당증의 상호 작용은
ROS 형성, 당분해 대사, 대식세포에 의한
전염증성 사이토카인 매개체의 방출을 촉진합니다(23).
따라서,
대식세포 분화 과정에서 대사 재프로그래밍에 더 집중하는 것은
염증과 대사 장애를 치료할 수 있는 잠재적 표적을 파악하는 데 필수적입니다.
인공호흡기 유발 폐 손상을 입은 쥐에서
미토콘드리아 반응성 산화 스트레스가 증가하여
NLRP3가 활성화되어 IL-1β와 TLR4 신호 전달과 결합된 폐 염증을 유발하는 것으로 보고되었습니다(24).
2.3 Sources of ROS production
Though the primary source of ROS in vivo is through aerobic respiration, cellular events, including peroxisomal β-oxidation of fatty acids, arginine metabolism, tissue-specific cellular enzymes and phagocytosis stimulation by pathogens also contributed to ROS production (3). Based on its source, cell type and environment, ROS signaling contributes to either normal physiological processes or metabolic dysfunction through inflammatory signaling. Diseased conditions, including diabetes mellitus, atherosclerosis and stroke, are known to be associated with redox balance (25).
The NADPH oxidase (NOX) family of proteins remains the primary cytosolic source of ROS, consist of seven different isoforms and comprises membrane and cytosolic components that are actively involved in the host response to various stimuli including bacterial and viral infections, cellular signaling and regulation of gene expression. Among the isoforms, NOX2 is well characterized for its role in phagocytic functions. Both NOX and inducible nitric oxide synthase (iNOS) are vital in generating enhanced ROS levels within phagocytes via oxidative burst to kill invading pathogens (26). In comparison to mitochondria, O2•- produced by NOX is dismutated into H2O2 by superoxide dismutase1 (SOD1), whereas nitric oxide produced by inducible nitric oxide synthase (iNOS) reacts with O2•- resulting in peroxynitrite production (19). For example, ROS drive hypoxia-inducible factor 1α (HIF1α) mediated GLUT1 expression, hexokinase activity, and resultant glycolysis in response to low oxygen tension as part of the angiogenic response (27) The co-localization of neutrophil phosphofructokinase 2 with NOX2 leads to its activation, resulting in NADP+ production as its byproduct, facilitating an enhanced glycolytic rate. The increase in glycolysis rate and enhancement of NOX2 activity and the relation between these processes are still under study (28).
Mitochondria are considered as the redox-active compartment within the cell, accounting for nearly 90% of oxygen utilization (1). They serve as a significant contributor to ROS in the form of O2•-. Mitochondrial SOD converts O2•- into H2O2, which in turn gets converted into HO• through the Fenton reaction, which in turn oxidizes biomolecules. SOD1 is constitutively expressed and regulates cytosolic O2•- levels (29). Factors like hyperoxia, oxidative stress, and inflammatory cytokines induce SOD2, whereas SOD3 are cell and tissue-specific and likely of significant importance in protection against stress factors from the extracellular environment. In addition, Grx (glutaredoxin), glutathione, and Trx (thioredoxin) systems play a predominant role in mitochondrial ROS buffering, like SOD (30).
2.3 ROS 생산의 원인
생체 내 ROS의 주요 원인은
호기성 호흡이지만,
지방산의 퍼옥시좀 β-산화,
아르기닌 대사,
조직 특이적 세포 효소,
병원체에 의한 식세포 자극을 포함한 세포 활동도 ROS 생산에 기여합니다(3).
그 원인과 세포 유형, 환경에 따라,
ROS 신호는
정상적인 생리 과정이나 염증 신호에 의한
대사 기능 장애에 기여합니다.
당뇨병, 동맥경화증, 뇌졸중 등의 질병은
산화 환원 균형과 관련이 있는 것으로 알려져 있습니다(25).
NADPH oxidase(NOX) 단백질군은
세포질에서 ROS의 주요 공급원으로 남아 있으며,
7개의 다른 이소형으로 구성되어 있으며,
박테리아 및 바이러스 감염, 세포 신호 전달, 유전자 발현 조절 등
다양한 자극에 대한 숙주 반응에 적극적으로 관여하는
막 및 세포질 구성 요소를 포함합니다.
이소폼 중에서도 NOX2는 식세포 기능에 관여하는 것으로 잘 알려져 있습니다.
NOX와 유도성 산화질소 합성효소(iNOS)는
모두 산화적 폭발을 통해
식세포 내에서 ROS 수준을 향상시켜
침입하는 병원체를 죽이는 데 필수적입니다(26).
미토콘드리아에 비해,
NOX에 의해 생성된 O2-는
슈퍼옥사이드 디스뮤타제1(SOD1)에 의해 H2O2로 분해되는 반면,
유도성 산화질소 합성효소(iNOS)에 의해 생성된 산화질소는
O2-와 반응하여 퍼옥시니트라이트를 생성합니다(19).
예를 들어,
ROS는
저산소증에 반응하여 저산소증 유도 인자 1α(HIF1α)에 의해 매개되는
GLUT1 발현, 헥소키나제 활성, 그리고 그 결과로 일어나는 당분해를 촉진하여
혈관 신생 반응의 일부로 작용합니다(27).
호중구 포스포프럭토키나제 2와 NOX2의 공동 국소화는
NOX2의 활성화를 유도하여,
그 부산물인 NADP+의 생성을 촉진함으로써 당분해 속도를 향상시킵니다.
글리콜 분해율의 증가와 NOX2 활동의 향상,
그리고 이 두 과정의 관계는 아직 연구 중입니다(28).
미토콘드리아는
세포 내 산화 환원 활성 구획으로 간주되며,
산소 활용의 거의 90%를 차지합니다(1).
미토콘드리아는
O2•-의 형태로 ROS에 중요한 기여를 합니다.
미토콘드리아 SOD는
O2-를 H2O2로 전환하고,
이 H2O2는 펜톤 반응을 통해
HO-로 전환되어 생체 분자를 산화시킵니다.
SOD1은
항상 발현되어 세포질 내 O2- 수준을 조절합니다(29).
고산소증, 산화 스트레스, 염증성 사이토카인과 같은 요인은
SOD2를 유도하는 반면,
SOD3는 세포 및 조직 특이적이며
세포 외 환경의 스트레스 요인으로부터 보호하는 데
중요한 역할을 할 가능성이 높습니다.
또한,
Grx(글루타레독신), 글루타티온, Trx(티오레독신) 시스템은
SOD(30)와 같이 미토콘드리아 ROS 완충에 중요한 역할을 합니다.
Under stress conditions, the ER tubular network holds a unique oxidizing environment, wherein redox signaling mediators play a vital role in ROS generation and mediate protein folding. Protein folding is highly sensitive to ER redox status, and dysregulation of disulfide bond formation in response to ER stress increases luminal oxidative stress, leading to a decline in ER function. During protein folding, protein disulfide isomerase (PDI) and endoplasmic reticulum oxidoreductase 1 (ERO1) introduce disulphide bonds into folded proteins, resulting in H2O2 formation (31). Protein disulfide isomerase introduces disulfide bonds onto protein substrates through thiol oxidation, resulting in a reduced state. However, PDI is reoxidized through ERO1, which transfers electrons from O2 through the flavin adenine dinucleotide cofactor, forming H2O2 (32). In addition to the PD1/ERO1 pathway, ROS are produced through NOX4, NADPH-P450 reductase (NPR), and GSH (33). NOX4 is reported to be consistently associated with ER, and NOX4 associated with p22phox utilizes NADH or NADPH as an electron donor to produce O2•-. They are also reported to interact with PDI, whereas the absence of PDI results in cell death (34). In macrophages, the interaction between p22phox and PDI was observed (26).
Peroxisomes, like mitochondria, are vital organelles that regulate crucial processes such as α- and β-oxidation, amino acid catabolism, glyoxylate metabolism, ketogenesis, polyamine oxidation and isoprenoid and cholesterol metabolism (35). Peroxisomal electron transfer leads to free electrons rather than ATP, which are transferred to H2O to form H2O2. In peroxisomes, H2O2 is produced by various oxygen-consuming oxidases, including D-amino acid oxidase, xanthine oxidase, d-aspartate oxidase, polyamine oxidase and acyl-CoA oxidase. In addition, peroxisomal oxidases and xanthine oxidases generate ROS and nitric oxide (36). The lysosomal electron transport chain generates HO• via proton translocation to maintain an optimal pH for acidic hydrolases (9).
Thus, ROS are produced as a byproduct of cellular events wherein the NOX family of proteins mediate the reduction of O2 to O2•- and phagocytes NOX accounts for an increased amount of O2•- and H2O2 production by respiratory burst (3). Upon stimulation, the membrane-bound catalytic core assembles with proteins from the cytosol (p47 phox, p67 phox and small G protein Rac), activating O2•- production. •NO produced by Nitric oxide synthase (NOS) can migrate through the cell membrane via diffusion and mediate several signaling pathways in a dose-dependent manner (37). However, inflammatory activation of iNOS by cytokines or lipopolysaccharides enhances cellular levels of •NO and results in inflammatory diseases and septic shock (38). Both oxidative and nitrosative stress can hinder the functioning of intracellular redox buffer systems, resulting in decreased antioxidant capacity of affected cells (39). Thus, a proper balance between ROS-RNS is essential in regulating immunological response. A schematic representation of types and sources of ROS are presented in Figures 1A, B, respectively.
스트레스가 가해지는 조건에서,
ER 관형 네트워크는 독특한 산화 환경을 유지하는데,
이 환경에서 산화 환원 신호 매개체가 ROS 생성에 중요한 역할을 하고
단백질 폴딩을 매개합니다.
단백질 폴딩은
ER 산화 환원 상태에 매우 민감하며,
ER 스트레스에 대한 반응으로
이황화 결합 형성의 조절 장애가 발생하면 관
내부의 산화 스트레스가 증가하여
ER 기능이 저하됩니다.
단백질 접힘 과정에서,
단백질 디설파이드 이소머라제(PDI)와 소포체 산화환원효소 1(ERO1)은
접힌 단백질에 디설파이드 결합을 도입하여
H2O2를 생성합니다(31).
단백질 디설파이드 이소머라제는
티올 산화를 통해 단백질 기질에 디설파이드 결합을 도입하여
환원 상태를 만듭니다.
그러나
PDI는 ERO1을 통해 재산화되어, O2에서 플라빈 아데닌 디뉴클레오티드 보조인자를 통해 전자를 전달하여 H2O2(32)를 형성합니다. PD1/ERO1 경로 외에도, ROS는 NOX4, NADPH-P450 환원효소(NPR), GSH(33)를 통해 생성됩니다. NOX4는 ER과 지속적으로 연관되어 있는 것으로 보고되고 있으며, p22phox와 연관된 NOX4는 NADH 또는 NADPH를 전자 공여체로 활용하여 O2•-를 생성합니다. 또한, PDI와 상호 작용하는 것으로 보고되고 있으며, PDI가 없으면 세포가 죽습니다(34). 대식세포에서 p22phox와 PDI 간의 상호 작용이 관찰되었습니다(26).
페록시좀은
미토콘드리아와 마찬가지로
α- 및 β-산화, 아미노산 이화,
글리옥실산 대사,
케톤 생성,
폴리아민 산화,
이소프레노이드 및 콜레스테롤 대사(35)와 같은
중요한 과정을 조절하는 필수적인 세포기관입니다.
과산화소체에서의 전자 전달은 ATP가 아닌 자유 전자를 생성하며,
이 자유 전자는 H2O로 전달되어 H2O2를 형성합니다.
과산화소체에서 H2O2는
D-아미노산 산화효소, 크산틴 산화효소, d-아스파르트산 산화효소, 폴리아민 산화효소, 아실-CoA 산화효소 등
다양한 산소 소비 산화효소에 의해 생성됩니다.
또한, 페록시좀 산화효소와 크산틴 산화효소는
ROS와 산화질소를 생성합니다(36).
리소좀 전자 수송 사슬은
산성 가수분해 효소에 대한 최적의 pH를 유지하기 위해
양성자 전좌를 통해 HO-를 생성합니다(9).
따라서
ROS는 세포 사건의 부산물로 생성되는데,
이때 NOX 단백질 군은 O2를 O2-로 환원시키는 과정을 매개하고,
식세포의 NOX는 호흡기 폭발(3)에 의해 증가된 양의 O2-와 H2O2 생산을 설명합니다.
자극이 가해지면,
막에 결합된 촉매 코어는 세포질(p47 phox, p67 phox, 작은 G 단백질 Rac)의 단백질과 결합하여
O2- 생산을 활성화합니다.
산화질소 합성효소(NOS)에 의해 생성된 •NO는
확산 작용을 통해 세포막을 통과할 수 있으며,
용량 의존적 방식으로 여러 가지 신호 전달 경로를 매개할 수 있습니다(37).
그러나
사이토카인이나 리포폴리사카라이드에 의한 iNOS의 염증 활성화는
세포 수준의 •NO를 증가시켜 염증성 질환과 패혈성 쇼크를 유발합니다(38).
산화 스트레스와 니트로사이드 스트레스는
모두 세포 내 산화 환원 완충 시스템의 기능을 방해하여,
영향을 받은 세포의 항산화 능력을 감소시킵니다(39).
따라서
면역 반응을 조절하는 데 있어
ROS-RNS 사이의 적절한 균형이 필수적입니다.
ROS의 유형과 원인에 대한 도식적 표현은
각각 그림 1A, B에 나와 있습니다.
2.4 Mechanisms of oxidative protein modifications
2.4.1 Oxidation of sulfur-containing and aromatic amino acids
Sulfur-containing biomolecules are crucial in protein folding, deactivation of reactive species, enzymes, redox signaling and other biochemical functions. Remarkably, most of the functions are associated with proteins and protein adducts, whereas its functions can be traced back to two amino acids, cysteine and methionine and their respective thiol or thioether functionality (40, 41). Oxidative stress targets the sulfhydryl group of cysteine and the methionine thioether group, resulting in increased post-translational modification events. Being sensitive to redox transformations, thiol, the side chain of cysteine, acquires different oxidation states. While thiol and disulfide are commonly known, growing evidence of protein modification also reports other oxygen derivatives, including sulfenic, sulfinic and sulfonic derivatives. Disulfide formation remains the most common thiol oxidation wherein the disulfide bonds are reasonably stable and stabilize protein structures via intra- and intermolecular disulfide bridges. Cysteine thiyl radical and sulfenic acid formation is reversible, and both intermediates are highly unstable. Both serve as precursors for several oxidized cysteine modifications (42). Sulfenic acids are highly reactive and play a prominent role in enzyme catalysis and cell signaling. They remain as key intermediates to other oxidation states, namely sulfinic and sulfonic acids. During the inflammatory process, immune cells, thiol or thiolate anion reaction with hypochlorous acid (HOCl) result in the formation of sulfenic acids (43).
Approximately 5% of cellular proteins remain in either sulfinic or sulfonic acid forms. The functional role of sulfinic acid modification has been reported mainly with the peroxiredoxin (Prxs) family (44), which reduces H2O2 and alkyl peroxides to water and alcohol. Under physiological conditions, in contrast to sulfenic acids, sulfinic derivatives do not react with thiols or undergo self-condensation reactions. Sulfinic acids are stable intermediates but oxidize readily to sulfonic acid (RSO3H), the most highly oxidized species of thiols and disulfides. Potent oxidizing agents, halogens, H2O2, and nitric acid can generate sulfonic acids from thiols (20). Introducing highly oxidized sulfur species can result in protein structural changes or inhibit enzyme activity that requires thiolate for catalysis. Alternatively, these reported cysteine oxidation products also serve as a prerequisite for proper protein function. Thus, the irreversible oxidation of cysteine to sulfinic and sulfonic acid can influence cellular homeostasis and protein functions in multiple ways. Oxidation of cysteine to sulfenic and sulfinic acid modifications can be reversed by S-glutathionylation, wherein glutaredoxin mediates sulfinic acid reduction, conjugation of sulfenic acid via S-glutathionylation, and deglutathionylation by glutaredoxin or sulfiredoxin. As noted by the N-end rule pathway, irreversible cysteine oxidation can also target a protein for degradation; for example, oxidation of N-terminal cysteine residues to sulfinic and sulfonic acid in specific mammalian proteins, such as GTPase-activating proteins (RGS) is required for arginylation by ATE1 R-transferases and subsequent ubiquitin-dependent degradation. Thus, the overoxidation of cysteine to sulfonic acid cannot be reversed, and the damaged proteins have to be degraded by the proteasome (45). Apart from cysteine, the other sulfur-containing amino acids that undergo oxidative modification include methionine, which is reduced to methionine sulfoxide by methionine sulfoxide reductases. Methionine sulfone formation resulting from further oxidation events is considered a stable modification.
Concerning aromatic amino acids, tyrosine remains the primary target of protein oxidation events due to its redox-active structure. Its phenolic side chain gets oxidized easily, forming an intermediary tyrosyl radical. Upon reaction with HO•, these radicals form 3-hydroxylysine, a neurotransmitter analogue 3,4 dihydroxyphenylalanine (DOPA) and, in interaction with another tyrosyl radical, forms a fluorescent protein crosslink dityrosine. Hydroxyl radicals, upon interaction with tryptophan and histidine, form hydroxytryptophan and 2-oxohistidine, respectively (46).
2.4 산화적 단백질 변형의 메커니즘
2.4.1 유황 함유 아미노산과 방향족 아미노산의 산화
유황 함유 생체 분자는
단백질 폴딩,
반응성 종의 비활성화, 효소,
산화 환원 신호 전달 및 기타 생화학 기능에 매우 중요합니다.
놀랍게도, 대부분의 기능은
단백질과 단백질 부가물과 관련이 있는 반면,
그 기능은 시스테인과 메티오닌이라는
두 가지 아미노산과 각각의 티올 또는 티오에테르 기능성(40, 41)으로 거슬러 올라갈 수 있습니다.
산화 스트레스는
시스테인의 설프히드릴기와 메티오닌 티오에테르기를 표적으로 하여
번역 후 변형 사건을 증가시킵니다.
산화 환원 변환에 민감한 시스테인의 측쇄인 티올은
다른 산화 상태를 얻습니다.
티올과 디설파이드가 일반적으로 알려져 있지만,
단백질 변형에 대한 증거가 늘어나면서
설펜산, 설핀산, 설폰산 유도체를 포함한 다른 산소 유도체도 보고되고 있습니다.
디설파이드 형성은
가장 일반적인 티올 산화 반응으로,
디설파이드 결합이 비교적 안정적이며
분자 내 및 분자 간 디설파이드 다리를 통해 단백질 구조를 안정화시킵니다.
시스테인 티올 라디칼과 설펜산 형성은 가역적이며,
두 중간체 모두 매우 불안정합니다.
둘 다 여러 가지 산화된 시스테인 변형의 전구체로 작용합니다(42).
설펜산은
반응성이 매우 높고,
효소 촉매 작용과 세포 신호 전달에 중요한 역할을 합니다.
설펜산은
다른 산화 상태의 주요 중간체, 즉
설핀산과 설폰산으로 남아 있습니다.
염증 과정에서 면역 세포, 티올 또는
티올산 음이온이 차아염소산(HOCl)과 반응하여
설펜산(43)이 형성됩니다.
세포 단백질의 약 5%가 설핀산 또는 설폰산 형태로 남아 있습니다.
설핀산 변형의 기능적 역할은
주로 과산화물 환원효소(Prxs) 계열(44)에서 보고되었으며,
이 효소는 H2O2와 알킬 과산화물을 물과 알코올로 환원시킵니다.
생리학적 조건 하에서 설펜산과는 달리,
설핀산 유도체는 티올과 반응하지 않거나
자가 축합 반응을 일으키지 않습니다.
설핀산은 안정된 중간체이지만,
티올과 이황화물의 가장 높은 산화 상태인 설폰산(RSO3H)으로 쉽게 산화됩니다.
강력한 산화제, 할로겐, H2O2, 질산은
티올(thiol)로부터 술폰산을 생성할 수 있습니다(20).
고도로 산화된 황 종을 도입하면
단백질 구조가 변화하거나 촉매 작용을 위해 티올레이트가 필요한 효소 활성이 억제될 수 있습니다.
또는, 이러한 보고된 시스테인 산화 생성물은
적절한 단백질 기능을 위한 전제 조건으로도 작용합니다.
따라서 시스테인의 설핀산과 설폰산으로의 비가역적 산화는 세포 항상성과 단백질 기능에 다양한 방식으로 영향을 미칠 수 있습니다. 시스테인의 설핀산과 설핀산으로의 산화는 S-글루타티온화에 의해 역전될 수 있는데, 이 과정에서 글루타레독신이 설핀산의 환원, S-글루타티온화를 통한 설핀산의 접합, 글루타레독신 또는 설피레독신에 의한 탈글루타티온화를 매개합니다. N 말단 규칙 경로에서 언급한 바와 같이, 돌이킬 수 없는 시스테인 산화는 단백질 분해를 위한 표적으로도 작용할 수 있습니다. 예를 들어, GTPase 활성화 단백질(RGS)과 같은 특정 포유류 단백질에서 N 말단 시스테인 잔기의 설핀산과 설폰산으로의 산화는 ATE1 R-전이효소에 의한 아르기닐화 및 이후의 유비퀴틴 의존적 분해를 위해 필요합니다. 따라서, 시스테인의 설폰산으로의 과산화는 되돌릴 수 없으며, 손상된 단백질은 프로테아좀에 의해 분해되어야 합니다(45). 시스테인 외에도, 산화 변형을 겪는 다른 황 함유 아미노산으로는 메티오닌이 있는데, 메티오닌은 메티오닌 설폭시드 환원효소에 의해 메티오닌 설폭시드로 환원됩니다. 추가적인 산화 작용으로 인해 메티오닌 설폰이 형성되는 것은 안정적인 변형으로 간주됩니다.
방향족 아미노산과 관련하여, 티로신은 산화 환원 활성 구조로 인해 단백질 산화 사건의 주요 표적이 되고 있습니다. 티로신의 페놀 측쇄는 쉽게 산화되어 중간 티로실 라디칼을 형성합니다. HO•와 반응하면, 이 라디칼은 3-하이드록시라이신, 신경전달물질 유사체 3,4-디하이드록시페닐알라닌(DOPA)을 형성하고, 다른 티로실 라디칼과 상호작용하면 형광 단백질 교차결합 디티로신을 형성합니다. 하이드록실 라디칼은 트립토판과 히스티딘과 상호작용하면 각각 하이드록시트립토판과 2-옥소히스티딘을 형성합니다(46).
2.4.2 Glycoxidation
Reported as a spontaneous non-enzymatic reaction, glycation involves the response of free-reducing sugars with lysine and arginine amino acid residues, DNA and lipids forming Amadori products. This product, in turn, undergoes irreversible rearrangement and dehydration reaction, leading to the formation of advanced glycation end products (AGEs) (47). Introduced by Louis-Camille Maillard in 1912, glycation results in loss of protein function and impaired tissue elasticity in the skin, blood vessels and tendons. Glycation reactions are reported to be enhanced during oxidative stress and hyperglycemia conditions, thus playing a pivotal role in the pathogenesis of diabetic complications and aging. The formation of AGEs does not entirely rely on oxidative conditions; more specifically, only selected AGEs are generated by oxidation. Formed by a combination of glycation and oxidation, a subset population of AGEs is termed glycoxidation products. Excessive generation of ROS from glucose autoxidation and covalent attachment of glucose molecules to circulating proteins results in the formation of AGEs (48). They serve as biomarkers for both oxidative and carbonyl stress. Carboxymethyl lysine (CML), reported as the most abundant AGEs in vivo, is formed by oxidative degradation of Amadori product fructoselysine. An alternative non-oxidative mechanism involves the reaction of α-dicarbonyl compound glyoxal and lysine, leading to CML formation via an isomerization mechanism. The oxidative degradation of carbohydrates, lipids, nucleotides, and serine mainly forms the precursor, glyoxal. Further oxidation reaction of this glyoxal results in the formation of α-oxoamide AGE glyoxylyl lysine. As it relies on oxidative processes, glyoxylyl lysine is considered an even more sensitive marker than CML. Apart from oxidatively formed glycoxidation products, some AGEs are formed by precursors generated by oxidation, for example, glucose oxidation or Amadori products to glucosone. Lysine-mediated cleavage of glucosone results in formyl lysine formation, and reports from recent studies confirm high levels of formaldehyde metabolism products (formyl lysine, formyl phosphate) in murine tissues (49).
2.4.2 당화작용
자연적으로 일어나는 비효소 반응으로 보고된 당화작용은
유리환원당과 리신 및 아르기닌 아미노산 잔기,
DNA 및 지질이 아마다리 생성물을 형성하는 반응을 수반합니다.
이 생성물은
차례로 돌이킬 수 없는 재배열과
탈수 반응을 거쳐
최종당화생성물(AGE)을 형성합니다(47).
1912년 루이 카밀 마야르(Louis-Camille Maillard)가 소개한 당화(glycation)는 단백질 기능의 상실과 피부, 혈관, 힘줄의 조직 탄력성 저하를 초래합니다. 당화 반응은 산화 스트레스와 고혈당 상태에 의해 강화되는 것으로 알려져 있으며, 따라서 당뇨 합병증과 노화의 병인에 중요한 역할을 합니다. AGE의 형성은 산화 조건에만 의존하지 않습니다. 좀 더 구체적으로 말하면, 선택된 AGE만이 산화에 의해 생성됩니다. 당화 작용과 산화의 결합에 의해 형성된 AGE의 하위 집합은 당산화 산물이라고 불립니다. 포도당의 자가 산화와 순환 단백질에 대한 포도당 분자의 공유 결합으로 인한 과도한 ROS 생성은 AGE의 형성을 초래합니다(48). 이들은 산화 스트레스와 카르보닐 스트레스의 바이오마커 역할을 합니다. 생체 내에서 가장 풍부한 AGE로 보고된 카르복시메틸라이신(CML)은 아마도리 생성물인 프럭토셀리신의 산화 분해에 의해 형성됩니다. 산화 반응이 아닌 다른 메커니즘으로는 α-디카르보닐 화합물인 글리옥살과 라이신의 반응이 있으며, 이성질체화 메커니즘을 통해 CML이 형성됩니다. 탄수화물, 지질, 뉴클레오티드, 세린의 산화 분해는 주로 글리옥살이라는 전구체를 형성합니다. 이 글리옥살의 추가적인 산화 반응은 α-옥소아미드 AGE 글리옥실라이신(glyoxylyl lysine)의 형성을 초래합니다. 산화 과정에 의존하기 때문에, 글리옥실라이신은 CML보다 훨씬 더 민감한 지표로 간주됩니다. 산화적으로 형성된 당산화물 외에도, 일부 AGE는 포도당 산화 또는 글루코손으로의 아마도리 생성물과 같은 산화에 의해 생성된 전구체에 의해 형성됩니다. 글루코손의 리신 매개 절단은 포르밀 리신의 형성을 초래하며, 최근 연구 보고서에 따르면 쥐 조직에서 높은 수준의 포름알데히드 대사 산물(포르밀 리신, 포르밀 포스페이트)이 확인되었습니다(49).
2.4.3 Lipid peroxidation
An increase in levels of ROS can impose direct damage to lipids. The most prevalent ROS reported to affect lipids profoundly include HO• and hydroperoxyl (HO2•). In biological systems, the Fenton reaction forms HO• through redox cycling. Hydroperoxyl radicals play an essential role in lipid peroxidation, wherein this protonated form of O2•- produces H2O2, which can react with redox-active metals, yielding HO•. Hydroperoxyl radicals (HO2•) are reported to be much stronger than O2•- and could initiate the oxidation of polyunsaturated phospholipids, thereby impairing membrane function (28). The lipid peroxidation process involves hydrogen abstraction from carbon with oxygen insertion, resulting in lipid peroxyl radicals and hydroperoxides (LOOH). More specifically, free oxygen radicals target lipids containing carbon-carbon double bond(s), especially polyunsaturated fatty acids (PUFAs). Under physiological or subtoxic lipid peroxidation rates, cells survive by upregulation of antioxidant pathways and proteins, whereas, at toxic concentration levels, cells induce apoptosis, eventually leading to cellular damage (32). Thus, lipid peroxidation events might facilitate disease progression and aging. During the process of lipid oxidation, several reactive carbonyl species (RCS) are formed, which include LOOH and different aldehydes formed as secondary products, such as malondialdehyde (MDA), propanal, hexanal, and 4-hydroxynonenal (4-HNE) (4).
Malondialdehyde is reported to be the most abundant secondary aldehyde generated by the decomposition of arachidonic acid and larger PUFAs. The reactivity of MDA is pH-dependent; thus, at physiological pH, it exists as an enolate ion with low reactivity. Upon a pH decrease, MDA enolizes to β-hydroxy acrolein with increased reactivity. Malondialdehyde initial reaction with proteins generates Schiff-base adducts referred to as advanced lipid peroxidation end-products (ALEs). Under oxidative stress conditions, acetaldehyde in the presence of MDA generates highly immunogenic malondialdehyde acetaldehyde (MAA) adducts. These adducts are of biological importance as they can alter the functional properties of biomolecules, resulting in disease progression. Protein Kinase C (PKC) plays a vital role in the intracellular signal transduction process, which involves cell proliferation and differentiation, inflammation and cytoskeletal organization. The binding of MAA adducts induces activation of PKC-α, a specific isoform in hepatic cells, resulting in increased secretion of urokinase-type plasminogen activator, causative of hepatic fibrosis (44). Another important example of α/β unsaturated RCS is 4-hydroxy-2- nonenal (4-HNE). They are reported to be highly reactive, wherein nucleophilic attack of cysteine and histidine forms stable Michael adducts (45). 4-HNE protein adducts can contribute to protein crosslinking and induce carbonyl stress. For example, 4-HNE is reported to modify membrane-associated protein, G-protein signaling 4 (RGS4) at cysteine residue during oxidative stress, thereby altering signaling events in stressed cells (46). However, based on their cellular level and pathways involved in lipid peroxidation products, MDA and 4-HNE pose a dual behaviour of either enhancing cell survival or promoting cell death.
2.4.3 지질 과산화
ROS 수치가 증가하면 지질에 직접적인 손상을 입힐 수 있습니다. 지질에 심각한 영향을 미치는 것으로 알려진 가장 일반적인 ROS에는 HO•와 하이드로퍼옥실(HO2•)이 있습니다. 생물학적 시스템에서 펜톤 반응은 산화 환원 순환을 통해 HO•를 형성합니다. 하이드로퍼옥실 라디칼은 지질 과산화에서 중요한 역할을 합니다. 이 양성자화된 형태의 O2-는 H2O2를 생성하고, 이 H2O2는 산화 환원 활성 금속과 반응하여 HO-를 생성합니다. 하이드로퍼옥실 라디칼(HO2-)은 O2-보다 훨씬 더 강력하며, 다중 불포화 인지질의 산화를 유발하여 막 기능을 손상시킬 수 있다고 보고되었습니다(28). 지질 과산화 과정은 탄소로부터 수소를 빼내고 산소를 삽입하는 과정을 포함하며, 그 결과 지질 과산화 라디칼과 과산화수소(LOOH)가 생성됩니다. 좀 더 구체적으로 말하자면, 자유 산소 라디칼은 탄소-탄소 이중 결합을 포함하는 지질, 특히 다중 불포화 지방산(PUFA)을 표적으로 합니다. 생리적 또는 아독성 지질 과산화 속도의 경우, 세포는 항산화 경로와 단백질의 상향 조절을 통해 생존하지만, 독성 농도 수준에서는 세포가 세포 사멸을 유도하여 결국 세포 손상을 초래합니다(32). 따라서, 지질 과산화 현상은 질병의 진행과 노화를 촉진할 수 있습니다. 지질 산화 과정에서 여러 반응성 카르보닐기(RCS)가 형성되는데, 여기에는 LOOH와 말론디알데히드(MDA), 프로판알데히드, 헥산알데히드, 4-하이드록시노네날(4-HNE)과 같은 2차 생성물로 형성되는 다양한 알데히드가 포함됩니다(4).
말론디알데히드는 아라키돈산과 더 큰 고도불포화지방산의 분해에 의해 생성되는 가장 풍부한 2차 알데히드로 알려져 있습니다. MDA의 반응성은 pH에 따라 달라집니다. 따라서 생리학적 pH에서는 반응성이 낮은 에놀레이트 이온으로 존재합니다. pH가 감소하면, MDA는 반응성이 증가하면서 β-하이드록시 아크롤레인으로 에놀레이트화됩니다. 단백질과 말론디알데히드의 초기 반응은 고급 지질 과산화 최종 산물(ALEs)이라고 불리는 쉬프-베이스 부가물을 생성합니다. 산화 스트레스 조건에서, MDA가 존재하는 아세트알데히드는 면역원성이 높은 말론디알데히드 아세트알데히드(MAA) 부가물을 생성합니다. 이러한 부가물은 생체 분자의 기능적 특성을 변화시켜 질병의 진행을 초래할 수 있기 때문에 생물학적으로 중요합니다. 단백질 키나아제 C(PKC)는 세포 증식과 분화, 염증, 세포 골격 조직과 관련된 세포 내 신호 전달 과정에서 중요한 역할을 합니다. MAA 부가물의 결합은 간세포의 특정 이소폼인 PKC-α의 활성화를 유도하여 간 섬유증의 원인인 우로키나제형 플라스미노겐 활성화제의 분비를 증가시킵니다(44). α/β 불포화 RCS의 또 다른 중요한 예는 4-하이드록시-2-노네날(4-HNE)입니다. 이 물질은 반응성이 매우 높은 것으로 알려져 있으며, 시스테인과 히스티딘의 친핵성 공격으로 안정적인 마이클 부가물이 형성됩니다(45). 4-HNE 단백질 부가물은 단백질 교차결합에 기여하고 카르보닐 응력을 유발할 수 있습니다. 예를 들어, 4-HNE는 산화 스트레스 동안 시스테인 잔기에서 막 관련 단백질인 G-단백질 신호 전달 4(RGS4)를 변형함으로써 스트레스 받은 세포의 신호 전달 과정을 변화시킨다고 보고되었습니다(46). 그러나, 지질 과산화 생성물과 관련된 세포 수준과 경로를 기반으로 볼 때, MDA와 4-HNE는 세포 생존을 강화하거나 세포 사멸을 촉진하는 이중적인 행동을 보입니다.
2.4.4 Protein carbonylation
ROS-mediated protein carbonylation events are characterized as the most common type of non-enzymatic post-translational modification (PTM). This stable modification is achieved by either direct oxidation of protein-bound amino acids, oxidative cleavage of the protein backbone and incorporation of carbonyls from glycoxidation or lipoxidation. ROS/reactive intermediates such as H2O2 and lipid hydroperoxides interact with specific amino acids, arginine, lysine, proline or threonine, causing protein-protein cross-linkages, resulting in protein denaturation and loss of activity. Various oxidation products have been reported so far, which include tryptophan forms kynurenine, nitrotryptophan; Phenylalanine forms 2,3- 2-, 3-, and 4-hydroxyphenylalanine, Dihydroxyphenylalanine; Histidine forms 2-Oxohistidine; Arginine and proline forms glutamic semialdehyde (50). Superoxide anion radical formation from O2 or HO• by the interaction of H2O2 with free iron (Fe2+) through the Fenton reaction results in the interaction of ROS with the amino acids mentioned above. Superoxide anion radical generated from O2 is converted by SOD to H2O2 and later into H2O by catalase, glutathione peroxidase or peroxiredoxin (47). The direct oxidation of amino acids, aminoadipic and glutamic semialdehyde contribute to approximately 60% of total protein carbonylation in the liver. Hydroxyl radical-mediated abstraction of hydrogen located next to the N6-amino function of lysine, metal-catalyzed oxidation of the carbon-centered radical, and hydrolysis of the resulting imine mediates aminoadipic semialdehyde formation. Concerning oxidative cleavage of the protein backbone, O2•- facilitates RO• formation at α-carbon next to a peptide bond. The RO• fragments either through the diamide pathway (homolytic cleavage of carbon-carbon bond) or the α-amidation pathway (carbon-nitrogen bond) (50). Protein carbonylation remains a valuable biomarker in aging and diseases, wherein they are shown to impair protein structure and function. In a carbonylated protein profiling study from lean and obese individuals with or without type 2 diabetes (T2D), 36 out of 158 unique carbonylated proteins were reported to be present only in obese patients with T2D. These identified proteins were found to play a vital role in intracellular signaling and angiogenesis, cell adhesion and cytoskeletal remodeling (51).
Highly oxidized proteins appear to be relatively poor substrates for degradation by ubiquitination. Thus, dysfunctional carbonylated proteins accumulate as covalently crossed protein aggregates, making them highly resistant to proteolysis, thereby affecting the functional integrity of cells during the aging and disease processes. On the contrary, proteins that have undergone mild oxidation are highly susceptible to proteasomal degradation due to exposure to hydrophobic amino acids by unfolding targeted protein domains. Hydrophobic surface patches remain the central motif recognized by the proteasome. Remarkably, 19S and 20S proteasome subunits are highly susceptible to carbonylation and HNE modification, suppressing their proteolytic activities (52). Protein carbonylation may also be beneficial by regulating and activating signaling pathways involved in antioxidant defense and cellular homeostasis. Carbonylation depends on the cellular redox environment, ROS abundance and its proximity to the proteins.
2.4.4 단백질 카르보닐화
ROS 매개 단백질 카르보닐화 사건은 가장 일반적인 비효소적 번역 후 변형(PTM)의 유형으로 특징지어집니다. 이 안정적인 변형은 단백질에 결합된 아미노산의 직접적인 산화, 단백질 골격의 산화 분해, 당산화 또는 지질산화로부터의 카르보닐의 통합에 의해 달성됩니다. ROS/반응성 중간체(예: H2O2 및 지질 과산화수소)는 특정 아미노산(아르기닌, 리신, 프롤린 또는 트레오닌)과 상호 작용하여 단백질-단백질 교차 결합을 일으켜 단백질 변성 및 활성의 상실을 초래합니다. 지금까지 트립토판이 키누레닌, 니트로트립토판으로 변하는 것, 페닐알라닌이 2,3-2-, 3-, 4-하이드록시페닐알라닌으로 변하는 것, 디하이드록시페닐알라닌으로 변하는 것, 히스티딘이 2-옥소히스티딘으로 변하는 것, 아르기닌과 프롤린이 글루타민산 반알데히드로 변하는 것(50) 등 다양한 산화 산물이 보고되었습니다. 펜톤 반응에 의해 H2O2와 자유 철(Fe2+)의 상호작용으로 O2 또는 HO•로부터 슈퍼옥사이드 음이온 라디칼이 생성되면, 앞서 언급한 아미노산과 ROS의 상호작용이 일어납니다. O2로부터 생성된 슈퍼옥사이드 음이온 라디칼은 SOD에 의해 H2O2로 전환되고, 이후 카탈라제, 글루타티온 퍼옥시다아제 또는 퍼옥시레독신(47)에 의해 H2O로 전환됩니다. 아미노산, 아미노아디프산, 글루탐산 반알데히드의 직접적인 산화는 간에서 전체 단백질 카르보닐화의 약 60%를 차지합니다. 리신의 N6-아미노 기능 옆에 위치한 수소의 수산기 라디칼 매개 추출, 탄소 중심 라디칼의 금속 촉매 산화, 그리고 그 결과물인 이미인의 가수분해는 아미노아디프산 반알데히드의 형성을 매개합니다. 단백질 골격의 산화적 절단에 관해서, O2•-는 펩티드 결합 옆의 α-탄소에서 RO• 형성을 촉진합니다. RO• 단편은 디아미드 경로(탄소-탄소 결합의 동해리 절단) 또는 α-아미드화 경로(탄소-질소 결합)를 통해 생성됩니다(50). 단백질 카르보닐화는 노화와 질병의 중요한 바이오마커로 남아 있으며, 단백질 구조와 기능을 손상시키는 것으로 나타났습니다. 제2형 당뇨병(T2D)이 있거나 없는 마른 체형과 비만인 개인을 대상으로 한 카르보닐화 단백질 프로파일링 연구에서, 158개의 고유 카르보닐화 단백질 중 36개가 T2D가 있는 비만인 환자에서만 존재하는 것으로 보고되었습니다. 이렇게 확인된 단백질은 세포 내 신호 전달과 혈관 신생, 세포 부착 및 세포 골격 재형성(51)에 중요한 역할을 하는 것으로 밝혀졌습니다.
고도로 산화된 단백질은 유비퀴틴화에 의한 분해에 상대적으로 취약한 기질로 보입니다. 따라서, 기능 장애가 있는 카르보닐화 단백질은 공유결합으로 교차된 단백질 응집체로 축적되어 단백질 분해에 대한 저항성이 높아져 노화 및 질병 과정에서 세포의 기능적 완전성에 영향을 미칩니다. 반대로, 약한 산화를 겪은 단백질은 표적 단백질 도메인의 전개에 의해 소수성 아미노산에 노출되어 프로테아좀 분해에 매우 취약합니다. 소수성 표면 패치는 프로테아좀이 인식하는 중심 모티프로 남아 있습니다. 놀랍게도, 19S 및 20S 프로테아좀 서브유닛은 카르보닐화 및 HNE 변형에 매우 민감하여, 그들의 단백질 분해 활성을 억제합니다(52). 단백질 카르보닐화는 항산화 방어 및 세포 항상성에 관여하는 신호 전달 경로를 조절하고 활성화함으로써 유익할 수도 있습니다. 카르보닐화는 세포의 산화 환원 환경, ROS의 풍부함, 그리고 단백질과의 근접성에 따라 달라집니다.
2.4.5 Nitrosylation
Nitric oxide (NO), also referred to as Nitric oxide radical (•NO), remains an important signaling molecule that exhibits pleiotropic functions like vasodilation, neurotransmission and pro-inflammatory signaling. Nitric oxide synthase (NOS) utilizes L-arginine as a substrate and, along with oxygen, produces citrulline and NO. They are reported to mediate both anti- and pro-oxidant mechanisms. In immune cells, nitric oxide limits ROS production via NADPH oxidase and pretreatment of cells with NO protects them against oxidative stress. Recent clinical trials support the beneficial effects of NO pretreatment in ischemia–reperfusion-mediated tissue injury (53). On the contrary, NO readily reacts with O2•- forming RNS, peroxynitrite (ONOO−) and nitrogen dioxide (NO2). In vivo, these RNS are reported to be potent oxidizing agents that can directly or reversibly modify cysteine residues through S-nitrosylation. Nitrogen dioxide radicals interact at the ortho-position of the tyrosine aromatic ring, resulting in the formation of irreversible modification of 3-nitrotyrosine. Both peroxynitrite and protein tyrosine nitration is reportedly involved in aging. Protein tyrosine nitration serves as a potential biomarker of disease progression. While initial experiments demonstrated the involvement of endogenous peroxynitrite and protein tyrosine nitration in apoptosis of motoneurons in culture, further work in ALS animal models confirmed the formation of 3-nitrotyrosine and protein-derived radicals in spinal cord motor neurons during disease progression (54, 55). Various protein oxidation events used as biomarkers in clinical settings are presented in Table 1.
2.4.5 니트로실화
산화질소 라디칼(•NO)이라고도 불리는 산화질소(NO)는 혈관 확장, 신경 전달, 전염증성 신호 전달과 같은 다발성 기능을 나타내는 중요한 신호 전달 분자입니다. 산화질소 합성효소(NOS)는 L-아르기닌을 기질로 활용하여 산소와 함께 시트룰린과 NO를 생성합니다. 이 물질은 항산화 및 산화 메커니즘을 모두 매개하는 것으로 알려져 있습니다. 면역 세포에서 산화질소는 NADPH 산화효소를 통해 ROS 생성을 제한하고, NO로 세포를 전처리하면 산화 스트레스로부터 세포를 보호합니다. 최근 임상 시험에서 허혈-재관류 매개 조직 손상(53)에 대한 NO 전처리의 유익한 효과가 입증되었습니다. 반대로, NO는 O2•-와 쉽게 반응하여 RNS, 퍼옥시니트라이트(ONOO−) 및 이산화질소(NO2)를 형성합니다. 생체 내에서, 이러한 RNS는 S-니트로실화를 통해 시스테인 잔기를 직접 또는 가역적으로 변형시킬 수 있는 강력한 산화제로 보고되고 있습니다. 이산화질소 라디칼은 티로신 방향족 고리의 오르토 위치에서 상호 작용하여 3-니트로티로신의 비가역적 변형을 일으킵니다. 퍼옥시니트라이트와 단백질 티로신 니트로화는 노화와 관련이 있는 것으로 보고되고 있습니다. 단백질 티로신 니트로화는 질병 진행의 잠재적 바이오마커 역할을 합니다. 초기 실험에서 배양된 운동신경세포의 세포사멸에 내인성 퍼옥시니트라이트와 단백질 티로신 니트로화가 관여한다는 사실이 입증되었지만, ALS 동물 모델에 대한 추가 연구에서 질병 진행 과정에서 척수 운동신경세포에 3-니트로티로신과 단백질 유래 라디칼이 형성된다는 사실이 확인되었습니다(54, 55). 임상 환경에서 바이오마커로 사용되는 다양한 단백질 산화 사건이 표 1에 나와 있습니다.
Table 1
Table 1 Oxidative modification of proteins as biomarkers in clinical studies.
3 Redox factors involved in immune cell activation
3.1 Redox factors in macrophage activation and function
Macrophages are a heterogeneous population of immune cells that play a vital role in tissue homeostasis in response to pathogen infection by phagocytosis and mediate tissue repair during injury. They rapidly recognize, engulf and destroy pathogens or apoptotic cells, which can be attributed to their plasticity and heterogeneity (70). Through polarization events, macrophages adopt either a pro-inflammatory phenotype classified as M1 macrophage or an anti-inflammatory M2 phenotype that mediates wound healing and inflammation resolution (71). Recently, it has been observed that excess molecular stimuli induce diverse and partially overlapping macrophage phenotypes that are distinct from M1 and M2. Macrophage activation by ROS, cytokines and commensal lipopolysaccharide (LPS) results in the activation of NF-κB and PI3K/AKT signaling pathways (72, 73). Thus, upregulated NF-κB increases pro-inflammatory chemokines and cytokine transcription, inducible NO synthase (iNOS) and HIF1 α. These signaling pathways, enzymes, and transcription factors are essential in maintaining macrophage activation and M1 polarization by driving metabolic reprogramming (17).
ROS regulate the intracellular signalosome within a constantly evolving cellular microenvironment, thus underlying its role in polarization and the specialized function of immune cell populations (74). M1 and M2 macrophages differ from resting macrophages in their phenotype and exhibit distinct metabolic profiles (75). M1 macrophage metabolism is characterized by aerobic glycolysis, changes in pentose phosphate pathway (PPP), FAS, and truncated TCA cycle, while M2 mostly depends on oxidative metabolism, fatty acid oxidation (FAO) and decreased glycolysis (76) (Figure 2).
면역세포 활성화에 관여하는 3가지 산화 환원 인자
3.1 대식세포 활성화와 기능에 관여하는 산화 환원 인자
대식세포는 병원체 감염에 대한 식세포 작용을 통해 조직 항상성에 중요한 역할을 하고, 손상 시 조직 복구를 매개하는 이질적인 면역 세포 집단입니다. 이들은 병원체나 세포자살 세포를 신속하게 인식하고, 포획하고, 파괴하는데, 이는 그들의 가소성과 이질성(70)에 기인합니다. 대식세포는 분극화를 통해 M1 대식세포로 분류되는 전염증성 표현형 또는 상처 치유와 염증 해결을 매개하는 항염증성 M2 표현형을 채택합니다(71). 최근에, 과도한 분자 자극이 M1과 M2와는 다른 다양하고 부분적으로 중복되는 대식세포 표현형을 유도한다는 사실이 관찰되었습니다. ROS, 사이토카인, 공생성 지질다당류(LPS)에 의한 대식세포 활성화는 NF-κB 및 PI3K/AKT 신호 전달 경로의 활성화를 초래합니다(72, 73). 따라서, NF-κB가 활성화되면 전염성 화학물질과 사이토카인 전사, 유도성 NO 합성효소(iNOS), HIF1α가 증가합니다. 이러한 신호 전달 경로, 효소, 전사 인자는 대사 재프로그래밍을 유도함으로써 대식세포 활성화와 M1 분화를 유지하는 데 필수적입니다(17).
ROS는 끊임없이 진화하는 세포 미세 환경 내에서 세포 내 신호 전달 경로를 조절하여 분화와 면역 세포 집단의 특수 기능에 대한 역할을 뒷받침합니다(74). M1과 M2 대식세포는 휴식 상태의 대식세포와 표현형이 다르며 뚜렷한 대사 프로파일을 나타냅니다(75). M1 대식세포의 대사는 호기성 당분해, 5탄당 인산경로(PPP)의 변화, FAS, 단축된 TCA 순환이 특징인 반면, M2는 주로 산화 대사, 지방산 산화(FAO), 감소된 당분해에 의존합니다(76) (그림 2).
Figure 2
Figure 2 Redox regulation of macrophage polarization. Superoxide (O2•-) generated by NOX or mitochondrial electron transfer chain is converted into H2O2 by superoxide dismutase (SOD), which balance between both pro-inflammatory (M1) and anti-inflammatory (M2) response of macrophage during polarization events. The figure was Created with BioRender.com.
3.2 Metabolic reprogramming in M1 macrophages
Glycolysis is a crucial metabolic event for M1 macrophages, and its inhibition affects typical functions of their inflammatory phenotype, including phagocytosis, ROS production, and pro-inflammatory cytokine secretion. Lipopolysaccharides and Toll-like receptor (TLRs) mediated differentiation of M1 macrophages are associated with a metabolic shift towards glycolysis, blocking it by glucose derivate like 2-deoxyglucose impairs both ROS and pro-inflammatory cytokine production. Numerous transcription factors are known to be involved in maintaining metabolic changes associated with M1 macrophages, and notable factors were discussed in detail. Reports demonstrated the involvement of HIF1α in activating inflammatory macrophage through the glycolysis mechanism (77). They act as modulators of the methylation status of hypoxia-responsive elements in the promoter regions. Enhanced expression of HIF1α reduces mitochondrial activity by suppressing electron transport chain enzymes, resulting in mitochondrial autophagy. HIF1α exaggerates glycolytic flux, thereby increasing the expression of glucose transporters (GLUT1 and GLUT3) and inflammatory mediators. More specifically, M1 macrophages rely on glycolysis and accumulation of succinate from the TCA cycle to stabilize HIF1α, which in turn activates the transcription of glycolytic genes sustaining glycolytic metabolism in M1 macrophages (36). Increased levels of HIF1α were evident along the differentiation of monocytes into tissue macrophages and play a prominent role in the uptake of bacteria by macrophages under hypoxic conditions and expression of tumour necrosis factor (TNFα) and nitric oxide (NO) through inducible NO synthetase (iNOS). In macrophages, overexpression of glucose receptors enhances glycolysis, which induces ROS production and pro-inflammatory mediators (78). In M1 macrophages, aerobic glycolysis induction depends entirely on redox-sensitive transcription factor HIF1α activated by the NF-κB pathway during inflammation. HIF1α interacts with pyruvate kinase isoenzyme M2 (PKM2), thereby mediating the transcription of glycolytic enzymes and inflammatory factors like IL-1β. ROS-like NO-mediated prolyl hydroxylases (PHDs) inhibition induces HIF1α (79). Further, NO reduces oxidative phosphorylation by nitrosylation, inhibiting proteins involved in the mitochondrial electron transport chain. Thus, ROS production positively regulates and maintains the shift towards aerobic glycolysis in M1 macrophages.
NADPH generated by PPP regulates the inflammatory response of M1 macrophages. They mediate ROS and NO production through NOX and iNOS, respectively, to kill invading pathogens and sustain the functionality of TRX and GSH antioxidant systems. Results presented by Nguyen and co-workers demonstrate that deletion of the TRX1 system impairs NLRP3 inflammasome formation and binding of NF- κB to target DNA in monocytes and macrophages (10). It leads to defective production of pro-inflammatory cytokines and ROS accumulation. Additionally, mitochondrial ROS-induced DNA damage aid in a significant drop in NAD+ levels in M1 macrophages. Suppression of PPP in macrophages attenuates LPS-induced inflammatory and oxidative stress response. Compared to glycolysis, intermediates of the TCA cycle, namely succinate and citrate, support biosynthesis in M1 macrophages (11). In LPS-stimulated macrophages, two breaks in the TCA cycle result in the accumulation of succinate and citrate, stabilizing HIF-1α and the subsequent increase in IL-1β transcription. Furthermore, succinate dehydrogenase-mediated oxidation of succinate and increased mitochondrial membrane potential drive ROS production (72). Another TCA cycle metabolite, citrate, gets transported into the cytosol and utilized for fatty acid synthesis to support membrane biogenesis and synthesis of pro-inflammatory lipid mediators, namely prostaglandins. In LPS-activated macrophages, ROS-dependent oxidation of unsaturated phospholipids results in glutamine utilization to feed the TCA cycle and lead to cytoplasmic accumulation of oxaloacetate. These metabolites stabilize HIF-1α, enhancing IL-1β secretion in atherosclerosis (13).
In M1 macrophages, ROS-mediated activation of NRF2 is essential for PPP maintenance and NADPH production, which is necessary for FAS, TRX, and GSH systems. IL-1β is produced as an inactive precursor in response to pathogens, and thus, further processing of it into a biologically active form requires the formation of multiprotein complexes termed inflammasomes. Mitochondrial ROS and its oxidation products are known to play a prominent role in inflammasome activation, whereas the role of NOX is highly dispensable (14).
3.3 Metabolic reprogramming in M2 macrophages
Unlike M1 macrophages, M2 macrophages hold an intact TCA cycle and enhanced mitochondrial OXPHOS. CD36 internalizes circulating lipoproteins and fatty acids, mediating fatty acid uptake and fueling OXPHOS. The increased cellular concentration of IL-4 and IL-13 drives M1 macrophages towards anti-inflammatory and healing phenotypes described as M2 macrophages. Tyrosine phosphorylation and signal transducer and activator of transcription 6 (STAT6) activation mediate polarization of macrophages into the M2 phase. IL-4 and IL-13 suppress pro-inflammatory cytokine production by upregulating transforming growth factor beta (TGF-β) activity (70). Adenosine 5′-monophosphate-activated protein kinase (AMPK) and peroxisome proliferator-activated receptor (PPAR) are found to play a vital role in the transition of macrophage polarization states through IL-13 and IL-4. AMPK inhibits NF-κB and stimulates OXPHOS and FAO, reducing HIF1α levels and inflammation and terminating aerobic glycolysis (80). AMPK negatively regulates LPS-induced inflammatory response in macrophages by inhibiting NF-κB activity and activating the PI3K/Akt signaling pathway. Enhanced IL-10 expression promotes TAM activation by PPARα/β, resulting in a polarization of M2 macrophages (12). The importance of FAO in M2 polarization was highlighted in several studies wherein blocking FAO with inhibitors against mitochondrial carnitine palmitoyl-transferase 1 inhibited the activation of M2 macrophages. Results from various studies confirm the PPARγ-mediated activation of M2 signature genes either through oleic acid and IL-4 stimulation or by promoting glutamine oxidation fueling OXPHOS (81). Modification in arginine metabolism emphasizes the transition from M1 to M2 polarization. Increased activity of iNOS mediates arginine metabolism to produce NO, which maintains the switch towards aerobic glycolysis in M1 macrophages. In the case of M2 macrophages, arginine gets metabolized into ornithine and urea due to increased transcription of arginase-1. Both urea and ornithine are essential in M2 macrophage proliferation and survival. Additionally, glutamine metabolism is particularly interesting since glutamine oxidation depletes extracellular glucose levels in an inflammatory environment, maintaining TCA activity and activating the glutamine–UDP-N-acetylglucosamine (GlcNAc) pathway to reinforce M2 polarization (82).
3.4 Redox regulation of neutrophil activation
Neutrophils are the first responders against invading pathogens through innate and humoral immunity. Activated neutrophils rely on glycolysis as their primary source of energy under both physiological and inflammatory environments; however, they can regulate their metabolism to carry out its effector functions, namely phagocytosis, oxidative burst, degranulation, extracellular trap formation and chemotaxis (83). Neutrophil extracellular trap (NETs) formation by neutrophils relies on glycolysis and PPP as a source of NADPH, resulting in free oxygen radical production. Superoxide anion radical, thus produced, further induces the formation of ROS and HOCl used by neutrophils in oxidative bursts following phagocytosis of invading pathogens. ROS promote several steps of NETosis, including releasing neutrophil elastase from granules by increasing membrane permeability and degradation of H1 linker and core histones resulting in chromatin decondensation (84). The morphological changes associated with NETosis were promoted by ROS, which in turn inactivated caspases to block apoptosis and trigger autophagy. Secondary oxidants, namely HOCl, mediate PMA-induced NETosis, and it entirely depends on NOX activity (Figure 3). The absence of extracellular Cl-, a substrate for Myeloperoxidase (MPO) in vitro results in decreased NET production (85). On the contrary, calcium ionophores induced NETosis are NOX independent and rely on mtROS. NOX-independent Netosis depends on calcium which in turn activates peptidyl arginase deiminases resulting in cellular hypercitrullination (86).
Figure 3
Figure 3 Reactive oxygen species production within the neutrophil phagosome. Post stimulation, oxygen reduction by NOX in the presence of NADPH produces superoxide (O2•-) within the phagosome. H2O2 produced by either the enzymatic or spontaneous dismutation of Superoxide further stimulates neutrophil granules, releasing myeloperoxidase (MPO) within the phagosome. MPO catalyses the oxidation of halides (Cl-) by interaction with H2O2, forming HOCl. The figure was Created with BioRender.com.
Neutrophil functions are highly influenced by cellular redox status, which includes both ROS/RNS production and cellular antioxidant systems (87). Enhanced ROS production is reported to comprise phagocytosis, resulting in dysregulated oxidative burst events and NET production. ROS levels determine the sensing of pathogens by neutrophils and their subsequent activation of NLRP3 inflammasome and cytokine synthesis (88). Additionally, chronically upregulated ROS and cytokine production lowers neutrophil migration by internalizing CXCR2, a membrane chemokine receptor. Neutrophil functions, namely oxidative burst and NET formation, were sustained by the glutathione system (GSH). Its basal activity was reported to be lower in neutrophils compared to other myeloid cells. Prolonged neutrophil activity and excessive production of MPO due to chronic nitrooxidative stress and inflammation lead to the depletion in GSH levels (89). Thus, depleted GSH levels in neutrophils affect their chemotaxis, transmigration and cytoskeletal reorganization, resulting in early apoptosis and impaired degranulation. Redox factors involved in macrophage and neutrophil function are presented in Table 2.
Table 2
Table 2 Redox mechanism influencing macrophage and neutrophil functions.
3.5 Immune cells oxidants mediated epigenetic regulation
Multiple transcriptional and epigenetic modification factors are known to be involved in macrophage differentiation and its activation. Epigenetic changes within macrophages allow them to switch between cellular programs affecting phenotype plasticity. DNA demethylation is known to be involved in the process of monocyte-to-macrophage differentiation. DNA demethylation affects specific genes that regulate actin cytoskeleton and phagocytosis (19). ROS-mediated oxidation of amino acid residues in histone H3 results in chromatin relaxation and accumulation of transcription factors. Cysteine residues at histone H3 sense redox changes and mediate further opening of chromatin structures (95). In LPS-stimulated macrophages, lipid peroxidation products are reported to form lysine adducts with H2, H3 and H4, including H3K23 and H3K27. These modifications at histone acetylation and methylation sites are known to be associated with the epigenetic patterning of cardiovascular diseases (96). LPS stimulation and TLR-4-dependent activation of inflammatory genes primarily depend on H3 and H4 acetylation. ROS-mediated post-translational modification of both class I and II histone deacetylases impair its enzymatic function, resulting in open chromatin structure.
Additionally, lipid peroxidation products mediate the carbonylation of HDACs, resulting in ubiquitination and proteasomal degradation of HDAC function, increased acetylation of histones in macrophages and release of pro-inflammatory cytokines. HDACs are predominant in regulating immunological pathways, more precisely in M1 activation. The difference in the expression pattern of almost all HDAC classes was observed in cells stimulated with LPS. Studies concerning macrophage stimulation with LPS result in an initial decrease in HDAC 4,5 and 7 expression, thereby leading to cyclooxygenase-2 gene activation (97). On the contrary, HDAC6 aids in the expression of pro-inflammatory genes in macrophages stimulated with LPS and thus, its inhibition limits macrophage activation. Concerning the link between DNA methylation and LPS stimulation, SOCS1, a negative regulator of cytokine signals, has been found. DNMT1-mediated hypermethylation of SOCS1 results in the loss of its activity, thereby enhancing the expression of LPS-induced Pro-inflammatory cytokines, namely TNF-α and IL-6. DNMT1 has also been reported to improve the demethylation and trimethylation events of H3K9 in regulator proteins like Notch1 and KIF4 and mediate their polarization towards M1 macrophages. Furthermore, DNA methyltransferase (DNMTs) is known to be involved in M2 differentiation and phenotypic regulation. Individuals with atherosclerosis and apolipoprotein E knockout mice fed an atherogenic diet displayed enhanced DNMT levels in macrophages (98). Since the activity of PPAR-γ is reduced, the macrophage transition from M1 to M2 is affected, and thus, the progression of atherosclerosis is marked by increased pro-inflammatory cytokine production. Oxidative stress is linked to increased histone acetyltransferase (HATs) activity of p300/CBP along with NFĸB DNA binding, promoting pro-inflammatory gene expression. In endothelial cells and hyperglycemic adipocytes, enhanced expression of HAT GCN5 and H3 acetylation is associated with increased ROS production, as confirmed in diabetic models. p300/CBP mediated acetylation of H3K9 at NOX2 promoter encourages ROS generation underlying complexity of epigenetic modifications in ROS balance and response (16).
Upon activation, neutrophils generate a range of ROS and O2•- generated were reported to damage proteins and are limited to the compartment site it has been generated as its rate of dismutation was enhanced by SOD to H2O2. Neutrophil heme protein myeloperoxidase utilizes H2O2 to oxidize the halides chloride (Cl-), iodide (I-), and bromide (Br-), or pseudohalide anion thiocyanate (SCN-), into hypochlorous acid (HOCl) (99). Being a potent oxidant, HOCl poses a high reactivity towards biological macromolecules targeting free and protein-associated cysteine, methionine residues and low molecular weight thiols, and its ability to diffuse from the site of generation is very short. Thus, HOCl reacts with amines and forms chloramines, which are less reactive than HOCl and diffuse further from the generation site. The reaction of HOCl and chloramines with cytosine produces 5-chlorocytosine (5-clC), directly incorporated into DNA as a chlorinated nucleotide (100). Studies carried out reported gene silencing, and no significant changes were observed in global methylation levels due to the incorporation of 5-clC. Chloramines are also reported to interact with histone amine groups, thereby preventing methylation or acetylation events (101). However, the lack of in vivo experimental evidence makes it unclear how 5-clC and chloramine levels are regulated under physiological conditions. Neutrophil oxidants can react with cellular targets, including small molecules and redox-sensitive components of epigenetic pathways. Intracellular availability of methionine and ascorbate was depleted by neutrophil oxidants and reported to impact methylation by disrupting S-adenosylmethionine (SAM) levels (102). Mass spectrometry analysis to investigate the oxidant effects of methylation reported impaired cytosine methylation on newly replicated DNA in the Jurkat T-lymphoma cell line upon sub-lethal level exposure to glycine chloramines (103). The study reported DNMT1 inhibition and depletion of SAM levels at doses, which had minimal effect on cell proliferation. Though H2O2 treatment inhibited DNMT1, it did not reduce SAM or global methylation levels (104). Further experimentation is required to determine whether the methylation and demethylation effects observed are heritable to subsequent generations.
4 Summary and conclusion
A sustained pro-oxidant cellular environment mediates the development and progression of various pathological conditions due to redox imbalance. Dysregulation in these redox environments decreases the activity of mitochondria, TCA cycle and immune cell metabolism (30). ROS and RNS are ubiquitous byproducts of cellular metabolism, and any disparity between their generation and degradation in aging and diseases results in oxidative and nitrosative stress. Oxidative stress can irreversibly damage cellular structures, including membrane lipids or lipoproteins, forming oxidation-specific epitopes (OSE) on damaged cells (105). This damage-associated molecular pattern is recognized and removed by innate immune cells, including macrophages and neutrophils, enabling cellular homeostasis. Excessive accumulation of these oxidation products triggers chronic inflammation and metabolic disorders, including atherosclerosis, diabetes and age-related macular degeneration (106). Immune cells function and survival are regulated by various redox factors, including the intracellular and extracellular concentration of ROS/RNS and cellular antioxidants, namely glutathione, thioredoxin and Nrf-2 (107). Considering diabetes, metabolic imbalance in these conditions is characterized by increased glycolytic flux, and ROS act as a secondary messenger and mediates metabolic shift towards pro-inflammatory macrophage phenotype. ROS were also reported to activate multiple pro-inflammatory signaling pathways, including MAPK, NLRP3 and NFκB, resulting in an epigenetic modification in hyperglycemic conditions (108). A crosstalk between these immune cells and endothelial cells in diseased conditions is reported to stimulate increased ROS formation and inflammatory phenotypes further. Thus, consideration should be paved towards ROS generated by macrophages and neutrophils to suppress inflammation in metabolic disorders.
The functioning of individual immune cells is under redox control and reported to be sensitive to intracellular and extracellular concentrations of ROS and influenced by the activity of cellular antioxidants. Redox mechanisms regulate and modulate various immune functions, including metabolic reprogramming of dendritic cells (DCs), T cells, B cells, and natural killer cells (NK), aiding in its activation and regulation (109). ROS are reported to be involved in diverse biological events, including Epithelial–mesenchymal transition (EMT). This transdifferentiation process is vital in invasion and metastasis phenomena during neoplastic progression. ROS regulate the integrin arrangement and urokinase plasminogen activator (uPA) pathway in extracellular matrix remodelling (110). ROS can influence the function of various proteins involved in the EMT process through reversible or irreversible oxidative modification of protein on free cysteine residues (111). Thus, targeting redox regulation to prevent EMT and tumor metastasis is promising.
Mechanistic insight into the specific immune response generated for oxidation-specific epitopes at functional levels should be studied, which can aid in understanding oxidative stress and its associated chronic inflammations. Though M1 and M2 represent the two extreme phenotype characteristics of macrophage activation stages regulated by redox metabolism, tissue-resident macrophages comprise a distinct subset and hold tissue-specific functions dependent on oxygen and nutrient supply, which can certainly influence redox status and metabolism (112). Thus, studies focusing on the characterization of redox proteome during an immune response are essential. Methodology focusing on in vivo localization and visualization of ROS and their sources will aid in a better understanding of this complex redox metabolism in health and diseases.
Author contributions
RRM: Conceptualization, Writing – original draft, Writing – review & editing. AP: Conceptualization, Project administration, Supervision, Writing – original draft, Writing – review & editing. PP: Writing – review & editing. JK: Conceptualization, Writing – original draft, Writing – review & editing.
Funding
The author(s) declare that financial support was received for the research, authorship, and/or publication of this article. This work was funded by [1] the European Regional Development Fund project “Plants as a tool for sustainable global development” (CZ.02.1.01/0.0/0.0/16_019/0000827); [2] state contract of the Ministry of Science and Higher Education of the Russian Federation “Genetic and epigenetic editing of tumor cells and microenvironment in order to block metastasis” no. 075-15-2021-1073 (to JK); [3] Tomsk State University Development Programme (Priority-2030).
Conflict of interest
The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.
The author(s) declared that they were an editorial board member of Frontiers, at the time of submission. This had no impact on the peer review process and the final decision.
Publisher’s note
All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.
Abbreviations
3-NT, 3-Nitrotyrosine; 4-HNE, 4-Hydroxy-2-nonenal; AD, Alzheimer’s disease; PD, Parkinson disease; CML, Carboxymethyl lysine; AGEs, Advanced glycation endproduct; DNPH 2,4-Dinitrophenylhydrazine; RA, Rheumatoid arthritis; RCS, Reactive carbonyl species; RNS, Reactive nitrogen species; ROS, Reactive oxygen species; T2D, Type 2 diabetes; VCAM1 Vascular cell adhesion protein 1; ICAM1, Intercellular Adhesion Molecule 1.
References
1. Checa J, Aran JM. Reactive oxygen species: drivers of physiological and pathological processes. JIR. (2020) 13:1057–73. doi: 10.2147/JIR.S275595
CrossRef Full Text | Google Scholar
2. Phaniendra A, Jestadi DB, Periyasamy L. Free radicals: properties, sources, targets, and their implication in various diseases. Ind J Clin Biochem. (2015) 30:11–26. doi: 10.1007/s12291-014-0446-0
CrossRef Full Text | Google Scholar
3. Prasad A, Manoharan RR, Sedlářová M, Pospíšil P. Free radical-mediated protein radical formation in differentiating monocytes. IJMS. (2021) 22:9963. doi: 10.3390/ijms22189963
CrossRef Full Text | Google Scholar
4. Manoharan RR, Sedlářová M, Pospíšil P, Prasad A. Detection and characterization of free oxygen radicals induced protein adduct formation in differentiating macrophages. Biochim Biophys Acta (BBA) Gen Subj. (2023) 1867:130324. doi: 10.1016/j.bbagen.2023.130324
CrossRef Full Text | Google Scholar
5. Sharifi-Rad M, Anil Kumar NV, Zucca P, Varoni EM, Dini L, Panzarini E, et al. Lifestyle, oxidative stress, and antioxidants: back and forth in the pathophysiology of chronic diseases. Front Physiol. (2020) 11:694. doi: 10.3389/fphys.2020.00694
CrossRef Full Text | Google Scholar
6. Sies H, Jones DP. Reactive oxygen species (ROS) as pleiotropic physiological signaling agents. Nat Rev Mol Cell Biol. (2020) 21:363–83. doi: 10.1038/s41580-020-0230-3
CrossRef Full Text | Google Scholar
7. Lambeth JD. Nox enzymes, ROS, and chronic disease: An example of antagonistic pleiotropy. Free Radical Biol Med. (2007) 43:332–47. doi: 10.1016/j.freeradbiomed.2007.03.027
CrossRef Full Text | Google Scholar
8. Zorov DB, Juhaszova M, Sollott SJ. Mitochondrial reactive oxygen species (ROS) and ROS-induced ROS release. Physiol Rev. (2014) 94:909–50. doi: 10.1152/physrev.00026.2013
CrossRef Full Text | Google Scholar
9. Fransen M, Nordgren M, Wang B, Apanasets O. Role of peroxisomes in ROS/RNS-metabolism: Implications for human disease. Biochim Biophys Acta (BBA) Mol Basis Dis. (2012) 1822:1363–73. doi: 10.1016/j.bbadis.2011.12.001
CrossRef Full Text | Google Scholar
10. Nguyen GT, Green ER, Mecsas J. Neutrophils to the ROScue: mechanisms of NADPH oxidase activation and bacterial resistance. Front Cell Infect Microbiol. (2017) 7:373. doi: 10.3389/fcimb.2017.00373