Skip to main content
Advertisement
  • Loading metrics

The role of cysteine peptidases in coronavirus cell entry and replication: The therapeutic potential of cathepsin inhibitors

Abstract

Over the last 2 decades, several coronaviruses (CoVs) have crossed the species barrier into humans, causing highly prevalent and severe respiratory diseases, often with fatal outcomes. CoVs are a large group of enveloped, single-stranded, positive-sense RNA viruses, which encode large replicase polyproteins that are processed by viral peptidases to generate the nonstructural proteins (Nsps) that mediate viral RNA synthesis. Papain-like peptidases (PLPs) and chymotrypsin-like cysteine 3C-like peptidase are essential for coronaviral replication and represent attractive antiviral drug targets. Furthermore, CoVs utilize the activation of their envelope spike glycoproteins by host cell peptidases to gain entry into cells. CoVs have evolved multiple strategies for spike protein activation, including the utilization of lysosomal cysteine cathepsins. In this review, viral and host peptidases involved in CoV cell entry and replication are discussed in depth, with an emphasis on papain-like cysteine cathepsins. Furthermore, important findings on cysteine peptidase inhibitors with regard to virus attenuation are highlighted as well as the potential of such inhibitors for future treatment strategies for CoV-related diseases.

Introduction

Coronaviruses (CoVs) belong to the Coronaviridae family within the Nidovirales order and are pleomorphic, enveloped, positive-strand RNA viruses with unusually large genomes, containing 27 to 32 kilobases. According to phylogenetic analyses, CoVs are classified into 4 genera, i.e., alpha, beta, gamma, and delta CoVs, which are further subdivided into several lineages. Alpha and beta CoVs are known to cause respiratory or intestinal infections in humans and other mammals, while gamma and delta CoVs predominantly affect birds [1, 2]. The first discovered CoV, known as infectious bronchitis virus, was isolated from chicken embryos already in 1937 [3]. Later, in the 1960s, the first 2 strains of human coronaviruses (HCoVs), HCoV-229E [4] and HCoV-OC43 [5], were discovered. Although they were endemic to the human population, accounting for 10% to 30% of common colds, they were not given much importance as they caused only mild symptoms, and more severe diseases developed only in elderly, newborn, and immunocompromised individuals [68]. However, this view changed at the beginning of the 21st century with the realization that CoVs are maintained within an animal reservoir from which their transmission to humans via intermediate host species is possible [9].

Such a spillover occurred in November 2002 in Foshan, Guangdong Province, China, where patients displayed symptoms of atypical pneumonia, which later became known as Severe Acute Respiratory Syndrome (SARS) [10]. A decade later, a novel CoV was identified in patients from the Arabian Peninsula in June of 2012 [11]. The clinical features of the underlying disease, called the Middle East Respiratory Syndrome (MERS), were similar to SARS, with additional gastrointestinal symptoms and frequent acute kidney failure. The differences might be explained by disparate preferences for the entry receptor [12], whereas dipeptidyl peptidase 4 (DPP4) is involved in Middle East Respiratory Syndrome Coronavirus (MERS-CoV) infection [13], and angiotensin-converting enzyme 2 (ACE2) is involved in Severe Acute Respiratory Syndrome Coronavirus (SARS-CoV) infection [14]. Both zoonotic viruses presumably originated in horseshoe bats and acquired adaptations to jump to humans from their intermediate hosts, including domestic Himalayan palm civet cats, raccoon dogs, Chinese ferret badgers (in the case of SARS-CoV), and dromedary camels (in the case of MERS-CoV) [1, 15, 16].

Since then, 2 additional strains of HCoV were found to be circulating in the human population, HCoV-NL63 [17] and HCoV-HKU1 [18], which result in milder symptoms and are, similarly to HCoV-229E and HCoV-OC43, well adapted to humans [19]. The emergence of another pandemic CoV zoonosis seemed very likely, as horseshoe bats were suggested to harbor viral variants capable of infecting human cells without the need for prior mutations [20]. Indeed, at the end of 2019, Wuhan, Hubei Province, China became the hotspot for the spread of a novel HCoV, initially termed as “novel coronavirus (2019-nCoV),” but subsequently renamed Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) [21]. It has been shown that SARS-CoV-2 shares an 80% sequence similarity to its predecessor SARS-CoV and only 50% with MERS-CoV [22]. Nevertheless, SARS-CoV-2 reached practically all parts of the globe, causing one of the most severe pandemics, Coronavirus Disease 2019 (COVID-19), in recent history [23].

COVID-19 spreads via droplets and contact with infected surfaces [24, 25], and the fecal–oral route of transmission is also possible [26]. The main symptoms of COVID-19 include fever, weakness, dry cough, myalgia, and dyspnea. The severity of COVID-19 widely ranges from a mild, uncomplicated illness to bilateral pneumonia with acute respiratory distress syndrome, sepsis, and multi-organ failure, including kidney failure, cardiac injury, and thrombosis [27, 28]. The elderly population is more susceptible to disease, and a significant number of patients display comorbidities such as diabetes, hypertension, and cardiovascular disease [29].

Besides symptomatic treatment, several approved therapeutics have been tested for the antiviral treatment of COVID-19, and many more are undergoing preclinical development. To stop successful entry into the cell, as well as viral replication and propagation, the inactivation of host or viral peptidase activities or even both is necessary. Spike surface envelope glycoprotein, one of four structural proteins [30], is needed for virion entry into target cells, which is facilitated by binding to cell receptors (e.g., ACE2, DPP4, or aminopeptidase N) and fusion of viral and cell membranes. The large ectodomain of the spike protein is a target for several host peptidases, which cleave it at the border of domains S1 and S2 (cleavage site S1/S2) and at the S2’ site [31]. The mechanism of their action is highly pH dependent and involves different enzymes at various stages of the viral life cycle. Furthermore, cleavage takes place on the cell membrane, in endosomes, extracellularly, or during protein biosynthesis in the endoplasmic reticulum/Golgi apparatus. Differences in cleavage events may modulate viral pathogenicity and host range as well as cell and tissue tropism [3133].

The viral genome encodes the cysteine papain-like peptidases (PLPs) and chymotrypsin-like cysteine 3C-like peptidase (3CLPro), also known as main peptidase (MPro) [34, 35], which are necessary for mediating viral genome transcription and replication. The viral RNA genome is translated into 2 large polyproteins, pp1a and pp1ab, which are, by their intrinsic peptidase activity, further cleaved into several nonstructural proteins (Nsps) as presented in Fig 1, building a replicase–transcriptase complex that is vital for viral transcription and replication [32, 33, 36, 53].

thumbnail
Fig 1.

Schematic representation of the SARS-CoV and SARS-CoV-2 genome (A) and the replicase polyproteins (B). The genomic RNA comprises 2 parts, of which the first part (ORF1a and ORF1b) directly translates into 2 polyproteins pp1a and pp1ab due to a −1 frameshift between the 2 ORFs (orange arrow). They are composed of 16 Nsps (Nsp1–Nsp16) that form the replication–transcription complex. The polyproteins are processed into mature Nsps by the 2 peptidases as indicated in panel B by black and white triangles, namely Mpro, which is Nsp5, cleaves the polyprotein at 11 sites and PLpro, which is part of the Nsp3, performs cleavage at 3 sites. Among the released Nsps are RdRp (Nsp12) and RNA helicase (Nsp13). The 3′ third of the genome encodes the structural and accessory proteins, including spike (S), envelope (E), membrane (M), and nucleocapsid (N) shown in green and gray. Nsp, nonstructural protein; ORF, open reading frame; RdRp, RNA-dependent RNA polymerase; SARS-CoV, Severe Acute Respiratory Syndrome Coronavirus; SARS-CoV-2, Severe Acute Respiratory Syndrome Coronavirus 2.

https://doi.org/10.1371/journal.ppat.1009013.g001

Many studies have recently been accelerated to provide key insights into the mechanisms underlying SARS-CoV-2 infection and to determine potential targets for antiviral interventions. Viral and host cysteine peptidases have become promising targets for antiviral treatment. In this review, we explore CoV-related promoting activities of viral and host peptidases with an emphasis on cysteine peptidases and discuss the possibilities of using their inhibitors to mitigate HCoV infections.

Viral and host peptidases in CoV cell entry and replication

Peptidases play important roles in key steps of CoV infection and replication (reviewed in [31, 36]). Both viral and host peptidases are involved in these processes (Table 1). The peptidases encoded in the viral genome are essential for replicase polyprotein processing and evasion of the host immune response, while host peptidases are involved in different steps of virion uptake into the host cell.

thumbnail
Table 1. The peptidases involved in the infection and replication of CoVs.

https://doi.org/10.1371/journal.ppat.1009013.t001

CoV-encoded peptidases

The genomes of CoVs encode the PLPs, papain-like (PL)1pro and PL2pro, within Nsp3 and 3CLpro within Nsp5 (reviewed in [15, 37]). 3CLpro, also named Mpro, is composed of 3 domains; domains 1 and 2 form the chymotrypsin-like fold with the substrate-binding site located in the cleft between the 2 domains, and domain 3 is required for dimer formation and affects catalytic activity by dynamically driven allostery. Mpro is active only as a dimer and processes the polyprotein at 11 distinct cleavage sites with 2 self-cleavage sites that are cut most efficiently. Both the substrate-binding and dimerization sites have been targeted for inhibitory drug development [3845]. CoV Mpro shares highly conserved substrate recognition pockets, which are responsible for cleaving the viral polyprotein and host factors involved in the innate immune response, including the signal transducer and activator of transcription 2 and the signaling protein nuclear factor-κB essential modulator [15, 4648]. Recently, conservation at the polyprotein cleavage sites has been confirmed. All 11 Mpro sites are highly conserved or identical, as only 12 out of the 306 residues of the SARS-CoV-2 Mpro sequence are different from those of SARS-CoV [49]. Additionally, a proteomic analysis of cellular responses indicated that SARS Mpro also affects cellular protein metabolism and modification, particularly in the ubiquitin proteasome pathway [50].

The other CoV peptidases, PLPs, are multifunctional enzymes that are involved in polyprotein processing but also play important roles in viral interactions with the host (reviewed in [37, 51]). They are primarily responsible for the cleavage of the N-terminal part of the polyprotein and catalyze replicase polyprotein processing at the cleavage sites between Nsp1/Nsp2, Nsp2/Nsp3, and Nsp3/Nsp4 [52]. CoVs encode either a single PLP (PLpro) or 2 PLPs (PL1pro and PL2pro) that process a total of 3 cleavage sites within the polyprotein [51, 53]. SARS-CoV encodes only 1 PLP domain (PLpro) within Nsp3, which belongs to the peptidase clan CA in the MEROPS peptidase classification. The active site contains a classic catalytic triad, composed of Cys112-His273-Asp287, and catalyzes viral polyprotein processing at the N-terminus of pp1a, releasing the Nsps Nsp1, Nsp2, and Nsp3 through recognition of a LeuXaaGlyGly (LXGG) motif [36, 54]. Beyond cleaving viral polyproteins, PLPs exert additional activities that promote virus replication. These enzymes resemble the structure of human ubiquitin-specific peptidases and are thereby known as viral ubiquitin-specific peptidases, often acting as deubiquitinating enzymes with the ability to remove the posttranslational modification of ubiquitin from target proteins [55]. The in vitro characterization of PLpro enzymatic activities reveals that PLpro can recognize and hydrolyze the cellular proteins ubiquitin and ubiquitin-like protein interferon-stimulated gene product 15 (ISG15), both bearing the LXGG recognition motif at their carboxyl terminus ([56]; reviewed in [36]). The dual deubiquitinase (ubiquitin cleaving) and deISGylase (ISG cleaving) activities of SARS-CoV PLpro antagonize the type I interferon (IFN-I) response and were suggested to be a key viral immune response mechanism [5759]. Additionally, SARS-CoV PLpro has been shown to inhibit interferon regulatory factor 3 phosphorylation early in the IFN-I signaling pathway [57, 58]. SARS-CoV PLpro is able to inhibit interferon regulatory factor 3-dependent IFN-I activation at later stages as well, by deubiquitinating interferon regulatory factor 3 and blocking the ability to induce IFN-β transcription, suggesting innate immune antagonistic activities of PLpro that may contribute to SARS-CoV pathogenesis [60].

Mpro and PLpro catalyze their own release and liberate other Nsps from the polyprotein. As such, the enzymatic activities of both viral peptidases are essential for the viral life cycle. Nevertheless, host cell peptidases are also required for viral infectivity (reviewed in [30]).

Host peptidases are essential for virion uptake

Host peptidases are critical for CoV entry and act either as receptors for the attachment of the virion spike protein to the target cell or as facilitators of virion envelope fusion with the target cell membrane. The metallo-carboxypeptidase ACE2 is the primary receptor for the SARS-CoV, HCoV-NL63, and novel SARS-CoV-2 [15, 61, 62]. The peptidase activity of ACE2 negatively regulates the renin–angiotensin system, which plays a key role in maintaining blood pressure homeostasis as well as fluid and salt balance in mammals (reviewed in [63]). Additionally, the unique binding of SARS-CoV spike protein to ACE2 induces ACE2 ectodomain shedding, which leads to the down-regulation of ACE2. Since normal ACE2 expression protects from lung injury, reduced ACE2 levels are believed to promote lung pathogenesis. ACE2 ectodomain shedding is mediated by the metallopeptidase ADAM17, a member of the a disintegrin and metalloproteinase (ADAM) family of peptidases, also named tumor necrosis factor alpha (TNF-α)-converting enzyme [6466]. The metallopeptidase aminopeptidase N (CD13) is the receptor for HCoV-229E, which utilizes the endosomal pathway for cell entry, in which lysosomal peptidases, including cathepsin L, are involved [67]. DPP4 (CD26) is a serine aminopeptidase that acts as a receptor for MERS-CoV and is also an essential determinant of MERS-CoV tropism [13, 68]. Following binding to the host cell receptor, additional factors affect the next steps of virion entry. For example, the transmembrane proteins tetraspanins (e.g., tetraspanin CD9) facilitate MERS-CoV entry by forming complexes between DPP4 and the type II transmembrane serine peptidases, which are important fusion-activating peptidases that facilitate virion entry and promote efficient infection [69]. After the spike protein binds to the cellular receptor, it is subjected to limited proteolysis at distinct cleavage sites located at the domain boundary (S1/S2) or within domain 2 upstream of the putative fusion peptide (S2’), which directly leads to membrane fusion.

Peptidases involved in CoV spike protein activation can act at different stages of the viral life cycle and include proprotein convertases (e.g., furin) during virus packaging, extracellular peptidases (e.g., elastase), cell surface peptidases (e.g., transmembrane serine peptidases) after receptor binding, and lysosomal peptidases (e.g., cathepsins L and B) after virion endocytosis (reviewed in [31, 70]). Proprotein convertase furin, a subtilisin-like serine peptidase, is not involved in activating MERS for virion entry but rather in determining cell tropism [71, 72]. Furthermore, a novel furin cleavage site at the S1/S2 boundary has been identified with a bioinformatics analysis of the novel SARS-CoV-2, indicating possible effects of this virus on cell tropism [73]. The extracellular serine peptidase elastase, which is secreted by neutrophils as part of an inflammatory response to viral infection, was shown to enhance SARS-CoV infection by proteolytic activation of the SARS-CoV spike protein [74]. The SARS-CoV spike protein can be proteolytically activated by a human airway and alveolar peptidase transmembrane peptidase/serine subfamily member 2 (TMPRSS2) as well as other transmembrane serine peptidases, including human airway trypsin-like peptidase [7577]. Furthermore, TMPRSS2 also activates spike protein for virion entry of MERS-CoV [78, 79] and other HCoVs, including HCoV-229E [80], HCoV-OC43, HCoV-HKU1 [81], and SARS-CoV-2 [61]. In addition to the TMPRSS2-facilitated direct entry from the cell membrane, the CoV also utilizes the endosomal pathway, where spike proteins are proteolytically activated by the lysosomal cysteine peptidases cathepsin L or cathepsin B. Among HCoVs, this is the case for the SARS-CoV [82] and MERS-CoV [86]. Between the 2 distinct virion entry pathways, namely the endosomal pathway and the direct cell surface entry, clinical isolates of HCoV-229E, HCoV-OC43, and HCoV-HKU1 have demonstrated that direct entry from the cell surface or early endosomes is preferred in vivo [80, 81]. Furthermore, factor Xa [83] and an unidentified leupeptin-sensitive peptidase [84] are other examples of peptidases that were shown to activate SARS-CoV spike protein for entry into host cells. Another cysteine peptidase, m-calpain, has been implicated in SARS-CoV replication, indicating an important role for m-calpain during the early steps of the SARS-CoV life cycle [85]. Taken together, host cell peptidases are indeed important for virion uptake into the host cell (Table 1), and host cysteine peptidases have been recognized as key players in virion entry and will be addressed in more detail in the following section.

Host papain-like cysteine cathepsins are involved in CoV uptake

CoVs, including SARS-CoVs, can employ 2 routes for host cell entry that are determined by the localization of the peptidases required for spike protein activation (Fig 2). Spike proteins can be activated by transmembrane serine peptidases at the cell membrane, resulting in fusion of the viral membrane with the plasma membrane [30]. Alternatively, a virion gains entry into a target cell by binding to the receptors on the cell surface (e.g., ACE2), resulting in its uptake into endosomes [87]. Endosomal vesicles within the target cells form highly dynamic multifunctional cellular compartments that contain multiple proteolytic enzymes, including lysosomal peptidases [88]. The main class of lysosomal peptidases comprises the cathepsins, which are subdivided into 3 subgroups according to the active site amino acid residue: aspartyl (cathepsins D and E), serine (cathepsins A and G), and cysteine cathepsins. The latter constitute the largest cathepsin family, with 11 peptidases in humans belonging to clan CA, family C1a: cathepsins B, C/DPP1, F, H, K, L, O, S, W, V, and Z/X [89]. They are delivered to early endosomes as inactive zymogens and are activated by the lower pH either through proteolytic processing by other endosomal/lysosomal hydrolases or through their interactions with glycosaminoglycans [9092]. Cysteine cathepsins are optimally active and stable at acidic pH and, generally, inactivated at neutral pH; although some of them, e.g., cathepsin S, maintain their proteolytic activity at neutral pH [91, 93]. Cathepsins were initially considered intracellular enzymes, responsible for the nonspecific bulk proteolysis in the acidic environment of the endosomal/lysosomal compartments [93, 94]. Later, a broad spectrum of more specific roles was proposed for cathepsins, under both physiological and pathological conditions [9499], including virus infections [30].

thumbnail
Fig 2. The routes of SARS-CoV entry into target cells.

The viral entry takes place either by (A) receptor-mediated endocytosis or by (B) membrane fusion. (A) Binding of SARS-CoV and SARS-CoV-2 to the cellular receptor ACE2 can result in virion uptake into endosomes. The spike protein is activated by cysteine peptidase cathepsin L in endo-lysosomes. (B) In membrane fusion, the spike protein activation is mediated by TMPRSS2 or other transmembrane serine peptidases at the cell surface, resulting in fusion of the viral membrane with the plasma membrane. Either way, the RNA genetic material of the virus is released, and the late stage of the life cycle subsequently takes place by RNA replication. ACE2, angiotensin-converting enzyme 2; SARS-CoV, Severe Acute Respiratory Syndrome Coronavirus; SARS-CoV-2, Severe Acute Respiratory Syndrome Coronavirus 2; TMPRSS2, transmembrane peptidase/serine subfamily member 2.

https://doi.org/10.1371/journal.ppat.1009013.g002

The role of cysteine cathepsins in the viral life cycle is particularly associated with virion entry. In non-enveloped viruses (i.e., reoviruses), the virion is converted into an infectious subvirion particle via partial proteolysis, mediated by cathepsins B, L, or S [171175]. In more detail, Ebert and colleagues showed that cathepsin L or cathepsin B is required for reovirus entry into murine fibroblasts L929 [172]. Two years later, Golden and colleagues showed that acid-independent cathepsin S mediated outer capsid processing in macrophage-like P388D cells supported infection by virions of strain Lang, but not strain c43 [175]. In the last decade, the proteolytic digestion by endosomal cathepsins has been described during entry of enveloped viruses, such as the Ebola virus and CoVs [30, 176]. Cathepsin L is most commonly associated with the activation of viral glycoproteins. It is known to process the SARS-CoV spike protein as well as many proteins of other HCoVs, including MERS-CoV and HCoV-229E [67, 79, 101]. The cysteine peptidase cathepsin B can also be involved in virion entry and exhibits distinct properties compared to cathepsin L and other cathepsins; it generally has a higher pH optimum and, while preferring an aromatic P2 residue, can process dibasic substrates [177, 178]. Cathepsin B was shown to play key roles in the lifecycle of the Ebola virus [176], Nipah virus [179], and feline CoV [180], catalytically activating viral membrane glycoproteins, which leads to fusion of the viral envelope with the endosomal membrane and virion release from endosomes into the cytoplasm [181, 182].

Interestingly, it has been shown that the SARS-CoV, but not HCoV-NL63, utilizes cathepsins to enter the host cell [100]. Furthermore, only cathepsin L, but not other cathepsins, primes the SARS-CoV for membrane fusion, in which the interaction with ACE2 is required [101, 102]. Based on these observations, a 3-step process of SARS-CoV and SARS-CoV-2 infection has been proposed (Fig 2): receptor binding, induced conformational changes in spike protein, and cathepsin L proteolysis within endosomes [101]. As mentioned above, the spike protein in the SARS-CoV exposes a peptidase-accessible loop between the S1 and S2 subunits, which can be targeted by different enzymes, including TMPRSS2, furin, and cathepsin L, thus activating the membrane fusion function of the spike protein. The SARS-CoV can use the endosomal cathepsins B or L for spike protein priming in TMPRSS2- cells [101]. However, spike protein priming by TMPRSS2, but not by cathepsins B and L, was proposed to be essential for virion entry into the primary target cells and for viral spread in the infected host [103105]. Nevertheless, a recent study [61] has demonstrated that SARS-CoV-2 infection was inhibited by a serine peptidase inhibitor, likely reflecting spike protein priming by other peptidases. Camostat mesylate, which is active against TMPRSS2 [104], only partially blocked SARS-CoV-2 spike protein-driven entry into Caco-2 cells and Vero-TMPRSS2+ cells, whereas full inhibition was attained in combination with E-64d, a broad inhibitor for cathepsins B, H, L, and calpain [106]. This suggests that SARS-CoV-2 spike protein can use both cathepsins B and L as well as TMPRSS2 for priming in these cells. Another study confirmed the important role of cysteine cathepsins in SARS-CoV-2 entry and identified cathepsin L as a critical lysosomal peptidase for virion entry into human embryonic kidney 293 cells stably expressing recombinant human ACE2, while the cathepsin B–specific inhibitor CA-074 [107] did not exhibit any marked effect on SARS-CoV-2 entry [77].

The impact of cysteine peptidases on the CoV host immune response

Studies on the effect of SARS-CoV-2 on the host immune response are limited; however, preliminary data show great similarity with other beta CoVs with similar genomic sequences [22, 108]. The antiviral immune response starts with the recognition of the virus and virally infected cells by the innate immune system. Viral RNA is recognized by pathogen-associated molecular pattern receptors, such as Toll-like receptors (TLRs) [109]. The activation of nucleic acid–sensing TLRs (i.e., TLR-3, TLR-7, and TLR-9) is a complex process and is thought to involve a combination of asparagine endopeptidase (AEP) legumain and/or multiple cathepsins [110113]. TLR trigger results in IFN-I secretion, which controls viral replication and supports the effective adaptive immune response [109]. Nevertheless, in senescent BALB/c mice administered with SARS-CoV intranasally, it was shown that SARS-CoV infection is able to delay the production of IFN-I, while the virus replicates rapidly, which results in diminished responses and numbers of immune effector cells [114, 115]. SARS-CoV-2 may develop even higher replication rates and lower induction of host IFN-I compared to those of SARS-CoV [116]. A recent clinical study showed that severely ill COVID-19 patients exhibited an impaired type IFN-I response characterized by low IFN production and activity, with subsequent down-regulation of IFN-stimulated genes [117].

To activate adaptive immunity, viral antigens must be presented to T cell receptors by infected cells and other antigen-presenting cells via the major histocompatibility complex class I or class II molecules. Together with other lysosomal peptidases and immunoproteasomes, cysteine cathepsins enable the formation of antigen epitopes or, by cleaving the invariant chain Ii, the priming of major histocompatibility complex II molecules and their trafficking to the cell surface [118120]. SARS-CoV-infected monocytes down-regulate the expression of cathepsins involved in antigen presentation and processing, suggesting a limited activation of a favorable adaptive immune response against SARS-CoV [121].

An increasing number of studies are reporting lymphopenia as one of the hallmarks of severe COVID-19, which is exacerbated with disease progression [117, 122125]. The most affected lymphocyte subsets are natural killer (NK) cells [126], cytotoxic CD8+ T lymphocytes (CTLs) [126, 127], and most CD4+ lymphocytes [127]. So far, several potential mechanisms that lead to lymphocyte deficiency in COVID-19 disease have been proposed, such as induced lymphocyte apoptosis, ACE2 receptor expression, or direct viral targeting of lymphoid organs [123]. In addition to apoptosis, inflammatory cell death (e.g., necroptosis or pyroptosis) cannot be excluded, as virulent HCoVs have been shown to promote RIPK3-dependent necroptosis and caspase-1-dependent pyroptosis [128]. Furthermore, functional assessments of NK cells and CTLs in COVID-19 patients showed that both their numbers and functions are compromised. NK cells and CTLs from COVID-19 patients exhibit increased expression of inhibitory receptor NKG2A and decreased expression of CD170a and effector molecules such as interferon gamma (IFNγ), TNF-α, and granzyme B, suggesting a state of exhaustion [126]. Therefore, virus-induced cytopathic effects that lead to immune insufficiency and dysregulation of immune response enable viral replication and cause tissue damage that is a major cause of COVID-19 pulmonary pathology. This has been confirmed by histologically examining lung tissue of severe cases of SARS-CoV-2 infection, which revealed infiltrates of inflammatory mononuclear cells and cytopathic-like changes [127].

In COVID-19 patients, transcriptomic analysis of peripheral blood mononuclear cells showed a significant increase in the expression of cathepsin L and B genes associated with the apoptotic pathway [129]. Cathepsins initiate apoptosis by cleavage of the proapoptotic molecule BH3-interacting domain death agonist (BID) and degradation of antiapoptotic Bcl-2 proteins. This results in the release of cytochrome c from mitochondria and the activation of caspases 3 and 7 [130, 131]. Furthermore, inhibitor studies suggest that cathepsins, in particular cathepsins B, C, and D, are key peptidases involved in inflammatory cell death [132]. The cytotoxic function of NK cells and CTLs is regulated by cathepsins. Perforin and granzymes are stored in their precursor forms in the cytotoxic granules of NK cells and CTLs. Upon cell activation, cathepsins C and H catalyze the conversion of progranzymes into granzymes and cathepsin L, although not redundantly, activates perforin. The function of these cathepsins is further regulated by the endogenous inhibitor cystatin F [133]; however, the status of cystatin F and cathepsins in cytotoxic cells in COVID-19 patients has not been evaluated so far.

Virus attenuation by cysteine peptidase inhibitors

Targeting peptidases involved in the infection and replication of CoVs represents an effective strategy to block viral replication. An attractive target for development of antiviral drugs directed against the SARS-CoV-2 and other CoVs is 3CLpro due to its essential role in processing polyproteins after translation. After the 2003 SARS-CoV outbreak, numerous inhibitors of 3CLpro peptidase were proposed, yet no new drug candidates have succeeded in entering clinical trials (reviewed in [34, 134, 135]). However, several peptidic and small molecular inhibitors have exhibited potent activities against SARS-CoV 3CLpro as reviewed in detail in the recent papers by others [136, 137]. Among other keto peptidomimetics, α-ketoamide inhibitors have been developed as potent inhibitors [45]. Similarly, the PLP domain of SARS-CoV Nsp3 has emerged as a viable drug target, resulting in the development of several SARS-CoV PLpro inhibitors (reviewed by [36]). Recently, naphthalene-based PLpro inhibitors were shown to be effective at halting SARS-CoV-2 PLpro activity as well as SARS-CoV-2 replication, which might represent a strategy for generating PLpro-targeted therapeutics to be used against SARS-CoV-2 [138]. Next, peptides derived from HKU1 CoV were shown as effective inhibitors of formyl peptide receptor that plays important role in recognition by phagocytic cells [138].

In addition to viral peptidases, host peptidases are important for virus infection. As mentioned above, CoVs, including SARS-CoV and novel SARS-CoV-2, utilize a cathepsin-dependent endosomal pathway in addition to the direct cell surface serine peptidase-mediated pathway [30, 61, 77]. This suggests that targeting cysteine cathepsins could be a promising strategy for the development of antiviral drugs directed against SARS-CoVs as well as other related CoV infections.

The potent effect of cysteine peptidase inhibitors on CoV replication was already shown in the 1990s (Table 2) when Colling and Grubb demonstrated the inhibition of CoV replication by cystatins C and D [139, 140]. Cystatins are general inhibitors of cysteine peptidases (reviewed in [141143]) and may inactivate either viral or host cysteine peptidases. Cystatin C potently affects HCoV-OC43 and HCoV-229E, inhibiting their replication by more than 99% at a concentration of 0.1 mM [140]. Cystatin D is specifically expressed in human saliva and tears [144] and inhibits CoV replication at physiological concentrations [139].

thumbnail
Table 2. A summary of cysteine cathepsin inhibitors tested for CoV inhibition.

https://doi.org/10.1371/journal.ppat.1009013.t002

Similarly, the irreversible epoxysuccinyl cell permeable inhibitor E-64d (Fig 3) also provided an effective strategy to block virion entry into cells, as it blocked SARS-CoV entry into Caco-2 and Vero-TMPRSS2+ cells as well as SARS-CoV-2 entry into human embryonic kidney 293/hACE2 cells. For the latter, virion uptake was blocked by E-64d by 92.5% [61, 77]. Furthermore, simultaneous treatment with the cysteine peptidase inhibitor E-64d and serine peptidase inhibitor camostat strongly inhibited virion entry and multistep growth of the SARS-CoV in airway epithelial Calu-3 cells and TMPRSS2-expressing HeLa cells [104]. The cysteine peptidase inhibitors K11777 and closely related vinylsulfones also target cathepsin-mediated virion entry into cells and therefore act as broad-spectrum antiviral agents [105]. In particular, K11777 is very potent, inhibiting SARS-CoV as well as Ebola virion cell entry at subnanomolar concentrations.

thumbnail
Fig 3. Structures of cysteine cathepsin inhibitors tested for CoV inhibition.

(A) The structures of cysteine cathepsin inhibitors. PDB code for cystatin C is 3GAX and for cystatin D is 1RN7. (B) The crystal structure of cathepsin L in complex with cathepsin L III inhibitor (PDB code 3OF8). Cathepsin L is shown in blue, cathepsin L III inhibitor is represented by sticks and colored green, and cathepsin L active site cysteine is marked yellow. The structures were prepared using ChemDraw (PerkinElmer, United States of America), and crystal structures were prepared using PyMol (Schrödinger, USA). CoV, coronavirus; PDB, Protein Data Bank.

https://doi.org/10.1371/journal.ppat.1009013.g003

Among the specific cysteine peptidase inhibitors, the targeting of cathepsin L has provided the most promising antiviral effects. The cathepsin L–specific inhibitor CID 23631927 was shown to block SARS-CoV entry in human embryonic kidney 293T cells [145]. Additionally, the activity-based probe, which labels active cathepsin L, confirmed the effect of the inhibitor on SARS-CoV entry into human cells [145]. Similarly, the cathepsin L–specific inhibitor SID 26681509 blocked SARS-CoV-2 entry into human embryonic kidney 293/hACE2 cells. In the same study, the cathepsin B–specific inhibitor CA-074 did not exert any marked effect on virion entry [77].

Another example of a drug that could impair the function of cathepsins involved in the cleavage of CoV spike protein is amantadine. As a prophylactic agent for influenza and Parkinson’s disease, it was shown to down-regulate cathepsin L expression at a concentration of 10 μM [146]. Additionally, amantadine acts as a lysosomotropic agent and disrupts the lysosomal pathway, interfering with the capacity of the virus to replicate [146]. Moreover, glycopeptide antibiotics such as teicoplanin and its derivatives dalbavancin, oritavancin, and telavacin underwent high-throughput screening of Food and Drug Administration (FDA)-approved drugs and were identified to prevent MERS-CoV, SARS-CoV, and Ebola pseudovirus cell entry due to cathepsin L inhibition [147, 148]. Teicoplanin was recently shown to prevent the entrance of SARS-CoV-2 pseudoviruses into the cytoplasm in low molecular concentration [183]. By screening a chemical library of compounds for blocking the entry of HIV-1 pseudotyped with SARS-CoV spike protein but not that of HIV-1 pseudotyped with vesicular stomatitis virus surface glycoprotein G, benzamide inhibitor SSAA09E1 was discovered as effective, again by inhibiting cathepsin L [82]. Small molecule inhibitors of cathepsin L that can impair virion entry and SARS-CoV spike protein cleavage were also identified in a high-throughput screening assay of peptides derived from the glycoproteins of the SARS-CoV and Ebola, Hendra, and Nipah viruses [149]. Among them, 5705213 and its derivative 7402683 inhibited cathepsin L-mediated cleavage of the recombinant SARS-CoV spike protein in a dose-dependent manner. Additionally, 5705213 in combination with the TMPRSS2 peptidase inhibitor completely blocked the entry of the pseudotypes bearing the SARS-CoV spike protein [149].

The entry of MERS-CoV into cells was also blocked by the furin inhibitor dec-RVKR-CMK (Decanoyl-Arg-Val-Lys-Arg-CMK) due to its inhibition of cathepsin L and TMPRSS2 enzymatic activity. In addition, dec-RVKR-CMK also inhibited cathepsin B, trypsin, and papain [72]. Another cathepsin L inhibitor that affects virus entry is MDL-28170 [101, 150]. This inhibitor blocked SARS-CoV infection of TMPRS2- HeLa-ACE2 cells driven by spike protein. Additionally, pseudotype SARS-CoV infection of TMPRS2- HeLa-ACE2 cells was inhibited by cathepsin L III inhibitor for 80% [104].

Together, various studies have identified an important role of cysteine cathepsins in virion entry (Table 2) and suggest that inhibition of cathepsins, especially cathepsin L, could be an effective strategy for the search/development of new drugs for the treatment of COVID-19. Although the role of cathepsin L seems pivotal, the role of other host cysteine peptidases in virus infection and replication should be further investigated to assess their individual contribution to virus spread.

The potential of novel selective cysteine peptidase inhibitors

Exogenous cysteine peptidase inhibitors can be of protein or nonprotein origin and isolated from animals, microorganisms, plants, fungi, neutralizing monoclonal antibodies and their fragments, propeptides, or synthetic small molecules. Based on the mechanism of inhibition, we distinguish between reversible and irreversible exogenous inhibitors [152]. The majority of cysteine peptidase inhibitors contain a reactive electrophilic functional group that reacts with cysteine’s thiol group in the active site of the enzyme, where the peptide fragment is crucial for selectivity and enzyme recognition (reviewed in [90, 152154]).

Reversible cysteine peptidase inhibitors can be peptide molecules such as propeptides, monoclonal antibodies, and small molecules, including aldehydes, methylketones, trifluoromethylketones, diketones, α-ketoacids, α-ketoesters, α-ketoamides, α-keto-β-aldehydes, nitriles, azapeptide nitriles, thiosemicarbazones, and biflavones (reviewed in [152155]). Numerous promising reversible cathepsin L inhibitors exist, including aldehyde derivative iCP (Napsul-Ile-Trp-CHO) [156, 157], thiosemicarbazone derivative KGP94 [158], tetrahydroquinoline oxocarbazate derivative CID 23631927, and SID 26681509, which block SARS-CoV entry into human cells [77, 145, 159, 160]. Certain compounds that are already used in clinical practice are candidates for repurposing, of which the reversible cysteine peptidase inhibitor nitroxoline shows promise [160, 161]. This well-established antimicrobial agent used for the treatment of urinary tract infections was identified as a potent cathepsin B inhibitor both in vitro and in vivo in various cell-based functional assays and tumor mice models [160162]. The antituberculosis agent rifampicin was also shown to inhibit cathepsins B, L, and H with inhibition constants in the low micromolar range [163]. In addition to classic small molecule inhibitors, organometallic complexes of quinolone, clioquinol, and nitroxoline with ruthenium have been identified as inhibitors of cysteine peptidases [162, 164].

The most studied irreversible inhibitors are epoxysuccinyl inhibitors, whose reactive electrophilic epoxide rings form covalent thioester bonds with the thiol group of the reactive active site. 1-[L-N-(trans-epoxysuccinyl)leucyl] amino-4-guanidinobutane (E-64) is the best known general epoxysuccinyl cysteine peptidase inhibitor, which was first isolated from Aspergillus japonicus [165]. Due to their high reactivity and lack of selectivity, E-64 and its cell-permeable derivative E-64d are not applicable to clinical practice. However, E-64 represents the lead compound for the development of new inhibitors with increased selectivity (reviewed in [90, 152, 153, 155]). As a derivative of E-64, the potent and selective cathepsin B inhibitor CA-074 was synthesized [107], and its low membrane permeability was improved by methylation of the carboxyl group [166]. CA-074Me is cell permeable and can be hydrolyzed to CA-074 within cells; however, such methylation decreases its selectivity for cathepsin B [166]. Based on the structure of E-64, the cathepsin L–specific inhibitor CLIK-148 was developed [167]. In addition to epoxysuccinyl inhibitors, irreversible cysteine peptidase inhibitors include epoxides, vinyl sulphones, aziridines, azomethylketones, halomethylketones, acyloxymethylketones, azapeptides, thiadiazoles, hydroxametes, β-lactams, and α,β-unsaturated carbonyl derivatives and others (reviewed in [152, 154]).

However, irreversible inhibitors are less suitable for use in clinical practice due to the presence of reactive electrophilic warheads, which can, in addition to targeting enzymes, nonspecifically react with nucleophiles from other off-target proteins, resulting in side effects. Furthermore, their bioavailability can be lower than that of reversible inhibitors due to their peptide scaffold (reviewed by [152154]). Therefore, new research is directed toward the development of selective reversible peptidase inhibitors, while selective irreversible inhibitors remain important research tools in elucidating the molecular mechanisms of cathepsins in disease [90]. However, as host peptidases are also involved in normal physiological process, the care must be taken to avoid side effects that may occur. These are expected to be similar as those seen in treatment of other diseases where peptidase inhibitors are already used in clinical practice. In particular, the impact of new treatment on host immune system has to be considered [90, 168].

Conclusions and future perspectives

CoVs have been recognized for more than 5 decades and represent a group of severely pathogenic viruses with a high risk to public health. Recurrent viral threats continue to emerge, most recently by the novel SARS-CoV-2, which developed into a pandemic of COVID-19 in 2020. So far, no approved vaccines or specific therapeutics exist for these pathogenic viruses, including SARS-CoV-2. Improved clinical outcomes of COVID-19 patients have been reported after treatment with the repurposed drug remdesivir, which inhibits viral RNA polymerases [169]. However, new antiviral drugs directed against the SARS-CoV-2 and other CoVs are urgently needed to provide adequate responses to future pandemic threats. Targeting viral and host peptidases involved in virion entry and replication could be a key for the development of new antiviral therapies. Peptidase inhibitors are already successfully used in clinical practice for treatment of other infectious diseases caused by viruses such as HIV, hepatitis, rhinovirus and other (reviewed in [170]). Nevertheless, a better understanding of the mechanisms by which host cysteine peptidases contribute to SARS-CoV-2 uptake into target cells would provide valuable new information for virus pathogenesis, vaccine design, and drug targeting. Peptidase inhibitors with dual inhibitory activity targeting both viral and host PLPs with an additional positive effect on the immune response may represent the needed multilayered quality of an effective anti-coronaviral therapeutic. The lysosomal cysteine peptidase cathepsin L appears to be critical for SARS-CoV as well as SARS-CoV-2 spike protein activation. As such, this cysteine peptidase is a promising therapeutic target for the development of antiviral drugs. Nevertheless, the contribution of other cysteine cathepsins should not be overlooked. Future work should focus on evaluating the CoV antiviral activity of a number of cathepsin inhibitors developed for treatment of various diseases, and on repurposing their application for treatment of viral infections, but in a way not to compromise the host antiviral immune response.

Acknowledgments

The authors sincerely acknowledge Eva Veronica Lasič for critically reviewing the manuscript.

References

  1. 1. Cui J, Li F, Shi ZL. Origin and evolution of pathogenic coronaviruses. Nat Rev Microbiol. 2019;17(3):181–192. pmid:30531947
  2. 2. Khan S, Siddique R, Shereen MA, Ali A, Liu J, Bai Q, et al. Emergence of a Novel Coronavirus, Severe Acute Respiratory Syndrome Coronavirus 2: biology and therapeutic options. J Clin Microbiol. 2020;58(5):e00187–e00120. pmid:32161092
  3. 3. Cunningham CH, Stuart HO. Cultivation of the virus of infectious bronchitis of chickens in embryonated chicken eggs. Am J Vet Res. 1947;8(27):209–212. pmid:20295891
  4. 4. Hamre D, Procknow JJ. A new virus isolated from the human respiratory tract. Proc Soc Exp Biol Med. 1966;121(1):190–193. pmid:4285768
  5. 5. Kapikian AZ, James HD, Kelly SJ, Dees JH, Turner HC, McIntosh K, et al. Isolation from Man of Avian Infectious Bronchitis Virus-Like Viruses (Coronaviruses) Similar to 229e Virus with Some Epidemiological Observations. J Infect Dis. 1969;119(3):282–290. pmid:4976345
  6. 6. Nickbakhsh S, Ho A, Marques DFP, McMenamin J, Gunson RN, Murcia PR. Epidemiology of seasonal coronaviruses: establishing the context for the emergence of Coronavirus Disease 2019. J Infect Dis. 2020;222(1):17–25. pmid:32296837
  7. 7. van der Hoek L, Pyrc K, Berkhout B. Human coronavirus NL63, a new respiratory virus. FEMS Microbiol Rev. 2006;30(5):760–773. pmid:16911043
  8. 8. Gagneur A, Vallet S, Talbot PJ, Legrand-Quillien M-C, Picard B, Payan C, et al. Outbreaks of human coronavirus in a pediatric and neonatal intensive care unit. Eur J Pediatr. 2008;167(12):1427–1434. pmid:18335238
  9. 9. Cheng VCC, Lau SKP, Woo PCY, Yuen KY. Severe acute respiratory syndrome coronavirus as an agent of emerging and reemerging infection. Clin Microbiol Rev. 2007;20(4):660–694. pmid:17934078
  10. 10. Xu RH, He J-F, Evans MR, Peng G-W, Field HE, Yu D-W, et al. Epidemiologic clues to SARS origin in China. Emerg Infect Dis. 2004;10(6):1030–1037. pmid:15207054
  11. 11. Zaki AM. Brief report: isolation of a novel Coronavirus from a man with pneumonia in Saudi Arabia. N Engl J Med. 2013;369(4):394–394.
  12. 12. Paules CI, Marston HD, Fauci AS. Coronavirus infections-more than just the common cold. JAMA. 2020;323(8):707–708. pmid:31971553
  13. 13. Raj VS, Mou H, Smits SL, Dekkers DHW, Müller MA, Dijkman R, et al. Dipeptidyl peptidase 4 is a functional receptor for the emerging human coronavirus-EMC. Nature. 2013;495(7440):251–254. pmid:23486063
  14. 14. Li WH, Moore MJ, Vasilieva N, Sui J, Wong SK, Berne MA, et al. Angiotensin-converting enzyme 2 is a functional receptor for the SARS coronavirus. Nature. 2003;426(6965):450–454. pmid:14647384
  15. 15. Perlman S, Netland J. Coronaviruses post-SARS: update on replication and pathogenesis. Nat Rev Microbiol. 2009;7(6):439–450. pmid:19430490
  16. 16. Su S, Wong G, Shi W, Liu J, Lai ACK, Zhou J, et al. Epidemiology, Genetic Recombination, and Pathogenesis of Coronaviruses. Trends Microbiol. 2016;24(6):490–502. pmid:27012512
  17. 17. van der Hoek L, Pyrc K, Jebbink MF. Vermeulen-Oost W, Berkhout RJM, Wolthers KC, et al. Identification of a new human coronavirus. Nat Med. 2004;10(4):368–373. pmid:15034574
  18. 18. Woo PC, Lau SKP, Chu C, Chan K, Tsoi H, Huang Y, et al. Characterization and complete genome sequence of a novel coronavirus, coronavirus HKU1, from patients with pneumonia. J Virol. 2005;79(2):884–895. pmid:15613317
  19. 19. Fung S-Y, Yuen K-S, Ye Z-W, Chan C-P, Jin D-Y. A tug-of-war between severe acute respiratory syndrome coronavirus 2 and host antiviral defence: lessons from other pathogenic viruses. Emerg Microbes Infect. 2020;9(1):558–570.
  20. 20. Menachery VD, Yount BL, Debbink K, Agnihothram S, Gralinski LE, Plante JA, et al. A SARS-like cluster of circulating bat coronaviruses shows potential for human emergence. Nat Med. 2015;21(12):1508–1513. pmid:26552008
  21. 21. Gorbalenya AE, Baker SC, Baric RS, de Groot RJ, Drosten C, Gulyaeva AA, et al. The species Severe acute respiratory syndrome-related coronavirus: classifying 2019-nCoV and naming it SARS-CoV-2. Nat Microbiol. 2020;5(4):536–544. pmid:32123347
  22. 22. Lu R, Zhao X, Li J, Niu P, Yang B, Wu H, et al. Genomic characterisation and epidemiology of 2019 novel coronavirus: implications for virus origins and receptor binding. Lancet. 2020;395(10224):565–574. pmid:32007145
  23. 23. Queen D, Harding K. COVID-19 a short-term challenge, telewound a lifetime change. Int Wound J. 2020;17(3):529. pmid:32339442
  24. 24. Ong SWX, Tan YK, Chia PY, Lee TH, Ng OT, Wong MSY, et al. Air, surface environmental, and personal protective equipment contamination by Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) from a symptomatic patient. JAMA. 2020;328:1610–1612.
  25. 25. van Doremalen N, Bushmaker T, Morris DH, Holbrook MG, Gamble A, Williamson BN, et al. Aerosol and surface stability of SARS-CoV-2 as compared with SARS-CoV-1. N Engl J Med. 2020;382(16):1564–1567. pmid:32182409
  26. 26. Chen Y, Chen L, Deng Q, Zhang G, Wu K, Ni L, et al. The presence of SARS-CoV-2 RNA in the feces of COVID-19 patients. J Med Virol. 2020;92(7):833–840. pmid:32243607
  27. 27. Li H, Liu Z, Ge JB. Scientific research progress of COVID-19/SARS-CoV-2 in the first five months. J Cell Mol Med. 2020.
  28. 28. Connors JM, Levy JH. COVID-19 and its implications for thrombosis and anticoagulation. Blood. 2020.
  29. 29. Wang D, Hu B, Hu C, Zhu F, Liu X, Zhang J, et al. Clinical Characteristics of 138 Hospitalized Patients With 2019 Novel Coronavirus-Infected Pneumonia in Wuhan, China. JAMA. 2020;323(11):1061–1069. pmid:32031570
  30. 30. Simmons G, Zmora P, Gierer S, Heurich A, Pöhlmann S. Proteolytic activation of the SARS-coronavirus spike protein: Cutting enzymes at the cutting edge of antiviral research. Antivir Res. 2013;100(3):605–614. pmid:24121034
  31. 31. Millet JK, Whittaker GR. Host cell proteases: Critical determinants of coronavirus tropism and pathogenesis. Virus Res. 2015;202:120–134. pmid:25445340
  32. 32. Yang H, Yang M, Ding Y, Liu Y, Lou Z, Zhou Z, et al. The crystal structures of severe acute respiratory syndrome virus main protease and its complex with an inhibitor. Proc Natl Acad Sci U S A. 2003;100(23):13190–13195. pmid:14585926
  33. 33. Thiel V, Herold J, Schelle B, Siddell SG. Viral replicase gene products suffice for coronavirus discontinuous transcription. J Virol. 2001;75(14):6676–6681. pmid:11413334
  34. 34. Pillaiyar T, Manickam M, Namasivayam V, Hayashi Y, Jung S-H. An Overview of Severe Acute Respiratory Syndrome-Coronavirus (SARS-CoV) 3CL Protease Inhibitors: Peptidomimetics and Small Molecule Chemotherapy. J Med Chem. 2016;59(14):6595–6628. pmid:26878082
  35. 35. Skariyachan S, Challapilli SB, Packirisamy S, Kumargowda ST, Sridhar VS. Recent aspects on the pathogenesis mechanism, animal models and novel therapeutic interventions for Middle East Respiratory Syndrome Coronavirus infections. Front Microbiol. 2019;10:569. pmid:30984127
  36. 36. Baez-Santos YM, St John SE, Mesecar AD. The SARS-coronavirus papain-like protease: structure, function and inhibition by designed antiviral compounds. Antivir Res. 2015;115:21–38. pmid:25554382
  37. 37. Lei J, Kusov Y, Hilgenfeld R. Nsp3 of coronaviruses: Structures and functions of a large multi-domain protein. Antivir Res. 2018;149:58–74. pmid:29128390
  38. 38. Fan KQ, Wei P, Feng Q, Chen S, Huang C, Ma L, et al. Biosynthesis, purification, and substrate specificity of severe acute respiratory syndrome coronavirus 3C-like proteinase. J Biol Chem. 2004;279(3):1637–1642. pmid:14561748
  39. 39. Yang HT, Bartlam M, Rao ZH. Drug design targeting the main protease, the Achilles’ heel of coronaviruses. Curr Pharm Des. 2006;12(35):4573–4590. pmid:17168763
  40. 40. Tomar S, Johnston ML, St John SE, Osswald HL, Nyalapatla PR, Paul LN, et al. Ligand-induced Dimerization of Middle East Respiratory Syndrome (MERS) Coronavirus nsp5 Protease (3CLpro): IMPLICATIONS FOR nsp5 REGULATION AND THE DEVELOPMENT OF ANTIVIRALS. J Biol Chem. 2015;290(32):19403–19422. pmid:26055715
  41. 41. Wang F, Chen C, Tan W, Yang K, Yang H. Structure of Main Protease from Human Coronavirus NL63: Insights for Wide Spectrum Anti-Coronavirus Drug Design. Sci Rep. 2016;6.
  42. 42. Ye G, Deng F, Shen Z, Luo R, Zhao L, Xiao S, et al. Structural basis for the dimerization and substrate recognition specificity of porcine epidemic diarrhea virus 3C-like protease. Virology. 2016;494:225–235. pmid:27128350
  43. 43. Cui W, Cui S, Chen C, Chen X, Wang Z, Yang H, et al. The crystal structure of main protease from mouse hepatitis virus A59 in complex with an inhibitor. Biochem Biophys Res Commun. 2019;511(4):794–799. pmid:30833083
  44. 44. Lim L, Gupta G, Roy A, Kang J, Srivastava S, Shi J, et al. Structurally- and dynamically-driven allostery of the chymotrypsin-like proteases of SARS, Dengue and Zika viruses. Prog Biophys Mol Biol. 2019;143:52–66. pmid:30217495
  45. 45. Zhang L, Lin D, Sun X, Curth U, Drosten C, Sauerhering L, et al. Crystal structure of SARS-CoV-2 main protease provides a basis for design of improved alpha-ketoamide inhibitors. Science. 2020;368(6489):409–412. pmid:32198291
  46. 46. Zhu X, Wang D, Zhou J, Pan T, Chen J, Yang Y, et al. Porcine Deltacoronavirus nsp5 Antagonizes Type I Interferon Signaling by Cleaving STAT2. J Virol. 2017;91(10).
  47. 47. Yang HT, Xie W, Xue X, Yang K, Ma J, Liang W, et al. Design of wide-spectrum inhibitors targeting coronavirus main proteases. PLoS Biol. 2005;3(10):1742–1752.
  48. 48. Zhou J, Fang L, Yang Z, Xu S, Lv M, Sun Z, et al. Identification of novel proteolytically inactive mutations in coronavirus 3C-like protease using a combined approach. FASEB J. 2019;33(12):14575–14587. pmid:31690127
  49. 49. Chen YW, Yiu CB, Wong KY. Prediction of the SARS-CoV-2 (2019-nCoV) 3C-like protease (3CL (pro)) structure: virtual screening reveals velpatasvir, ledipasvir, and other drug repurposing candidates. F1000Res. 2020;9:129. pmid:32194944
  50. 50. Lai C-C, Jou M-J, Huang S-Y, Li S-W, Wan L, Tsai F-J, et al. Proteomic analysis of up-regulated proteins in human promonocyte cells expressing severe acute respiratory syndrome coronavirus 3C-like protease. Proteomics. 2007;7(9):1446–1460. pmid:17407183
  51. 51. Mielech AM, Chen Y, Mesecar AD, Baker SC. Nidovirus papain-like proteases: multifunctional enzymes with protease, deubiquitinating and deISGylating activities. Virus Res. 2014;194:184–190. pmid:24512893
  52. 52. Niemeyer D, Mösbauer K, Klein EM, Sieberg A, Mettelman RC, Mielech AM, et al. The papain-like protease determines a virulence trait that varies among members of the SARS-coronavirus species. PLoS Pathog. 2018;14(9):e1007296. pmid:30248143
  53. 53. Barretto N, Jukneliene D, Ratia K, Chen Z, Mesecar AD, Baker SC. The papain-like protease of severe acute respiratory syndrome coronavirus has deubiquitinating activity. J Virol. 2005;79(24):15189–15198. pmid:16306590
  54. 54. Harcourt BH, Jukneliene D, Kanjanahaluethai A, Bechill J, Severson KM, Smith CM, et al. Identification of severe acute respiratory syndrome coronavirus replicase products and characterization of papain-like protease activity. J Virol. 2004;78(24):13600–13612. pmid:15564471
  55. 55. Ratia K, Saikatendu KS, Santarsiero BD, Barretto N, Baker SC, Stevens RC, et al. Severe acute respiratory syndrome coronavirus papain-like protease: Structure of a viral deubiquitinating enzyme. Proc Natl Acad Sci U S A. 2006;103(15):5717–5722. pmid:16581910
  56. 56. Freitas BT, Durie IA, Murray J, Longo JE, Miller HC, Crich D, et al. Characterization and Noncovalent Inhibition of the Deubiquitinase and deISGylase Activity of SARS-CoV-2 Papain-Like Protease. ACS Infect Dis. 2020;6(8):2099–2109. pmid:32428392
  57. 57. Devaraj SG, Wang N, Chen Z, Chen Z, Tseng M, Barretto N, et al. Regulation of IRF-3-dependent innate immunity by the papain-like protease domain of the severe acute respiratory syndrome coronavirus. J Biol Chem. 2007;282(44):32208–32221. pmid:17761676
  58. 58. Frieman M, Ratia K, Johnston RE, Mesecar AD, Baric RS. Severe Acute Respiratory Syndrome Coronavirus Papain-Like Protease Ubiquitin-Like Domain and Catalytic Domain Regulate Antagonism of IRF3 and NF-kappa B Signaling. J Virol. 2009;83(13):6689–6705. pmid:19369340
  59. 59. Daczkowski CM, Dzimianski JV, Clasman JR, Goodwin O, Mesecar AD, Pegan SD. Structural Insights into the Interaction of Coronavirus Papain-Like Proteases and Interferon-Stimulated Gene Product 15 from Different Species. J Mol Biol. 2017;429(11):1661–1683. pmid:28438633
  60. 60. Matthews K, Schäfer A, Pham A, Frieman M. The SARS coronavirus papain like protease can inhibit IRF3 at a post activation step that requires deubiquitination activity. Virol J. 2014;11.
  61. 61. Hoffmann M, Kleine-Weber H, Schroeder S, Krüger N, Herrler T, Erichsen S, et al. SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and Is Blocked by a Clinically Proven Protease Inhibitor. Cell. 2020;181(2):271–280.e8. pmid:32142651
  62. 62. Pyrc K, Berkhout B, van der Hoek L. The novel human coronaviruses NL63 and HKU1. J Virol. 2007;81(7):3051–3057. pmid:17079323
  63. 63. Kuba K, T Ohto-Nakanishi Imai Y., Penninger JM. Trilogy of ACE2: a peptidase in the renin-angiotensin system, a SARS receptor, and a partner for amino acid transporters. Pharmacol Ther. 2010;128(1):119–128. pmid:20599443
  64. 64. Lambert DW, Yarski M, Warner FJ, Thornhill P, Parkin ET, Smith AI, et al. Tumor necrosis factor-alpha 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(34):30113–30119. pmid:15983030
  65. 65. Haga S, Yamamoto N, Nakai-Murakami C, Osawa Y, Tokunaga K, Sata T, et al. Modulation of TNF-alpha-converting enzyme by the spike protein of SARS-CoV and ACE2 induces TNF-alpha production and facilitates viral entry. Proc Natl Acad Sci U S A. 2008;105(22):7809–7814. pmid:18490652
  66. 66. 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(2):1293–1307. pmid:24227843
  67. 67. Kawase M, Shirato K, Matsuyama S, Taguchi F. Protease-Mediated Entry via the Endosome of Human Coronavirus 229E. J Virol. 2009;83(2):712–721.
  68. 68. Widagdo W, Sooksawasdi S, Ayudhya N, Hundie GB, Haagmans BL. Host Determinants of MERS-CoV Transmission and Pathogenesis. Viruses. 2019;11(3):280.
  69. 69. Earnest JT, Hantak MP, Li K, McCray PB, Perlman S, Gallagher T. The tetraspanin CD9 facilitates MERS-coronavirus entry by scaffolding host cell receptors and proteases. PLoS Pathog. 2017;13(7):e1006546. pmid:28759649
  70. 70. Li F. Structure, Function, and Evolution of Coronavirus Spike Proteins. Annu Rev Virol. 2016;3:37–261.
  71. 71. Park J-E, Li K, Barlan A, Fehr AR, Perlman S, McCray PB Jr, et al. Proteolytic processing of Middle East respiratory syndrome coronavirus spikes expands virus tropism. Proc Natl Acad Sci U S A. 2016;113(43):12262–12267. pmid:27791014
  72. 72. Matsuyama S, Shirato K, Kawase M, Terada Y, Kawachi K, Fukushi S, et al. Middle East respiratory syndrome coronavirus spike protein is not activated directly by cellular furin during viral entry into target cells. J Virol. 2018;92(19):e00683–e00618. pmid:30021905
  73. 73. Wang Q, Qiu Y, Li J-Y, Zhou Z-J, Liao C-H, Ge X-Y. A Unique Protease Cleavage Site Predicted in the Spike Protein of the Novel Pneumonia Coronavirus (2019-nCoV) Potentially Related to Viral Transmissibility. Virol Sin. 2020;35(3):337–339. pmid:32198713
  74. 74. Belouzard S, Madu I, Whittaker GR. Elastase-mediated Activation of the Severe Acute Respiratory Syndrome Coronavirus Spike Protein at Discrete Sites within the S2 Domain. J Biol Chem. 2010;285(30):22756–22761.
  75. 75. 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(2):873–882. pmid:21068237
  76. 76. Bertram S, Glowacka I, Müller MA, Lavender H, Gnirss K, Nehlmeier I, et al. Cleavage and Activation of the Severe Acute Respiratory Syndrome Coronavirus Spike Protein by Human Airway Trypsin-Like Protease. J Virol. 2011;85(24):13363–13372. pmid:21994442
  77. 77. Ou X, Liu Y, Lei X, Li P, Mi D, Ren L, et al. Characterization of spike glycoprotein of SARS-CoV-2 on virus entry and its immune cross-reactivity with SARS-CoV. Nat Commun. 2020;11(1):1620. pmid:32221306
  78. 78. Gierer S, Bertram S, Kaup F, Wrensch F, Heurich A, Krämer-Kühl A, et al. The Spike Protein of the Emerging Betacoronavirus EMC Uses a Novel Coronavirus Receptor for Entry, Can Be Activated by TMPRSS2, and Is Targeted by Neutralizing Antibodies. J Virol. 2013;87(10):5502–5511. pmid:23468491
  79. 79. Shirato K, Kawase M, Matsuyama S. Middle East Respiratory Syndrome Coronavirus Infection Mediated by the Transmembrane Serine Protease TMPRSS2. J Virol. 2013;87(23):12552–12561. pmid:24027332
  80. 80. Shirato K, Kanou K, Kawase M, Matsuyama S. Clinical Isolates of Human Coronavirus 229E Bypass the Endosome for Cell Entry. J Virol. 2017;91(1):e01387–e01316. pmid:27733646
  81. 81. Shirato K, Kawase M, Matsuyama S. Wild-type human coronaviruses prefer cell-surface TMPRSS2 to endosomal cathepsins for cell entry. Virology. 2018;517:9–15. pmid:29217279
  82. 82. Adedeji AO, Severson W, Jonsson C, Singh K, Weiss SR, Sarafianos SG. Novel Inhibitors of Severe Acute Respiratory Syndrome Coronavirus Entry That Act by Three Distinct Mechanisms. J Virol. 2013;87(14):8017–8028. pmid:23678171
  83. 83. Du L, Kao RY, Zhou Y, He Y, Zhao G, Wong C, et al. Cleavage of spike protein of SARS coronavirus by protease factor Xa is associated with viral infectivity. Biochem Biophys Res Commun. 2007;359(1):174–179. pmid:17533109
  84. 84. Simmons G, Bertram S, Glowacka I, Steffen I, Chaipan C, Agudelo J, et al. Different host cell proteases activate the SARS-coronavirus spike-protein for cell-cell and virus-cell fusion. Virology. 2011;413(2):265–274. pmid:21435673
  85. 85. Schneider M, Ackermann K, Stuart M, Wex C, Protzer U, Schätzl HM, et al. Severe Acute Respiratory Syndrome Coronavirus Replication Is Severely Impaired by MG132 due to Proteasome-Independent Inhibition of M-Calpain. J Virol. 2012;86(18):10112–10122. pmid:22787216
  86. 86. Kleine-Weber H, Elzayat MT, Hoffmann M, Pöhlmann S. Functional analysis of potential cleavage sites in the MERS-coronavirus spike protein. Sci Rep. 2018;8. pmid:29311689
  87. 87. Ashraf H. Cathepsin enzyme provides clue to SARS infection. Drug Discov Today. 2005;10(21):1409.
  88. 88. Muller S, Dennemarker J, Reinheckel T. Specific functions of lysosomal proteases in endocytic and autophagic pathways. Biochim Biophy Acta. 2012;1824(1):34–43.
  89. 89. Rawlings ND, Barrett AJ, Bateman A. MEROPS: the peptidase database. Nucleic Acids Res. 2010;38:D227–D233. pmid:19892822
  90. 90. Turk B. Targeting proteases: successes, failures and future prospects. Nat Rev Drug Discov. 2006;5(9):785–799. pmid:16955069
  91. 91. Turk V, Stoka V, Vasiljeva O, Renko M, Sun T, Turk B, et al. Cysteine cathepsins: From structure, function and regulation to new frontiers. Biochimica Biophy Acta. 2012;1824(1):68–88.
  92. 92. Chwieralski CE, Welte T, Buhling F. Cathepsin-regulated apoptosis. Apoptosis. 2006;11(2):143–149. pmid:16502253
  93. 93. Turk V, Turk B, Turk D. Lysosomal cysteine proteases: facts and opportunities. EMBO J. 2001;20(17):4629–4633. pmid:11532926
  94. 94. Turk B, Turk D, Turk V. Lysosomal cysteine proteases: more than scavengers. BBA-Protein Struct M. 2000;1477(1–2):98–111.
  95. 95. Berdowska I. Cysteine proteases as disease markers. Clin Chim Acta. 2004;342(1–2):41–69. pmid:15026265
  96. 96. Pislar A, Nanut MP, Kos J. Lysosomal cysteine peptidases—Molecules signaling tumor cell death and survival. Semin Cancer Biol. 2015;35:168–179. pmid:26255843
  97. 97. Vasiljeva O, Reinheckel T, Peters C, Turk D, Turk V, Turk B. Emerging roles of cysteine cathepsins in disease and their potential as drug targets. Curr Pharm Des. 2007;13(4):387–403. pmid:17311556
  98. 98. Mohamed MM, Sloane BF. Cysteine cathepsins: multifunctional enzymes in cancer. Nat Rev Cancer. 2006;6(10):764–775. pmid:16990854
  99. 99. Pislar A, Kos J. Cysteine Cathepsins in Neurological Disorders. Mol Neurobiol. 2014;49(2):1017–1030. pmid:24234234
  100. 100. Huang IC, Bosch BJ, Li W, Farzan M, Rottier P, Choe H. SARS-CoV, but not HCoV-NL63, utilizes cathepsins to infect cells: viral entry. The Nidoviruses. 2006;581:335–338.
  101. 101. Simmons G, Gosalia DN, Rennekamp AJ, Reeves JD, Diamond SL, Bates P. Inhibitors of cathepsin L prevent severe acute respiratory syndrome coronavirus entry. Proc Natl Acad Sci U S A. 2005;102(33):11876–11881. pmid:16081529
  102. 102. Huang I-C, Bosch BJ, Li F, Li W, Lee KH, Ghiran S, et al. SARS coronavirus, but not human coronavirus NL63, utilizes cathepsin L to infect ACE2-expressing cells. J Biol Chem. 2006;281(6):3198–3203. pmid:16339146
  103. 103. Iwata-Yoshikawa N, Okamura T, Shimizu Y, Hasegawa H, Takeda M, Nagata N. TMPRSS2 Contributes to Virus Spread and Immunopathology in the Airways of Murine Models after Coronavirus Infection. J Virol. 2019;93(6):e01815–e01818. pmid:30626688
  104. 104. Kawase M, Shirato K, van der Hoek L, Taguchi F, Matsuyama S. Simultaneous Treatment of Human Bronchial Epithelial Cells with Serine and Cysteine Protease Inhibitors Prevents Severe Acute Respiratory Syndrome Coronavirus Entry. J Virol. 2012;86(12):6537–6545. pmid:22496216
  105. 105. Zhou Y, Vedantham P, Lu K, Agudelo J, R Carrion Jr, Nunneley JW, et al. Protease inhibitors targeting coronavirus and filovirus entry. Antivir Res. 2015;116:76–84. pmid:25666761
  106. 106. Wilcox D, Mason RW. Inhibition of Cysteine Proteinases in Lysosomes and Whole Cells. Biochem J. 1992;285:495–502. pmid:1637341
  107. 107. Murata M, Miyashita S, Yokoo C, Tamai M, Hanada K, Hatayama K, et al. Novel epoxysuccinyl peptides. Selective inhibitors of cathepsin B, in vitro. FEBS Lett. 1991;280(2):307–310. pmid:2013328
  108. 108. Prompetchara E, Kettoy C, Palaga T. Immune responses in COVID-19 and potential vaccines: Lessons learned from SARS and MERS epidemic. Asian Pac J Allergy Immunol. 2020;38(1):1–9. pmid:32105090
  109. 109. Kawai T, Akira S. Innate immune recognition of viral infection. Nat Immunol. 2006;7(2):131–137. pmid:16424890
  110. 110. Ewald SE, Engel A, Lee J, Wang M, Bogyo M, Barton GM. Nucleic acid recognition by Toll-like receptors is coupled to stepwise processing by cathepsins and asparagine endopeptidase. J Exp Med. 2011;208(4):643–651. pmid:21402738
  111. 111. Sepulveda FE, Maschalidi S, Colisson R, Heslop L, Ghirelli C, Sakka E, et al. Critical role for asparagine endopeptidase in endocytic Toll-like receptor signaling in dendritic cells. Immunity. 2009;31(5):737–748. pmid:19879164
  112. 112. Ewald SE, Lee BL, Lau L, Wickliffe KE, Shi G-P, Chapman HA, et al. The ectodomain of Toll-like receptor 9 is cleaved to generate a functional receptor. Nature. 2008;456(7222):658–662. pmid:18820679
  113. 113. Maschalidi S, Hässler S, Blanc F, Sepulveda FE, Tohme M, Chignard M, et al. Asparagine Endopeptidase Controls Anti-Influenza Virus Immune Responses through TLR7 Activation. PLoS Pathog. 2012;8(8):e1002841. pmid:22916010
  114. 114. Kindler E, Thiel V. SARS-CoV and IFN: Too Little, Too Late. Cell Host Microbe. 2016;19(2):139–141. pmid:26867172
  115. 115. Chen J, Lau YF, Lamirande EW, Paddock CD, Bartlett JH, Zaki SR, et al. Cellular Immune Responses to Severe Acute Respiratory Syndrome Coronavirus (SARS-CoV) Infection in Senescent BALB/c Mice: CD4(+) T Cells Are Important in Control of SARS-CoV Infection. J Virol. 2010;84(3):1289–1301. pmid:19906920
  116. 116. Chu H, Chan JF, Wang Y, Yuen TT, Chai Y, Hou Y, et al. Comparative replication and immune activation profiles of SARS-CoV-2 and SARS-CoV in human lungs: an ex vivo study with implications for the pathogenesis of COVID-19. Clin Infect Dis. 2020;71(6):1400–1409. pmid:32270184
  117. 117. Gruber C. Impaired interferon signature in severe COVID-19. Nat Rev Immunol. 2020.
  118. 118. Nakagawa TY, Rudensky AY. The role of lysosomal proteinases in MHC class II-mediated antigen processing and presentation. Immunol Rev. 1999;172:121–129. pmid:10631942
  119. 119. Jakoš T, Pišlar A, Pečar Fonović U, Kos J. Lysosomal peptidases in innate immune cells: implications for cancer immunity. Cancer Immunol Immunother. 2020;69(2):275–283. pmid:31813053
  120. 120. Villadangos JA, Bryant RAR, Deussing J, Driessen C, Lennon‐Duménil A‐ M, Riese RJ, et al. Proteases involved in MHC class II antigen presentation. Immunol Rev. 1999;172:109–120. pmid:10631941
  121. 121. Hu W, Yen Y-T, Singh S, Kao C-L. BA Wu-Hsieh. SARS-CoV Regulates Immune Function-Related Gene Expression in Human Monocytic Cells. Viral Immunol. 2012;25(4):277–288. pmid:22876772
  122. 122. Qin C, Zhou L, Hu Z, Zhang S, Yang S, Tao Y, et al. Dysregulation of immune response in patients with COVID-19 in Wuhan. China Clin Infect Dis. 2020;71(15):762–768. pmid:32161940
  123. 123. Tan L, Wang Q, Zhang D, Ding J, Huang Q, Tang Y-Q, et al. Lymphopenia predicts disease severity of COVID-19: a descriptive and predictive study. Signal Transduct Target Ther. 2020;5(1):33. pmid:32296069
  124. 124. Huang C, Wang Y, Li X, Ren L, Zhao J, Hu Y, et al. Clinical features of patients infected with 2019 novel coronavirus in Wuhan. China Lancet. 2020;395(10223):497–506. pmid:31986264
  125. 125. Giamarellos-Bourboulis EJ, Netea MG, Rovina N, Akinosoglou K, Antoniadou A, Antonakos N, et al. Complex Immune Dysregulation in COVID-19 Patients with Severe Respiratory Failure. Cell Host Microbe. 2020;27(6):992–1000. pmid:32320677
  126. 126. Zheng M, Gao Y, Wang G, Song G, Liu S, Sun D, et al. Functional exhaustion of antiviral lymphocytes in COVID-19 patients. Cell Mol Immunol. 2020;17(5):533–535. pmid:32203188
  127. 127. Xu B, Fan C-Y, Wang A-L, Zou Y-L, Yu Y-H, He C, et al. Suppressed T cell-mediated immunity in patients with COVID-19: A clinical retrospective study in Wuhan. China J Infect. 2020;81(1):e51–e60. pmid:32315725
  128. 128. Yue Y, Nabar NR, Shi C-S, Kamenyeva O, Xiao X, Hwang I-Y, et al. SARS-Coronavirus Open Reading Frame-3a drives multimodal necrotic cell death. Cell Death Dis. 2018;9.
  129. 129. Xiong Y, Liu Y, Cao L, Wang D, Guo M, Jiang A, et al. Transcriptomic characteristics of bronchoalveolar lavage fluid and peripheral blood mononuclear cells in COVID-19 patients. Emerg Microbes Infect. 2020;9(1):761–770.
  130. 130. Stoka V, Turk B, Schendel SL, Kim TH, Cirman T, Snipas SJ, et al. Lysosomal protease pathways to apoptosis—Cleavage of Bid, not pro-caspases, is the most likely route. J Biol Chem. 2001;276(5):3149–3157. pmid:11073962
  131. 131. Cirman T, Oresić K, Mazovec GD, Turk V, Reed JC, Myers RM, et al. Selective disruption of lysosomes in HeLa cells triggers apoptosis mediated by cleavage of bid by multiple papain-like lysosomal cathepsins. J Biol Chem. 2004;279(5):3578–3587. pmid:14581476
  132. 132. Brojatsch J, Lima H, Palliser D, Jacobson LS, Muehlbauer SM, Furtado R, et al. Distinct cathepsins control necrotic cell death mediated by pyroptosis inducers and lysosome-destabilizing agents. Cell Cycle. 2015;14(7):964–972. pmid:25830414
  133. 133. Nanut MP, Sabotič J, Švajger U, Jewett A, Kos J. Cystatin F Affects Natural Killer Cell Cytotoxicity. Front Immunol. 2017;8:1459. pmid:29180998
  134. 134. Liang PH. Characterization and inhibition of SARS-coronavirus main protease. Curr Top Med Chem. 2006;6(4):361–376. pmid:16611148
  135. 135. Nakajima T, Kajikawa Y, Hirose T, Tokiwa N, Hanada K, Fukuhara T. Surgical-orthodontic approach to skeletal class III malocclusion. Analysis of 45 cases with evaluation of curved oblique osteotomy and sliding osteotomy. Int J Oral Surg. 1978;7(4):274–280. pmid:100438
  136. 136. Li Q, Kang C. Progress in Developing Inhibitors of SARS-CoV-2 3C-Like Protease. Microorganisms. 2020:8(8).
  137. 137. Liu Y, Liang C, Xin L, Ren X, Tian L, Ju X, et al. The development of Coronavirus 3C-Like protease (3CL(pro)) inhibitors from 2010 to 2020. Eur J Med Chem. 2020;206:112711. pmid:32810751
  138. 138. Mills JS. Peptides derived from HIV-1, HIV-2, Ebola virus, SARS coronavirus and coronavirus 229E exhibit high affinity binding to the formyl peptide receptor. Biochim Biophys Acta. 2006;1762(7):693–703. pmid:16842982
  139. 139. Collins AR, Grubb A. Cystatin D, a natural salivary cysteine protease inhibitor, inhibits coronavirus replication at its physiologic concentration. Oral Microbiol Immunol. 1998;13(1):59–61. pmid:9573825
  140. 140. Collins AR, Grubb A. Inhibitory Effects of Recombinant Human Cystatin-C on Human Coronaviruses. Antimicrob Agents Chemother. 1991;35(11):2444–2446. pmid:1804023
  141. 141. Breznik B, Mitrović A, Lah TT, Kos J. Cystatins in cancer progression: More than just cathepsin inhibitors. Biochimie. 2019;166:233–250. pmid:31071357
  142. 142. Magister S, Kos J. Cystatins in Immune System. J Cancer. 2013;4(1):45–56. pmid:23386904
  143. 143. Prunk M, Perišić Nanut M, Sabotič J, Kos J. Cystatins, cysteine peptidase inhibitors, as regulators of immune cell cytotoxicity. Period Biol. 2016;118(4):353–362.
  144. 144. Balbín M, Hall A, Grubb A, Mason RW, López-Otín C, Abrahamson M. Structural and functional characterization of two allelic variants of human cystatin D sharing a characteristic inhibition spectrum against mammalian cysteine proteinases. J Biol Chem. 1994;269(37):23156–23162. pmid:8083219
  145. 145. Shah PP, Wang T, Kaletsky RL, Myers MC, Purvis JE, Jing H, et al. A Small-Molecule Oxocarbazate Inhibitor of Human Cathepsin L Blocks Severe Acute Respiratory Syndrome and Ebola Pseudotype Virus Infection into Human Embryonic Kidney 293T cells. Mol Pharmacol. 2010;78(2):319–324. pmid:20466822
  146. 146. Smieszek SP, Przychodzen BP, Polymeropoulos MH. Amantadine disrupts lysosomal gene expression: A hypothesis for COVID19 treatment. Int J Antimicrob Agents. 2020;106004.
  147. 147. Zhou N, Pan T, Zhang J, Li Q, Zhang X, Bai C, et al. Glycopeptide Antibiotics Potently Inhibit Cathepsin L in the Late Endosome/Lysosome and Block the Entry of Ebola Virus, Middle East Respiratory Syndrome Coronavirus (MERS-CoV), and Severe Acute Respiratory Syndrome Coronavirus (SARS-CoV). J Biol Chem. 2016;291(17):9218–9232. pmid:26953343
  148. 148. Rabaan AA, Alahmed SH, Bazzi AM, Alhani HM. A review of candidate therapies for Middle East respiratory syndrome from a molecular perspective. J Med Microbiol. 2017;66(9):1261–1274. pmid:28855003
  149. 149. Elshabrawy HA, Fan J, Haddad CS, Ratia K, Broder CC, Caffrey M, et al. Identification of a Broad-Spectrum Antiviral Small Molecule against Severe Acute Respiratory Syndrome Coronavirus and Ebola, Hendra, and Nipah Viruses by Using a Novel High-Throughput Screening Assay. J Virol. 2014;88(8):4353–4365. pmid:24501399
  150. 150. Bertram S, Dijkman R, Habjan M, Heurich A, Gierer S, Glowacka I, et al. TMPRSS2 Activates the Human Coronavirus 229E for Cathepsin-Independent Host Cell Entry and Is Expressed in Viral Target Cells in the Respiratory Epithelium. J Virol. 2013;87(11):6150–6160. pmid:23536651
  151. 151. Simmons G, Rennekamp AJ, Bates P. Proteolysis of SARS-associated coronavirus spike glycoprotein. The Nidoviruses. 2006;581:235–240.
  152. 152. Kos J, Mitrovic A, Mirkovic B. The current stage of cathepsin B inhibitors as potential anticancer agents. Future Med Chem. 2014;6(11):1355–1371. pmid:25163003
  153. 153. Pogorzelska A, Zolnowska B, Bartoszewski R. Cysteine cathepsins as a prospective target for anticancer therapies-current progress and prospects. Biochimie. 2018;151:85–106. pmid:29870804
  154. 154. Dana D, Pathak SK. A Review of Small Molecule Inhibitors and Functional Probes of Human Cathepsin L. Molecules. 2020;25(3).
  155. 155. Khaket TP, Kwon TK, Kang SC. Cathepsins: Potent regulators in carcinogenesis. Pharmacol Ther. 2019;198:1–19. pmid:30763594
  156. 156. Rebbaa A, Chu F, Sudha T, Gallati C, Dier U, Dyskin E, et al. The Anti-angiogenic Activity of NSITC, a Specific Cathepsin L Inhibitor. Anticancer Res. 2009;29(11):4473–4481. pmid:20032394
  157. 157. Yasuma T, Oi S, Choh N, Nomura T, Furuyama N, Nishimura A, et al. Synthesis of peptide aldehyde derivatives as selective inhibitors of human cathepsin L and their inhibitory effect on bone resorption. J Med Chem. 1998;41(22):4301–4308. pmid:9784105
  158. 158. Sudhan DR, Siemann DW. Cathepsin L inhibition by the small molecule KGP94 suppresses tumor microenvironment enhanced metastasis associated cell functions of prostate and breast cancer cells. Clin Exp Metastasis. 2013;30(7):891–902. pmid:23748470
  159. 159. Traugott U, Snyder S, Norton WT, Raine CS. Characterization of antioligodendrocyte serum. Ann Neurol. 1978;4(5):431–439. pmid:367262
  160. 160. Mirković B, Markelc B, Butinar M, Mitrović A, Sosič I, Gobec S, et al. Nitroxoline impairs tumor progression in vitro and in vivo by regulating cathepsin B activity. Oncotarget. 2015;6(22):19027–19042. pmid:25848918
  161. 161. Mirković B, Renko M, Turk S, Sosič I, Jevnikar Z, Obermajer N, et al. Novel mechanism of cathepsin B inhibition by antibiotic nitroxoline and related compounds. ChemMedChem. 2011;6(8):1351–1356. pmid:21598397
  162. 162. Mitrovic A, Kos J. Nitroxoline: repurposing its antimicrobial to antitumor application. Acta Biochim Pol. 2019;66(4):521–531. pmid:31834689
  163. 163. Kamboj RC, Raghav N, Mittal A, Khurana S, Sadana R, Singh H. Effects of some antituberculous and anti-leprotic drugs on cathepsins B, H and L. Indian J Clin Biochem. 2003;18(2):39–47. pmid:23105391
  164. 164. Mitrović A, Kljun J, Sosič I, Gobec S, Turelb I, Kos J. Clioquinol-ruthenium complex impairs tumour cell invasion by inhibiting cathepsin B activity. Dalton Trans. 2016;45(42):16913–16921. pmid:27711842
  165. 165. Itoh Y, Shimura H, Ito M, Watanabe N, Yamagishi M, Tamai M, et al. A novel hepatoprotective gamma-lactone, MH-031. I. Discovery, isolation, physico-chemical properties and structural elucidation. J Antibiot (Tokyo). 1991;44(8):832–837.
  166. 166. Buttle DJ, Murata M, Knight CG, Barrett AJ, et al. CA074 methyl ester: a proinhibitor for intracellular cathepsin B. Arch Biochem Biophys. 1992;299(2):377–380. pmid:1444478
  167. 167. Katunuma N, Murata E, Kakegawa H, Matsui A, Tsuzuki H, Tsuge H, et al. Structure based development of novel specific inhibitors for cathepsin L and cathepsin S in vitro and in vivo. FEBS Lett. 1999;458(1):6–10. pmid:10518923
  168. 168. Hardes K. Steinmetzer T. The Antiviral Potential of Host Protease Inhibitors, in Activation of Viruses by Host Proteases, E. Böttcher-Friebertshäuser, W. Garten, and H.D. Klenk, editors. 2018;279–325.
  169. 169. Holshue ML, DeBolt C, Lindquist S, Lofy KH, Wiesman J, Bruce H, et al. First Case of 2019 Novel Coronavirus in the United States. N Engl J Med. 2020;382(10):929–936. pmid:32004427
  170. 170. Agbowuro AA, Huston WM, Gamble AB, Tyndall JDA. Proteases and protease inhibitors in infectious diseases. Med Res Rev. 2018;38(4):1295–1331. pmid:29149530
  171. 171. Baer GS, Ebert DH, Chung CJ, Erickson AH, Dermody TS. Mutant cells selected during persistent reovirus infection do not express mature cathepsin L and do not support reovirus disassembly. J Virol. 1999;73:9532–9543. pmid:10516062
  172. 172. Ebert DH, Deussing J, Peters C, Dermody TS. Cathepsin L and cathepsin B mediate reovirus disassembly in murine fibroblast cells. J Biol Chem. 2002;277:24609–24617. pmid:11986312
  173. 173. Golden JW, Linke J, Schmechel S, Thoemke K, Schiff LA. Addition of exogenous protease facilitates reovirus infection in many restrictive cells. J Virol. 2002;76:7430–7443. pmid:12097555
  174. 174. Jane-Valbuena J, Breun LA, Schiff LA, Nibert ML. Sites and determinants of early cleavages in the proteolytic processing pathway of reovirus surface protein sigma 3. J Virol. 2002;76:5184–5197. pmid:11967333
  175. 175. Golden JW, Bahe JA, Lucas WT, Nibert ML, Schiff LA. Cathepsin S supports acid-independent infection by some reoviruses. J Biol Chem. 2004;279:8547–8557. pmid:14670972
  176. 176. Chandran K, Sullivan NJ, Felbor U, Whelan SP, Cunningham JM. Endosomal proteolysis of the Ebola virus glycoprotein is necessary for infection. Science. 2005;308:1643–1645. pmid:15831716
  177. 177. Asboth B, Majer Z, Polgar L. Cysteine proteases: the S2P2 hydrogen bond is more important for catalysis than is the analogous S1P1 bond. FEBS Lett. 1988;233:339–341. pmid:3384097
  178. 178. Choe Y, Leonetti F, Greenbaum DC, Lecaille F, Bogyo M, Bromme D, et al. Substrate profiling of cysteine proteases using a combinatorial peptide library identifies functionally unique specificities. J Biol Chem. 2006;281:12824–12832. pmid:16520377
  179. 179. Diederich S, Thiel L, Maisner A. Role of endocytosis and cathepsin-mediated activation in Nipah virus entry. Virology. 2008;375:391–400. pmid:18342904
  180. 180. Regan AD, Shraybman R, Cohen RD, Whittaker GR. Differential role for low pH and cathepsin-mediated cleavage of the viral spike protein during entry of serotype II feline coronaviruses. Vet Microbiol. 2008;132:235–248. pmid:18606506
  181. 181. Nanbo A, Imai M, Watanabe S, Noda T, Takahashi K, Neumann G, et al. Ebolavirus is internalized into host cells via macropinocytosis in a viral glycoprotein-dependent manner. PLoS Pathog. 2010;6.
  182. 182. Gnirss K, Kuhl A, Karsten C, Glowacka I, Bertram S, Kaup F, et al. Cathepsins B and L activate Ebola but not Marburg virus glycoproteins for efficient entry into cell lines and macrophages independent of TMPRSS2 expression. Virology. 2012;424:3–10. pmid:22222211
  183. 183. Alshaeri HK, Natto ZS. A contemporary look at COVID-19 medications: available and potentially effective drugs. Eur Rev Med Pharmacol Sci. 2020;24(17):9188–9195. pmid:32965013