Next Article in Journal
Targeted Protein-Specific Multi-Epitope-Based Vaccine Designing against Human Cytomegalovirus by Using Immunoinformatics Approaches
Previous Article in Journal
Vaccines and Vaccine Adjuvants for Infectious Diseases and Autoimmune Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

S Protein, ACE2 and Host Cell Proteases in SARS-CoV-2 Cell Entry and Infectivity; Is Soluble ACE2 a Two Blade Sword? A Narrative Review

1
Department of Anesthesiology and Critical Care Medicine, Laleh Hospital, Tehran 1467684595, Iran
2
Department of Chemical Engineering, Imperial College London, London SW7 2AZ, UK
3
Alliance Retina Consultants, La Mesa, CA 91942, USA
*
Author to whom correspondence should be addressed.
Vaccines 2023, 11(2), 204; https://doi.org/10.3390/vaccines11020204
Submission received: 1 December 2022 / Revised: 7 January 2023 / Accepted: 12 January 2023 / Published: 17 January 2023
(This article belongs to the Special Issue Neutralizing Antibody Response against SARS-CoV-2)

Abstract

:
Since the spread of the deadly virus SARS-CoV-2 in late 2019, researchers have restlessly sought to unravel how the virus enters the host cells. Some proteins on each side of the interaction between the virus and the host cells are involved as the major contributors to this process: (1) the nano-machine spike protein on behalf of the virus, (2) angiotensin converting enzyme II, the mono-carboxypeptidase and the key component of renin angiotensin system on behalf of the host cell, (3) some host proteases and proteins exploited by SARS-CoV-2. In this review, the complex process of SARS-CoV-2 entrance into the host cells with the contribution of the involved host proteins as well as the sequential conformational changes in the spike protein tending to increase the probability of complexification of the latter with angiotensin converting enzyme II, the receptor of the virus on the host cells, are discussed. Moreover, the release of the catalytic ectodomain of angiotensin converting enzyme II as its soluble form in the extracellular space and its positive or negative impact on the infectivity of the virus are considered.

1. Introduction

Deciphering the pattern of SARS-CoV-2 entry into the host cells, a crucial step in the virus infectivity, delineates a way to propose effective measures to interfere with the pathogenesis of COVID-19 successfully. Describing SARS-CoV-2 entry into cells by plainly describing the picking up of the virus, bound to its receptor, by the cells would be a rather simplification of a complicated process actually associated with the elaborate interactions among extra- and intra-cellular proteases, angiotensin converting enzyme II (ACE2) and SARS-CoV-2 spike (S) protein [1,2]. The docking of the S protein to its receptor, ACE2, a type I transmembrane mono-carboxypeptidase of renin angiotensin system (RAS), eventually leads to the fusion of the virus membrane with the host cell plasma membrane to release the viral genome into the cell cytoplasm [3]. Additionally, the fusion may occur at the membrane of endosomal compartment after clathrin-mediated endocytosis of the virus-ACE2 complex [4,5]. Moreover, there have been non-endosomal clathrin-independent models introduced for SARS-CoV-2 cell entry [4]. Although the fascinating ACE2-dependent entry has extensively been explored as the major model of SARS-CoV-2 entrance into the cells, other ACE2-independent models such as those mediated by Fc gamma receptors (FcγR) in monocytes have also been suggested [6,7,8,9]. It is also noticeable that the catalytic subunit of ACE2 released as a soluble molecule (sACE2) into the extracellular fluid with the ability to entrap the roaming virions in the circulation may negatively affect the infectivity of SARS-CoV-2 [10]. Paradoxically, sACE2 activity in the plasma was considered to have a positive correlation with the severity and mortality of COVID-19 [11]. In this context, the contribution of A Distegrin and Metalloproteinase 17 (ADAM17) to cleavage of ACE2 is worth investigating (see Section 4 and Section 9) [10].
Additionally, the priming effects of a constellation of proteolytic proteins in different models of SARS-CoV-2 cell entry should not be ignored [12], among those the transmembrane serine protease type II (TMPRSS2), cathepsins, furin and metalloproteases are of utmost importance [13,14,15,16]. To make it more complicated, the proteases themselves are secreted as inactive zymogens containing an inhibitor prodomain which covers the active site. The activation of proteases requires that this prodomain be dissociated in a meticulously regulated fashion [17]. The prodomain of TMPRSS2 is needed to be autocleaved [18] while in the case of cathepsins, cleavage of the inhibitory prodomain by distinct peptidases is pH-dependent [19]. In addition, furin requires pH- and Ca2+-dependent autoproteolytic removal of an 83-amino acid propeptide during its intracellular trafficking from the trans-Golgi network (TGN)/endosomal system to its destination on the cell membrane [20]. ADAM17, mostly residing in endoplasmic reticulum as an inactive metalloprotease, loses its prodomain through cleavage by a pro-protein convertase such as furin in Golgi apparatus [21].
In this review, the complex nature of SARS-CoV-2 cell entry with special attention to the role of proteinases and the updates of contribution of dysregulation of renin-angiotensin system (RAS) to the pathogenesis of COVID-19 at molecular and cellular level are discussed. Additionally, the paradox of the function of the soluble form of ACE2 and its relevant therapeutic implications are described.

2. S Protein Structure

S protein of SARS-CoV-2 (of 1273 amino acid residues), a class I fusion and homo-trimeric glycoprotein, is composed of two subunits, S1 and S2, in each of its monomers (Figure 1). The former is in charge of docking of the virion to ACE2 and the latter promotes fusion of viral and host cell lipid outer membranes through a fusogenic process [22]. From N- toward C-terminal ends, S1 comprises a signaling sequence, an N-terminal domain (NTD), a receptor binding domain (RBD) including its receptor binding motif (RBM) and two subdomains (SD1 and SD2) [23,24]. S2, a rather conserved molecule in SARS-CoV and SARS-CoV-2, consists of a fusion peptide (FP), the heptad repeat 1 (HR1), the heptad repeat 2 (HR2), a transmembrane domain (TM) and a cytoplasmic tail (CT) [22,25] (Figure 1). Absent in SARS-CoV, a polybasic amino acid sequence, RRAR (Arg-Arg-Ala-Arg), called the furin-cleavage site is located at the S1/S2 boundary, the deletion of which attenuates SARS-CoV-2 replication in the respiratory cells [26]. It has been shown that the presence of the RRAR motif enables SARS-CoV-2 to evade endosomal interferon-induced transmembrane (IFITM) proteins [27].

3. Binding of ACE2 with RBD Increases the Chance of Further Complexification of ACE2 and SARS-CoV-2

Binding of RBD with ACE2 is the first step of virus cell entry. Glycosylation of ACE2 and RBD residues may block this process [29]. However, RBDs in a trimeric S protein are not always accessible for this complexification and may be packed together [30]. Constitutively, the RBD position within the S1 structure shifts dynamically through an intrinsic hinge-like movement between “open or up (U)” and “closed or down (D)” states [23] (Figure 2). RBD in U status moves outward the apex of S1 domain to be ACE2-accessible while in D position it hides within the protein apex to be ACE2-inaccessible [31] (Figure 1 and Figure 2). As to the trimeric nature of S protein and considering U or D states of RBD, the apex of the trimeric S1 might harbor symmetric three open (UUU) or three closed (DDD) RBDs as well as intermediate asymmetric conformations of one open/two closed (UDD) or two open/one closed (UUD) ACE2-binding domains. The predominant RBD conformation (UUU, DDD or other intermediates) differs in the variants of the SARS-CoV-2 [32]. Recent studies attributed the escalated transmissibility and virulence of the new mutant variants to the abundancy and stability of RBD in the U state [32,33]. Moreover, according to cryo-electron microscopy findings, docking of the single open RBD to ACE2 induces a rotation of the same domain which provokes its center of mass within the UDD complex to move approximately 5.5 Å away from the axis of the whole trimeric S protein molecule. This rotation is accompanied by moving of the three NTDs in the trimer of about 1.5–3.0 Å [34] which results in a reduction in the contact area among S1 subunits and the neighboring S2 core. This causes successive gaining of U state in the neighboring two D-positioned RBDs within the trimeric S protein [34] (Figure 1). To make it more clear, it should be noticed that in ACE2-unbound form of closed RBD, there are π˗π interactions among some residues of S1 in one protomer and Tyr(Y)837 of S2 in the neighboring protomer in addition to a salt bridge between Asp(D)614 of S1 and Lys(K)854 of the S2 which are disordered after ACE2 binding [34]. In D614G (Asp614 to Glycine) mutation, the hydrogen bond between D614 residue of SD2 (of S1 subunit in one protomer) and Threonine(T)859 residue of S2 subunit in the neighboring protomer is deleted which induces S1 to move away from S-trimer axis [23,35]. These conformational changes bringing about successive expression of more U states results in escalating the probability of complexification, not the affinity, of RBD with ACE2 [35]. Clarifying this issue needs to consider that the binding affinity (according to its definition [36]) depends on the quantity and type of the mutations crowded in RBD which determine the density and dynamicity of polar interactions of RBM interfacial and non-interfacial amino acid residues with the counterpart residues of ACE2 [37,38].

4. Proteases Play a Pivotal Role in Various Models of SARS-CoV-2 Cell Entry

Various models of SARS-CoV-2 cell entry have been introduced [2,4,6] all of which end up with fusion of the virus membrane and the host cell plasma membrane [2]. Docking of RBD to ACE2 initiates an allosteric destabilization of the polybasic motif (RRAR) at S1/S2 interface. From this point on, proteolytic processing of S protein ultimately results in fusogenic configuration of S protein enabling virus membrane to intermingle with the host cell membrane [43,44,45]. An in vitro study revealed that serine proteases, furin and TMPRSS2, could enhance SARS-CoV-2 replication efficiency and cytopathology [46]. Furin (see Section 5) contributes to the processing of RRAR motif, the furin-cleavage site, at S1/S2 boundary of SARS-CoV-2 [47]. This cleavage has been correlated with syncytia formation of the infected lung epithelial cells [27]. Remaining non-covalently attached to S1 after the furin cleavage [48,49], S2 following formation of RBD-ACE2 complex is further proteolytically cleaved at S2′ segment located proximal to fusion peptide (FP) [34,50]. The cleavage at S2′ is mediated by the host cell plasma membrane serine protease (TMPRSS2, see Section 6) as well as other tissue proteases [51,52,53,54,55,56]. This promotes shedding of S1 subunit and exposure of FP [57]. Containing hydrophobic residues, FP is responsible for insertion of S protein into the host cell plasma membrane through gaining a wedge-shaped conformation in a Ca2+-dependent process [50,58]. Prior to fusion, sufficient energy is provided by a transition of the conformational shape of S protein from metastable prefusion into irreversible postfusion status [2]. This ultimately leads to a trimeric hairpin kneeling reconfiguration of HR1 and HR2 domains of S2 in order to bring the lipid membrane of the virus into the vicinity of the host cell plasma membrane [59,60,61].
Given that the presence of furin-cleavage site in S protein enhances TMPRSS2-mediated cell entry [27], different studies debate whether the cleavage at S1/S2 boundary is essential for SARS-CoV-2 penetration to the respiratory cells or cell-cell fusion as a means of direct spreading of the virus among neighboring cells [3,62]. A study on 293T cells infected with SARS-CoV-2, the presence of ACE2 but not TMPRSS2 was sufficient for syncytia formation albeit the destabilization of S1/S2 cleavage at the multibasic residues contributed to effective cell-cell fusion [63]. As a matter of fact, it was demonstrated that the presence of furin is not absolutely necessary for this proteolysis because other cellular proteases may also take over its function, yet furin enhances the process substantially [64,65]. Intriguingly, Tang et al., showed that S1/S2 cleavage is a prerequisite for TMPRSS cleavage of S2′ in case SARS-CoV-2 enters the host cell through the cell surface plasma membrane but not for the virus invasion via endosomal internalization [52]. For any route of the virus entry, the importance of the proteolytic destabilizing cleavage of S1/S2 junction (see Section 3, gaining U from D states in RBDs) and subsequent proteases-induced transition of S2 subunit to fusogenic configuration should be emphasized [66]. In silico and cell culture studies showed peptidomimetic binding compounds as well as lactoferricin and lactoferrin hinder the infectivity of SARS-CoV-2 significantly through inhibition of TMPRSS2-induced priming of S protein [67,68]. However, a recent study on HEK293T cells revealed that both TMPRSS2 and metalloproteases, particularly ADAM17, individually or co-operatively, promote activation of S2′ segment of S protein and syncytia formation in infection by SARS-CoV-2 and its emerging variants of concern [69]. Another study on various cell lines demonstrated cell-specific characteristic of this metalloprotease-dependent SARS-CoV-2 cell entry. This also requires the presence of furin-cleavage motif that is unique to SARS-CoV-2 not seen in SARS-CoV or MERS-CoV [70]. In addition, a cell free assay revealed that ADAM17 and ADAM10 (another member of ADAM family) are able to prime S protein, as well [71].
The endocytic cell entry of SARS-CoV-2 follows either a canonical (clathrin-mediated) or non-canonical (non-clathrin-mediated) pathway; the former proceeds through budding of clathrin-coated vesicles while the latter comprises caveolae-mediated, flotillin-1-dependent and glycosylphosphatidyl inositol-anchored protein endocytic pathways as well as macropinocytosis [4]. Moreover, the mandatory well-orchestrated role that the host Cathepsin B/L play in activation of S protein in both endosomal clathrin-dependent or non-endosomal clathrin-independent routes of entry must not be ignored, either [5,6,14,72]. In these two endocytic processes, activation of Cathepsins needs endosomal acidification interlinked with NAADP-elicited Ca2+ loss from the endosome which is mediated by vacuolar H+-ATPase (V-ATPase) [73,74,75,76,77].

5. Furin, a Ubiquitously Expressed Multifunctional Protease

Furin, (a type I transmembrane serine endopeptidase) trafficking between plasma membrane and endosomes is reserved predominantly in trans-Golgi network (TGN) and contributes to a quite large array of human physiologic, pathologic and metabolic processes as well as bacterial (anthrax) and viral (avian influenza, HIV) pathogenesis [78,79,80,81]. This endopeptidase comprises a signal peptide, an inhibitory prodomain, a subtilisin-like catalytic domain, a middle P domain, a cysteine-rich domain, a transmembrane peptide and a cytoplasmic tail. Full activation of furin, as a pre-proprotein, requires that the molecule be autocleaved at a primary (Arg-Thr-Lys-Arg107↓Asp108) as well as a secondary (Arg-Gly-Val-Thr-Lys-Arg75↓Ser76) site; the latter results in the release of the inhibitory prodomain fragment [82].
Furin exists in two forms: a membrane-bound and a circulating soluble form secreted by the cells; the higher levels of the latter increases the risk of metabolic syndrome, diabetes and hypertension [83]. This enzyme, ubiquitously expressed in tissues, is a member of Ca2+-dependent prohormone/proprotein convertase (PC) family involved in the proteolytic maturation of many hormones, neuropeptides, immune system molecular effectors, growth factors, von Willebrand factor, receptors (e.g., Notch receptors) and other enzymes such as ADAM17 (see Section 9) and metalloproteinases (MMPs) [84,85,86,87,88]. Furin proteolytic modification of its substrates occurs through cleavage at the C-terminal end of multibasic motifs (Arg-X-Arg/Lys/Arg-Arg↓ homologous to RRAR residues at S1/S2 boundary; see Section 2) in the substrates in the presence of Ca2+ [89]. Through this cleavage of the similar motif at S1/S2 interface of SARS-CoV-2, furin plays a pivotal role in the virus infectivity/pathogenicity [45]. Revealing the importance of this process, a cell culture study demonstrated that host cell-mediated O-glycosylation of S protein at proline681 (P681) proximal to the RRAR motif weakens furin-mediated cleavage and attenuates SARS-CoV-2 infectivity. Mutation of P681 in alfa and delta variants of this virus which mitigates O-glycosylation increases furin cleavage and potentiates the distribution of the virus and syncytia formation [90]. Moreover, a mutant variant of SARS-CoV-2 with deleted furin cleavage site showed attenuated infectivity, yet this variant compared to the wild type virus needed higher levels of anti-RBD antibody to be neutralized. This negates its usefulness for manufacturing an effective vaccine [91]. In a clinical study on a limited number of patients with COVID-19, the level of circulating furin was found to be correlated with the disease severity [92].
Optimum pH for full activation of furin is 7.0 though more than 50% of its maximum activity, depending on the substrates, occurs at pH between 6.5 and 8.0 [93]. Moreover, furin requires a calcium concentration of 10−3 M for its full function [79]. Assuming that extracellular ionized Ca concentration is about 10−3 M compared to its cytoplasmic content of about 10−7 M (reaches to 1–5 × 10−4 M in intracellular Ca2+ stores or escalates ultimately to 10−6 M to 10−5 M in stimulation of cell signaling [94,95]), it is obvious that nearly full proteolytic activity of furin in exocytic and endocytic pathways occurs extracellularly or in TNG/endosomal compartments and not in the cytoplasm [79,93,96]. In addition, a soluble active form of furin has long been recognized [97]. Given that the favorable ionized Ca2+ concentration for furin function is provided in extracellular space, it is implied that soluble furin isoform might potentiate SARS-CoV-2 S protein cleavage in a juxtacrine fashion [98].

6. TMPRSS2, Its Soluble Protease Domain and Calcium

TMPRSS2, a transcript product of an androgen-regulated polymorphic gene, tmprss2, [99,100] is expressed on the plasma membrane of several cell types [101]. This serine protease exists as two isoforms (1 and 2); the former contains 529 (with an extended cytoplasmic N-terminus) and the latter 492 residues [102]. TMPRSS2 co-expressed with ACE2 in the lungs and some other tissues, plays a pivotal role for entrance of SARS-CoV and SARS-CoV-2 into the host cells through priming the virus S protein (see Section 2 and Section 5) and cleavage of ACE2 (see Section 8) [53,103]. TMPRSS2 as a type II transmembrane protease is composed of an intracellular N-terminus, a single pass transmembrane domain and an extracellular segment. The latter contains three domains: a low density lipoprotein receptor type A (LDLRA), a class A scavenger receptor cysteine-rich (SRCR) and the C-terminal canonical serine protease domain (PDTMP) [102,104] (Figure 3).
Each of LDLRA and SRCR domains, collectively called the stem region which contributes to substrate recognition as well as protein-protein interactions and ligand binding, harbor Ca2+ binding sites [102,105,106]. PDTMP cleaves a variety of peptide substrates at Arginine and Lysine residues [106] which in the case of ACE2 occurs at the cluster of these two amino acids located in residues 697–716, noticing that this proteolysis is required for TMPRSS2-mediated cathepsin-independent virus entrance to the host cell [107]. Moreover, activation of TMPRSS2 (isoform 2) requires autocleavage of its own molecule at Arg255 in an as yet obscure condition, which may result in the release of PDTMP as a soluble form (sPDTMP) into extracellular space [108,109]. However, the majority of PDTMP remains linked to the rest of the molecule through a disulfide bond [102,110]. sPDTMP may interact with other cell surface proteins (in an autocrine manner), the extracellular matrix and proteins located on the neighboring cells (juxtacrine manner) [102]. Mutagenetic studies have demonstrated that SRCR, but not LDLRA, plays an essential role in proteolytic activity of PDTMP. Therefore, it might be a debatable issue whether PDTMP after dissociation from the stem region retains its proteolytic function or gains a yet unraveled distinct activity [111]. It is worth reminding that Ca2+ binding to or unbinding from a distinct domain of a protein determines the task the domain should fulfil [112] or affects the structural stabilization of the proteinases and substrates affinity [113,114]. It means that this bivalent ion participates in the regulation of the functions of SRCR and LDLRA domains rather than the proteolytic activity of PDTMP by itself after the latter is released as sPDTMP to the extracellular space losing its support by SRCR.

7. Major Activity of Cathepsins Occurs in the Membrane-Bound Intracellular Organelles

Human cathepsins, categorized in three families and 15 classes, function as lysosomal proteases and contribute to many physiological processes. These proteases are involved in the process of autophagy, the disequilibrium of which leads to numerous pathologies. Cathepsin activity is pH-dependent: acidic pH (3.5–6) enhances while neutral pH attenuates and alkaline pH inactivates their function [115]. However, some cathepsins including cathepsin K and H have been found to function at pH 7 in a stable manner implicating their ability to display activity outside lysosomes [116]. However, the majority of cathepsins, including cathepsin B and L that contribute to the activation of the S protein, undergo processing and maturation in the lysosomes rather than within the cytoplasm or extracellular space, otherwise they might induce apoptotic or necroptotic pathways or provoke inflammation in addition to their physiological role in tissue repair, extracellular matrix degradation and remodeling [115,117]. Even though TMPRSS2 and cathepsins may both contribute to activation of SARS-CoV-2 S protein, the former displays a more dominant function than the latter [118].

8. Membrane-Bound and Soluble ACE2

ACE2 contributes to infectivity and the entry of SARS-CoV-2 into the host cells while physiologically plays a crucial role in protection of the lungs, heart and other tissues against a variety of inflammatory or hypoxia-induced insults [119,120,121,122]. The non-homogenous distribution of ACE2, yet in a limited scale compared to ACE, was shown in a wide variety of tissues with the highest expression in the small intestine, heart, testis, thyroid glands and adipose tissue; intermediate expression in the lungs, colon, liver, bladder and adrenal glands; and the lowest content in the blood, spleen, bone marrow, brain, blood vessels and muscles [123]. Beyond the genetic traits determined by loci near immune related genes such as IFNAR2 and CXCR6 [124], the distinct genetic variants of highly polymorphic Ace2 (gene located on chromosome Xp22.2 encoding for ACE2 protein [125]) in different populations have also been shown to affect the susceptibility to and severity of COVID-19 in some cohorts of patients [126,127,128,129].
ACE2, a type I transmembrane protein of 805 residues, comprises an extracellular segment (residues 19–740) including a signal peptide of 18 amino acids, an N-terminal claw-like protease domain (PDACE2, residues 19–615) and a collectrin-like domain (residues 616–740) including ferredoxin-like fold “Neck” domain (residues 616–726) attached to a long transmembrane domain (residues 741–763) ending with a cytosolic C-terminal tail [130,131,132,133,134] (Figure 4). The cytosolic tail containing a conserved endocytosis short linear motif had been demonstrated to play a role in SARS-CoV-S protein-induced shedding of ACE2 (see Section 9), TNF-α production and SARS-CoV infection [135,136] although a combined in silico and in vitro study strengthened by confocal imaging denied any role for C-terminal tail of ACE2 in SARS-CoV and SARS-CoV-2 cell entry [137]. The collectrin-like domain contributes to dimerization of two ACE2 molecules (assuming ACE2-A and ACE2-B) through interacting with Arg652, Glu653, Ser709, Arg710 and Asp713 in ACE2-A with Tyr641, Tyr633, Asn638, Glu639, Asn636 and Arg716 in ACE2-B [130].
Physiologically, PDACE2 with its catalytic site removes one amino acid from the C-terminal end of angiotensin I (Ang I) and angiotensin II (Ang II) to turn them into angiotensin (1–9) and angiotensin (1–7), respectively [138]. An X-ray study uncovered the presence of a wide and deep cleft within the metallopeptidase catalytic domain of ACE2 with two subdomains I and II comprising its wall [139]. The catalytic efficiency of ACE2 for hydrolysis of Ang II is 400 times more than that of Ang I (as the precursor of Ang II) [138]. This means that any deficiency of ACE2 affects hydrolysis of Ang II 400 times more than that of its precursor, Ang I. Physiologically, this enzymatic hydrolysis leads to an increase in angiotensin(1–7)/Ang II concentration ratio resulting in attenuation of proinflammatory and organ-damaging effects of Ang II/AT1R pathway [140]. Conversely, deficiency of ACE2, which leads to an increase in tissue or plasma Ang II, may deprive the cell of one of its anti-inflammatory and protective tools [119]. An animal study on Ace2 (ACE2 gene) knockout mice proved that cardiac contractility is impaired, hypoxia-inducible genes such as BINP3 (encoding pro-apoptotic Bcl2 interacting protein) are upregulated and Ang II increases in the kidneys, heart and plasma [141,142]. Reciprocally, Ang II, while upregulates angiotensin converting enzyme (ACE), was shown, both in vitro and in vivo, to downregulate ACE2 expression at both mRNA and protein levels rather through a cellular inflammatory reaction induced by AT1R mediated ERK/p38MAPK pathway [143,144]. Furthermore, as a positive feedback effect, Ang II through stimulation of AT1R promotes shedding of the membrane bound ACE2 (mACE2) ectodomain off the cell surface into the plasma as soluble ACE2 (sACE2; residues 18-708) [145] (see above in this section) by inducing a sheddase protein, ADAM17 (see Section 9) [146,147]. Both of these two effects, which results in a decrease in mACE2 and an increase in circulating sACE2, can be blocked by Ang II type I blockers (ARBs), e.g., losartan [143,144,147]. Consequently, the concentration of sACE2 increases in pathologies associated with higher AT1R stimulation such as myocardial infarction, heart failure, metabolic syndrome and diabetes mellitus [148,149,150].

9. ADAM17, the Sheddase of ACE2

Regulated selective cleavage and release of the extracellular segment of many of transmembrane proteins into the extracellular space called “ectodomain shedding” modifies a diverse array of trans- and cis-signaling pathways [151]. ADAM17, the first identified shedding protease, is a Zn2+-dependent metalloproteinase and a member of the family of membrane-anchored ADAMs, which plays a role in both innate and humoral adaptive immunity, among the others [152]. This sheddase promotes proinflammatory effects through dissociating a variety of ligand proteins including ACE2, TNF-α and transmembrane CX3CL as its substrate as well as transmembrane receptors including IL6Ra and TNF receptors I and II [146,153,154,155]. This metalloprotease exists in two forms: the full length pro-ADAM17 of 100KDa and the mature form of 80KDa lacking the inhibitory prodomain: the latter comprises almost two thirds of its total cell content which resides predominantly in perinuclear space along with TNF-α [156,157]. ADAM17 as a type I transmembrane multidomain proteinase comprises an N-terminal signal sequence, an inhibitory prodomain, a metalloproteinase catalytic domain, a disintegrin-like domain, a cysteine-rich and membrane proximal domains (MPD, involved in substrate recognition) attached to a single transmembrane domain ending into a cytoplasmic tail [158,159] (Figure 5). The transmembrane domain, but not the cytoplasmic domain, contributes to the rapid activation of this metalloprotease by different signaling pathways [160]. However, phosphorylation of the cytoplasmic domain of ADAM17 by mitogen activating protein kinase (MAPK) network including p38MAPK and extracellular signal-regulated kinase (ERK) as well as Polo-like kinase2 (PLK2) keep this metalloprotease proteolytically functional through prevention of its dimerization and dampening of its binding to tissue inhibitor of metalloproteinase3 (TIMP3) at the cell surface [161,162]; TIMP3 plays an inhibitory role for the dimerized ADAM17 on the cell surface [163].
A proteolytically inactive rhomboid protein (iRhom2) encoded by Rhbdf2 gene involved in innate immunity against viral infections [165], regulates trafficking of pro-ADAM17 from endoplasmic reticulum to Golgi apparatus where furin detaches ADAM17 inhibitory prodomain to make it mature, substrate specific and be prepared for expression on the cell surface in complex with iRhom2 [166,167]. Infection with DNA and RNA viruses (influenza A, RSV) upregulates Rhbdf2 and ADAM17 genes [167]. However, an in vivo study revealed that ADAM17-dependent ectodomain shedding is a saturable phenomenon: it does not increase significantly when ADAM17 expression rises above a certain level [168]. It must be mentioned that this metalloprotease is rapidly and reversibly switched “on” and “off” through conformational changes making its catalytic domain accessible and inaccessible, respectively, yet independent of removal of the inhibitory pro-domain or dissociation of TIMP3 [169].
Given that ADAM17 small interfering RNA (siRNA) is capable of attenuating Ang II-mediated inflammation in vascular smooth muscle cells [170] it is implied that along with other downstream mediators, ADAM17 is also involved in Ang II-induced inflammatory responses. Additionally, Ang II-mediated stimulation of AT1R, a G-protein coupled receptor (GPCR), provokes oxidative stress, induces phospholipase C (PLC)-mediated protein kinase C (PKC) activation and increases calcium influx [171,172,173,174]. Oxidative stress in tumor cells and platelets could previously be found to activate ADAM17 with pro-inflammatory effects (see previous paragraph) [175,176]. Reciprocally, ADAM17 in a mouse model could also increase NADPH oxidase 4 (Nox4) activity resulting in oxidative stress [177] which by inducing pro-inflammatory genes is intertwined with inflammation [178,179]. It is, however, noticeable that ADAM17 positively regulates Thioredoxin-1 (Trx-1) activity as the key effector of intracellular reducing system and downregulates ADAM17 as a negative feedback effect [180,181].
Furthermore, the PKC pathway, depending on the nature of its activator, affects ADAM17-dependent ectodomain shedding. Short-term (minutes) activation of protein kinase C (PKC) by phorbol ester (the strongest non-physiologic stimulator of PKC) increases ADAM17 content on the cell plasma membrane and the sheddase activity while prolonged (hours) exposure to phorbol ester downregulates mature form of ADAM17 on the cell surface abolishing the ectodomain shedding without reducing the total cell content of ADAM17 (mature plus pro-ADAM17) [21]. Conversely, ligand-activated GPCRs, such as thrombin-mediated protease activated receptor1 (PAR1), with the potential to activating of PKC signaling [182,183], induces sheddase function without any change in the content of ADAM17 on the plasma membrane [21]. Moreover, Ang II, via AT1R, was found to upregulate PAR1 and PAR2 in rat aorta which lead to pro-inflammatory responses accompanied by raising IL-6 and monocyte chemoattractant protein-1 (MCP-1), the substrates of ADAM17 [156,184,185]. Along with the natural homo- and hetero-merization of GPCRs [172], an in vitro study revealed synergistic interaction between AT1R and PAR1 [186], implicating a positive influence of these receptors in activating ADAM17. Moreover, AT1R stimulation alone is associated with a rise in IL-1β which by itself promotes ectodomain shedding by ADAM17 [187,188]. Considering that Ang II via AT1R could induce PKCδ/p38MAPK [189] which also activates ADAM17 [190,191], it is logically implied that AT1R also potentiates ADAM17-induced ectodomain shedding through p38MAPK pathway.
Ca2+ influx and calmodulin inhibition stimulate ADAM10 [192]. Given that ADAM17 is involved in processing of pro-α2δ-1 and α2δ-3 subunits of voltage gated calcium channels which enhances calcium influx [193] it is implied that ADAM17 also indirectly contributes to activating ADAM10 in ectodomain shedding.

10. Shedding of ACE2 Ectodomain and sACE2 in SARS-CoV-2

As was mentioned previously, ADAM10 and ADAM17, the predominant sheddases, contribute to cleaving mACE2 off the cell surface which lead to a vast array of physiological and pathological cis- and trans-signaling [155]. Cleavage of the ectodomain of mACE2 occurs through a basal low level constitutive or a metalloproteinase-dependent fashion [194]. A cell culture study showed that inhibition of both ADAM17 and ADAM10, widely found on pneumocyte type I and II, results in ACE2 increment on the surface of these cells [71]. Consistently, an animal model of diabetes proved that ADAM17 gene knockdown and overexpression could reduce and raise mACE2 shedding off the cardiomyocytes, respectively [195].
Oxidative stress, MAPK network, PKC and Ca2+ signaling activation in various viral infections including SARS-CoV and SARS-CoV-2 has already been described [196,197,198,199,200,201,202,203]. Ang II-mediated AT1R stimulation activates these signaling pathways, as well [204,205]. All these signaling pathways directly promote ADAM17-dependent ACE2 shedding which by itself decreases angiotensin(1–7)/Ang II concentration ratio and promotes Ang II/AT1R pathway which also induces ADAM17 sheddase function (see previous section). Additionally, COVID-19 has been shown to promote ADAM17 expression both at the protein and transcriptional level [206].
In this context, neither should it be forgotten that the downregulation of mACE2 in COVID19 can be due to internalization of the virus-ACE2 complex [207,208,209,210] nor should the cleaving effects of TMPRSS2 along with ADAM17 be ignored. These two proteases compete in shedding, yet at different sites, of ACE2: ADAM17 and TMPRSS2 cleave Arginine and Lysine amino acids in residues 652–659 and 697–716, respectively [107]. However, TMPRSS2-mediated dissociation of mACE2 does not result in the release of sACE2 [10].
It was also shown that docking of the SARS-CoV S protein with ACE2 could induce detachment of the latter off the cell membrane by ADAM17, which can be blocked by the TNF protease inhibitor 2 (TAPI-2) [135]. Concordantly, SARS-CoV infection was shown to downregulate mACE2 in Vero E6 and the cells of the lungs in mice [211,212]. Anti-ADAM17 compounds after SARS-CoV infection could, in vitro and in vivo, inhibit ACE2 shedding, reduce sACE2 and suppress infectivity of the virus [213]. Consistently, a cell culture study confirmed that SARS-CoV-2 S protein induces metalloprotease-dependent ACE2 shedding [69]. Kornilov et al., in the early months after the COVID-19 pandemic, considered the elevated plasma level of sACE2 as a biomarker or possible cause of more severe SARS-CoV-2 infection seen in pathologies associated with higher levels of sACE2, e.g., diabetes, heart failure, metabolic syndrome and myocardial infarction as well as in the old, men and postmenopausal women [214]. Whatever the mechanism is, cleavage of mACE2 reduces its presence on the cell membrane and raises its soluble form (sACE2, containing PDACE2 including RBD docking domain) in extracellular space including airway surface liquid covering the lung epithelial cells [215,216].

11. sACE2: Attenuating Inflammatory Responses?

Both classical (ACE/Ang II/AT1R) and its counter-acting non-classical (ACE2/angiotensin(1–7)/Mas receptor) pathways of RAS exert their effects as an autocrine or paracrine signaling system at the local tissue level and also as an endocrine system signaling in the whole body [217]. In sepsis, upregulation of the classical endocrine RAS pathway to a definite level leads to life-saving responses, yet overstimulation of the local arm of the same classical RAS aggravates tissue inflammation which might result in acute respiratory distress syndrome (ARDS) attenuated through counter-activation of non-classic arm of local RAS [218]. Similarly, growing evidence suggests the contribution of dysregulated RAS to the pathogenesis of COVID-19 [219]. Plasma Ang II has been reported to increase early in critical SARS-CoV-2-induced lung injury which reaches high levels on admission and decreases steadily during the later days till discharge of the patient. It shows a positive correlation with the severity of the disease [220,221,222,223]. Furthermore, as sACE2 retains its catalytic activity [215] its coincidental increment with the decreasing of circulating Ang II level might be regarded as a sensible physiological response attenuating the adverse pathological effects of early overwhelming surge of Ang II. Concordantly, Lier et al., in a limited number of critical COVID-19 patients discovered that plasma concentration of sACE2 and angiotensin (1–7) increases while circulating Ang II decreases. They explained these to be likely the result of raised detachment of pulmonary cell mACE2 and its release into the circulation as sACE2 [224]. Similarly, in a recent study, sACE2 and Ang II were found to increase and decrease, respectively in severe COVID-19 [225].
Additionally, Kragstrup et al., in a study on COVID-19 patients determined the association of increasing sACE2 plasma content in the first week with the severity of the disease which could predict the worse outcome in 28 days post-infection. They concluded with uncertainty that increased sACE2 could play a role in spreading the virus to uninfected tissues [226]. Almost consistently, the findings of a recent human study of COVID-19 demonstrated that shedding of mACE2 was higher in all the patients compared to healthy controls on the day of admission and soared significantly to much higher levels in severe than moderate groups on the first post-admission week [227]. In this context, raised sACE2 may also be the result of ADAM-17 overactivation in more severe cases of COVID-19 which leads to an exceeding amount of mACE2 shedding. In a transgenic mouse model, monoclonal antibody against ADAM17 potently raised SARS-CoV-2 viral load due to decreased shedding of mACE2 which along with retaining anti-viral defense capabilities through increased Ifnb and Isg56 genes expression, remarkably ameliorated the post-infection inflammatory responses [228]. This is implied that COVID-19-induced lung-destructive inflammation might erupt rather via aggravation of ADAM-17-mediated release of mACE2 and TNF-α which results in the dysregulation of local RAS in favor of Ang II/AT1R axis and promoting inflammatory responses.
It is noticeable that Patel et al., found persistently elevated sACE2 activity in recovered patients from 35 to 114 post-COVID-infection days which was positively correlated with the severity of the disease [229]. Nevertheless, Lundström et al., reported that sACE2 concentration increases in COVID-19 patients compared to healthy people (5.0 vs. 1.4 ng/mL; 3.5 fold). They also found that early increment in plasma sACE2 content in the first 14 days of COVID-19 was transient and decreased significantly within four months post-infection, yet remained more than that of the control group. They also described a positive correlation between raised sACE2 and the prolongation of the symptom, yet with insignificant association with the mortality, severity indices (care level, any need for ventilator support) or total number of comorbidities [230]. Intriguingly, the rise in sACE2 concentration in this study showed no meaningful relationship with the plasma content of CRP, IL-6, TNF-α and plasminogen activator inhibitor-1 but weakly to moderately correlated with monocyte count, vWF concentration, D-Dimer and coagulation factor VIII [230]. It is evident from Lundström’s and another studies that, in COVID-19, sACE2 concentration rises (3.5 fold) less than that of IL-6 (about 6 fold up to a mean of 32 pg/mL [230] compared to 5 pg/mL in healthy individuals [231]). Considering over-activation of ADAM17 in COVID-19 [10] and its role in both shedding of mACE2 and IL-6-induced inflammatory responses [232] it seems that the lack of strong correlation between sACE2 level and inflammatory markers (IL-6) might be due to the saturable nature of ADAM17-dependent shedding of mACE2 (see Section 9) [168]; sACE2 does not rise to the level of increasing the activation of ADAM17 which is responsible for the raised IL-6-dependent signaling, as well. Consistently, in a mathematical model it was demonstrated that high levels of ADAM17 activity could only modestly deplete ACE2 off the surface of pancreatic β-cells [233]. Moreover, COVID-19-induced inflammatory responses erupts rather through multiple cascades ending into a multifaceted vasculopathic cytokine upregulation [234,235].
Interestingly, Daniell et al., discovered that despite the more abundant sACE2 in the plasma, its activity (in both plasma and saliva) and plasma angiotensin(1–7) concentration decreased significantly in COVID-19; the activity recovered gradually through the time [236]. The authors found a moderate correlation between the decline in angiotensin(1–7) concentration and lower sACE2 activity and attributed the discrepancy between plasma sACE2 content and activity to the presence of some other unknown mechanisms which may negatively regulate this soluble carboxypeptidase activity, mostly probable due to binding of sACE2 with the circulating SARS-CoV-2 S protein [236]. However, as fluorogenic Mca-APK(Dnp) mimicking Ang II is not a selective substrate for ACE2, using this compound in this study might have diminished the accuracy of the activity assay [237]. Moreover, Mca-APK(Dnp) based ACE2 function assay may generate conflicting results due to the presence of NaCl in the media preparation [238]. In addition, a reduction in plasma angiotensin(1–7) may also result from its hydrolysis via plasma angiotensin converting enzyme (ACE) or other peptidases rather than due to provisional diminished sACE2-dependent Ang II hydrolysis [239,240]. Moreover, there are some concerns about the specificity of angiotensin (1–7) assays using commercially available ELISA tests and the method of preparation of the plasma sample [241]. It is also noticeable that sACE2 catalytic activity is not expected to vanish after docking with RBD as the active enzymatic and RBD-binding sites in this soluble carboxypeptidase are located far from each other [216]. However, this concept has also been debated because docking to RBD could slightly close the functional cleft (see above in ACE2 section) within the ACE2 catalytic domain to some degree, hence this might modestly reduce its peptidase activity [242] (Figure 4 and Figure 6). Nevertheless, Lambert et al., demonstrated the catalytic ability of sACE2 [194] and beyond this Kiselev et al., revealed that docking to SARS-CoV-2-RBD potentiates multifunctional ACE2 ability to catalyze des-Arg9-bradykinin, with more modest accelerating effect on hydrolyzing Ang II [243]. It should also be mentioned that some inaccuracies might have inadvertently occurred as Mca-APK(Dnp) used in Kieslev’s study is a non-specific substrate of ACE2 [244]. However, it shows that sACE2 in docking with RBD, also attenuates inflammatory responses not only through hydrolyzing Ang II but via catalysis of des-Arg9-bradykinin which is involved in inflammation through kinin B1-receptor [245,246].
Altogether, it is implied that the rise in sACE2 in COVID-19 is the result, not the cause, of SARS-CoV-2-induced activation of a cascade of proinflammatory pathways accompanied by ADAM-17 stimulation which lead to the release of mACE2 as a soluble form into the circulation. It seems to act as a protective means to attenuate systemic classical ACE/Ang II/AT1R proinflammatory pathway in COVID-19.

12. sACE2: Increasing SARS-CoV-2 Infectivity?

Cytoplasmic tail signaling of mACE2 was found to have no influence on SARS-CoV-2 cell entry [137]. Retaining the amino acid residues involved in docking to RBD, sACE2 lacking membrane anchors is able to bind with the S protein of SARS-CoV-2 floating in the extracellular space [216]. This seems to prevent circulating SARS-CoV-2 from attaching to mACE2 on the surface of the encountered cells [249,250]. On the other hand, it has been hypothesized that sACE2 contributes to the entry of circulating SARS-CoV-2 into the cells with lower mACE2 content [251]. Although may be fascinating, both negating and supporting reports have made this issue unresolved [216,252].
Yeung et al., discovered that sACE2-S protein complex may facilitate SARS-CoV-2 cell entry through receptor-mediated endocytosis which is much more enhanced by the fusion of this complex with AT1R, vasopressin or vasopressin receptor, AVPR1B. They also demonstrated that sACE2 in low concentration (up to 100 ng/mL) dose dependently increases and in high concentration (25–100 μg/mL) decreases SARS-CoV-2 cell entry, in vitro [253]. Their findings support several reports showing the efficacy of recombinant human sACE2 or ACE2-containing exosomes in inhibiting SARS-CoV-2 cell entry via blocking RBD to bind with mACE2 [254,255].
Attaching at the “Neck” domain, mACE2 molecules naturally form homodimers with higher binding affinity (12–22 nM) to RBD compared to its monomer (77 nM) [256,257] (Figure 7). Assuming the trimeric nature of the S protein and homodimeric configuration of mACE2, it is mathematically speculated that interaction of two trimeric S proteins and three dimeric mACE2 should be involved in full fusion of SARS-CoV-2 and its receptor [258]. As ADAM17 cleaves mACE2 particularly at residues 652–659 [107] included in the Neck-domain (residues 616–726) [247], it is not unexpected, yet needed to be studied, that sACE2 exist in a monomeric form. It has been discovered that as sACE2 monomer dissociate rapidly from RBD, occupation of the second and the third RBDs in the trimeric S protein by homo-dimer and homo-trimer of this soluble carboxypeptidase neutralizes the virus more effectively [249,257]. This encouraged experts to produce a stable mutant dimeric and a homo-trimeric variant of sACE2 with preserved peptidase activity but higher binding affinity to RBD (about 600 pM for the dimeric and 60 pM for the trimeric sACE2) to hinder mACE2-RBD complexification [257,259]. In order to block S proteins on each SARS-CoV-2 particle with high efficiency, multimerized sACE2-fusion proteins have been manufactured through molecular engineering [260].
A combined hydrogen/deuterium exchange mass spectrometry (HDX-MS) and molecular dynamic simulation study revealed flexibility and dynamics of some distal hotspots in S protein relative to the site of RBD-ACE2 complex [261]. Concordantly, mACE2 complexed with RBD in “up” conformation was found to change the dynamics of the RBD leading to a continuous swing motion of S1-mACE2 complex on the top of the trimeric S protein. This declines the surface between S1 of mACE2-attached protomer and S2 subunits of the other two unattached protomers leading the trimeric S protein to untwist (Figure 1 and Figure 2) so that the two closed RBDs sequentially gain “up” status [34,262] (see also Section 3 and Figure 2). This continuous swing motion also enhances RBD and NTD dynamics in the attached S1 and leads to transition of S2 subunit of the same protomer toward the stable postfusion status to facilitate SARS-CoV2 cell entry [262]. It is needed to study whether mACE2 point of attachment to the plasma membrane plays a role as a fulcrum in this scenario and whether docking of monomeric sACE2 to RBD might induce the same swing motion and conformational changes.

13. Clinical Evidence Showing Benefit of Recombinant Human Soluble ACE2

Reminding that SARS-CoV-2 cell entry eventually leads to downregulation of mACE2, it is supposed that administration of sACE2 results in suppression of mACE2 shedding through blocking RBD docking sites [263]. Hence, it is persuasively expected that recombinant human soluble ACE2 (rhsACE2) is able to rebalance local Ang II/angiotensin(1–7) ratio in favor of non-classical anti-inflammatory ACE2/angiotensin(1–7)/MasR axis to reduce massive release of destructive cytokines [264,265,266]. During the last decade, pharmacodynamic and pharmacokinetic characteristics of rhsACE2, its safety in healthy people [267] and subsiding effect in acute respiratory distress syndrome (ARDS) have been extensively studied [268,269,270].
There are several reports including the ones published by Roshanravan et al. [271] and Basit et al. [272], which show the promising effect of sACE2 (as a decoy receptor) or truncated ACE2 (amino acid sequence 21–119 of N-terminal region) on reducing virulence of SARS-CoV-2 and its variants of concern in preclinical experiments [273,274,275]. Even some experts tried to produce sACE2-IgG conjugate to overcome the short half-life of sACE2 in the circulation and to improve its activity against SARS-CoV-2 infection [276]. In another study, Zhang et al., found that rhsACE2-Fc created by fusion of rhsACE2 to crystallizable fragment (Fc) of the N-terminal of human IgG with long half-life and high affinity to SARS-CoV-2-RBD could dose-dependently neutralize SARS-CoV-2 and its variants of concern both prophylactically and therapeutically. No infection promotion was observed after administration of this product [277]. Furthermore, rhsACE2-Fc could hinder S-protein-dependent giant cell formation and protect human bronchial cells against the invasion of the virus. This product could potently suppress the invasion of the mutants of SARS-CoV-2 generated through sequential passage in mammalian cells [277]. A recent animal study revealed how the impact of RAS dysregulation due to loss of mACE2 which leads to oxidative stress might eventuate into lung injury through activation of Ang II/AT1R/NF-κB/NOX1&2 signaling pathway [278]; the deadly signaling cascade that had hypothetically been proposed as the pathophysiologic basis of lung inflammation in early weeks of SARS-CoV-2 pandemic [279]. According to this study, rhsACE2 was found to be able to cleave Ang I and Ang II to angiotensin (1–9) and angiotensin (1–7), respectively. Moreover, this product could diminish proinflammatory factors such as TNF-α, IL-1β, IL-6 and myeloperoxidase activity in broncho-alveolar lavage fluid and ameliorate SARS-CoV-2-RBD-aggravated acute lung injury [278]. Additionally, considering the mouth as the most accessible gate of entry of SARS-CoV-2-contaminated droplets, an ACE2-containing chewing gum has successfully been tried to reduce the high viral load in the tongue, buccal and gingival epithelial cells through entrapping the virions in the saliva and hindering the attachment of RBD with mACE2 on the mouth epithelial cells [280].

14. Future Directions

It is evident that precise recognition of fascinating SARS-CoV-2 entry into the cells, as the key component of its infectivity, provides a vast array of data enriching scientists to defeat against the incessant threat of this deadly virus. The vaccines, even though showed appreciable efficacy to reduce the worldwide burden of the COVID-19 pandemics, may lose anti-SARS-CoV-2 neutralizing potency in long term as the rate of the virus mutation in S protein hotspots against which vaccines are targeted, overwhelms the laboratories’ potentials to release up to date immunogenic, yet safe, particles. In addition, the ability of health care organizations to disseminate the vaccines evenly around the world could not be relied on [281,282]. The currently available anti-COVID drugs may hypothetically rather do harm than good; a debating subject needed to be clarified precisely over the course of time [283,284,285,286,287]. As a matter of fact it seems that COVID-19, beyond being a seemingly ineradicable infectious disease, is rather a dys-homeostatic phenomenon initiated exactly as the virus rushes to penetrate into the cell through exploiting the homeostatic regulatory pathways. Hindering the virus cell entry seems to be an astonishing task to do in abolishing the infectivity and simultaneous stopping emergence of newly mutated variants of the virus. In this context, rebalancing the signaling pathways derailed in the course of the virus entrance by repurposing safe drugs and products is worth noticing as it plays a pivotal role in surviving the patients without considering the presence or absence of the virus itself or whether it is mutated or not.

15. Conclusions

SARS-CoV-2 cell entry, a meticulously orchestrated process, necessitates that elaborate ongoing interactions be regulated among multiple factors. The virus, by inducing host cell proteases, facilitates its fusion with the plasma membrane and, vice versa, the effector proteins of the cell play a crucial regulatory role in promoting required conformational changes of the virus key molecules. Among those, the mutual interactions of the S protein and mACE2, both as nanomachines, are of paramount importance, blocking of which suppresses the virus infectivity. Moreover, the role of the cell key proteinases, TMPRSS2, furin, Cathepsins as well as ADAM17, in different models of SARS-CoV-2 cell entry should not be ignored. Inhibiting these interacting molecules may limit the pathological reactions of COVID-19. Additionally, it is also noticeable that the disordered homeostatic signaling cascades such as RAS, GPCR-mediated and protease-dependent pathways may aggravate the clinical course of SARS-CoV-2 infection, rebalancing of which may restrict the pathogenesis and hence the mortality as well as morbidity of the disease. There are a growing number of studies showing the abnormal predominance of pro-inflammatory ACE/Ang II/AT1R/Nox over anti-inflammatory ACE2/Angiotensin(1–7)/MasR pathways as the probable cause of chaotic inflammatory responses in COVID-19. In this context, regulating effects of ACE inhibitors and ARBs have undergone variable in vitro and in vivo studies followed by clinical trials with controversial results. Similarly, administration of rhsACE2 which, at first glance, may act as a vehicle enabling SARS-CoV-2 to spread to other tissues enhancing the severity of the disease, may be promising in blocking the complexification of S protein with mACE2 while could also return the consequences of the pathological disorganized pathways into normal organized cascades to recover tissues physiological functions.

Author Contributions

Conceptualization, R.N.; methodology, R.N., M.F.T. and D.J.N.; software, R.N.; validation, R.N., M.F.T. and D.J.N.; investigation, R.N.; resources, R.N.; data curation, M.F.T. and D.J.N.; writing—original draft preparation, R.N., M.F.T. and D.J.N.; writing—review and editing, R.N.; supervision and project administration, R.N., M.F.T. and D.J.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data was created in this study.

Acknowledgments

This article is dedicated to the honorable women and men in Iran, the outstanding people who have proved to be worthy of living in dignity, pride and peace along a 7000-year long history with excellence shining brilliantly through the engraved Charter of the first world’s Declaration of Human Rights; a milestone document by Cyrus the Great. The authors would like to extend their sincere gratitude and greatest appreciation to Fatemeh Zare-Mirakabad ([email protected]), and Milad Besharatifard ([email protected]), Department of Mathematics and Computer Sciences, Amirkabir University of Technology, for their remarkable contribution to visualizing the 3D structure of the proteins obtained from the Protein Data Bank (RCSB.org) by using free (open-source) version of PyMol molecular graphics system (http://www.pymol.org) originally authored by Warren L. Delano [288].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, J.; Petitjean, S.J.; Koehler, M.; Zhang, Q.; Dumitru, A.C.; Chen, W.; Derclaye, S.; Vincent, S.P.; Soumillion, P.; Alsteens, D. Molecular interaction and inhibition of SARS-CoV-2 binding to the ACE2 receptor. Nat. Commun. 2020, 11, 1–10. [Google Scholar] [CrossRef] [PubMed]
  2. Jackson, C.B.; Farzan, M.; Chen, B.; Choe, H. Mechanisms of SARS-CoV-2 entry into cells. Nat. Rev. Mol. Cell Biol. 2022, 23, 3–20. [Google Scholar] [CrossRef] [PubMed]
  3. Shang, J.; Wan, Y.; Luo, C.; Ye, G.; Geng, Q.; Auerbach, A.; Li, F. Cell entry mechanisms of SARS-CoV-2. Proc. Natl. Acad. Sci. USA 2020, 117, 11727–11734. [Google Scholar] [CrossRef] [PubMed]
  4. Glebov, O.O. Understanding SARS-CoV-2 endocytosis for COVID-19 drug repurposing. FEBS J. 2020, 287, 3664–3671. [Google Scholar] [CrossRef]
  5. Bayati, A.; Kumar, R.; Francis, V.; McPherson, P.S. SARS-CoV-2 infects cells after viral entry via clathrin-mediated endocytosis. J. Biol. Chem. 2021, 296, 100306. [Google Scholar] [CrossRef]
  6. Knyazev, E.; Nersisyan, S.; Tonevitsky, A. Endocytosis and transcytosis of SARS-CoV-2 across the intestinal epithelium and other tissue barriers. Front. Immunol. 2021, 12, 636966. [Google Scholar] [CrossRef]
  7. Wang, H.; Yang, P.; Liu, K.; Guo, F.; Zhang, Y.; Zhang, G.; Jiang, C. SARS coronavirus entry into host cells through a novel clathrin-and caveolae-independent endocytic pathway. Cell Res. 2008, 18, 290–301. [Google Scholar] [CrossRef] [Green Version]
  8. Shen, X.-R.; Geng, R.; Li, Q.; Chen, Y.; Li, S.-F.; Wang, Q.; Min, J.; Yang, Y.; Li, B.; Jiang, R.-D. ACE2-independent infection of T lymphocytes by SARS-CoV-2. Signal Transduct. Target. Ther. 2022, 7, 1–11. [Google Scholar] [CrossRef]
  9. Junqueira, C.; Crespo, Â.; Ranjbar, S.; de Lacerda, L.B.; Lewandrowski, M.; Ingber, J.; Parry, B.; Ravid, S.; Clark, S.; Schrimpf, M.R. FcγR-mediated SARS-CoV-2 infection of monocytes activates inflammation. Nature 2022, 606, 576–584. [Google Scholar] [CrossRef]
  10. Zipeto, D.; Palmeira, J.d.F.; Argañaraz, G.A.; Argañaraz, E.R. ACE2/ADAM17/TMPRSS2 interplay may be the main risk factor for COVID-19. Front. Immunol. 2020, 11, 2642. [Google Scholar] [CrossRef]
  11. Fagyas, M.; Fejes, Z.; Sütő, R.; Nagy, Z.; Székely, B.; Pócsi, M.; Ivády, G.; Bíró, E.; Bekő, G.; Nagy, A. Circulating ACE2 activity predicts mortality and disease severity in hospitalized COVID-19 patients. Int. J. Infect. Dis. 2022, 115, 8–16. [Google Scholar] [CrossRef]
  12. Fuentes-Prior, P. Priming of SARS-CoV-2 S protein by several membrane-bound serine proteinases could explain enhanced viral infectivity and systemic COVID-19 infection. J. Biol. Chem. 2021, 296, 100135. [Google Scholar] [CrossRef]
  13. Oubahmane, M.; Hdoufane, I.; Bjij, I.; Lahcen, N.A.; Villemin, D.; Daoud, R.; Allali, A.E.; Cherqaoui, D. Host cell proteases mediating SARS-CoV-2 entry: An overview. Curr. Top. Med. Chem. 2022, 22, 1776–1792. [Google Scholar]
  14. Stevens, C.S.; Oguntuyo, K.Y.; Lee, B. Proteases and variants: Context matters for SARS-CoV-2 entry assays. Curr. Opin. Virol. 2021, 50, 49–58. [Google Scholar] [CrossRef]
  15. Guizani, I.; Fourti, N.; Zidi, W.; Feki, M.; Allal-Elasmi, M. SARS-CoV-2 and pathological matrix remodeling mediators. Inflamm. Res. 2021, 70, 847–858. [Google Scholar] [CrossRef]
  16. Benlarbi, M.; Laroche, G.; Fink, C.; Fu, K.; Mulloy, R.P.; Phan, A.; Ariana, A.; Stewart, C.M.; Prévost, J.; Beaudoin-Bussières, G. Identification of a SARS-CoV-2 host metalloproteinase-dependent entry pathway differentially used by SARS-CoV-2 and variants of concern Alpha, Delta, and Omicron. bioRxiv 2022. [Google Scholar]
  17. Boon, L.; Ugarte-Berzal, E.; Vandooren, J.; Opdenakker, G. Protease propeptide structures, mechanisms of activation, and functions. Crit. Rev. Biochem. Mol. Biol. 2020, 55, 111–165. [Google Scholar] [CrossRef]
  18. Fraser, B.J.; Beldar, S.; Seitova, A.; Hutchinson, A.; Mannar, D.; Li, Y.; Kwon, D.; Tan, R.; Wilson, R.P.; Leopold, K. Structure and activity of human TMPRSS2 protease implicated in SARS-CoV-2 activation. Nat. Chem. Biol. 2022, 18, 963–971. [Google Scholar] [CrossRef]
  19. Verma, S.; Dixit, R.; Pandey, K.C. Cysteine proteases: Modes of activation and future prospects as pharmacological targets. Front. Pharmacol. 2016, 7, 107. [Google Scholar] [CrossRef] [Green Version]
  20. Anderson, E.D.; VanSlyke, J.K.; Thulin, C.D.; Jean, F.; Thomas, G. Activation of the furin endoprotease is a multiple-step process: Requirements for acidification and internal propeptide cleavage. EMBO J. 1997, 16, 1508–1518. [Google Scholar] [CrossRef] [Green Version]
  21. Lorenzen, I.; Lokau, J.; Korpys, Y.; Oldefest, M.; Flynn, C.M.; Künzel, U.; Garbers, C.; Freeman, M.; Grötzinger, J.; Düsterhöft, S. Control of ADAM17 activity by regulation of its cellular localisation. Sci. Rep. 2016, 6, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Wang, M.-Y.; Zhao, R.; Gao, L.-J.; Gao, X.-F.; Wang, D.-P.; Cao, J.-M. SARS-CoV-2: Structure, biology, and structure-based therapeutics development. Front. Cell. Infect. Microbiol. 2020, 10, 587269. [Google Scholar] [CrossRef] [PubMed]
  23. Berger, I.; Schaffitzel, C. The SARS-CoV-2 spike protein: Balancing stability and infectivity. Cell Res. 2020, 30, 1059–1060. [Google Scholar] [CrossRef] [PubMed]
  24. Lan, J.; Ge, J.; Yu, J.; Shan, S.; Zhou, H.; Fan, S.; Zhang, Q.; Shi, X.; Wang, Q.; Zhang, L. Structure of the SARS-CoV-2 spike receptor-binding domain bound to the ACE2 receptor. Nature 2020, 581, 215–220. [Google Scholar] [CrossRef] [Green Version]
  25. Shah, P.; Canziani, G.A.; Carter, E.P.; Chaiken, I. The case for S2: The potential benefits of the S2 subunit of the SARS-CoV-2 spike protein as an immunogen in fighting the COVID-19 pandemic. Front. Immunol. 2021, 12, 637651. [Google Scholar] [CrossRef]
  26. Wu, Y.; Zhao, S. Furin cleavage sites naturally occur in coronaviruses. Stem Cell Res. 2021, 50, 102115. [Google Scholar] [CrossRef]
  27. Peacock, T.P.; Goldhill, D.H.; Zhou, J.; Baillon, L.; Frise, R.; Swann, O.C.; Kugathasan, R.; Penn, R.; Brown, J.C.; Sanchez-David, R.Y. The furin cleavage site in the SARS-CoV-2 spike protein is required for transmission in ferrets. Nat. Microbiol. 2021, 6, 899–909. [Google Scholar] [CrossRef]
  28. Buchanan, C.J.; Gaunt, B.; Harrison, P.J.; Yang, Y.; Liu, J.; Khan, A.; Giltrap, A.M.; Le Bas, A.; Ward, P.N.; Gupta, K. Pathogen-sugar interactions revealed by universal saturation transfer analysis. Science 2022, 377, eabm3125. [Google Scholar] [CrossRef]
  29. Mehdipour, A.R.; Hummer, G. Dual nature of human ACE2 glycosylation in binding to SARS-CoV-2 spike. Proc. Natl. Acad. Sci. USA 2021, 118, e2100425118. [Google Scholar] [CrossRef]
  30. Zhu, C.; He, G.; Yin, Q.; Zeng, L.; Ye, X.; Shi, Y.; Xu, W. Molecular biology of the SARs-CoV-2 spike protein: A review of current knowledge. J. Med. Virol. 2021, 93, 5729–5741. [Google Scholar] [CrossRef]
  31. Khare, S.; Azevedo, M.; Parajuli, P.; Gokulan, K. Conformational changes of the receptor binding domain of sars-cov-2 spike protein and prediction of a b-cell antigenic epitope using structural data. Front. Artif. Intell. 2021, 4, 31. [Google Scholar] [CrossRef]
  32. Yurkovetskiy, L.; Wang, X.; Pascal, K.E.; Tomkins-Tinch, C.; Nyalile, T.P.; Wang, Y.; Baum, A.; Diehl, W.E.; Dauphin, A.; Carbone, C. Structural and functional analysis of the D614G SARS-CoV-2 spike protein variant. Cell 2020, 183, 739–751.e738. [Google Scholar] [CrossRef]
  33. Giron, C.C.; Laaksonen, A.; Barroso da Silva, F.L. Up state of the SARS-COV-2 spike homotrimer favors an increased virulence for new variants. Front. Med. Technol. 2021, 3, 694347. [Google Scholar] [CrossRef]
  34. Benton, D.J.; Wrobel, A.G.; Xu, P.; Roustan, C.; Martin, S.R.; Rosenthal, P.B.; Skehel, J.J.; Gamblin, S.J. Receptor binding and priming of the spike protein of SARS-CoV-2 for membrane fusion. Nature 2020, 588, 327–330. [Google Scholar] [CrossRef]
  35. Korber, B.; Fischer, W.M.; Gnanakaran, S.; Yoon, H.; Theiler, J.; Abfalterer, W.; Hengartner, N.; Giorgi, E.E.; Bhattacharya, T.; Foley, B. Tracking changes in SARS-CoV-2 spike: Evidence that D614G increases infectivity of the COVID-19 virus. Cell 2020, 182, 812–827.e819. [Google Scholar] [CrossRef]
  36. Jordan, M. The meaning of affinity and the importance of identity in the designed world. Interactions 2010, 17, 6–11. [Google Scholar] [CrossRef]
  37. Delgado, J.M.; Duro, N.; Rogers, D.M.; Tkatchenko, A.; Pandit, S.A.; Varma, S. Molecular basis for higher affinity of SARS-CoV-2 spike RBD for human ACE2 receptor. Proteins Struct. Funct. Bioinform. 2021, 89, 1134–1144. [Google Scholar] [CrossRef]
  38. Wu, L.; Zhou, L.; Mo, M.; Liu, T.; Wu, C.; Gong, C.; Lu, K.; Gong, L.; Zhu, W.; Xu, Z. SARS-CoV-2 Omicron RBD shows weaker binding affinity than the currently dominant Delta variant to human ACE2. Signal Transduct. Target. Ther. 2022, 7, 1–3. [Google Scholar] [CrossRef]
  39. Stalls, V.; Lindenberger, J.; Gobeil, S.M.-C.; Henderson, R.; Parks, R.; Barr, M.; Deyton, M.; Martin, M.; Janowska, K.; Huang, X. Cryo-EM structures of SARS-CoV-2 Omicron BA. 2 spike. Cell Rep. 2022, 39, 111009. [Google Scholar] [CrossRef]
  40. Gobeil, S.M.-C.; Janowska, K.; McDowell, S.; Mansouri, K.; Parks, R.; Manne, K.; Stalls, V.; Kopp, M.F.; Henderson, R.; Edwards, R.J. D614G mutation alters SARS-CoV-2 spike conformation and enhances protease cleavage at the S1/S2 junction. Cell Rep. 2021, 34, 108630. [Google Scholar] [CrossRef]
  41. Henderson, R.; Edwards, R.J.; Mansouri, K.; Janowska, K.; Stalls, V.; Gobeil, S.; Kopp, M.; Li, D.; Parks, R.; Hsu, A.L. Controlling the SARS-CoV-2 spike glycoprotein conformation. Nat. Struct. Mol. Biol. 2020, 27, 925–933. [Google Scholar] [CrossRef] [PubMed]
  42. Gobeil, S.M.-C.; Janowska, K.; McDowell, S.; Mansouri, K.; Parks, R.; Stalls, V.; Kopp, M.F.; Manne, K.; Li, D.; Wiehe, K. Effect of natural mutations of SARS-CoV-2 on spike structure, conformation, and antigenicity. Science 2021, 373, eabi6226. [Google Scholar] [CrossRef] [PubMed]
  43. Huang, Y.; Yang, C.; Xu, X.-f.; Xu, W.; Liu, S.-w. Structural and functional properties of SARS-CoV-2 spike protein: Potential antivirus drug development for COVID-19. Acta Pharmacol. Sin. 2020, 41, 1141–1149. [Google Scholar] [CrossRef] [PubMed]
  44. Saputri, D.S.; Li, S.; Van Eerden, F.J.; Rozewicki, J.; Xu, Z.; Ismanto, H.S.; Davila, A.; Teraguchi, S.; Katoh, K.; Standley, D.M. Flexible, functional, and familiar: Characteristics of SARS-CoV-2 spike protein evolution. Front. Microbiol. 2020, 11, 2112. [Google Scholar] [CrossRef]
  45. Xia, S.; Lan, Q.; Su, S.; Wang, X.; Xu, W.; Liu, Z.; Zhu, Y.; Wang, Q.; Lu, L.; Jiang, S. The role of furin cleavage site in SARS-CoV-2 spike protein-mediated membrane fusion in the presence or absence of trypsin. Signal Transduct. Target. Ther. 2020, 5, 1–3. [Google Scholar] [CrossRef]
  46. Hou, Y.J.; Okuda, K.; Edwards, C.E.; Martinez, D.R.; Asakura, T.; Dinnon, I.I.I.K.H.; Kato, T.; Lee, R.E.; Yount, B.L.; Mascenik, T.M. SARS-CoV-2 reverse genetics reveals a variable infection gradient in the respiratory tract. Cell 2020, 182, 429–446.e414. [Google Scholar] [CrossRef]
  47. Örd, M.; Faustova, I.; Loog, M. The sequence at Spike S1/S2 site enables cleavage by furin and phospho-regulation in SARS-CoV2 but not in SARS-CoV1 or MERS-CoV. Sci. Rep. 2020, 10, 1–10. [Google Scholar] [CrossRef]
  48. Xia, X. Domains and functions of spike protein in Sars-Cov-2 in the context of vaccine design. Viruses 2021, 13, 109. [Google Scholar] [CrossRef]
  49. Lemmin, T.; Kalbermatter, D.; Harder, D.; Plattet, P.; Fotiadis, D. Structures and dynamics of the novel S1/S2 protease cleavage site loop of the SARS-CoV-2 spike glycoprotein. J. Struct. Biol. X 2020, 4, 100038. [Google Scholar] [CrossRef]
  50. Yu, S.; Zheng, X.; Zhou, B.; Li, J.; Chen, M.; Deng, R.; Wong, G.; Lavillette, D.; Meng, G. SARS-CoV-2 spike engagement of ACE2 primes S2′ site cleavage and fusion initiation. Proc. Natl. Acad. Sci. USA 2022, 119, e2111199119. [Google Scholar] [CrossRef]
  51. Sternberg, A.; Naujokat, C. Structural features of coronavirus SARS-CoV-2 spike protein: Targets for vaccination. Life Sci. 2020, 257, 118056. [Google Scholar] [CrossRef]
  52. Tang, T.; Jaimes, J.A.; Bidon, M.K.; Straus, M.R.; Daniel, S.; Whittaker, G.R. Proteolytic activation of SARS-CoV-2 spike at the S1/S2 boundary: Potential role of proteases beyond furin. ACS Infect. Dis. 2021, 7, 264–272. [Google Scholar] [CrossRef]
  53. Bestle, D.; Heindl, M.R.; Limburg, H.; Pilgram, O.; Moulton, H.; Stein, D.A.; Hardes, K.; Eickmann, M.; Dolnik, O.; Rohde, C. TMPRSS2 and furin are both essential for proteolytic activation of SARS-CoV-2 in human airway cells. Life Sci. Alliance 2020, 3, e202000786. [Google Scholar] [CrossRef]
  54. Juraszek, J.; Rutten, L.; Blokland, S.; Bouchier, P.; Voorzaat, R.; Ritschel, T.; Bakkers, M.J.; Renault, L.L.; Langedijk, J.P. Stabilizing the closed SARS-CoV-2 spike trimer. Nat. Commun. 2021, 12, 1–8. [Google Scholar] [CrossRef]
  55. Hoffmann, M.; Kleine-Weber, H.; Schroeder, S.; Krüger, N.; Herrler, T.; Erichsen, S.; Schiergens, T.S.; Herrler, G.; Wu, N.-H.; Nitsche, A. SARS-CoV-2 cell entry depends on ACE2 and TMPRSS2 and is blocked by a clinically proven protease inhibitor. Cell 2020, 181, 271–280.e278. [Google Scholar] [CrossRef]
  56. Belouzard, S.; Madu, I.; Whittaker, G.R. Elastase-mediated activation of the severe acute respiratory syndrome coronavirus spike protein at discrete sites within the S2 domain. J. Biol. Chem. 2010, 285, 22758–22763. [Google Scholar] [CrossRef] [Green Version]
  57. Fan, X.; Cao, D.; Kong, L.; Zhang, X. Cryo-EM analysis of the post-fusion structure of the SARS-CoV spike glycoprotein. Nat. Commun. 2020, 11, 1–10. [Google Scholar] [CrossRef]
  58. Koppisetti, R.K.; Fulcher, Y.G.; Van Doren, S.R. Fusion peptide of SARS-CoV-2 spike rearranges into a wedge inserted in bilayered micelles. J. Am. Chem. Soc. 2021, 143, 13205–13211. [Google Scholar] [CrossRef]
  59. Arbeitman, C.R.; Rojas, P.; Ojeda-May, P.; Garcia, M.E. The SARS-CoV-2 spike protein is vulnerable to moderate electric fields. Nat. Commun. 2021, 12, 1–13. [Google Scholar] [CrossRef]
  60. Cai, Y.; Zhang, J.; Xiao, T.; Peng, H.; Sterling, S.M.; Walsh, R.M., Jr.; Rawson, S.; Rits-Volloch, S.; Chen, B. Distinct conformational states of SARS-CoV-2 spike protein. Science 2020, 369, 1586–1592. [Google Scholar] [CrossRef]
  61. Walls, A.C.; Tortorici, M.A.; Bosch, B.-J.; Frenz, B.; Rottier, P.J.; DiMaio, F.; Rey, F.A.; Veesler, D. Cryo-electron microscopy structure of a coronavirus spike glycoprotein trimer. Nature 2016, 531, 114–117. [Google Scholar] [CrossRef] [Green Version]
  62. Hoffmann, M.; Kleine-Weber, H.; Pöhlmann, S. A multibasic cleavage site in the spike protein of SARS-CoV-2 is essential for infection of human lung cells. Mol. Cell 2020, 78, 779–. [Google Scholar] [CrossRef]
  63. Hörnich, B.F.; Großkopf, A.K.; Schlagowski, S.; Tenbusch, M.; Kleine-Weber, H.; Neipel, F.; Stahl-Hennig, C.; Hahn, A.S. SARS-CoV-2 and SARS-CoV spike-mediated cell-cell fusion differ in their requirements for receptor expression and proteolytic activation. J. Virol. 2021, 95, e00002–e00021. [Google Scholar] [CrossRef] [PubMed]
  64. Papa, G.; Mallery, D.L.; Albecka, A.; Welch, L.G.; Cattin-Ortolá, J.; Luptak, J.; Paul, D.; McMahon, H.T.; Goodfellow, I.G.; Carter, A. Furin cleavage of SARS-CoV-2 Spike promotes but is not essential for infection and cell-cell fusion. PLoS Pathog. 2021, 17, e1009246. [Google Scholar] [CrossRef] [PubMed]
  65. Duan, L.; Zheng, Q.; Zhang, H.; Niu, Y.; Lou, Y.; Wang, H. The SARS-CoV-2 spike glycoprotein biosynthesis, structure, function, and antigenicity: Implications for the design of spike-based vaccine immunogens. Front. Immunol. 2020, 11, 2593. [Google Scholar] [CrossRef] [PubMed]
  66. Hoffmann, M.; Pöhlmann, S. How SARS-CoV-2 makes the cut. Nat. Microbiol. 2021, 6, 828–829. [Google Scholar] [CrossRef]
  67. Ohradanova-Repic, A.; Skrabana, R.; Gebetsberger, L.; Tajti, G.; Baráth, P.; Ondrovičová, G.; Praženicová, R.; Jantova, N.; Hrasnova, P.; Stockinger, H. Blockade of TMPRSS2-mediated priming of SARS-CoV-2 by lactoferricin. Front. Immunol. 2022, 13, 958581. [Google Scholar] [CrossRef]
  68. Wettstein, L.; Knaff, P.M.; Kersten, C.; Müller, P.; Weil, T.; Conzelmann, C.; Müller, J.A.; Brückner, M.; Hoffmann, M.; Pöhlmann, S. Peptidomimetic inhibitors of TMPRSS2 block SARS-CoV-2 infection in cell culture. Commun. Biol. 2022, 5, 1–12. [Google Scholar] [CrossRef]
  69. Harte, J.V.; Wakerlin, S.L.; Lindsay, A.J.; McCarthy, J.V.; Cloeman-Vaughan, C. Metalloprotease-Dependent S2’-Activation Promotes Cell–Cell Fusion and Syncytiation of SARS-CoV-2. Viruses 2022, 14, 2094. [Google Scholar] [CrossRef]
  70. Yamamoto, M.; Gohda, J.; Kobayashi, A.; Tomita, K.; Hirayama, Y.; Koshikawa, N.; Seiki, M.; Semba, K.; Akiyama, T.; Kawaguchi, Y. Metalloproteinase-dependent and TMPRSS2-independent cell surface entry pathway of SARS-CoV-2 requires the furin cleavage site and the S2 domain of spike protein. Mbio 2022, 13, e00519–e00522. [Google Scholar] [CrossRef]
  71. Jocher, G.; Grass, V.; Tschirner, S.K.; Riepler, L.; Breimann, S.; Kaya, T.; Oelsner, M.; Hamad, M.S.; Hofmann, L.I.; Blobel, C.P. ADAM10 and ADAM17 promote SARS-CoV-2 cell entry and spike protein-mediated lung cell fusion. EMBO Rep. 2022, 23, e54305. [Google Scholar] [CrossRef]
  72. Xiu, S.; Dick, A.; Ju, H.; Mirzaie, S.; Abdi, F.; Cocklin, S.; Zhan, P.; Liu, X. Inhibitors of SARS-CoV-2 entry: Current and future opportunities. J. Med. Chem. 2020, 63, 12256–12274. [Google Scholar] [CrossRef]
  73. Petersen, O.H.; Gerasimenko, O.V.; Gerasimenko, J.V. Endocytic uptake of SARS-CoV-2: The critical roles of pH, Ca2+, and NAADP. Function 2020, 1, zqaa003. [Google Scholar] [CrossRef]
  74. Khan, N.; Chen, X.; Geiger, J.D. Role of endolysosomes in severe acute respiratory syndrome coronavirus-2 infection and coronavirus disease 2019 pathogenesis: Implications for potential treatments. Front. Pharmacol. 2020, 11, 595888. [Google Scholar] [CrossRef]
  75. Padmanabhan, P.; Desikan, R.; Dixit, N.M. Targeting TMPRSS2 and Cathepsin B/L together may be synergistic against SARS-CoV-2 infection. PLoS Comput. Biol. 2020, 16, e1008461. [Google Scholar] [CrossRef]
  76. Hashimoto, R.; Sakamoto, A.; Deguchi, S.; Yi, R.; Sano, E.; Hotta, A.; Takahashi, K.; Yamanaka, S.; Takayama, K. Dual inhibition of TMPRSS2 and Cathepsin B prevents SARS-CoV-2 infection in iPS cells. Mol. Ther. Nucleic Acids 2021, 26, 1107–1114. [Google Scholar] [CrossRef]
  77. Zhao, M.-M.; Yang, W.-L.; Yang, F.-Y.; Zhang, L.; Huang, W.-J.; Hou, W.; Fan, C.-F.; Jin, R.-H.; Feng, Y.-M.; Wang, Y.-C. Cathepsin L plays a key role in SARS-CoV-2 infection in humans and humanized mice and is a promising target for new drug development. Signal Transduct. Target. Ther. 2021, 6, 1–12. [Google Scholar] [CrossRef]
  78. Schapiro, F.B.; Soe, T.T.; Mallet, W.G.; Maxfield, F.R. Role of cytoplasmic domain serines in intracellular trafficking of furin. Mol. Biol. Cell 2004, 15, 2884–2894. [Google Scholar] [CrossRef] [Green Version]
  79. Thomas, G. Furin at the cutting edge: From protein traffic to embryogenesis and disease. Nat. Rev. Mol. Cell Biol. 2002, 3, 753–766. [Google Scholar] [CrossRef] [Green Version]
  80. Fitzgerald, K. Furin protease: From SARS Cov-2 to anthrax, diabetes, and hypertension. Perm. J. 2020, 24, 187. [Google Scholar] [CrossRef]
  81. Gu, M.; Rappapor, J.; Leppla, S.H. Furin is important but not essential for the proteolytic maturation of gp160 of HIV-1. FEBS Lett. 1995, 365, 95–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Gawlik, K.; Shiryaev, S.A.; Zhu, W.; Motamedchaboki, K.; Desjardins, R.; Day, R.; Remacle, A.G.; Stec, B.; Strongin, A.Y. Autocatalytic activation of the furin zymogen requires removal of the emerging enzyme’s N-terminus from the active site. PLoS ONE 2009, 4, e5031. [Google Scholar] [CrossRef] [PubMed]
  83. Fernandez, C.; Rysä, J.; Almgren, P.; Nilsson, J.; Engström, G.; Orho-Melander, M.; Ruskoaho, H.; Melander, O. Plasma levels of the proprotein convertase furin and incidence of diabetes and mortality. J. Intern. Med. 2018, 284, 377–387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Rose, M.; Duhamel, M.; Rodet, F.; Salzet, M. The role of proprotein convertases in the regulation of the function of immune cells in the oncoimmune response. Front. Immunol. 2021, 12, 667850. [Google Scholar] [CrossRef]
  85. Srour, N.; Lebel, A.; McMahon, S.; Fournier, I.; Fugère, M.; Day, R.; Dubois, C.M. TACE/ADAM-17 maturation and activation of sheddase activity require proprotein convertase activity. FEBS Lett. 2003, 554, 275–283. [Google Scholar] [CrossRef] [Green Version]
  86. Sato, H.; Kinoshita, T.; Takino, T.; Nakayama, K.; Seiki, M. Activation of a recombinant membrane type 1-matrix metalloproteinase (MT1-MMP) by furin and its interaction with tissue inhibitor of metalloproteinases (TIMP)-2. FEBS Lett. 1996, 393, 101–104. [Google Scholar] [CrossRef] [Green Version]
  87. Pesu, M.; Watford, W.T.; Wei, L.; Xu, L.; Fuss, I.; Strober, W.; Andersson, J.; Shevach, E.M.; Quezado, M.; Bouladoux, N. T-cell-expressed proprotein convertase furin is essential for maintenance of peripheral immune tolerance. Nature 2008, 455, 246–250. [Google Scholar] [CrossRef] [Green Version]
  88. Rawley, O.; Lillicrap, D. Functional roles of the von Willebrand factor propeptide. Hämostaseologie 2021, 41, 063–068. [Google Scholar] [CrossRef]
  89. Vankadari, N. Structure of furin protease binding to SARS-CoV-2 spike glycoprotein and implications for potential targets and virulence. J. Phys. Chem. Lett. 2020, 11, 6655–6663. [Google Scholar] [CrossRef]
  90. Zhang, L.; Mann, M.; Syed, Z.A.; Reynolds, H.M.; Tian, E.; Samara, N.L.; Zeldin, D.C.; Tabak, L.A.; Ten Hagen, K.G. Furin cleavage of the SARS-CoV-2 spike is modulated by O-glycosylation. Proc. Natl. Acad. Sci. USA 2021, 118, e2109905118. [Google Scholar] [CrossRef]
  91. Johnson, B.A.; Xie, X.; Bailey, A.L.; Kalveram, B.; Lokugamage, K.G.; Muruato, A.; Zou, J.; Zhang, X.; Juelich, T.; Smith, J.K. Loss of furin cleavage site attenuates SARS-CoV-2 pathogenesis. Nature 2021, 591, 293–299. [Google Scholar] [CrossRef]
  92. Kocyigit, A.; Sogut, O.; Durmus, E.; Kanimdan, E.; Guler, E.M.; Kaplan, O.; Yenigun, V.B.; Eren, C.; Ozman, Z.; Yasar, O. Circulating furin, IL-6, and presepsin levels and disease severity in SARS-CoV-2–infected patients. Sci. Prog. 2021, 104 (Suppl. S2), 00368504211026119. [Google Scholar] [CrossRef]
  93. Declercq, J.; Creemers, J.W.M. Furin. In Handbook of Proteolytic Enzymes, 3rd ed.; Rawlings, N.D., Saavedra, G.S., Eds.; Academic Press: London, UK, 2012; pp. 3281–3285. [Google Scholar]
  94. Bagur, R.; Hajnóczky, G. Intracellular Ca2+ sensing: Its role in calcium homeostasis and signaling. Mol. Cell 2017, 66, 780–788. [Google Scholar] [CrossRef] [Green Version]
  95. Atchison, D.K.; Beierwaltes, W.H. The influence of extracellular and intracellular calcium on the secretion of renin. Pflügers Arch. Eur. J. Physiol. 2013, 465, 59–69. [Google Scholar] [CrossRef] [Green Version]
  96. Thier, S.O. Potassium physiology. Am. J. Med. 1986, 80, 3–7. [Google Scholar] [CrossRef]
  97. Vidricaire, G.; Denault, J.-B.; Leduc, R. Characterization of a secreted form of human furin endoprotease. Biochem. Biophys. Res. Commun. 1993, 195, 1011–1018. [Google Scholar] [CrossRef]
  98. Lukassen, S.; Chua, R.L.; Trefzer, T.; Kahn, N.C.; Schneider, M.A.; Muley, T.; Winter, H.; Meister, M.; Veith, C.; Boots, A.W. SARS-CoV-2 receptor ACE 2 and TMPRSS 2 are primarily expressed in bronchial transient secretory cells. EMBO J. 2020, 39, e105114. [Google Scholar] [CrossRef]
  99. Qiao, Y.; Wang, X.-M.; Mannan, R.; Pitchiaya, S.; Zhang, Y.; Wotring, J.W.; Xiao, L.; Robinson, D.R.; Wu, Y.-M.; Tien, J.C.-Y. Targeting transcriptional regulation of SARS-CoV-2 entry factors ACE2 and TMPRSS2. Proc. Natl. Acad. Sci. USA 2021, 118, e2021450118. [Google Scholar] [CrossRef]
  100. Rokni, M.; Heidari Nia, M.; Sarhadi, M.; Mirinejad, S.; Sargazi, S.; Moudi, M.; Saravani, R.; Rahdar, S.; Kargar, M. Association of TMPRSS2 gene polymorphisms with COVID-19 severity and mortality: A case-control study with computational analyses. Appl. Biochem. Biotechnol. 2022, 194, 3507–3526. [Google Scholar] [CrossRef]
  101. Dong, M.; Zhang, J.; Ma, X.; Tan, J.; Chen, L.; Liu, S.; Xin, Y.; Zhuang, L. ACE2, TMPRSS2 distribution and extrapulmonary organ injury in patients with COVID-19. Biomed. Pharmacother. 2020, 131, 110678. [Google Scholar] [CrossRef]
  102. Thunders, M.; Delahunt, B. Gene of the month: TMPRSS2 (transmembrane serine protease 2). J. Clin. Pathol. 2020, 73, 773–776. [Google Scholar] [CrossRef] [PubMed]
  103. Shulla, A.; Heald-Sargent, T.; Subramanya, G.; Zhao, J.; Perlman, S.; Gallagher, T. A transmembrane serine protease is linked to the severe acute respiratory syndrome coronavirus receptor and activates virus entry. J. Virol. 2011, 85, 873–882. [Google Scholar] [CrossRef] [PubMed]
  104. Hussain, M.; Jabeen, N.; Amanullah, A.; Baig, A.A.; Aziz, B.; Shabbir, S.; Raza, F.; Uddin, N. Molecular docking between human TMPRSS2 and SARS-CoV-2 spike protein: Conformation and intermolecular interactions. AIMS Microbiol. 2020, 6, 350. [Google Scholar] [CrossRef] [PubMed]
  105. Reichhardt, M.P.; Loimaranta, V.; Lea, S.M.; Johnson, S. Structures of SALSA/DMBT1 SRCR domains reveal the conserved ligand-binding mechanism of the ancient SRCR fold. Life Sci. Alliance 2020, 3, e201900502. [Google Scholar] [CrossRef] [PubMed]
  106. Paoloni-Giacobino, A.; Chen, H.; Peitsch, M.C.; Rossier, C.; Antonarakis, S.E. Cloning of the TMPRSS2 gene, which encodes a novel serine protease with transmembrane, LDLRA, and SRCR domains and maps to 21q22. 3. Genomics 1997, 44, 309–320. [Google Scholar] [CrossRef]
  107. Heurich, A.; Hofmann-Winkler, H.; Gierer, S.; Liepold, T.; Jahn, O.; Pöhlmann, S. TMPRSS2 and ADAM17 cleave ACE2 differentially and only proteolysis by TMPRSS2 augments entry driven by the severe acute respiratory syndrome coronavirus spike protein. J. Virol. 2014, 88, 1293–1307. [Google Scholar] [CrossRef] [Green Version]
  108. Afar, D.E.; Vivanco, I.; Hubert, R.S.; Kuo, J.; Chen, E.; Saffran, D.C.; Raitano, A.B.; Jakobovits, A. Catalytic cleavage of the androgen-regulated TMPRSS2 protease results in its secretion by prostate and prostate cancer epithelia. Cancer Res. 2001, 61, 1686–1692. [Google Scholar]
  109. David, A.; Parkinson, N.; Peacock, T.P.; Pairo-Castineira, E.; Khanna, T.; Cobat, A.; Tenesa, A.; Sancho-Shimizu, V.; Casanova, J.-L.; Abel, L. A common TMPRSS2 variant has a protective effect against severe COVID-19. Curr. Res. Transl. Med. 2022, 70, 103333. [Google Scholar] [CrossRef]
  110. Wettstein, L.; Kirchhoff, F.; Münch, J. The transmembrane protease TMPRSS2 as a therapeutic target for COVID-19 treatment. Int. J. Mol. Sci. 2022, 23, 1351. [Google Scholar] [CrossRef]
  111. Epstein, R.J. The secret identities of TMPRSS2: Fertility factor, virus trafficker, inflammation moderator, prostate protector and tumor suppressor. Tumor Biol. 2021, 43, 159–176. [Google Scholar] [CrossRef]
  112. Görlach, A.; Bertram, K.; Hudecova, S.; Krizanova, O. Calcium and ROS: A mutual interplay. Redox Biol. 2015, 6, 260–271. [Google Scholar] [CrossRef] [Green Version]
  113. Papaleo, E.; Fantucci, P.; De Gioia, L. Effects of calcium binding on structure and autolysis regulation in trypsins. A molecular dynamics investigation. J. Chem. Theory Comput. 2005, 1, 1286–1297. [Google Scholar] [CrossRef]
  114. Liu, S.-Q.; Tao, Y.; Meng, Z.-H.; Fu, Y.-X.; Zhang, K.-Q. The effect of calciums on molecular motions of proteinase K. J. Mol. Model. 2011, 17, 289–300. [Google Scholar] [CrossRef]
  115. Patel, S.; Homaei, A.; El-Seedi, H.R.; Akhtar, N. Cathepsins: Proteases that are vital for survival but can also be fatal. Biomed. Pharmacother. 2018, 105, 526–532. [Google Scholar] [CrossRef]
  116. Yadati, T.; Houben, T.; Bitorina, A.; Shiri-Sverdlov, R. The ins and outs of cathepsins: Physiological function and role in disease management. Cells 2020, 9, 1679. [Google Scholar] [CrossRef]
  117. Gomes, C.P.; Fernandes, D.E.; Casimiro, F.; da Mata, G.F.; Passos, M.T.; Varela, P.; Mastroianni-Kirsztajn, G.; Pesquero, J.B. Cathepsin L in COVID-19, From Pharmacological Evidences to Genetics. Front. Cell. Infect. Microbiol. 2020, 10, 589505. [Google Scholar] [CrossRef]
  118. Ou, T.; Mou, H.; Zhang, L.; Ojha, A.; Choe, H.; Farzan, M. Hydroxychloroquine-mediated inhibition of SARS-CoV-2 entry is attenuated by TMPRSS2. PLoS Pathog. 2021, 17, e1009212. [Google Scholar] [CrossRef]
  119. Turner, A. ACE2 cell biology, regulation, and physiological functions. In The Protective Arm of the Renin Angiotensin System (RAS): Functional Aspects and Therapeutic Implications; Unger, T., Steckelings, U.M., Eds.; Academic Press: London, UK, 2015; pp. 185–189. [Google Scholar]
  120. Scialo, F.; Daniele, A.; Amato, F.; Pastore, L.; Matera, M.G.; Cazzola, M.; Castaldo, G.; Bianco, A. ACE2: The major cell entry receptor for SARS-CoV-2. Lung 2020, 198, 867–877. [Google Scholar] [CrossRef]
  121. Kuba, K.; Imai, Y.; Ohto-Nakanishi, T.; Penninger, J.M. Trilogy of ACE2: A peptidase in the renin–angiotensin system, a SARS receptor, and a partner for amino acid transporters. Pharmacol. Ther. 2010, 128, 119–128. [Google Scholar] [CrossRef]
  122. Hu, H.; Zhang, R.; Dong, L.; Chen, E.; Ying, K. Overexpression of ACE2 prevents hypoxia-induced pulmonary hypertension in rats by inhibiting proliferation and immigration of PASMCs. Eur. Rev. Med. Pharm. Sci. 2020, 24, 3968–3980. [Google Scholar]
  123. Li, M.-Y.; Li, L.; Zhang, Y.; Wang, X.-S. Expression of the SARS-CoV-2 cell receptor gene ACE2 in a wide variety of human tissues. Infect. Dis. Poverty 2020, 9, 23–29. [Google Scholar] [CrossRef] [PubMed]
  124. Neimi, M.E.K.; Karjalainen, J.; Liao, R.G.; Neale, B.M.; Daly, M.; Ganna, A.; Pathak, G.A.; Andrews, S.J.; Kanai, M.; Veerapen, K.; et al. Mapping the human genetic architecture of COVID-19. Nature 2021, 600, 472–477. [Google Scholar]
  125. Luo, Y.; Liu, C.; Guan, T.; Li, Y.; Lai, Y.; Li, F.; Zhao, H.; Maimaiti, T.; Zeyaweiding, A. Association of ACE2 genetic polymorphisms with hypertension-related target organ damages in south Xinjiang. Hypertens. Res. 2019, 42, 681–689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Cao, Y.; Li, L.; Feng, Z.; Wan, S.; Huang, P.; Sun, X.; Wen, F.; Huang, X.; Ning, G.; Wang, W. Comparative genetic analysis of the novel coronavirus (2019-nCoV/SARS-CoV-2) receptor ACE2 in different populations. Cell Discov. 2020, 6, 1–4. [Google Scholar] [CrossRef] [Green Version]
  127. Singh, H.; Choudhari, R.; Nema, V.; Khan, A.A. ACE2 and TMPRSS2 polymorphisms in various diseases with special reference to its impact on COVID-19 disease. Microb. Pathog. 2021, 150, 104621. [Google Scholar] [CrossRef]
  128. Li, J.; Wang, Y.; Liu, Y.; Zhang, Z.; Zhai, Y.; Dai, Y.; Wu, Z.; Nie, X.; Du, L. Polymorphisms and mutations of ACE2 and TMPRSS2 genes are associated with COVID-19, a systematic review. Eur. J. Med. Res. 2022, 27, 1–10. [Google Scholar] [CrossRef]
  129. Lanjanian, H.; Moazzam-Jazi, M.; Hedayati, M.; Akbarzadeh, M.; Guity, K.; Sedaghati-Khayat, B.; Azizi, F.; Daneshpour, M.S. SARS-CoV-2 infection susceptibility influenced by ACE2 genetic polymorphisms: Insights from Tehran Cardio-Metabolic Genetic Study. Sci. Rep. 2021, 11, 1–13. [Google Scholar] [CrossRef]
  130. Senapati, S.; Banerjee, P.; Bhagavatula, S.; Kushwaha, P.P.; Kumar, S. Contributions of human ACE2 and TMPRSS2 in determining host–pathogen interaction of COVID-19. J. Genet. 2021, 100, 1–16. [Google Scholar] [CrossRef]
  131. Devaux, C.A.; Rolain, J.-M.; Raoult, D. ACE2 receptor polymorphism: Susceptibility to SARS-CoV-2, hypertension, multi-organ failure, and COVID-19 disease outcome. J. Microbiol. Immunol. Infect. 2020, 53, 425–435. [Google Scholar] [CrossRef]
  132. Badawi, S.; Ali, B.R. ACE2 Nascence, trafficking, and SARS-CoV-2 pathogenesis: The saga continues. Hum. Genom. 2021, 15, 1–14. [Google Scholar]
  133. Gheblawi, M.; Wang, K.; Viveiros, A.; Nguyen, Q.; Zhong, J.-C.; Turner, A.J.; Raizada, M.K.; Grant, M.B.; Oudit, G.Y. Angiotensin-converting enzyme 2: SARS-CoV-2 receptor and regulator of the renin-angiotensin system: Celebrating the 20th anniversary of the discovery of ACE2. Circ. Res. 2020, 126, 1456–1474. [Google Scholar] [CrossRef]
  134. Donoghue, M.; Hsieh, F.; Baronas, E.; Godbout, K.; Gosselin, M.; Stagliano, N.; Donovan, M.; Woolf, B.; Robison, K.; Jeyaseelan, R. A novel angiotensin-converting enzyme–related carboxypeptidase (ACE2) converts angiotensin I to angiotensin 1-9. Circ. Res. 2000, 87, e1–e9. [Google Scholar] [CrossRef]
  135. Haga, S.; Yamamoto, N.; Nakai-Murakami, C.; Osawa, Y.; Tokunaga, K.; Sata, T.; Yamamoto, N.; Sasazuki, T.; Ishizaka, Y. Modulation of TNF-α-converting enzyme by the spike protein of SARS-CoV and ACE2 induces TNF-α production and facilitates viral entry. Proc. Natl. Acad. Sci. USA 2008, 105, 7809–7814. [Google Scholar] [CrossRef] [Green Version]
  136. Kliche, J.; Kuss, H.; Ali, M.; Ivarsson, Y. Cytoplasmic short linear motifs in ACE2 and integrin β3 link SARS-CoV-2 host cell receptors to mediators of endocytosis and autophagy. Sci. Signal. 2021, 14, eabf1117. [Google Scholar] [CrossRef]
  137. Karthika, T.; Joseph, J.; Das, V.; Nair, N.; Charulekha, P.; Roji, M.D.; Raj, V.S. SARS-CoV-2 cellular entry is independent of the ACE2 cytoplasmic domain signaling. Cells 2021, 10, 1814. [Google Scholar] [CrossRef]
  138. Vickers, C.; Hales, P.; Kaushik, V.; Dick, L.; Gavin, J.; Tang, J.; Godbout, K.; Parsons, T.; Baronas, E.; Hsieh, F. Hydrolysis of biological peptides by human angiotensin-converting enzyme-related carboxypeptidase. J. Biol. Chem. 2002, 277, 14838–14843. [Google Scholar] [CrossRef]
  139. Towler, P.; Staker, B.; Prasad, S.G.; Menon, S.; Tang, J.; Parsons, T.; Ryan, D.; Fisher, M.; Williams, D.; Dales, N.A. ACE2 X-ray structures reveal a large hinge-bending motion important for inhibitor binding and catalysis. J. Biol. Chem. 2004, 279, 17996–18007. [Google Scholar] [CrossRef] [Green Version]
  140. Villalobos, L.A.; Hipólito-Luengo, S.; Ramos-González, M.; Cercas, E.; Vallejo, S.; Romero, A.; Romacho, T.; Carraro, R.; Sánchez-Ferrer, C.F.; Peiró, C. The angiotensin-(1-7)/mas axis counteracts angiotensin II-dependent and-independent pro-inflammatory signaling in human vascular smooth muscle cells. Front. Pharmacol. 2016, 7, 482. [Google Scholar] [CrossRef] [Green Version]
  141. Crackower, M.A.; Sarao, R.; Oudit, G.Y.; Yagil, C.; Kozieradzki, I.; Scanga, S.E.; Oliveira-dos-Santos, A.J.; da Costa, J.; Zhang, L.; Pei, Y. Angiotensin-converting enzyme 2 is an essential regulator of heart function. Nature 2002, 417, 822–828. [Google Scholar] [CrossRef]
  142. Zhang, J.; Ney, P.A. Role of BNIP3 and NIX in cell death, autophagy, and mitophagy. Cell Death Differ. 2009, 16, 939–946. [Google Scholar] [CrossRef] [Green Version]
  143. Koka, V.; Huang, X.R.; Chung, A.C.; Wang, W.; Truong, L.D.; Lan, H.Y. Angiotensin II up-regulates angiotensin I-converting enzyme (ACE), but down-regulates ACE2 via the AT1-ERK/p38 MAP kinase pathway. Am. J. Pathol. 2008, 172, 1174–1183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Gallagher, P.E.; Ferrario, C.M.; Tallant, E.A. MAP kinase/phosphatase pathway mediates the regulation of ACE2 by angiotensin peptides. Am. J. Physiol. Cell Physiol. 2008, 295, C1169–C1174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Kumar, S.A.; Wei, H. Comparative docking studies to understand the binding affinity of nicotine with soluble ACE2 (sACE2)-SARS-CoV-2 complex over sACE2. Toxicol. Rep. 2020, 7, 1366–1372. [Google Scholar]
  146. Eguchi, S.; Kawai, T.; Scalia, R.; Rizzo, V. Understanding angiotensin II type 1 receptor signaling in vascular pathophysiology. Hypertension 2018, 71, 804–810. [Google Scholar] [CrossRef]
  147. Patel, V.B.; Clarke, N.; Wang, Z.; Fan, D.; Parajuli, N.; Basu, R.; Putko, B.; Kassiri, Z.; Turner, A.J.; Oudit, G.Y. Angiotensin II induced proteolytic cleavage of myocardial ACE2 is mediated by TACE/ADAM-17: A positive feedback mechanism in the RAS. J. Mol. Cell. Cardiol. 2014, 66, 167–176. [Google Scholar] [CrossRef]
  148. Chu, K.Y.; Leung, P.S. Angiotensin II in type 2 diabetes mellitus. Curr. Protein Pept. Sci. 2009, 10, 75–84. [Google Scholar] [CrossRef]
  149. Ferder, L.; Inserra, F.; Martínez-Maldonado, M. Inflammation and the metabolic syndrome: Role of angiotensin II and oxidative stress. Curr. Hypertens. Rep. 2006, 8, 191–198. [Google Scholar] [CrossRef]
  150. Ortiz-Pérez, J.T.; Riera, M.; Bosch, X.; De Caralt, T.M.; Perea, R.J.; Pascual, J.; Soler, M.J. Role of circulating angiotensin converting enzyme 2 in left ventricular remodeling following myocardial infarction: A prospective controlled study. PLoS ONE 2013, 8, e61695. [Google Scholar] [CrossRef] [Green Version]
  151. Hayashida, K.; Bartlett, A.H.; Chen, Y.; Park, P.W. Molecular and cellular mechanisms of ectodomain shedding. Anat. Rec. Adv. Integr. Anat. Evol. Biol. 2010, 293, 925–937. [Google Scholar] [CrossRef] [Green Version]
  152. Marczynska, J.; Ozga, A.; Wlodarczyk, A.; Majchrzak-Gorecka, M.; Kulig, P.; Banas, M.; Michalczyk-Wetula, D.; Majewski, P.; Hutloff, A.; Schwarz, J. The role of metalloproteinase ADAM17 in regulating ICOS Ligand–mediated humoral immune responses. J. Immunol. 2014, 193, 2753–2763. [Google Scholar] [CrossRef] [Green Version]
  153. Grötzinger, J.; Lorenzen, I.; Düsterhöft, S. Molecular insights into the multilayered regulation of ADAM17: The role of the extracellular region. Biochim. Biophys. Acta Mol. Cell Res. 2017, 1864, 2088–2095. [Google Scholar] [CrossRef]
  154. Scheller, J.; Chalaris, A.; Garbers, C.; Rose-John, S. ADAM17: A molecular switch to control inflammation and tissue regeneration. Trends Immunol. 2011, 32, 380–387. [Google Scholar] [CrossRef]
  155. Dreymueller, D.; Pruessmeyer, J.; Groth, E.; Ludwig, A. The role of ADAM-mediated shedding in vascular biology. Eur. J. Cell Biol. 2012, 91, 472–485. [Google Scholar] [CrossRef]
  156. De Queiroz, T.M.; Lakkappa, N.; Lazartigues, E. ADAM17-mediated shedding of inflammatory cytokines in hypertension. Front. Pharmacol. 2020, 11, 1154. [Google Scholar] [CrossRef]
  157. Schlöndorff, J.; Becherer, J.D.; Blobel, C.P. Intracellular maturation and localization of the tumour necrosis factor α convertase (TACE). Biochem. J. 2000, 347, 131–138. [Google Scholar] [CrossRef]
  158. Zunke, F.; Rose-John, S. The shedding protease ADAM17: Physiology and pathophysiology. BBA Mol. Cell Res. 2017, 1864, 2059–2070. [Google Scholar] [CrossRef]
  159. Lorenzen, I.; Lokau, J.; Düsterhöft, S.; Trad, A.; Garbers, C.; Scheller, J.; Rose-John, S.; Grötzinger, J. The membrane-proximal domain of A Disintegrin and Metalloprotease 17 (ADAM17) is responsible for recognition of the interleukin-6 receptor and interleukin-1 receptor II. FEBS Lett. 2012, 586, 1093–1100. [Google Scholar] [CrossRef]
  160. Weskamp, G.; Tüshaus, J.; Li, D.; Feederle, R.; Maretzky, T.; Swendemann, S.; Falck-Pedersen, E.; McIlwain, D.R.; Mak, T.W.; Salmon, J.E. ADAM17 stabilizes its interacting partner inactive Rhomboid 2 (iRhom2) but not inactive Rhomboid 1 (iRhom1). J. Biol. Chem. 2020, 295, 4350–4358. [Google Scholar] [CrossRef] [Green Version]
  161. Al-Salihi, M.A.; Lang, P.A. iRhom2: An emerging adaptor regulating immunity and disease. Int. J. Mol. Sci. 2020, 21, 6570. [Google Scholar] [CrossRef]
  162. Saad, M.I.; Alhayyani, S.; McLeod, L.; Yu, L.; Alanazi, M.; Deswaerte, V.; Tang, K.; Jarde, T.; Smith, J.A.; Prodanovic, Z. ADAM 17 selectively activates the IL-6 trans-signaling/ERK MAPK axis in KRAS-addicted lung cancer. EMBO Mol. Med. 2019, 11, e9976. [Google Scholar] [CrossRef]
  163. Fan, D.; Kassiri, Z. Biology of tissue inhibitor of metalloproteinase 3 (TIMP3), and its therapeutic implications in cardiovascular pathology. Front. Physiol. 2020, 11, 661. [Google Scholar] [CrossRef] [PubMed]
  164. Ingram, R.N.; Orth, P.; Strickland, C.L.; Le, H.V.; Madison, V.; Beyer, B.M. Stabilization of the autoproteolysis of TNF-alpha converting enzyme (TACE) results in a novel crystal form suitable for structure-based drug design studies. Protein Eng. Des. Sel. 2006, 19, 155–161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Luo, W.-W.; Li, S.; Li, C.; Zheng, Z.-Q.; Cao, P.; Tong, Z.; Lian, H.; Wang, S.-Y.; Shu, H.-B.; Wang, Y.-Y. iRhom2 is essential for innate immunity to RNA virus by antagonizing ER-and mitochondria-associated degradation of VISA. PLoS Pathog. 2017, 13, e1006693. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Adrain, C.; Zettl, M.; Christova, Y.; Taylor, N.; Freeman, M. Tumor necrosis factor signaling requires iRhom2 to promote trafficking and activation of TACE. Science 2012, 335, 225–228. [Google Scholar] [CrossRef] [Green Version]
  167. Chao-Chu, J.; Murtough, S.; Zaman, N.; Pennington, D.J.; Blaydon, D.C.; Kelsell, D.P. iRHOM2: A regulator of palmoplantar biology, inflammation, and viral susceptibility. J. Investig. Dermatol. 2021, 141, 722–726. [Google Scholar] [CrossRef]
  168. Yoda, M.; Kimura, T.; Tohmonda, T.; Morioka, H.; Matsumoto, M.; Okada, Y.; Toyama, Y.; Horiuchi, K. Systemic overexpression of TNFα-converting enzyme does not lead to enhanced shedding activity in vivo. PLoS ONE 2013, 8, e54412. [Google Scholar] [CrossRef] [Green Version]
  169. Le Gall, S.M.; Maretzky, T.; Issuree, P.D.; Niu, X.-D.; Reiss, K.; Saftig, P.; Khokha, R.; Lundell, D.; Blobel, C.P. ADAM17 is regulated by a rapid and reversible mechanism that controls access to its catalytic site. J. Cell Sci. 2010, 123, 3913–3922. [Google Scholar] [CrossRef]
  170. Zeng, S.-y.; Yang, L.; Hong, C.-l.; Lu, H.-q.; Yan, Q.-j.; Chen, Y.; Qin, X.-p. Evidence That ADAM17 mediates the protective action of CGRP against angiotensin II-induced inflammation in vascular smooth muscle cells. Mediat. Inflamm. 2018, 2018, 2109352. [Google Scholar] [CrossRef]
  171. Guo, D.F.; Sun, Y.L.; Hamet, P.; Inagami, T. The angiotensin II type 1 receptor and receptor-associated proteins. Cell Res. 2001, 11, 165–180. [Google Scholar] [CrossRef] [Green Version]
  172. Nejat, R.; Sadr, A.S.; Torshizi, M.F.; Najafi, D.J. GPCRs of diverse physiologic and pathologic effects with fingerprints in COVID-19. Biol. Life Sci. Forum 2021, 7, 19. [Google Scholar]
  173. Fuller, A.J.; Hauschild, B.C.; Gonzalez-Villalobos, R.; Awayda, M.S.; Imig, J.D.; Inscho, E.W.; Navar, L.G. Calcium and chloride channel activation by angiotensin II-AT1 receptors in preglomerular vascular smooth muscle cells. Am. J. Physiol. Ren. Physiol. 2005, 289, F760–F767. [Google Scholar] [CrossRef] [Green Version]
  174. Vivar, R.; Soto, C.; Copaja, M.; Mateluna, F.; Aranguiz, P.; Munoz, J.P.; Chiong, M.; Garcia, L.; Letelier, A.; Thomas, W.G. Phospholipase C/protein kinase C pathway mediates angiotensin II-dependent apoptosis in neonatal rat cardiac fibroblasts expressing AT1 receptor. J. Cardiovasc. Pharmacol. 2008, 52, 184–190. [Google Scholar] [CrossRef]
  175. Fischer, O.M.; Hart, S.; Gschwind, A.; Prenzel, N.; Ullrich, A. Oxidative and osmotic stress signaling in tumor cells is mediated by ADAM proteases and heparin-binding epidermal growth factor. Mol. Cell. Biol. 2004, 24, 5172–5183. [Google Scholar] [CrossRef] [Green Version]
  176. Brill, A.; Chauhan, A.K.; Canault, M.; Walsh, M.T.; Bergmeier, W.; Wagner, D.D. Oxidative stress activates ADAM17/TACE and induces its target receptor shedding in platelets in a p38-dependent fashion. Cardiovasc. Res. 2009, 84, 137–144. [Google Scholar] [CrossRef] [Green Version]
  177. Ford, B.M.; Eid, A.A.; Göőz, M.; Barnes, J.L.; Gorin, Y.C.; Abboud, H.E. ADAM17 mediates Nox4 expression and NADPH oxidase activity in the kidney cortex of OVE26 mice. Am. J. Physiol. Ren. Physiol. 2013, 305, F323–F332. [Google Scholar] [CrossRef] [Green Version]
  178. Chatterjee, S. Oxidative stress, inflammation, and disease. In Oxidative Stress and Biomaterials; Elsevier: Amsterdam, The Netherlands, 2016; pp. 35–58. [Google Scholar]
  179. Forcados, G.E.; Muhammad, A.; Oladipo, O.O.; Makama, S.; Meseko, C.A. Metabolic implications of oxidative stress and inflammatory process in SARS-CoV-2 pathogenesis: Therapeutic potential of natural antioxidants. Front. Cell. Infect. Microbiol. 2021, 11, 654813. [Google Scholar] [CrossRef]
  180. De Costa, R.A.; Granato, D.C.; Trino, L.D.; Yokoo, S.; Carnielli, C.M.; Kawahara, R.; Domingues, R.R.; Pauletti, B.A.; Neves, L.X.; Santana, A.G. ADAM17 cytoplasmic domain modulates Thioredoxin-1 conformation and activity. Redox Biol. 2020, 37, 101735. [Google Scholar] [CrossRef]
  181. Granato, D.C.; e Costa, R.A.; Kawahara, R.; Yokoo, S.; Aragao, A.Z.; Domingues, R.R.; Pauletti, B.A.; Honorato, R.V.; Fattori, J.; Figueira, A.C.M. Thioredoxin-1 negatively modulates ADAM17 activity through direct binding and indirect reductive activity. Antioxid. Redox Signal. 2018, 29, 717–734. [Google Scholar] [CrossRef]
  182. Fang, Q.; Liu, X.; Abe, S.; Kobayashi, T.; Wang, X.; Kohyama, T.; Hashimoto, M.; Wyatt, T.; Rennard, S. Thrombin induces collagen gel contraction partially through PAR1 activation and PKC-ε. Eur. Respir. J. 2004, 24, 918–924. [Google Scholar] [CrossRef]
  183. Yang, C.-C.; Hsiao, L.-D.; Shih, Y.-F.; Hsu, C.-K.; Hu, C.-Y.; Yang, C.-M. Thrombin induces COX-2 and PGE2 expression via PAR1/PKCalpha/MAPK-dependent NF-kappaB activation in human tracheal smooth muscle cells. Mediat. Inflamm. 2022, 2022, 4600029. [Google Scholar] [CrossRef]
  184. He, R.-Q.; Tang, X.-F.; Zhang, B.-L.; Li, X.-D.; Hong, M.-N.; Chen, Q.-Z.; Han, W.-Q.; Gao, P.-J. Protease-activated receptor 1 and 2 contribute to angiotensin II-induced activation of adventitial fibroblasts from rat aorta. Biochem. Biophys. Res. Commun. 2016, 473, 517–523. [Google Scholar] [CrossRef] [PubMed]
  185. Menghini, R.; Fiorentino, L.; Casagrande, V.; Lauro, R.; Federici, M. The role of ADAM17 in metabolic inflammation. Atherosclerosis 2013, 228, 12–17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Zamel, I.A.; Palakkott, A.; Ashraf, A.; Iratni, R.; Ayoub, M.A. Interplay between angiotensin II type 1 receptor and thrombin receptor revealed by bioluminescence resonance energy transfer assay. Front. Pharmacol. 2020, 11, 1283. [Google Scholar] [CrossRef]
  187. Hall, K.C.; Blobel, C.P. Interleukin-1 stimulates ADAM17 through a mechanism independent of its cytoplasmic domain or phosphorylation at threonine 735. PLoS ONE 2012, 7, e31600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Pearl, M.H.; Grotts, J.; Rossetti, M.; Zhang, Q.; Gjertson, D.W.; Weng, P.; Elashoff, D.; Reed, E.F.; Chambers, E.T. Cytokine profiles associated with angiotensin II type 1 receptor antibodies. Kidney Int. Rep. 2019, 4, 541–550. [Google Scholar] [CrossRef] [Green Version]
  189. Cardoso, V.G.; Gonçalves, G.L.; Costa-Pessoa, J.M.; Thieme, K.; Lins, B.B.; Casare, F.A.M.; de Ponte, M.C.; Camara, N.O.S.; Oliveira-Souza, M. Angiotensin II-induced podocyte apoptosis is mediated by endoplasmic reticulum stress/PKC-δ/p38 MAPK pathway activation and trough increased Na+/H+ exchanger isoform 1 activity. BMC Nephrol. 2018, 19, 179–191. [Google Scholar] [CrossRef] [Green Version]
  190. Zhu, G.; Liu, J.; Wang, Y.; Jia, N.; Wang, W.; Zhang, J.; Liang, Y.; Tian, H.; Zhang, J. ADAM17 mediates hypoxia-induced keratinocyte migration via the p38/MAPK pathway. BioMed Res. Int. 2021, 2021, 8328216. [Google Scholar] [CrossRef]
  191. Xu, P.; Derynck, R. Direct activation of TACE-mediated ectodomain shedding by p38 MAP kinase regulates EGF receptor-dependent cell proliferation. Mol. Cell 2010, 37, 551–566. [Google Scholar] [CrossRef]
  192. Horiuchi, K.; Le Gall, S.; Schulte, M.; Yamaguchi, T.; Reiss, K.; Murphy, G.; Toyama, Y.; Hartmann, D.; Saftig, P.; Blobel, C.P. Substrate selectivity of epidermal growth factor-receptor ligand sheddases and their regulation by phorbol esters and calcium influx. Mol. Biol. Cell 2007, 18, 176–188. [Google Scholar] [CrossRef] [Green Version]
  193. Kadurin, I.; Dahimene, S.; Page, K.M.; Ellaway, J.I.; Chaggar, K.; Troeberg, L.; Nagase, H.; Dolphin, A.C. ADAM17 mediates proteolytic maturation of voltage-gated calcium channel auxiliary α2δ subunits, and enables calcium current enhancement. Function 2022, 3, zqac013. [Google Scholar] [CrossRef]
  194. Lambert, D.W.; Yarski, M.; Warner, F.J.; Thornhill, P.; Parkin, E.T.; Smith, A.I.; Hooper, N.M.; Turner, A.J. Tumor necrosis factor-α convertase (ADAM17) mediates regulated ectodomain shedding of the severe-acute respiratory syndrome-coronavirus (SARS-CoV) receptor, angiotensin-converting enzyme-2 (ACE2). J. Biol. Chem. 2005, 280, 30113–30119. [Google Scholar] [CrossRef] [Green Version]
  195. Cheng, J.; Xue, F.; Cheng, C.; Sui, W.; Zhang, M.; Qiao, L.; Ma, J.; Ji, X.; Chen, W.; Yu, X. ADAM17 knockdown mitigates while ADAM17 overexpression aggravates cardiac fibrosis and dysfunction via regulating ACE2 shedding and myofibroblast transformation. Front. Pharmacol. 2022, 13, 997916. [Google Scholar] [CrossRef]
  196. Kumar, R.; Khandelwal, N.; Thachamvally, R.; Tripathi, B.N.; Barua, S.; Kashyap, S.K.; Maherchandani, S.; Kumar, N. Role of MAPK/MNK1 signaling in virus replication. Virus Res. 2018, 253, 48–61. [Google Scholar] [CrossRef]
  197. Pleschka, S. RNA viruses and the mitogenic Raf/MEK/ERK signal transduction cascade. Biol. Chem. 2008, 389, 1273–1282. [Google Scholar] [CrossRef]
  198. Huang, C.; Feng, F.; Shi, Y.; Li, W.; Wang, Z.; Zhu, Y.; Yuan, S.; Hu, D.; Dai, J.; Jiang, Q. Protein Kinase C inhibitors reduce SARS-CoV-2 replication in cultured cells. Microbiol. Spectr. 2022, 10, e0105622. [Google Scholar] [CrossRef]
  199. Liu, M.; Yang, Y.; Gu, C.; Yue, Y.; Wu, K.K.; Wu, J.; Zhu, Y. Spike protein of SARS-CoV stimulates cyclooxygenase-2 expression via both calcium-dependent and calcium-independent protein kinase C pathways. FASEB J. 2007, 21, 1586–1596. [Google Scholar] [CrossRef] [Green Version]
  200. Wieczfinska, J.; Kleniewska, P.; Pawliczak, R. Oxidative stress-related mechanisms in SARS-CoV-2 infections. Oxidative Med. Cell. Longev. 2022, 2022, 5589089. [Google Scholar] [CrossRef]
  201. Suhail, S.; Zajac, J.; Fossum, C.; Lowater, H.; McCracken, C.; Severson, N.; Laatsch, B.; Narkiewicz-Jodko, A.; Johnson, B.; Liebau, J. Role of oxidative stress on SARS-CoV (SARS) and SARS-CoV-2 (COVID-19) infection: A review. Protein J. 2020, 39, 644–656. [Google Scholar] [CrossRef]
  202. Berlansky, S.; Sallinger, M.; Grabmayr, H.; Humer, C.; Bernhard, A.; Fahrner, M.; Frischauf, I. Calcium signals during SARS-CoV-2 infection: Assessing the potential of emerging therapies. Cells 2022, 11, 253. [Google Scholar] [CrossRef]
  203. Vollbracht, C.; Kraft, K. Oxidative stress and hyper-inflammation as major drivers of severe COVID-19 and long COVID: Implications for the benefit of high-dose intravenous vitamin C. Front. Pharmacol. 2022, 13, 899198. [Google Scholar] [CrossRef]
  204. Aiello, E.; Cingolani, H. Angiotensin II stimulates cardiac L-type Ca2+ current by a Ca2+-and protein kinase C-dependent mechanism. Am. J. Physiol. Heart Circ. Physiol. 2001, 280, H1528–H1536. [Google Scholar] [CrossRef] [PubMed]
  205. Verma, K.; Pant, M.; Paliwal, S.; Dwivedi, J.; Sharma, S. An insight on multicentric signaling of angiotensin II in cardiovascular system: A Recent Update. Front. Pharmacol. 2021, 12, 734917. [Google Scholar] [CrossRef] [PubMed]
  206. Palacios, Y.; Ruiz, A.; Ramon-Luing, L.A.; Ocaña-Guzman, R.; Barreto-Rodriguez, O.; Sanchez-Moncivais, A.; Tecuatzi-Cadena, B.; Regalado-García, A.G.; Pineda-Gudiño, R.D.; García-Martínez, A. Severe COVID-19 patients show an increase in soluble TNFR1 and ADAM17, with a relationship to mortality. Int. J. Mol. Sci. 2021, 22, 8423. [Google Scholar] [CrossRef] [PubMed]
  207. Vieira, C.; Nery, L.; Martins, L.; Jabour, L.; Dias, R.; Simões e Silva, A.C. Downregulation of membrane-bound angiotensin converting enzyme 2 (ACE2) receptor has a pivotal role in COVID-19 immunopathology. Curr. Drug Targets 2021, 22, 254–281. [Google Scholar] [CrossRef] [PubMed]
  208. Ramos, S.G.; da Cruz Rattis, B.A.; Ottaviani, G.; Celes, M.R.N.; Dias, E.P. ACE2 down-regulation may act as a transient molecular disease causing RAAS dysregulation and tissue damage in the microcirculatory environment among COVID-19 patients. Am. J. Pathol. 2021, 191, 1154–1164. [Google Scholar] [CrossRef]
  209. Yalcin, H.C.; Sukumaran, V.; Al-Ruweidi, M.K.A.; Shurbaji, S. Do changes in ace-2 expression affect SARS-CoV-2 virulence and related complications: A closer look into membrane-bound and soluble forms. Int. J. Mol. Sci. 2021, 22, 6703. [Google Scholar] [CrossRef]
  210. Tseng, Y.H.; Yang, R.C.; Lu, T.S. Two hits to the renin-angiotensin system may play a key role in severe COVID-19. Kaohsiung J. Med. Sci. 2020, 36, 389–392. [Google Scholar] [CrossRef]
  211. Tang, Q.; Wang, Y.; Ou, L.; Li, J.; Zheng, K.; Zhan, H.; Gu, J.; Zhou, G.; Xie, S.; Zhang, J. Downregulation of ACE2 expression by SARS-CoV-2 worsens the prognosis of KIRC and KIRP patients via metabolism and immunoregulation. Int. J. Biol. Sci. 2021, 17, 1925. [Google Scholar] [CrossRef]
  212. Bourgonje, A.R.; Abdulle, A.E.; Timens, W.; Hillebrands, J.L.; Navis, G.J.; Gordijn, S.J.; Bolling, M.C.; Dijkstra, G.; Voors, A.A.; Osterhaus, A.D. Angiotensin-converting enzyme 2 (ACE2), SARS-CoV-2 and the pathophysiology of coronavirus disease 2019 (COVID-19). J. Pathol. 2020, 251, 228–248. [Google Scholar] [CrossRef]
  213. Haga, S.; Nagata, N.; Okamura, T.; Yamamoto, N.; Sata, T.; Yamamoto, N.; Sasazuki, T.; Ishizaka, Y. TACE antagonists blocking ACE2 shedding caused by the spike protein of SARS-CoV are candidate antiviral compounds. Antivir. Res. 2010, 85, 551–555. [Google Scholar] [CrossRef]
  214. Kornilov, S.A.; Lucas, I.; Jade, K.; Dai, C.L.; Lovejoy, J.C.; Magis, A.T. Plasma levels of soluble ACE2are associated with sex, Metabolic Syndrome, and its biomarkers in a large cohort, pointing to a possible mechanism for increased severity in COVID-19. Crit. Care 2020, 24, 1–3. [Google Scholar] [CrossRef]
  215. Jia, H.P.; Look, D.C.; Tan, P.; Shi, L.; Hickey, M.; Gakhar, L.; Chappell, M.C.; Wohlford-Lenane, C.; McCray, P.B., Jr. Ectodomain shedding of angiotensin converting enzyme 2 in human airway epithelia. Am. J. Physiol. Lung Cell. Mol. Physiol. 2009, 297, L84–L96. [Google Scholar] [CrossRef] [Green Version]
  216. Wang, J.; Zhao, H.; An, Y. ACE2 Shedding and the Role in COVID-19. Front. Cell. Infect. Microbiol. 2022, 11, 789180. [Google Scholar] [CrossRef]
  217. Paul, M.; Poyan Mehr, A.; Kreutz, R. Physiology of local renin-angiotensin systems. Physiol. Rev. 2006, 86, 747–803. [Google Scholar] [CrossRef]
  218. Bitker, L.; Burrell, L.M. Classic and nonclassic renin-angiotensin systems in the critically ill. Crit. Care Clin. 2019, 35, 213–227. [Google Scholar] [CrossRef]
  219. Tepasse, P.-R.; Vollenberg, R.; Steinebrey, N.; König, S. The Dysregulation of the Renin–Angiotensin System in COVID-19 Studied by Serum Proteomics: Angiotensinogen increases with disease severity. Molecules 2022, 27, 2495. [Google Scholar] [CrossRef]
  220. Camargo, R.L.; Bombassaro, B.; Monfort-Pires, M.; Mansour, E.; Palma, A.C.; Ribeiro, L.C.; Ulaf, R.G.; Bernardes, A.F.; Nunes, T.A.; Agrela, M.V. Plasma Angiotensin II is increased in critical coronavirus disease 2019. Front. Cardiovasc. Med. 2022, 145, 112420. [Google Scholar] [CrossRef]
  221. Wu, Z.; Hu, R.; Zhang, C.; Ren, W.; Yu, A.; Zhou, X. Elevation of plasma angiotensin II level is a potential pathogenesis for the critically ill COVID-19 patients. Crit. Care 2020, 24, 1–3. [Google Scholar] [CrossRef]
  222. Liu, Y.; Yang, Y.; Zhang, C.; Huang, F.; Wang, F.; Yuan, J.; Wang, Z.; Li, J.; Li, J.; Feng, C. Clinical and biochemical indexes from 2019-nCoV infected patients linked to viral loads and lung injury. Sci. China Life Sci. 2020, 63, 364–374. [Google Scholar] [CrossRef] [Green Version]
  223. Osman, I.O.; Melenotte, C.; Brouqui, P.; Million, M.; Lagier, J.-C.; Parola, P.; Stein, A.; La Scola, B.; Meddeb, L.; Mege, J.-L. Expression of ACE2, soluble ACE2, angiotensin I, angiotensin II and angiotensin-(1-7) is modulated in COVID-19 patients. Front. Immunol. 2021, 12, 2350. [Google Scholar] [CrossRef]
  224. van Lier, D.; Kox, M.; Santos, K.; van der Hoeven, H.; Pillay, J.; Pickkers, P. Increased blood angiotensin converting enzyme 2 activity in critically ill COVID-19 patients. ERJ Open Res. 2021, 7, 00848–02020. [Google Scholar] [CrossRef] [PubMed]
  225. Elrayess, M.; Zedan, H.T.; Alattar, R.A.; Abusriwil, H.; Al-Ruweidi, M.A.A.; Almuraikhy, S.; Parengal, J.; Alhariri, B.; Yassine, H.S.; Hssain, A.A.; et al. Soluble ACE2 and angiotensin II levels are modulated in hypertensive COVID-19 patients treated with different antihypertension drugs. Blood Press. 2022, 31, 80–90. [Google Scholar] [CrossRef] [PubMed]
  226. Kragstrup, T.W.; Singh, H.S.; Grundberg, I.; Nielsen, A.L.-L.; Rivellese, F.; Mehta, A.; Goldberg, M.B.; Filbin, M.R.; Qvist, P.; Bibby, B.M. Plasma ACE2 predicts outcome of COVID-19 in hospitalized patients. PLoS ONE 2021, 16, e0252799. [Google Scholar] [CrossRef] [PubMed]
  227. Mariappan, V.; Ranganadin, P.; Shanmugam, L.; Rao, S.; Pillai, A.B. Early shedding of membrane-bounded ACE2 could be an indicator for disease severity in SARS-CoV-2. Biochimie 2022, 201, 139–147. [Google Scholar] [CrossRef]
  228. Hedges, J.F.; Snyder, D.T.; Robison, A.; Grifka-Walk, H.M.; Blackwell, K.; Shepardson, K.; Kominsky, D.; Rynda-Apple, A.; Walcheck, B.; Jutila, M.A. An ADAM17-neutralizing antibody reduces inflammation and mortality while increasing viral burden in a COVID-19 mouse model. Front. Immunol. 2022, 13, 918881. [Google Scholar] [CrossRef]
  229. Patel, S.K.; Juno, J.A.; Lee, W.S.; Wragg, K.M.; Hogarth, P.M.; Kent, S.J.; Burrell, L.M. Plasma ACE2 activity is persistently elevated following SARS-CoV-2 infection: Implications for COVID-19 pathogenesis and consequences. Eur. Respir. J. 2021, 57, 2003730. [Google Scholar] [CrossRef]
  230. Lundström, A.; Ziegler, L.; Havervall, S.; Rudberg, A.S.; Von Meijenfeldt, F.; Lisman, T.; Mackman, N.; Sandén, P.; Thålin, C. Soluble angiotensin-converting enzyme 2 is transiently elevated in COVID-19 and correlates with specific inflammatory and endothelial markers. J. Med. Virol. 2021, 93, 5908–5916. [Google Scholar] [CrossRef]
  231. Said, E.A.; Al-Reesi, I.; Al-Shizawi, N.; Jaju, S.; Al-Balushi, M.S.; Koh, C.Y.; Al-Jabri, A.A.; Jeyaseelan, L. Defining IL-6 levels in healthy individuals: A meta-analysis. J. Med. Virol. 2021, 93, 3915–3924. [Google Scholar] [CrossRef]
  232. Yan, I.; Schwarz, J.; Lücke, K.; Schumacher, N.; Schumacher, V.; Schmidt, S.; Rabe, B.; Saftig, P.; Donners, M.; Rose-John, S. ADAM17 controls IL-6 signaling by cleavage of the murine IL-6Rα from the cell surface of leukocytes during inflammatory responses. J. Leukoc. Biol. 2016, 99, 749–760. [Google Scholar] [CrossRef]
  233. Pedersen, K.B.; Chodavarapu, H.; Porretta, C.; Robinson, L.K.; Lazartigues, E. Dynamics of ADAM17-mediated shedding of ACE2 applied to pancreatic islets of male db/db mice. Endocrinology 2015, 156, 4411–4425. [Google Scholar] [CrossRef] [Green Version]
  234. Contini, C.; Caselli, E.; Martini, F.; Maritati, M.; Torreggiani, E.; Seraceni, S.; Vesce, F.; Perri, P.; Rizzo, L.; Tognon, M. COVID-19 is a multifaceted challenging pandemic which needs urgent public health interventions. Microorganisms 2020, 8, 1228. [Google Scholar] [CrossRef]
  235. Quinaglia, T.; Shabani, M.; Breder, I.; Silber, H.A.; Lima, J.A.; Sposito, A.C. Coronavirus disease-19: The multi-level, multi-faceted vasculopathy. Atherosclerosis 2021, 322, 39–50. [Google Scholar] [CrossRef]
  236. Daniell, H.; Nair, S.K.; Shi, Y.; Wang, P.; Montone, K.T.; Shaw, P.A.; Choi, G.H.; Ghani, D.; Weaver, J.; Rader, D.J. Decrease in Angiotensin-Converting Enzyme activity but not concentration in plasma/lungs in COVID-19 patients offers clues for diagnosis/treatment. Mol. Ther. Methods Clin. Dev. 2022, 26, 266–278. [Google Scholar] [CrossRef]
  237. Carrera, E.; Linares, A.; Joachim, A.; Jean-Baptiste, M.; Speth, R. MCA-APK (Dnp) is not a selective substrate of angiotensin-converting enzyme-2 (1067.6). FASEB J. 2014, 28, 1067. [Google Scholar] [CrossRef]
  238. Elased, K.M.; Cunha, T.S.; Gurley, S.B.; Coffman, T.M.; Morris, M. New mass spectrometric assay for angiotensin-converting enzyme 2 activity. Hypertension 2006, 47, 1010–1017. [Google Scholar] [CrossRef] [Green Version]
  239. Chappell, M.C.; Pirro, N.T.; Sykes, A.; Ferrario, C.M. Metabolism of angiotensin-(1–7) by angiotensin-converting enzyme. Hypertension 1998, 31, 362–367. [Google Scholar] [CrossRef] [Green Version]
  240. Chappell, M.C. The angiotensin-(1-7) axis: Formation and metabolism pathways. In Angiotensin-(1-7): A Comprehensive Review; Dos Santos, R.A.S., Ed.; Springer: Cham, Switzerland, 2019; pp. 1–26. [Google Scholar]
  241. Chappell, M.C.; Pirro, N.T.; South, A.M.; Gwathmey, T.M. Concerns on the Specificity of Commercial ELISAs for the Measurement of Angiotensin (1–7) and Angiotensin II in Human Plasma. Hypertension 2021, 77, e29–e31. [Google Scholar] [CrossRef]
  242. Maiti, B.K. Bioengineered angiotensin-converting-enzyme-2: A potential therapeutic option against SARS-CoV-2 infection. J. Hum. Hypertens. 2022, 36, 488–492. [Google Scholar] [CrossRef]
  243. Kiseleva, A.A.; Troisi, E.M.; Hensley, S.E.; Kohli, R.M.; Epstein, J.A. SARS-CoV-2 spike protein binding selectively accelerates substrate-specific catalytic activity of ACE2. J. Biochem. 2021, 170, 299–306. [Google Scholar] [CrossRef]
  244. Pedersen, K.B.; Sriramula, S.; Chhabra, K.H.; Xia, H.; Lazartigues, E. Species-specific inhibitor sensitivity of angiotensin-converting enzyme 2 (ACE2) and its implication for ACE2 activity assays. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2011, 301, R1293–R1299. [Google Scholar] [CrossRef]
  245. Oz, M.; Lorke, D.E. Multifunctional angiotensin converting enzyme 2, the SARS-CoV-2 entry receptor, and critical appraisal of its role in acute lung injury. Biomed. Pharmacother. 2021, 136, 111193. [Google Scholar] [CrossRef] [PubMed]
  246. Sodhi, C.P.; Wohlford-Lenane, C.; Yamaguchi, Y.; Prindle, T.; Fulton, W.B.; Wang, S.; McCray, P.B., Jr.; Chappell, M.; Hackam, D.J.; Jia, H. Attenuation of pulmonary ACE2 activity impairs inactivation of des-Arg9 bradykinin/BKB1R axis and facilitates LPS-induced neutrophil infiltration. Am. J. Physiol. Lung Cell. Mol. Physiol. 2018, 314, L17–L31. [Google Scholar] [CrossRef] [PubMed]
  247. Yan, R.; Zhang, Y.; Li, Y.; Xia, L.; Guo, Y.; Zhou, Q. Structural basis for the recognition of SARS-CoV-2 by full-length human ACE2. Science 2020, 367, 1444–1448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Chen, Y.; Sun, L.; Ullah, I.; Beaudoin-Bussières, G.; Anand, S.P.; Hederman, A.P.; Tolbert, W.D.; Sherburn, R.; Nguyen, D.N.; Marchitto, L. Engineered ACE2-Fc counters murine lethal SARS-CoV-2 infection through direct neutralization and Fc-effector activities. Sci. Adv. 2022, 8, eabn4188. [Google Scholar] [CrossRef] [PubMed]
  249. Anand, S.P.; Chen, Y.; Prévost, J.; Gasser, R.; Beaudoin-Bussières, G.; Abrams, C.F.; Pazgier, M.; Finzi, A. Interaction of human ACE2 to membrane-bound SARS-CoV-1 and SARS-CoV-2 S glycoproteins. Viruses 2020, 12, 1104. [Google Scholar] [CrossRef]
  250. Tai, W.; He, L.; Zhang, X.; Pu, J.; Voronin, D.; Jiang, S.; Zhou, Y.; Du, L. Characterization of the receptor-binding domain (RBD) of 2019 novel coronavirus: Implication for development of RBD protein as a viral attachment inhibitor and vaccine. Cell. Mol. Immunol. 2020, 17, 613–620. [Google Scholar] [CrossRef] [Green Version]
  251. Rahman, M.M.; Hasan, M.; Ahmed, A. Potential detrimental role of soluble ACE2 in severe COVID-19 comorbid patients. Rev. Med. Virol. 2021, 31, 1–12. [Google Scholar] [CrossRef]
  252. Kiani, A.; Dhuli, K.; Anpilogov, K.; Bressan, S.; Dautaj, A.; Dundar, M.; Beccari, T.; Ergoren, M.; Bertelli, M. Natural compounds as inhibitors of SARS-CoV-2 endocytosis: A promising approach against COVID-19. Acta Bio-medica: Atenei Parmensis 2020, 91, e2020008. [Google Scholar]
  253. Yeung, M.L.; Teng, J.L.L.; Jia, L.; Zhang, C.; Huang, C.; Cai, J.-P.; Zhou, R.; Chan, K.-H.; Zhao, H.; Zhu, L. Soluble ACE2-mediated cell entry of SARS-CoV-2 via interaction with proteins related to the renin-angiotensin system. Cell 2021, 184, 2212–2228.e2212. [Google Scholar] [CrossRef]
  254. Zoufaly, A.; Poglitsch, M.; Aberle, J.H.; Hoepler, W.; Seitz, T.; Traugott, M.; Grieb, A.; Pawelka, E.; Laferl, H.; Wenisch, C. Human recombinant soluble ACE2 in severe COVID-19. Lancet Respir. Med. 2020, 8, 1154–1158. [Google Scholar] [CrossRef]
  255. Cocozza, F.; Névo, N.; Piovesana, E.; Lahaye, X.; Buchrieser, J.; Schwartz, O.; Manel, N.; Tkach, M.; Théry, C.; Martin-Jaular, L. Extracellular vesicles containing ACE2 efficiently prevent infection by SARS-CoV-2 Spike protein-containing virus. J. Extracell. Vesicles 2020, 10, e12050. [Google Scholar] [CrossRef]
  256. Zhu, J.; Su, Y.; Tang, Y. Disrupting ACE2 dimerization mitigates the infection by SARS-CoV-2 Pseudovirus. Front. Virol. 2022, 2, 916700. [Google Scholar] [CrossRef]
  257. Xiao, T.; Lu, J.; Zhang, J.; Johnson, R.I.; McKay, L.G.; Storm, N.; Lavine, C.L.; Peng, H.; Cai, Y.; Rits-Volloch, S. A trimeric human angiotensin-converting enzyme 2 as an anti-SARS-CoV-2 agent. Nat. Struct. Mol. Biol. 2021, 28, 202–209. [Google Scholar] [CrossRef]
  258. Nesci, S. SARS-CoV-2 first contact: Spike–ACE2 interactions in COVID-19. Chem. Biol. Drug Des. 2021, 98, 207–211. [Google Scholar] [CrossRef]
  259. Chan, K.K.; Dorosky, D.; Sharma, P.; Abbasi, S.A.; Dye, J.M.; Kranz, D.M.; Herbert, A.S.; Procko, E. Engineering human ACE2 to optimize binding to the spike protein of SARS coronavirus 2. Science 2020, 369, 1261–1265. [Google Scholar] [CrossRef]
  260. Kayabolen, A.; Akcan, U.; Ozturan, D.; Ulbegi-Polat, H.; Sahin, G.; Degirmenci, N.; Bayraktar, C.; Soyler, G.; Sarayloo, E.; Nurtop, E. Protein scaffold-based multimerization of soluble ACE2 efficiently blocks SARS-CoV-2 infection in vitro and in vivo. Adv. Sci. 2021, 9, 2201294. [Google Scholar] [CrossRef]
  261. Raghuvamsi, P.V.; Tulsian, N.K.; Samsudin, F.; Qian, X.; Purushotorman, K.; Yue, G.; Kozma, M.M.; Hwa, W.Y.; Lescar, J.; Bond, P.J. SARS-CoV-2 S protein: ACE2 interaction reveals novel allosteric targets. Elife 2021, 10, e63646. [Google Scholar] [CrossRef]
  262. Xu, C.; Wang, Y.; Liu, C.; Zhang, C.; Han, W.; Hong, X.; Wang, Y.; Hong, Q.; Wang, S.; Zhao, Q. Conformational dynamics of SARS-CoV-2 trimeric spike glycoprotein in complex with receptor ACE2 revealed by cryo-EM. Sci. Adv. 2021, 7, eabe5575. [Google Scholar] [CrossRef]
  263. Feng, F.; Chen, J.; Zhao, J.; Li, Y.; Li, M.; Sun, C. Killing two birds with one stone by administration of soluble ACE2: A promising strategy to treat both cardiovascular diseases and SARS-CoV-2 infection. Viruses 2021, 13, 2243. [Google Scholar] [CrossRef]
  264. Basu, R.; Poglitsch, M.; Yogasundaram, H.; Thomas, J.; Rowe, B.H.; Oudit, G.Y. Roles of angiotensin peptides and recombinant human ACE2 in heart failure. J. Am. Coll. Cardiol. 2017, 69, 805–819. [Google Scholar] [CrossRef]
  265. Iwasaki, M.; Saito, J.; Zhao, H.; Sakamoto, A.; Hirota, K.; Ma, D. Inflammation triggered by SARS-CoV-2 and ACE2 augment drives multiple organ failure of severe COVID-19: Molecular mechanisms and implications. Inflammation 2021, 44, 13–34. [Google Scholar] [CrossRef] [PubMed]
  266. Rodrigues Prestes, T.R.; Rocha, N.P.; Miranda, A.S.; Teixeira, A.L.; Simoes-e-Silva, A.C. The anti-inflammatory potential of ACE2/angiotensin-(1-7)/Mas receptor axis: Evidence from basic and clinical research. Curr. Drug Targets 2017, 18, 1301–1313. [Google Scholar] [CrossRef] [PubMed]
  267. Haschke, M.; Schuster, M.; Poglitsch, M.; Loibner, H.; Salzberg, M.; Bruggisser, M.; Penninger, J.; Krähenbühl, S. Pharmacokinetics and pharmacodynamics of recombinant human angiotensin-converting enzyme 2 in healthy human subjects. Clin. Pharmacokinet. 2013, 52, 783–792. [Google Scholar] [CrossRef] [PubMed]
  268. Imai, Y.; Kuba, K.; Rao, S.; Huan, Y.; Guo, F.; Guan, B.; Yang, P.; Sarao, R.; Wada, T.; Leong-Poi, H. Angiotensin-converting enzyme 2 protects from severe acute lung failure. Nature 2005, 436, 112–116. [Google Scholar] [CrossRef]
  269. Gu, H.; Xie, Z.; Li, T.; Zhang, S.; Lai, C.; Zhu, P.; Wang, K.; Han, L.; Duan, Y.; Zhao, Z. Angiotensin-converting enzyme 2 inhibits lung injury induced by respiratory syncytial virus. Sci. Rep. 2016, 6, 1–10. [Google Scholar] [CrossRef] [Green Version]
  270. Zhang, H.; Baker, A. Recombinant human ACE2: Acing out angiotensin II in ARDS therapy. Crit. Care 2017, 21, 1–3. [Google Scholar] [CrossRef] [Green Version]
  271. Roshanravan, N.; Ghaffari, S.; Hedayati, M. Angiotensin converting enzyme-2 as therapeutic target in COVID-19. Diabetes Metab. Syndr. Clin. Res. Rev. 2020, 14, 637–639. [Google Scholar] [CrossRef]
  272. Basit, A.; Ali, T.; Rehman, S.U. Truncated human angiotensin converting enzyme 2; a potential inhibitor of SARS-CoV-2 spike glycoprotein and potent COVID-19 therapeutic agent. J. Biomol. Struct. Dyn. 2021, 39, 3605–3614. [Google Scholar] [CrossRef]
  273. Ferrari, M.; Mekkaoui, L.; Ilca, F.T.; Akbar, Z.; Bughda, R.; Lamb, K.; Ward, K.; Parekh, F.; Karattil, R.; Allen, C. Characterization of a novel ACE2-based therapeutic with enhanced rather than reduced activity against SARS-CoV-2 variants. J. Virol. 2021, 95, e0068521. [Google Scholar] [CrossRef]
  274. Vitiello, A.; Ferrara, F. Pharmacotherapy based on ACE2 targeting and COVID-19 infection. Int. J. Mol. Sci. 2022, 23, 6644. [Google Scholar] [CrossRef]
  275. Krishnamurthy, S.; Lockey, R.; Kolliputi, N. Soluble ACE2 as a potential therapy for COVID-19. Am. J. Physiol. Cell Physiol. 2021, 320, C279–C281. [Google Scholar] [CrossRef]
  276. Iwanaga, N.; Cooper, L.; Rong, L.; Maness, N.J.; Beddingfield, B.; Qin, Z.; Crabtree, J.; Tripp, R.A.; Yang, H.; Blair, R. ACE2-IgG1 fusions with improved in vitro and in vivo activity against SARS-CoV-2. Iscience 2022, 25, 103670. [Google Scholar] [CrossRef]
  277. Zhang, Z.; Zeng, E.; Zhang, L.; Wang, W.; Jin, Y.; Sun, J.; Huang, S.; Yin, W.; Dai, J.; Zhuang, Z. Potent prophylactic and therapeutic efficacy of recombinant human ACE2-Fc against SARS-CoV-2 infection in vivo. Cell Discov. 2021, 7, 1–15. [Google Scholar] [CrossRef]
  278. Zhang, L.; Zhang, Y.; Qin, X.; Jiang, X.; Zhang, J.; Mao, L.; Jiang, Z.; Jiang, Y.; Liu, G.; Qiu, J. Recombinant ACE2 protein protects against acute lung injury induced by SARS-CoV-2 spike RBD protein. Crit. Care 2022, 26, 172–193. [Google Scholar] [CrossRef]
  279. Nejat, R.; Sadr, A. Are losartan and imatinib effective against SARS-CoV2 pathogenesis? A pathophysiologic-based in silico study. In Silico Pharmacol. 2021, 9, 1–22. [Google Scholar] [CrossRef]
  280. Daniell, H.; Nair, S.K.; Esmaeili, N.; Wakade, G.; Shahid, N.; Ganesan, P.K.; Islam, M.R.; Shepley-McTaggart, A.; Feng, S.; Gary, E.N. Debulking SARS-CoV-2 in saliva using angiotensin converting enzyme 2 in chewing gum to decrease oral virus transmission and infection. Mol. Ther. 2022, 30, 1966–1978. [Google Scholar] [CrossRef]
  281. Malik, J.A.; Ahmed, S.; Mir, A.; Shinde, M.; Bender, O.; Alshammari, F.; Ansari, M.; Anwar, S. The SARS-CoV-2 mutations versus vaccine effectiveness: New opportunities to new challenges. J. Infect. Public Health 2022, 15, 228–240. [Google Scholar] [CrossRef]
  282. Thakur, S.; Sasi, S.; Pillai, S.; Nag, A.; Shukla, D.; Singhal, R.; Phalke, S.; Velu, G. SARS-CoV-2 mutations and their impact on diagnostics, therapeutics and vaccines. Front. Med. 2022, 9, 815389. [Google Scholar] [CrossRef]
  283. Mohammadzadeh, N.; Asl, N.S.M.; Forouharnejad, K.; Ghadimi, K.; Parsa, S.; Mohammadi, S.; Omidi, A. Mechanism and adverse effects of COVID-19 drugs: A basic review. Int. J. Physiol. Pathophysiol. Pharmacol. 2021, 13, 102. [Google Scholar]
  284. Khalatbari, A.; Aghazadeh, Z.; Ji, C. Adverse effects of anti-Covid-19 drug candidates and alcohol on cellular stress responses of hepatocytes. Hepatol. Commun. 2022, 6, 1262. [Google Scholar] [CrossRef]
  285. Liu, D.; Zeng, X.; Ding, Z.; Lv, F.; Mehta, J.; Wang, X. Adverse cardiovascular effects of anti-COVID-19 drugs. Front. Pharmacol. 2021, 12, 699949. [Google Scholar] [CrossRef]
  286. Izcovich, A.; Siemieniuk, R.A.; Bartoszko, J.J.; Ge, L.; Zeraatkar, D.; Kum, E.; Qasim, A.; Khamis, A.M.; Rochwerg, B.; Agoritsas, T. Adverse effects of remdesivir, hydroxychloroquine and lopinavir/ritonavir when used for COVID-19, systematic review and meta-analysis of randomised trials. BMJ Open 2022, 12, e048502. [Google Scholar] [CrossRef]
  287. Consortium, W.S.T. Remdesivir and three other drugs for hospitalised patients with COVID-19: Final results of the WHO Solidarity randomised trial and updated meta-analyses. Lancet 2022, 399, 1941–1953. [Google Scholar]
  288. Delano, W.L. PyMol: An open-source molecular graphics tool. CCP4 Newsl. Protein Crystallogr. 2002, 40, 82–92. [Google Scholar]
Figure 1. SARS-CoV-2 spike protein homotrimeric structure (C) (PDB ID: 7QUS [28]). Monomers (visualized as green, red and blue structures) are added one by one, from (A) to (C), to show how each monomer is folded and the dimeric as well as the trimeric structures are inter-twisted. It makes boundary identification of S1/S2 interface in a three-dimensional view difficult to show, yet has been depicted here by green curve arrows for simplification. The territory of S1/S2 cleavage site is also pointed (blue arrow). It is noticeable how in “closed, D” conformation (see the text, Section 3), RBD in the first monomer (A) is embedded in other S protein residues when the second (B) and third (C) monomers are added. This complex feature of S protein molecule leads RBDs to pack together as inaccessible domains in the “D” conformation (see the text below). NTD: N-terminal domain; RBD: receptor binding domain.
Figure 1. SARS-CoV-2 spike protein homotrimeric structure (C) (PDB ID: 7QUS [28]). Monomers (visualized as green, red and blue structures) are added one by one, from (A) to (C), to show how each monomer is folded and the dimeric as well as the trimeric structures are inter-twisted. It makes boundary identification of S1/S2 interface in a three-dimensional view difficult to show, yet has been depicted here by green curve arrows for simplification. The territory of S1/S2 cleavage site is also pointed (blue arrow). It is noticeable how in “closed, D” conformation (see the text, Section 3), RBD in the first monomer (A) is embedded in other S protein residues when the second (B) and third (C) monomers are added. This complex feature of S protein molecule leads RBDs to pack together as inaccessible domains in the “D” conformation (see the text below). NTD: N-terminal domain; RBD: receptor binding domain.
Vaccines 11 00204 g001
Figure 2. Linear representation of different domains of monomeric S protein and “open, U” and “closed, D” conformations of S protein. The defining of the boundary of S1 and S2 subdomains (arrows on the left side) is just a simplification (see Figure 1). Sequential opening of RBDs of the trimeric S protein after docking of the first open RBD to ACE2 is displayed. It results in an untwisting movement of the monomers away from each other after binding of each “U” RBD to ACE2 (see the text). Green helices at the top of S protein represents RBD-binding site of ACE2. NTD: N-terminal domain; RBD: receptor binding domain; RBM: receptor binding motif; SD1 and SD2: subdomains 1 and 2; FP: fusion peptide; HR1 and HR2: heptad repeat 1 and 2; TM: transmembrane domain; IC: cytoplasmic domain; DDD: three closed RBDs (PDB ID: 7UB5 [39], PyMol molecular graphics system); UDD: one open/two closed RBDs (PDB ID: 7KEB, [40], PyMol molecular graphics system); UUD: two open/one closed RBDs (PDB ID: 6X2B [41], PyMol graphics system); UUU: three open RBDs (PDB ID: 8CSA [42], PyMol molecular graphics system).
Figure 2. Linear representation of different domains of monomeric S protein and “open, U” and “closed, D” conformations of S protein. The defining of the boundary of S1 and S2 subdomains (arrows on the left side) is just a simplification (see Figure 1). Sequential opening of RBDs of the trimeric S protein after docking of the first open RBD to ACE2 is displayed. It results in an untwisting movement of the monomers away from each other after binding of each “U” RBD to ACE2 (see the text). Green helices at the top of S protein represents RBD-binding site of ACE2. NTD: N-terminal domain; RBD: receptor binding domain; RBM: receptor binding motif; SD1 and SD2: subdomains 1 and 2; FP: fusion peptide; HR1 and HR2: heptad repeat 1 and 2; TM: transmembrane domain; IC: cytoplasmic domain; DDD: three closed RBDs (PDB ID: 7UB5 [39], PyMol molecular graphics system); UDD: one open/two closed RBDs (PDB ID: 7KEB, [40], PyMol molecular graphics system); UUD: two open/one closed RBDs (PDB ID: 6X2B [41], PyMol graphics system); UUU: three open RBDs (PDB ID: 8CSA [42], PyMol molecular graphics system).
Vaccines 11 00204 g002
Figure 3. Linear representation of distinct domains and three dimensional shape of human TMPRSS2 (isoform 2) (PDB ID: 7MEQ [18], PyMol molecular graphics system).
Figure 3. Linear representation of distinct domains and three dimensional shape of human TMPRSS2 (isoform 2) (PDB ID: 7MEQ [18], PyMol molecular graphics system).
Vaccines 11 00204 g003
Figure 4. Three dimensional structure of the protease domain of ACE2 (PDB ID: 6M0J [24], PyMol molecular graphics system). The catalytic cleft, between subdomains I and II, involved in converting Ang II to angiotensin (1–7), is noticeable. The N-terminal helix in blue is the territory of binding of ACE2 with S protein. Residue 614 is the point of attachment of the catalytic domain on the way to C-terminal with other domains of ACE2; the neck domain, the transmembrane domain and the cytoplasmic tail.
Figure 4. Three dimensional structure of the protease domain of ACE2 (PDB ID: 6M0J [24], PyMol molecular graphics system). The catalytic cleft, between subdomains I and II, involved in converting Ang II to angiotensin (1–7), is noticeable. The N-terminal helix in blue is the territory of binding of ACE2 with S protein. Residue 614 is the point of attachment of the catalytic domain on the way to C-terminal with other domains of ACE2; the neck domain, the transmembrane domain and the cytoplasmic tail.
Vaccines 11 00204 g004
Figure 5. Linear representation of distinct domains, schematic and three dimensional shape of ADAM17 (PDB ID: 2DDF [164], PyMol molecular graphics system). Cys-rich: cysteine rich domain; MPD: membrane proximal domain; TMD: transmembrane domain.
Figure 5. Linear representation of distinct domains, schematic and three dimensional shape of ADAM17 (PDB ID: 2DDF [164], PyMol molecular graphics system). Cys-rich: cysteine rich domain; MPD: membrane proximal domain; TMD: transmembrane domain.
Vaccines 11 00204 g005
Figure 6. Three dimensional structure of open (catalytically active full length, greenish-blue) (PDB ID: 6M1D, [247], PyMol molecular graphics system) and closed (catalytically inactive, greenish-blue) (PDB ID: 7RPV, [248], PyMol molecular graphics system) forms of ACE2. As has been debated, complexification of RBD (light green) and ACE2 protease domain (the “closed” figure) makes the catalytic cleft shrink to some degree. RBD: receptor binding domain; TM: transmembrane domain.
Figure 6. Three dimensional structure of open (catalytically active full length, greenish-blue) (PDB ID: 6M1D, [247], PyMol molecular graphics system) and closed (catalytically inactive, greenish-blue) (PDB ID: 7RPV, [248], PyMol molecular graphics system) forms of ACE2. As has been debated, complexification of RBD (light green) and ACE2 protease domain (the “closed” figure) makes the catalytic cleft shrink to some degree. RBD: receptor binding domain; TM: transmembrane domain.
Vaccines 11 00204 g006
Figure 7. The three dimensional structure of an ACE2 homodimer molecule bound with two RBDs (yellow and orange) (PDB ID: 6M17, [247], PyMol molecular graphics system). ACE2 monomers are in red and greenish-blue colors. The “Neck” region is involved in homodierization of monomeric form of ACE2 molecules. PD: ACE2 peptidase domain, RBD: receptor biding domain of SARS-CoV-2 S protein, TMD: ACE2 trans-membrane domain.
Figure 7. The three dimensional structure of an ACE2 homodimer molecule bound with two RBDs (yellow and orange) (PDB ID: 6M17, [247], PyMol molecular graphics system). ACE2 monomers are in red and greenish-blue colors. The “Neck” region is involved in homodierization of monomeric form of ACE2 molecules. PD: ACE2 peptidase domain, RBD: receptor biding domain of SARS-CoV-2 S protein, TMD: ACE2 trans-membrane domain.
Vaccines 11 00204 g007
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

Nejat, R.; Torshizi, M.F.; Najafi, D.J. S Protein, ACE2 and Host Cell Proteases in SARS-CoV-2 Cell Entry and Infectivity; Is Soluble ACE2 a Two Blade Sword? A Narrative Review. Vaccines 2023, 11, 204. https://doi.org/10.3390/vaccines11020204

AMA Style

Nejat R, Torshizi MF, Najafi DJ. S Protein, ACE2 and Host Cell Proteases in SARS-CoV-2 Cell Entry and Infectivity; Is Soluble ACE2 a Two Blade Sword? A Narrative Review. Vaccines. 2023; 11(2):204. https://doi.org/10.3390/vaccines11020204

Chicago/Turabian Style

Nejat, Reza, Maziar Fayaz Torshizi, and David J. Najafi. 2023. "S Protein, ACE2 and Host Cell Proteases in SARS-CoV-2 Cell Entry and Infectivity; Is Soluble ACE2 a Two Blade Sword? A Narrative Review" Vaccines 11, no. 2: 204. https://doi.org/10.3390/vaccines11020204

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