Next Article in Journal
Modulatory Roles of AHR, FFAR2, FXR, and TGR5 Gene Expression in Metabolic-Associated Fatty Liver Disease and COVID-19 Outcomes
Previous Article in Journal
Genetic Reassortment in a Child Coinfected with Two Influenza B Viruses, B/Yamagata Lineage and B/Victoria-Lineage Strains
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Host Cell Proteases Involved in Human Respiratory Viral Infections and Their Inhibitors: A Review

by
Bailey Lubinski
1 and
Gary R. Whittaker
2,*
1
Department of Microbiology & Immunology, College of Veterinary Medicine, Cornell University, Ithaca, NY 14850, USA
2
Department of Microbiology & Immunology and Public & Ecosystem Health, College of Veterinary Medicine, Cornell University, Ithaca, NY 14850, USA
*
Author to whom correspondence should be addressed.
Viruses 2024, 16(6), 984; https://doi.org/10.3390/v16060984
Submission received: 13 May 2024 / Revised: 6 June 2024 / Accepted: 11 June 2024 / Published: 19 June 2024
(This article belongs to the Section Human Virology and Viral Diseases)

Abstract

:
Viral tropism is most commonly linked to receptor use, but host cell protease use can be a notable factor in susceptibility to infection. Here we review the use of host cell proteases by human viruses, focusing on those with primarily respiratory tropism, particularly SARS-CoV-2. We first describe the various classes of proteases present in the respiratory tract, as well as elsewhere in the body, and incorporate the targeting of these proteases as therapeutic drugs for use in humans. Host cell proteases are also linked to the systemic spread of viruses and play important roles outside of the respiratory tract; therefore, we address how proteases affect viruses across the spectrum of infections that can occur in humans, intending to understand the extrapulmonary spread of SARS-CoV-2.

1. Introduction

Among the hurdles that viruses must overcome is the inability to actively adapt to new environments. One way to overcome this hurdle is to utilize host components to modify and activate viral proteins. The human body contains hundreds of proteases, enzymes that process the peptide bonds of both proproteins and proteins through hydrolytic cleavage. Proteases are essential for countless basic human life processes such as hormone and neurotransmitter activation and release. Likewise, proteolytic cleavage can be important for several steps in the life cycles of viruses, such as host cell entry, uncoating, and virion formation. While many viruses encode their own proteases—which are excellent drug targets—the ability to co-opt host cell proteases can be a major advantage, but one that has received comparably little attention for drug development. For respiratory viruses, the proteases present in the respiratory tract can be used for various viral functions, which can impact pathogenesis and cell tropism. As host proteases do not undergo the same selective evolutionary pressure as viral proteases, inhibitors that target host proteases may elicit a slower viral evolutionary response towards developing drug resistance. Identifying potent inhibitors of respiratory viruses has long been a goal for researchers and has become more of a priority with the advent of the COVID-19 pandemic. SARS-CoV-2, the causative agent of COVID-19, uses several proteases that were previously well known to be involved in the life cycles of other respiratory viruses, as well as illuminating new factors for viral infection. This has allowed researchers to explore foundational research on those viruses and use it as a guide for current drug development. This review will primarily focus on the proteases implicated in SARS-CoV-2 infection, the inhibitors used to disrupt that infection, and the background of these proteases and protease inhibitors in other respiratory viruses.

2. Respiratory Proteases

Proteases recognize cleavage substrates through a variety of physical properties. The requirements that a substrate must meet in order to be cleaved vary by the protease of interest. Some proteases, for example, digestive enzymes such as trypsin, are quite general in their requirements and, thus, retain the ability to cleave a broad array of proteins. Others such as furin are much more specific. Among the properties that determine cleavability are the steric availability of the cleavage site to the protease, as well as the charge, polarity, and glycosylation status of the amino acids within and surrounding the cleavage site. Substrate cleavage sites are typically described using a numbering system introduced by Schechter and Berger [1]. The amino acids on the N terminal side of the cleavage site are numbered P1, P2, etc., with the numbering increasing as distance from the scissile bond increases. On the C terminal side, amino acids are similarly numbered, but as P1′, P2′, etc., as depicted in Figure 1.
Respiratory proteases conduct a variety of important roles in the respiratory tract. They are involved in ciliary function, ion transport, mucus expression, and the immune response, and their dysregulation is tied to multiple disease states [2]. Proteases are categorized by their catalytic residue, and there are four main types of proteases found in the respiratory tract: serine, cysteine, metallo-, and aspartyl proteases. An excellent review of host cell proteases and their role in the cell’s immune response to respiratory virus infections has previously been published [3]. In this review, we shall focus on serine, cysteine, and metalloproteases and their well-characterized roles in facilitating the entry into the cell and the replication of respiratory viruses.
Proteases from all three of these categories are co-opted by SARS-CoV-2, and their presence in the host has important implications for both transmission of the virus and pathology. The expression of ACE2 and TMPRSS2 that is highly homologous to human ACE2 and TMPRSS2 is known to be an important susceptibility factor in animal infection with SARS-CoV-2 [4,5,6,7,8,9,10]. Additionally, many of the human proteases are polymorphic. Genetic variants of furin, TMPRSS2, and ACE2 are all present in the human population at significant levels, and several studies have found that these variants have an impact on COVID-19 presentation [11,12,13,14,15]. This highlights that continued study of host proteases is essential for a more comprehensive understanding of viral infections and the factors that increase the potential for zoonotic spillover events.

3. Serine Proteases

Serine proteases catalyze the cleavage of proteins via a nucleophile on a serine amino acid in the catalytic site of the enzyme. This serine can be a part of a catalytic triad or dyad. There are 50 different serine protease classes, further divided into 14 clans [16]. A wide variety of serine proteases have been found to proteolytically cleave the glycoproteins of respiratory viruses. Table 1 depicts selected well-studied proteases, their cleavage preferences, and the viruses with which they are associated. Serine proteases are the most common class of protease present in the lungs and, thus, are often utilized by viruses.

3.1. Furin

Furin is a proprotein convertase that minimally cleaves at an R-X-X-R amino acid motif, a polybasic region that must be within an accessible portion of a protein. Furin is present in both the trans-Golgi network and endosomes, and it is essential for the processing of many host substrates crucial to homeostasis and the immune response. While the canonically recognized cleavage site is R-X-X-R, the “furin cleavage site”, or FCS, is actually much more complex, with furin interacting with its cleavage substrate at 30 amino acids spanning the cleavage site, with some amino acids interacting more strongly than others. A study looking at the interaction of furin with 130 known furin cleavage sites found that the properties of 20 of these amino acids were essential determinants of furin’s ability to cleave [53]. The canonical R-X-X-R motif lies within an eight-amino-acid-long core region of this larger motif. Within the core region, positively charged residues and flexibility are requirements, with specific amino acids needed at certain positions. Outside of the core region, the polarity of amino acids and accessibility to solvents are also determining factors of cleavage. Recent work on MHV has shown that what may appear to be minor changes in the FCS can have a pronounced effect [54]. In addition to specific amino acid side-chain requirements, it has been found that the glycosylation patterns of proteins (both N- and O-linked) can sometimes abrogate furin cleavage [55]. Furin cleavage requires a pH of 6–8.5 and the presence of Ca2+ in order to be able to cleave [56].
Furin cleavage of influenza hemagglutinin (HA) has been extensively studied as a pathogenicity determinant in the context of highly pathogenic avian influenza [57]. The HA cleavage site is essential for influenza infectivity as, when cleaved, the N-terminus of the fusion peptide is exposed. However, the cleavage of HA exhibits a certain amount of variability. Individual strains of influenza (low pathogenicity or high pathogenicity) may carry either a monobasic or polybasic cleavage site, which determines whether or not their HA can undergo furin cleavage or needs to be cleaved by another protease, such as HAT, TMPRSS2, and other trypsin-like proteases. The presence of a monobasic or polybasic cleavage site is the defining characteristic that separates low-pathogenic viruses from highly pathogenic viruses [58]. Strains of influenza with a monobasic cleavage site are reliant on the presence of trypsin-like proteases on the surface of cells during cell entry. These proteases are only expressed in certain cells; therefore, infection is restricted to those cells. Strains with a polybasic cleavage site are able to utilize furin as virions are being made; therefore, HA is already fusion-competent upon its release and the virions are able to infect cells that do not express trypsin-like proteases. While furin is a ubiquitous protease, it should be noted, however, that cells often express very low amounts of furin, and pathogenesis may be linked to expression levels in different tissues and cell types, e.g., endothelial cells [59]. Also, while other animal species do have furin homologs, there may be differences in the levels and localization of expression that could lead to alternate effects of polybasic cleavage sites on pathogenicity in those animals. However, in both ferrets and poultry, the ability of furin to process HA has been shown to be a determinant of the pathogenicity of highly pathogenic avian influenza (HPAI) [60,61].
Viral envelope proteins are typically cleaved at a single position, but in some cases, there are two distinct cleavage sites (see examples in Figure 2). This was first noted for SARS-CoV-1 and then with other coronaviruses. Depending on the individual virus, furin cleavage may be considered a priming event, before the activation protease subsequently cleaves adjacent to the fusion peptide, as with influenza HA, HIV-1 Env, and for paramyxo- and pneumoviruses—this activation is typically through non-furin proteases (see below). Respiratory syncytial virus (RSV, or human orthopneumovirus) is distinct in having two furin cleavage sites, a property shared with MERS-CoV, although the proteases used may be cell-type specific [25,62].
For SARS-CoV-2, furin cleavage of the spike (S) protein is important for the transmission and pathogenesis of the virus, the formation of syncytia, and the available entry pathway into the cell [63]. Furin processing allows for S to enter more easily into an “up” conformation that promotes receptor binding and acts as a “priming” event during virion formation (see Figure 3) [64]. This priming is essential for further proteolytic processing by either TMPRSS2 or other proteases. However, in the absence of this priming event, virion glycoproteins can instead be cleaved by cathepsins following uptake into the late endosome, leading to fusion with the endosomal membrane, as will be discussed further below. This entry pathway takes more time but allows for the virus to enter a wider range of cells. However, the furin/TMPRSS2 pathway has an additional benefit in that it allows for SARS-CoV-2 virions to evade elements of the immune system such as the interferon-induced transmembrane proteins present in the endosome [24]. Additionally, furin is known to be important for the formation of syncytia during SARS-CoV-2 infection. Syncytia are formed when S is trafficked to the host cell membrane during the course of infection. S then can fuse the cell membrane with the membrane of a neighboring cell, a process which is hypothesized to allow for cell-to-cell transmission of the virus and the evasion of the immune system [65].
As SARS-CoV-2 continued to spread and evolve in the human population, the mutations that accumulated around the FCS underscored the difference between evolutionary pressures in vivo and in vitro. While the virus readily loses the FCS via deletions and point mutations in vitro, the FCS has been strongly conserved in vivo [66]. Though several globally dominating variants have emerged with mutations near the FCS, these mutations have never completely interfered with furin processing. Rather, these mutations seem to take on the role of regulating the level of furin cleavage that occurs (P681H), with some mutations increasing the level of cleavage (P681R), as well as these point mutations having the ability to alter O-glycosylation of the FCS [67,68,69,70,71]. While S entering the “up conformation” is advantageous in many ways, it does make the S more susceptible to immune recognition as well as making it less stable. This stability decrease has been lessened by other mutations in S such as D614G that may allow the S to undergo a higher level of cleavage with fewer disadvantages [72]. Modification of the FCS via point mutations also occurred extensively as different isolates of the prototype coronavirus MHV were isolated, and in certain other coronaviruses, (OC43) modulation of the FCS may occur through an indel, in addition to point mutations [73,74]. Thus, modulation of furin cleavage through natural selection is a powerful means of regulating CoV tropism and pathogenesis; the virus continues to need to balance the efficiency of the transmission benefit of furin cleavage with the need to evade the immune response, now that most people have some level of pre-existing immunity.
For coronaviruses, furin-mediated cleavage (via priming at S1/S2) is linked to fusion activation at S2′, but it may also serve a more general role in virion maturation. In this case, furin may be playing an analogous role to that of other viral glycoproteins such as HIV-1 Env and filovirus GP. For the filoviruses EBOV and MARV, furin cleavage precedes the critical cathepsin-mediated cleavage events needed for virus entry, and it is interesting to note that the FCS is found in a quite different position in the GP protein sequence, implying a distinct maturation process between EBOV and MARV [75]. A similar role for furin is found for the DENV prM protein, where it controls the proportion of mature vs immature particles that form, which in turn may affect ADE and viral pathogenesis [76,77,78]. prM that is not cleaved by furin during virion formation can still be cleaved in the endosome during ADE by cathepsins. This alternative cathepsin cleavage route also can take over when SARS-CoV-2 S is not cleaved by furin, leading to furin cleavage being important for infection but not essential.

3.2. Coagulation Cascade Proteases (Plasmin, Thrombin, Factor Xa)

It is important to keep in mind that the host cell components that viruses take advantage of have important roles in maintaining homeostasis. The Blood Coagulation Cascade, which relies on several different proteases, is essential to stop blood loss following injury. Among the contributors to the cascade are thrombin, Factor Xa, and plasmin. These proteases can be responsible for cleaving other proteins within the pathway, or they can activate various cell types to respond to injury via the cleavage of Protease Activated Receptors (PARs), which are ubiquitously expressed [79]. This cascade exists in a carefully regulated balance, and viral disruption of this balance in either direction can lead to pathology. Respiratory viral infections are known to increase the risk of deep vein thrombosis and pulmonary embolism, as well as other serious vascular complications [80]. This can be caused by either the upregulation of coagulant factors or the downregulation of anticoagulant proteins. Several respiratory viruses can utilize the coagulation proteases for glycoprotein cleavage. This has been posited to form a positive feedback loop, where the glycoprotein is cleaved, leading to higher cell entry, which leads to the immune response of increased production of coagulation factors, which in turn leads to increased glycoprotein cleavage [21].
While influenza virus infection in the human respiratory tract is now known to be driven by trypsin-like proteases, the initial breakthrough for our understanding identified plasmin as the key activating protease [81]. The virus strain studied was the neuro-adapted isolate WSN, which had adapted to use plasmin during passage to make it trypsin-independent for growth in cell culture and in the murine brain [82]. Some clinical strains of influenza are also able to utilize plasmin for HA activation [42]. Early studies on SARS-CoV-1 showed a role for plasmin-activated cleavage, primarily at S2′ (R797), along with trypsin and TMPRSS11a; for SARS-CoV-2, plasmin has been indicated as one of the many proteases capable of activating spike [21,44]. Notably, RSV also is activated by plasmin, along with trypsin and thrombin [43].
Influenza HA activation, which is normally dependent on trypsin-like proteases or furin, is thought to lead to dependence on Factor Xa when the virus is adapted to grow in eggs, which is still a common method of growing influenza stocks (e.g., vaccine seeds) [83]. In contrast, SARS-CoV-2 does not appear to need adaptation to allow the use of a range of systemic secreted proteases such as plasmin, thrombin, and Factor Xa [21]. However, the impact of these proteases during an in vivo infection has yet to be fully elucidated.

3.3. Elastase

In vivo, elastase is produced mainly in the pancreas but also exists in a neutrophil-secreted form as part of the host’s innate immune defenses. The use of elastase as an activating protease has been reported for SARS-CoV-1, which may play a factor in promoting the initial inflammatory response to infection, but the situation for SARS-CoV-2 is less clear [36]. As with any virus-activating protease, the enzyme exists in tissue as part of a tightly regulated system, with neutrophil elastase proposed to be involved in ACE2 shedding, and so it has a protective effect for COVID [84]. Interestingly, engineering influenza HA to be elastase-cleaved has been used as a means to generate a live-attenuated vaccine virus [85]. Elastase-dependent influenza is attenuated, as elastase is not present in high enough levels during infection to make up for the lack of TMPRSS2 or furin cleavage during virion formation. This highlights the fact that it is not enough to have a proteolytic cleavage site on the glycoprotein; the cleavage site needs to be able to be cleaved by specific proteases that are present in specific locations during infection.

3.4. Type II Transmembrane Serine Proteases (TTSPs)

Type II Transmembrane Serine Proteases (TTSPs) all have a serine protease domain of chymotrypsin with a catalytic triad of histidine, aspartic acid, and serine [86]. Their catalytic triad leads to these proteases being considered trypsin-like. TTSPs are attractive proteases for viruses due to their location on the surface of epithelial cells, allowing for the possibility of glycoprotein processing during cell entry. There are at least nine TTSPs that are found in the respiratory tract, and they are often involved in viral glycoprotein processing. Low pathogenic strains of influenza are known to rely on TTSPs for glycoprotein activation. Table 2 depicts select TTSPs that are known to be utilized by respiratory viruses.

3.5. TMPRSS2

TMPRSS2, or Transmembrane Serine Protease 2, is a member of the hepsin/TMPRSS subfamily of TTSPS, which consists of at least seven similar proteases [86]. It cleaves at sterically available monobasic sites, where an arginine or lysine is present. The role of TMRPSS2 has yet to be fully elucidated, with it having proposed roles in fertility, inflammation moderation, prostate protection, and tumor suppression [107]. TMPRSS2 is present throughout the respiratory tract in the lung, bronchus, larynx, trachea, vocal folds, buccal mucosa, nasal mucosa, and tonsils [108]. The level of TMPRSS2 expression is tightly regulated in each tissue type, with over-expression of TMPRSS2 being associated with pathological states such as cancer. Inhibition of TMPRSS2 occurs naturally through host-made factors such as plasminogen inhibitor 1 and hepatocyte growth factor activator inhibitor (HAI-2). However, TMPRSS2 is also associated with the coagulation cascade proteases discussed above. TMPRRS2, trypsin, matriptase, and Factor Xa are all able to activate PAR2, which leads to the increased expression of matrix metalloproteases. These matrix metalloproteases are also known to be involved in viral life cycles; thus, even when TMPRSS2 does not have a direct interaction with the virus, it may still affect the outcome of infection through other mechanisms.
Its involvement in many respiratory virus life cycles can be readily explained by its high level of expression throughout the respiratory tract. TMPRSS2 is involved in the direct proteolytic activation of many viral glycoproteins and was first identified as a viral cofactor for influenza [91]. TMPRSS2 processes the influenza HA during virion formation, an essential step that renders HA fusion competent. TMPRSS2 also acts as a proteolytic activator during the cell entry of several coronaviruses, including SARS-CoV-1, SARS-CoV-2, HCoV-229E, and MERS-CoV. The activation of these viral glycoproteins by TMPRSS2 is an essential step for entry at the cell surface as it allows these glycoproteins to change into their fusion-competent conformations. TMPRSS2′s cleavage specificity is also broad enough that it is able to cleave cleavage sites that are traditionally considered to be “furin cleavage sites”. TMPRSS2 is interestingly also involved in the cleavage of ACE2, an additional host cell protease that will be discussed in depth later in this paper. This processing of ACE2 has been found to augment SARS-CoV-1’s cell entry, in addition to TMPRSS2’s processing of the SARS-CoV-1 S [109]. Interestingly, while the S1/S2 cleavage site of SARS-CoV-2 has mutated throughout the course of the COVID-19 pandemic, the S2′ site, which TMPRSS2 is able to cleave, has remained unaltered [110,111].

4. Cysteine Proteases–Cathepsins

Cathepsins are a type of cysteine protease that resemble papain, leading to them also being referred to as papain-like proteases. Their active site consists of a catalytic dyad with a histidine residue and a cysteine amino acid acting as the nucleophile [112]. Cathepsins are generally endosomal proteases but can also be found in other locations. For example, cathepsin L is present in endosomes and lysosomes but is also secreted outside of the cell. The level of cathepsin L in the blood is correlated to the level of pathogenesis following viral infection and is also elevated during other noninfectious pathological states [113]. Cathepsins need to be exposed to an acidic environment to be activated, and they prefer acidic and reducing conditions for cleavage. These proteases participate in a variety of virus life cycles and often are integral for alternative entry pathways. Cathepsins cleave viral glycoproteins in these so-called alternative pathways for several respiratory viruses, along with coronaviruses such as SARS-CoV-1, SARS-CoV-2, and MERS-CoV. For SARS-CoV-2, later variants show a preference for the cathepsin pathway in cell culture, despite this shift not occurring in vivo or in human airway organoids [114,115,116]. Beyond just glycoprotein cleavage, cathepsins also are used by viruses to modulate the immune response through several different pathways [117].
As a part of the Mammalian Orthoreovirus (MRV) life cycle, the virions’ capsids must be proteolytically digested away to become Infectious Subvirion Particles (ISVPs), which are then able to penetrate the endosomal membrane and gain entry into the cell [118]. This proteolytic digestion of the capsid has been attributed to several different cathepsins, making cysteine proteases essential for MRV infection. Specific point mutations in the MRV protein that undergo cleavage have been found to increase the efficiency of cathepsin cleavage [119,120]. However, despite proteolytic cleavage being considered the rate-limiting step of infection, few strains of MRV have incorporated these mutations, and viruses with these mutations lose titer more rapidly at elevated temperatures when compared to wild-type virus [121]. This reaffirms that proteolytic cleavage requires a careful balance between cleavage efficiency and overall viral fitness. Cathepsin cleavage is intimately tied to the route of entry for reovirus, and has been reviewed in Mainou and Dermody 2012 [122].
Many other virus systems take advantage of cathepsin cleavage in endosomes. Perhaps the most notable of these is the Ebola virus, whereby, following non-specific binding at the cell surface and internalization, cathepsin L and cathepsin B degrade the filovirus GP down to an 18kDa “stub”, which is the fusion-active component upon engagement with the NPC-1 receptor [123]. Another not classically respiratory virus (albeit transmitted via the oro-nasal route) is Nipah virus; in this case, the F protein enters recycling endosomes where it encounters cathepsins as part of the virus assembly pathway [124,125]. Cathepsins are also involved in RSV infection, via cleavage of the G (attachment) protein, and influenza infection via unspecified means [126]. Table 3 lists 6 different cathepsins, along with their cleavage determinants and the respiratory viruses that use them.

5. Metalloproteases

Metalloproteases differ from serine and cysteine proteases in that instead of using an amino acid as a nucleophile in their hydrolytic reaction, metalloproteases use a water molecule that has been activated by a metallic cation. Some metalloproteases have been found to cleave viral glycoproteins, such as ADAM and MT-MMP. However, this function is less well studied from the perspective of viruses than other groups of proteases.
Many metalloproteases also impact the viral life cycle in less direct ways. Matrix metalloproteases (MMPs) in particular are important players in many pathways that regulate the immune response and wound repair. MMPS are regulated by Tissue Inhibitors of Metalloproteases (TIMPs), and the interruption of this careful balance between activity and inhibition is associated with many disease states [142]. RSV is known to induce the expression of many MMPs. Drugs targeting these proteases were found to enhance viral clearance, inhibit syncytia formation, and prevent viral replication [52,143]. MMP-9 specifically is also implicated in infection with other respiratory viruses, such as influenza, rhinovirus, and human parainfluenza virus [144,145,146]. The exact mechanism of action of these MMPs in enhancing viral infection is unknown and remains mainly as hypotheses, based on either MMPs’ role in restructuring the respiratory tract or their role in the immune response.
MMPs were directly shown to act on the coronavirus spike, first through a study of MHV where MMP and ADAM family proteases acted in a strain-dependent manner, and they have more recently been added to the list of activating proteases for SARS-CoV-2 [147,148,149]. However, COVID-19 infections are also impacted by the role that MMPs play in the immune system. SARS-CoV-2 infections are known to increase the level of MMPs, which can lead to tissue damage and contribute to a cytokine storm reaction [150,151,152,153,154]. The role of MMPs is complex and involves multiple biological pathways. An excellent review of MMPs and COVID-19, as well as other coronavirus infections, has been published by Salomão et al. [155].

ACE2

Another metalloprotease of interest is Angiotensin-Converting Enzyme II (ACE2), which, as the name suggests, regulates angiotensin II levels through hydrolytic cleavage. Its position at the cell surface allows it to function as a receptor for SARS-CoV-2. The S1 subunit of the spike protein binds to ACE2, which begins the process of the cell being infected [156]. This action does not directly cleave the spike protein, but it does allow for the spike protein to be cleaved by other proteases, such as the previously discussed TMPRSS2 [157]. This use of proteases as receptors, and not for their proteolytic capacity, is also seen in several other coronaviruses, and the pharmacological intervention in this receptor binding interaction is a major area of research. Drugs targeting the SARS-CoV-2 spike and ACE2 are not focused on blocking the proteolytic activity of ACE2. Rather, research focuses on ACE2 decoys that will bind the spike protein before the virus arrives at a cell, thereby preventing successful receptor binding and infection [158]. Recently ACE2 was also revealed to act as a regulator of TMPRSS2 activity. This regulation occurs through a non-catalytic mechanism but has been shown to increase TMPRSS2 activation of influenza A and MERS-CoV [159]. When we think of the roles of host proteases in viral infection, it is tempting to think of them as having a straightforward interaction with viral glycoproteins, i.e., ACE2 acts as a receptor for SARS-CoV-2 S. However, we must remember that these proteases exist in a complex and ever-changing environment, and they have important roles that they serve within the host. The relationship between proteases and viral infections cannot be simple, because the system that this relationship exists in is not simple. Table 4 lists several metalloproteases utilized by respiratory viruses, and illustrates some of the different roles that these proteases can fulfill in the viral life cycle.

6. Protease Inhibitors

Protease inhibitors, like proteases themselves, are a broad and diverse group. Inhibitors play an essential role in maintaining homeostasis, preventing proteases from being overly active. Due to this role, protease inhibitors have been found in many life forms, leading to a large group of existing inhibitors to explore as drug candidates. Inhibitors can be classified by the type of protease they inhibit or, like other enzyme inhibitors, by their mode of action. When an inhibitor binds to the enzyme through a covalent interaction, this is considered an irreversible reaction. These inhibitors fall into either the less specific group of affinity-labeling inhibitors or the more specific suicide inhibitors, which require the protease to enzymatically act upon them to achieve inhibition [164]. For reversible inhibitors, inhibitors noncovalently interact with the protease, and the location of that interaction determines whether it is competitive, noncompetitive, or uncompetitive inhibition. Competitive inhibitors occupy the active site of the protease, noncompetitive inhibitors bind elsewhere on the protein, and uncompetitive inhibitors only bind to enzymes that are already bound to their substrates (see Figure 4). Inhibitors can be proteins, peptides, small molecules, or antibodies. They can directly interact with the protease, or they can interact with molecules that proteases need to be active, i.e., Ca2+ chelators that inhibit furin.
Because the kinetics of inhibition differ between the inhibitor types, whether an inhibitor is reversible or irreversible is a determinant of its appeal as an antiviral. Irreversible inhibitors are attractive because of how long they are active and the ease with which they can be designed, but they have more potential for negative side effects, especially if they are widely active on many proteases [165]. Both the length of activity and the potential for negative side effects stem from the fact that the inhibition is irreversible. Once a protease is inhibited, a cell will need to make new protease in order to regain proteolytic function. An irreversible inhibitor that is broadly acting could disrupt many processes involved in homeostasis. Therefore, irreversible inhibitors have a higher requirement for specificity and subcellular targeting design than reversible inhibitors. Competitive reversible inhibitors are also more straightforward to design, due to their need to fit into the active site of the enzyme, but they require a high concentration to compete with other substrates. This can be a barrier when trying to inhibit the cleavage of viral proteins in subcellular compartments, as you need to get a large quantity of the inhibitor into those specific compartments. Uncompetitive and noncompetitive reversible inhibitors do not have this same issue with concentration; however, these types of inhibitors are complicated to design. However, the problems of irreversible and competitive inhibitors both have the same solution, cellular and subcellular drug targeting. Directing inhibitors to the correct location would allow the concentration to reach a high level in the area of interest without the need for an extreme dose to the rest of the body. Ideally, the inhibitor would remain inert until it reached its target destination, lowering the chances of adverse side effects. Research is quickly advancing, both in drug modifications that direct drugs to be taken up into specific cell compartments, as well as in the creation of nanoparticle carriers that allow for precise drug targeting, as well as having other benefits [166,167].
The human genome encodes for numerous protease inhibitors, as maintaining a careful balance of proteolytic activity is essential for homeostasis. However, these endogenous inhibitors also carry out important roles in the immune response to viral infections. The viruses that are inhibited by endogenous inhibitors are diverse and include coronaviruses, flaviviruses, and filoviruses, among others [168]. This inhibition can be a direct interruption of the catalytic mechanism of cleavage, a rerouting of proteases to an alternate cellular compartment, or an interaction with the protease in one of its other functional domains. A comprehensive review of these endogenous inhibitors has been carried out by Lotke, Petersen, and Sauter [168]. Of particular interest among these inhibitors are serpins.
Serine protease inhibitors (serpins) are present in most types of life. These inhibitors mostly inhibit serine proteases but select serpins can also inhibit other types of proteases. Serpins contain cleavage motifs of the proteases that they inhibit, and upon their cleavage, they undergo a conformational change that covalently traps the protease in an irreversible interaction [169]. This conformational change is irreversible; thus, they are suicide inhibitors. Interestingly, poxviruses, along with some plant viruses and herpesviruses, have been found to encode their own serpins. Poxvirus serpins are the most highly studied and have been found to exhibit anti-apoptotic and anti-inflammatory roles that can promote viral replication [170]. Serp-1, a serpin encoded in the myxoma virus (MyxV) genome, has been tested as a therapeutic target for several inflammatory disorders as well as viral infections with notable success, as outlined in the review by Varkoly et al. [171]. By inhibiting the urokinase-type plasminogen activator receptor, complement receptors, and some of the thrombolytic enzymes (such as the previously discussed plasmin), it can prevent the overactivation of the immune response [172,173,174]. The success of using a viral serpin, which was originally made to support a viral infection, to then treat disease caused by viral infections is highly promising and supports the notion that the viral use of host proteases exists in a balance which we can seek to disrupt.

7. Current Therapeutic Strategies Targeting Coronaviruses, Influenza Viruses, and Para-Myxoviruses

The COVID-19 pandemic prompted a resurgence of therapeutic development for respiratory viruses, and while some drugs targeting SARS-CoV-2 are now approved for use in humans, many additional drugs are currently approved in development, including inhibitors of the viral polymerase and viral Mpro and 3CL-like proteases [175]. Recent oral inhibitors now available include Paxlovid®, and while this remains in use, Molnupiravir® and the injectable Remdesivir® are no longer widely recommended, and most therapeutic monoclonal antibodies have been discontinued. Currently, the main therapeutics targeting influenza are the DAAs Tamiflu® and Relenza® (targeting the viral neuraminidase), with Baloxavir® also now available (targeting the viral polymerase). However, numerous influenza virus strains are emerging that are resistant to Tamiflu® and Relenza®, rendering the treatment ineffective. For paramyxoviruses and pneumoviruses, ribavirin has been utilized for RSV treatment and shows similar in vitro neutralization of HMPV [176], but there are significant questions concerning its efficacy in clinical settings. Based on successes with RSV, humanized monoclonal antibodies, the primary prophylactic treatment for HRSV, are currently in development for HMPV, but the cost of such treatment is extremely high. Host-targeted approaches are also in development and have the notable potential benefit of limiting the emergence of antiviral-resistant viruses.
  • Inhibition of furin as an antiviral strategy
Given its importance for viral infection, furin has received notable attention as a therapeutic option [177]. The first generation of furin inhibitors were peptide-based and lacked specificity, and were followed by serpins such as AT-PDX, which have high specificity and increased potential as therapeutic agents. Recent studies with bioavailable small molecule furin inhibitors such as BOS-981 that effectively inhibit SARS-CoV-2 show promise as candidate therapeutics, especially when combined with TTSP inhibitors (see below) [178]. Ca2+ chelators, serpins, and proprotein convertase inhibitors have all been successful at inhibiting furin. Most known inhibitors of furin have inhibitory effects on other proteases as well, so the development of new, targeted furin inhibitors is still an area of intense interest. However, as some viruses are able to utilize multiple proteases for glycoprotein cleavage, it is important to design treatments that carefully balance the need to prevent off-target effects while still targeting the diverse set of proteases that viruses utilize.
  • Inhibition of TTSPs as an antiviral strategy
Inhibition of TTSPs is among the most promising alternative host-targeted approaches under consideration as therapeutics or prophylactics, for SARS-CoV-2 or other respiratory pathogens. The use of TTSP inhibitors was originally investigated for the treatment of influenza, as an alternative to inhibitors that target the influenza protease neuraminidase [179,180,181]. Camostat, a TTSP inhibitor repurposed as an antiviral, is currently in clinical trials for COVID-19—although currently with only limited success [182]. Notably, camostat has relatively low potency and is taken orally, which may contribute to its limited success in targeting respiratory disease. Impeding proteolytic activation of the viral HA via the serine protease inhibitor aprotinin [183], which has anti-fibrinolytic properties, has also been explored, including preclinical development and some limited clinical trials. The partial success of these trials reinforces the concept of targeting host cell-mediated proteolytic activation of viral glycoproteins as a viable therapeutic strategy. However, the development of aprotinin has limitations. First, being proteinaceous, indications are that aprotinin has a very short biological half-life, on the order of 0.7 h [184]. Second, it is currently unclear exactly which proteases are active in the human respiratory tract, and so which proteases need to be targeted. While aprotinin is somewhat broadly acting, it is likely to only target a subset of proteases. Importantly, it is a bovine-derived product, likely contributing to its toxicity and removal from the market [185]. Therefore, the development of alternative protease-targeted therapeutics in addition to aprotinin is likely to significantly expand the therapeutic options available to treat respiratory viruses.
As proof of principle for this approach, a lead protease inhibitor comprising the Kunitz type, serine protease inhibitor SPINT2, also known as HAI-2, was tested [186]. SPINT2 is a 25 kDa protein found on the plasma membrane of a number of tissues, where it is mainly bound to the serine protease matriptase. Both matriptase and SPINT2 are initially located on the plasma membrane but are proteolytically released into the extracellular space. SPINT2 contains two Kunitz-type inhibitor domains that potently modulate the proteolytic activity of several trypsin-like serine proteases, including matriptase and hepatocyte growth factor activator; inhibition of these two proteases is thought to be the primary biological role of SPINT2. The properties of SPINT2, therefore, made it an excellent proof of principle to efficiently inhibit viral replication, including both influenza and HMPV [185,187]. Small molecule approaches previously discovered included an initial lead compound, IN-1 (N-0100), as an influenza therapeutic [94]. This was developed into N-0385, a small molecule peptidomimetic with a ketobenzothiazole warhead. In mice, N-0385 treatment reduced morbidity and mortality associated with influenza H1N1 and H3N2 infection, as well as proving highly effective in mice, to achieve progress towards the goal of prophylactic inhibition of a broad range of respiratory viruses [188].

8. Conclusions

The human genome encodes for hundreds of diverse proteolytic enzymes that are vital for the maintenance of homeostasis. These proteases are sometimes co-opted for virion maturation, cell entry, and immune modulation, among other things. These proteases provide an additional area to focus on in antiviral development. The COVID-19 pandemic has helped further the field of antiviral development but has also revealed complexities in the relationship between viruses and proteases that make designing antivirals more challenging. However, the field is readily rising to the challenge, and the field of antivirals targeting host cell proteases remains one of great interest.

Author Contributions

Writing—review and editing, B.L. and G.R.W.; supervision, G.R.W. All authors have read and agreed to the published version of the manuscript.

Funding

Work in the author’s lab on this topic is recently funded by the National Institute of Health (NIH, grant number 1R56AI181271—01) to G.W.

Acknowledgments

We would like to thank all of the members of the Whittaker lab for their support, with a special acknowledgment to Annette Choi for many helpful discussions.

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the writing of the manuscript.

References

  1. Schechter, I.; Berger, A. On the size of the active site in proteases. I. Papain. Biochem. Biophys. Res. Commun. 1967, 27, 157–162. [Google Scholar] [CrossRef]
  2. McKelvey, M.C.; Brown, R.; Ryan, S.; Mall, M.A.; Weldon, S.; Taggart, C.C. Proteases, Mucus, and Mucosal Immunity in Chronic Lung Disease. Int. J. Mol. Sci. 2021, 22, 5018. [Google Scholar] [CrossRef]
  3. Majchrzak, M.; Poręba, M. The roles of cellular protease interactions in viral infections and programmed cell death: A lesson learned from the SARS-CoV-2 outbreak and COVID-19 pandemic. Pharmacol. Rep. 2022, 74, 1149–1165. [Google Scholar] [CrossRef]
  4. Hobbs, E.C.; Reid, T.J. Animals and SARS-CoV-2: Species susceptibility and viral transmission in experimental and natural conditions, and the potential implications for community transmission. Transbound. Emerg. Dis. 2021, 68, 1850–1867. [Google Scholar] [CrossRef]
  5. Carossino, M.; Izadmehr, S.; Trujillo, J.D.; Gaudreault, N.N.; Dittmar, W.; Morozov, I.; Balasuriya, U.B.R.; Cordon-Cardo, C.; García-Sastre, A.; Richt, J.A. ACE2 and TMPRSS2 distribution in the respiratory tract of different animal species and its correlation with SARS-CoV-2 tissue tropism. Microbiol. Spectr. 2024, 12, e03270-23. [Google Scholar] [CrossRef]
  6. Lean, F.Z.X.; Núñez, A.; Spiro, S.; Priestnall, S.L.; Vreman, S.; Bailey, D.; James, J.; Wrigglesworth, E.; Suarez-Bonnet, A.; Conceivao, C.; et al. Differential susceptibility of SARS-CoV-2 in animals: Evidence of ACE2 host receptor distribution in companion animals, livestock and wildlife by immunohistochemical characterisation. Transbound Emerg Dis. 2022, 69, 2275–2286. [Google Scholar] [CrossRef]
  7. Martins, M.; Boggiatto, P.M.; Buckley, A.; Cassmann, E.D.; Falkenberg, S.; Caserta, L.C.; Fernandes, M.H.V.; Kanipe, C.; Lager, K.; Palmer, M.V.; et al. From Deer-to-Deer: SARS-CoV-2 is efficiently transmitted and presents broad tissue tropism and replication sites in white-tailed deer. PLoS Pathog. 2022, 18, e1010197. [Google Scholar] [CrossRef]
  8. Klestova, Z. Possible spread of SARS-CoV-2 in domestic and wild animals and body temperature role. Virus Res. 2023, 327, 199066. [Google Scholar] [CrossRef]
  9. Shi, J.; Wen, Z.; Zhong, G.; Yang, H.; Wang, C.; Huang, B.; Liu, R.; He, X.; Shuai, L.; Sun, Z.; et al. Susceptibility of ferrets, cats, dogs, and other domesticated animals to SARS–coronavirus 2. Science 2020, 368, 1016–1020. [Google Scholar] [CrossRef]
  10. Barua, A.; Grot, N.; Plawski, A. The basis of mink susceptibility to SARS-CoV-2 infection. J. Appl. Genet. 2022, 63, 543–555. [Google Scholar] [CrossRef]
  11. Pandey, R.K.; Srivastava, A.; Mishra, R.K.; Singh, P.P.; Chaubey, G. Novel genetic association of the Furin gene polymorphism rs1981458 with COVID-19 severity among Indian populations. Sci. Rep. 2024, 14, 7822. [Google Scholar] [CrossRef]
  12. Al-Mulla, F.; Mohammad, A.; Al Madhoun, A.; Haddad, D.; Ali, H.; Eaaswarkhanth, M.; John, S.E.; Nizam, R.; Channanath, A.; Abu-Farha, M.; et al. ACE2 and FURIN variants are potential predictors of SARS-CoV-2 outcome: A time to implement precision medicine against COVID-19. Heliyon 2021, 7, e06133. [Google Scholar] [CrossRef]
  13. Sienko, J.; Marczak, I.; Kotowski, M.; Bogacz, A.; Tejchman, K.; Sienko, M.; Kotifs, K. Association of ACE2 Gene Variants with the Severity of COVID-19 Disease—A Prospective Observational Study. Int. J. Environ. Res. Public Health 2022, 19, 12622. [Google Scholar] [CrossRef]
  14. David, A.; Parkinson, N.; Peacock, T.P.; Pairo-Castineira, E.; Khanna, T.; Cobat, A.; Tenesa, A.; Sancho-Shmizu, V.; Casanova, J.L.; Abel, L.; et al. A common TMPRSS2 variant has a protective effect against severe COVID-19. Curr. Res. Transl. Med. 2022, 70, 103333. [Google Scholar] [CrossRef]
  15. Coto, E.; Albaiceta, G.M.; Amado-Rodríguez, L.; García-Clemente, M.; Cuesta-Llavona, E.; Vázquez-Coto, D.; Alonso, B.; Iglesias, S.; Melón, S.; Alvarez-Argüelles, M.E.; et al. FURIN gene variants (rs6224/rs4702) as potential markers of death and cardiovascular traits in severe COVID-19. J. Med. Virol. 2022, 94, 3589–3595. [Google Scholar] [CrossRef]
  16. Rawlings, N.D.; Barrett, A.J. Chapter 559—Introduction: Serine Peptidases and Their Clans. In Handbook of Proteolytic Enzymes, 3rd ed.; Rawlings, N.D., Salvesen, G., Eds.; Academic Press: Cambridge, MA, USA, 2013; pp. 2491–2523. Available online: https://www.sciencedirect.com/science/article/pii/B9780123822192005597 (accessed on 13 July 2023).
  17. Andersson, M.K.; Enoksson, M.; Gallwitz, M.; Hellman, L. The extended substrate specificity of the human mast cell chymase reveals a serine protease with well-defined substrate recognition profile. Int. Immunol. 2009, 21, 95–104. [Google Scholar] [CrossRef]
  18. Nygaard, R.M.; Golden, J.W.; Schiff, L.A. Impact of Host Proteases on Reovirus Infection in the Respiratory Tract. J. Virol. 2012, 86, 1238–1243. [Google Scholar] [CrossRef]
  19. Brown, M.A.; Stenberg, L.M.; Stenflo, J. Coagulation Factor Xa. In Handbook of Proteolytic Enzymes; Academic Press: Cambridge, MA, USA, 2013; pp. 2908–2915. [Google Scholar]
  20. Du, L.; Kao, R.Y.; Zhou, Y.; He, Y.; Zhao, G.; Wong, C.; Jiang, S.; Yuen, K.Y.; Jin, D.Y.; Zheng, B.J. Cleavage of spike protein of SARS coronavirus by protease factor Xa is associated with viral infectivity. Biochem. Biophys. Res. Commun. 2007, 359, 174–179. [Google Scholar] [CrossRef]
  21. Kastenhuber, E.R.; Mercadante, M.; Nilsson-Payant, B.; Johnson, J.L.; Jaimes, J.A.; Muecksch, F.; Weisblum, Y.; Bram, Y.; Chandar, V.; Whittaker, G.R.; et al. Coagulation factors directly cleave SARS-CoV-2 spike and enhance viral entry. van der Meer JW, editor. eLife 2022, 11, e77444. [Google Scholar] [CrossRef]
  22. Hosaka, M.; Nagahama, M.; Kim, W.S.; Watanabe, T.; Hatsuzawa, K.; Ikemizu, J.; Murakami, K.; Nakayama, K. Arg-X-Lys/Arg-Arg motif as a signal for precursor cleavage catalyzed by furin within the constitutive secretory pathway. J. Biol. Chem. 1991, 266, 12127–12130. [Google Scholar] [CrossRef]
  23. Kawaoka, Y.; Webster, R.G. Sequence requirements for cleavage activation of influenza virus hemagglutinin expressed in mammalian cells. Proc. Natl. Acad. Sci. USA 1988, 85, 324–328. [Google Scholar] [CrossRef]
  24. 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.; et al. 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]
  25. Millet, J.K.; Whittaker, G.R. Host cell entry of Middle East respiratory syndrome coronavirus after two-step, furin-mediated activation of the spike protein. Proc. Natl. Acad. Sci. USA 2014, 111, 15214–15219. [Google Scholar] [CrossRef]
  26. Collins, P.L.; Huang, Y.T.; Wertz, G.W. Nucleotide sequence of the gene encoding the fusion (F) glycoprotein of human respiratory syncytial virus. Proc. Natl. Acad. Sci. USA 1984, 81, 7683–7687. [Google Scholar] [CrossRef]
  27. Ortmann, D.; Ohuchi, M.; Angliker, H.; Shaw, E.; Garten, W.; Klenk, H.D. Proteolytic cleavage of wild type and mutants of the F protein of human parainfluenza virus type 3 by two subtilisin-like endoproteases, furin and Kex2. J. Virol. 1994, 68, 2772–2776. [Google Scholar] [CrossRef]
  28. Watanabe, M.; Hirano, A.; Stenglein, S.; Nelson, J.; Thomas, G.; Wong, T.C. Engineered serine protease inhibitor prevents furin-catalyzed activation of the fusion glycoprotein and production of infectious measles virus. J. Virol. 1995, 69, 3206–3210. [Google Scholar] [CrossRef]
  29. Ueo, A.; Kubota, M.; Shirogane, Y.; Ohno, S.; Hashiguchi, T.; Yanagi, Y. Lysosome-Associated Membrane Proteins Support the Furin-Mediated Processing of the Mumps Virus Fusion Protein. J. Virol. 2020, 94, e00050-20. [Google Scholar] [CrossRef]
  30. Borgoño, C.A.; Gavigan, J.A.; Alves, J.; Bowles, B.; Harris, J.L.; Sotiropoulou, G.; Diamandis, E.P. Defining the extended substrate specificity of kallikrein 1-related peptidases. Biol. Chem. 2007, 388, 1215–1225. [Google Scholar] [CrossRef]
  31. Milewska, A.; Falkowski, K.; Kulczycka, M.; Bielecka, E.; Naskalska, A.; Mak, P.; Lesner, A.; Ochman, M.; Urlik, M.; Diamandis, E.; et al. Kallikrein 13 serves as a priming protease during infection by the human coronavirus HKU1. Sci. Signal. 2020, 13, eaba9902. [Google Scholar] [CrossRef]
  32. Li, H.X.; Hwang, B.Y.; Laxmikanthan, G.; Blaber, S.I.; Blaber, M.; Golubkov, P.A.; Ren, P.; Iverson, B.I.; Georgiou, G. Substrate specificity of human kallikreins 1 and 6 determined by phage display. Protein Sci. 2008, 17, 664–672. [Google Scholar] [CrossRef]
  33. Hamilton, B.S.; Whittaker, G.R. Cleavage Activation of Human-adapted Influenza Virus Subtypes by Kallikrein-related Peptidases 5 and 12. J. Biol. Chem. 2013, 288, 17399–17407. [Google Scholar] [CrossRef]
  34. Harris, J.L.; Backes, B.J.; Leonetti, F.; Mahrus, S.; Ellman, J.A.; Craik, C.S. Rapid and general profiling of protease specificity by using combinatorial fluorogenic substrate libraries. Proc. Natl. Acad. Sci. USA 2000, 97, 7754–7759. [Google Scholar] [CrossRef]
  35. Golden, J.W.; Schiff, L.A. Neutrophil elastase, an acid-independent serine protease, facilitates reovirus uncoating and infection in U937 promonocyte cells. Virol. J. 2005, 2, 48. [Google Scholar] [CrossRef]
  36. 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]
  37. Matsuyama, S.; Ujike, M.; Morikawa, S.; Tashiro, M.; Taguchi, F. Protease-mediated enhancement of severe acute respiratory syndrome coronavirus infection. Proc. Natl. Acad. Sci. USA 2005, 102, 12543–12547. [Google Scholar] [CrossRef]
  38. Seidah, N.G. Chapter 730—Proprotein Convertase 5. In Handbook of Proteolytic Enzymes, 3rd ed.; Rawlings, N.D., Salvesen, G., Eds.; Academic Press: Cambridge, MA, USA, 2013; pp. 3305–3310. Available online: https://www.sciencedirect.com/science/article/pii/B9780123822192007304 (accessed on 4 May 2024).
  39. Basak, A.; Zhong, M.; Munzer, J.S.; Chrétien, M.; Seidah, N.G. Implication of the proprotein convertases furin, PC5 and PC7 in the cleavage of surface glycoproteins of Hong Kong, Ebola and respiratory syncytial viruses: A comparative analysis with fluorogenic peptides. Biochem. J. 2001, 353 Pt 3, 537–545. [Google Scholar] [CrossRef]
  40. Castellino, F.J. Chapter 648—Plasmin. In Handbook of Proteolytic Enzymes, 3rd ed.; Rawlings, N.D., Salvesen, G., Eds.; Academic Press: Cambridge, MA, USA, 2013; pp. 2958–2968. Available online: https://www.sciencedirect.com/science/article/pii/B9780123822192006487 (accessed on 13 July 2023).
  41. Xue, F.; Seto, C.T. Selective Inhibitors of the Serine Protease Plasmin:  Probing the S3 and S3‘ Subsites Using a Combinatorial Library. J. Med. Chem. 2005, 48, 6908–6917. [Google Scholar] [CrossRef]
  42. Tse, L.V.; Marcano, V.C.; Huang, W.; Pocwierz, M.S.; Whittaker, G.R. Plasmin-Mediated Activation of Pandemic H1N1 Influenza Virus Hemagglutinin Is Independent of the Viral Neuraminidase. J. Virol. 2013, 87, 5161–5169. [Google Scholar] [CrossRef]
  43. Dubovi, E.J.; Geratz, J.D.; Tidwell, R.R. Enhancement of respiratory syncytial virus-induced cytopathology by trypsin, thrombin, and plasmin. Infect Immun. 1983, 40, 351–358. [Google Scholar] [CrossRef]
  44. Kam, Y.W.; Okumura, Y.; Kido, H.; Ng, L.F.P.; Bruzzone, R.; Altmeyer, R. Cleavage of the SARS Coronavirus Spike Glycoprotein by Airway Proteases Enhances Virus Entry into Human Bronchial Epithelial Cells In Vitro. PLoS ONE 2009, 4, e7870. [Google Scholar] [CrossRef]
  45. Hou, Y.; Yu, T.; Wang, T.; Ding, Y.; Cui, Y.; Nie, H. Competitive cleavage of SARS-CoV-2 spike protein and epithelial sodium channel by plasmin as a potential mechanism for COVID-19 infection. Am. J. Physiol.—Lung Cell. Mol. Physiol. 2022, 323, L569–L577. [Google Scholar] [CrossRef]
  46. Gallwitz, M.; Enoksson, M.; Thorpe, M.; Hellman, L. The Extended Cleavage Specificity of Human Thrombin. PLoS ONE 2012, 7, e31756. [Google Scholar] [CrossRef]
  47. Rodriguez, T.; Dobrovolny, H.M. Quantifying the effect of trypsin and elastase on in vitro SARS-CoV infections. Virus Res. 2021, 299, 198423. [Google Scholar] [CrossRef]
  48. Kirchdoerfer, R.N.; Wang, N.; Pallesen, J.; Wrapp, D.; Turner, H.L.; Cottrell, C.A.; Corbett, K.S.; Graham, B.S.; McLellan, J.S.; Ward, A.B. Stabilized coronavirus spikes are resistant to conformational changes induced by receptor recognition or proteolysis. Sci. Rep. 2018, 8, 15701. [Google Scholar] [CrossRef]
  49. Schowalter, R.M.; Smith, S.E.; Dutch, R.E. Characterization of Human Metapneumovirus F Protein-Promoted Membrane Fusion: Critical Roles for Proteolytic Processing and Low pH. J. Virol. 2006, 80, 10931–10941. [Google Scholar] [CrossRef]
  50. Henrickson, K.J. Parainfluenza Viruses. Clin. Microbiol. Rev. 2003, 16, 242–264. [Google Scholar] [CrossRef]
  51. Thorpe, M.; Fu, Z.; Chahal, G.; Akula, S.; Kervinen, J.; De Garavilla, L.; Hellman, L. Extended cleavage specificity of human neutrophil cathepsin G: A low activity protease with dual chymase and tryptase-type specificities. PLoS ONE 2018, 13, e0195077. [Google Scholar] [CrossRef]
  52. Foronjy, R.F.; Taggart, C.C.; Dabo, A.J.; Weldon, S.; Cummins, N.; Geraghty, P. Type-I interferons induce lung protease responses following respiratory syncytial virus infection via RIG-I-like receptors. Mucosal Immunol. 2015, 8, 161–175. [Google Scholar] [CrossRef]
  53. Tian, S. A 20 Residues Motif Delineates the Furin Cleavage Site and its Physical Properties May Influence Viral Fusion. Biochem. Insights 2009, 2, BCI.S2049. [Google Scholar] [CrossRef]
  54. Choi, A.; Kots, E.D.; Singleton, D.T.; Weinstein, H.; Whittaker, G.R. Analysis of the molecular determinants for furin cleavage of the spike protein S1/S2 site in defined strains of the prototype coronavirus murine hepatitis virus (MHV). Virus Res. 2024, 340, 199283. [Google Scholar] [CrossRef]
  55. Gram Schjoldager, K.T.B.; Vester-Christensen, M.B.; Goth, C.K.; Petersen, T.N.; Brunak, S.; Bennett, E.P.; Levery, S.B.; Clausen, H. A Systematic Study of Site-specific GalNAc-type O-Glycosylation Modulating Proprotein Convertase Processing. J. Biol. Chem. 2011, 286, 40122–40132. [Google Scholar] [CrossRef]
  56. Molloy, S.S.; Bresnahan, P.A.; Leppla, S.H.; Klimpel, K.R.; Thomas, G. Human furin is a calcium-dependent serine endoprotease that recognizes the sequence Arg-X-X-Arg and efficiently cleaves anthrax toxin protective antigen. J. Biol. Chem. 1992, 267, 16396–16402. [Google Scholar] [CrossRef]
  57. Böttcher-Friebertshäuser, E.; Klenk, H.D.; Garten, W. Activation of influenza viruses by proteases from host cells and bacteria in the human airway epithelium. Pathog. Dis. 2013, 69, 87–100. [Google Scholar] [CrossRef]
  58. Lu, X.; Shi, Y.; Gao, F.; Xiao, H.; Wang, M.; Qi, J.; Gao, G.F. Insights into Avian Influenza Virus Pathogenicity: The Hemagglutinin Precursor HA0 of Subtype H16 Has an Alpha-Helix Structure in Its Cleavage Site with Inefficient HA1/HA2 Cleavage. J. Virol. 2012, 86, 12861–12870. [Google Scholar] [CrossRef]
  59. Mayer, G.; Boileau, G.; Bendayan, M. Sorting of Furin in Polarized Epithelial and Endothelial Cells: Expression Beyond the Golgi Apparatus. J. Histochem. Cytochem. 2004, 52, 567–579. [Google Scholar] [CrossRef]
  60. Luczo, J.M.; Stambas, J.; Durr, P.A.; Michalski, W.P.; Bingham, J. Molecular pathogenesis of H5 highly pathogenic avian influenza: The role of the haemagglutinin cleavage site motif. Rev. Med. Virol. 2015, 25, 406–430. [Google Scholar] [CrossRef]
  61. Schrauwen, E.J.A.; Herfst, S.; Leijten, L.M.; Van Run, P.; Bestebroer, T.M.; Linster, M.; Bodewes, R.; Kreijtz, J.H.C.M.; Rimmelzwaan, G.F.; Osterhaus, A.D.M.E.; et al. The Multibasic Cleavage Site in H5N1 Virus Is Critical for Systemic Spread along the Olfactory and Hematogenous Routes in Ferrets. J. Virol. 2012, 86, 3975–3984. [Google Scholar] [CrossRef]
  62. Krzyzaniak, M.A.; Zumstein, M.T.; Gerez, J.A.; Picotti, P.; Helenius, A. Host Cell Entry of Respiratory Syncytial Virus Involves Macropinocytosis Followed by Proteolytic Activation of the F Protein. PLoS Pathog. 2013, 9, e1003309. [Google Scholar] [CrossRef]
  63. Takeda, M. Proteolytic activation of SARS-CoV-2 spike protein. Microbiol. Immunol. 2022, 66, 15–23. [Google Scholar] [CrossRef]
  64. Gobeil, S.M.C.; Janowska, K.; McDowell, S.; Mansouri, K.; Parks, R.; Manne, K.; Stalls, V.; Kopp, M.F.; Henderson, R.; Edwards, R.J.; et al. 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]
  65. Rajah, M.M.; Bernier, A.; Buchrieser, J.; Schwartz, O. The Mechanism and Consequences of SARS-CoV-2 Spike-Mediated Fusion and Syncytia Formation. J. Mol. Biol. 2022, 434, 167280. [Google Scholar] [CrossRef]
  66. Sasaki, M.; Uemura, K.; Sato, A.; Toba, S.; Sanaki, T.; Maenaka, K.; Hall, W.W.; Orba, Y.; Sawa, H. SARS-CoV-2 variants with mutations at the S1/S2 cleavage site are generated in vitro during propagation in TMPRSS2-deficient cells. PLoS Pathog. 2021, 17, e1009233. [Google Scholar] [CrossRef]
  67. Saito, A.; Irie, T.; Suzuki, R.; Maemura, T.; Nasser, H.; Uriu, K.; Kosugi, Y.; Shirakawa, K.; Sadamasu, K.; Kimura, I.; et al. Enhanced Fusogenicity and Pathogenicity of SARS-CoV-2 Delta P681R mutation. Nature 2022, 602, 300–306. [Google Scholar] [CrossRef]
  68. Lubinski, B.; Frazier, L.E.; Phan, M.V.T.; Bugembe, D.L.; Cunningham, J.L.; Tang, T.; Daniel, S.; Cotton, M.; Jaimes, J.A.; Whittaker, G.R. Spike Protein Cleavage-Activation in the Context of the SARS-CoV-2 P681R Mutation: An Analysis from Its First Appearance in Lineage A.23.1 Identified in Uganda. Microbiol. Spectr. 2022, 10, e01514-22. [Google Scholar] [CrossRef]
  69. Lubinski, B.; Fernandes, M.H.V.; Frazier, L.; Tang, T.; Daniel, S.; Diel, D.G.; Jaimes, J.A.; Whittaker, G.R. Functional evaluation of the P681H mutation on the proteolytic activation of the SARS-CoV-2 variant B.1.1.7 (Alpha) spike. iScience 2022, 25, 103589. [Google Scholar] [CrossRef]
  70. Liu, Y.; Liu, J.; Johnson, B.A.; Xia, H.; Ku, Z.; Schindewolf, C.; Widen, S.G.; An, Z.; Weaver, S.C.; Menchery, V.D.; et al. Delta spike P681R mutation enhances SARS-CoV-2 fitness over Alpha variant. Cell Rep. 2022, 39, 110829. [Google Scholar] [CrossRef]
  71. 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]
  72. Gellenoncourt, S.; Saunders, N.; Robinot, R.; Auguste, L.; Rajah, M.M.; Kervevan, J.; Jeger-Madiot, R.; Staropoli, I.; Planchais, C.; Mouquet, H.; et al. The Spike-Stabilizing D614G Mutation Interacts with S1/S2 Cleavage Site Mutations To Promote the Infectious Potential of SARS-CoV-2 Variants. J. Virol. 2022, 96, e01301-22. [Google Scholar] [CrossRef]
  73. Lau, S.K.P.; Li, K.S.M.; Li, X.; Tsang, K.Y.; Sridhar, S.; Woo, P.C.Y. Fatal Pneumonia Associated with a Novel Genotype of Human Coronavirus OC43. Front. Microbiol. 2022, 12, 795449. [Google Scholar] [CrossRef]
  74. Stodola, J.K.; Dubois, G.; Le Coupanec, A.; Desforges, M.; Talbot, P.J. The OC43 human coronavirus envelope protein is critical for infectious virus production and propagation in neuronal cells and is a determinant of neurovirulence and CNS pathology. Virology 2018, 515, 134–149. [Google Scholar] [CrossRef]
  75. Lee, J.E.; Fusco, M.L.; Hessell, A.J.; Oswald, W.B.; Burton, D.R.; Saphire, E.O. Structure of the Ebola virus glycoprotein bound to an antibody from a human survivor. Nature 2008, 454, 177–182. [Google Scholar] [CrossRef]
  76. Murray, J.M.; Aaskov, J.G.; Wright, P.J. Processing of the dengue virus type 2 proteins prM and C-prM. J. Gen. Virol. 1993, 74, 175–182. [Google Scholar] [CrossRef]
  77. Dejnirattisai, W.; Jumnainsong, A.; Onsirisakul, N.; Fitton, P.; Vasanawathana, S.; Limpitikul, W.; Puttikhunt, C.; Edwards, C.; Duangchinda, T.; Supasa, S.; et al. Cross-Reacting Antibodies Enhance Dengue Virus Infection in Humans. Science 2010, 328, 745–748. [Google Scholar] [CrossRef]
  78. Luo, Y.Y.; Feng, J.J.; Zhou, J.M.; Yu, Z.Z.; Fang, D.Y.; Yan, H.J.; Zeng, G.C.; Jiang, L.F. Identification of a novel infection-enhancing epitope on dengue prM using a dengue cross-reacting monoclonal antibody. BMC Microbiol. 2013, 13, 194. [Google Scholar] [CrossRef]
  79. Schreuder, H.; Matter, H. Serine Proteinases from the Blood Coagulation Cascade. In Structural Biology in Drug Discovery; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2020; pp. 395–422. Available online: https://onlinelibrary.wiley.com/doi/abs/10.1002/9781118681121.ch17 (accessed on 9 April 2024).
  80. Franczuk, P.; Tkaczyszyn, M.; Kulak, M.; Domenico, E.; Ponikowski, P.; Jankowska, E.A. Cardiovascular Complications of Viral Respiratory Infections and COVID-19. Biomedicines 2022, 11, 71. [Google Scholar] [CrossRef]
  81. Lazarowitz, S.G.; Choppin, P.W. Enhancement of the infectivity of influenza A and B viruses by proteolytic cleavage of the hemagglutinin polypeptide. Virology 1975, 68, 440–454. [Google Scholar] [CrossRef]
  82. Sun, X.; Tse, L.V.; Ferguson, A.D.; Whittaker, G.R. Modifications to the Hemagglutinin Cleavage Site Control the Virulence of a Neurotropic H1N1 Influenza Virus. J. Virol. 2010, 84, 8683–8690. [Google Scholar] [CrossRef]
  83. Gotoh, B.; Ogasawara, T.; Toyoda, T.; Inocencio, N.M.; Hamaguchi, M.; Nagai, Y. An endoprotease homologous to the blood clotting factor X as a determinant of viral tropism in chick embryo. EMBO J. 1990, 9, 4189–4195. [Google Scholar] [CrossRef]
  84. Kummarapurugu, A.B.; Hawkridge, A.M.; Ma, J.; Osei, S.; Martin, R.K.; Zheng, S.; Voynow, J.A. Neutrophil elastase decreases SARS-CoV-2 spike protein binding to human bronchial epithelia by clipping ACE-2 ectodomain from the epithelial surface. J. Biol. Chem. 2023, 299, 104820. [Google Scholar] [CrossRef]
  85. Masic, A.; Booth, J.S.; Mutwiri, G.K.; Babiuk, L.A.; Zhou, Y. Elastase-Dependent Live Attenuated Swine Influenza A Viruses Are Immunogenic and Confer Protection against Swine Influenza A Virus Infection in Pigs. J. Virol. 2009, 83, 10198–10210. [Google Scholar] [CrossRef]
  86. Bugge, T.H.; Antalis, T.M.; Wu, Q. Type II Transmembrane Serine Proteases. J. Biol. Chem. 2009, 284, 23177–23181. [Google Scholar] [CrossRef]
  87. Béliveau, F.; Désilets, A.; Leduc, R. Probing the substrate specificities of matriptase, matriptase-2, hepsin and DESC1 with internally quenched fluorescent peptides. FEBS J. 2009, 276, 2213–2226. [Google Scholar] [CrossRef]
  88. Zmora, P.; Blazejewska, P.; Moldenhauer, A.S.; Welsch, K.; Nehlmeier, I.; Wu, Q.; Schneider, H.; Pöhlmann, S.; Bertram, S. DESC1 and MSPL Activate Influenza A Viruses and Emerging Coronaviruses for Host Cell Entry. J. Virol. 2014, 88, 12087–12097. [Google Scholar] [CrossRef]
  89. Yasuoka, S.; Ohnishi, T.; Kawano, S.; Tsuchihashi, S.; Ogawara, M.; Masuda, K.; Yamaoka, K.; Takahashi, M.; Sano, T. Purification, characterization, and localization of a novel trypsin-like protease found in the human airway. Am. J. Respir. Cell Mol. Biol. 1997, 16, 300–308. [Google Scholar] [CrossRef]
  90. Wysocka, M.; Spichalska, B.; Lesner, A.; Jaros, M.; Brzozowski, K.; Łęgowska, A.; Rolka, K. Substrate specificity and inhibitory study of human airway trypsin-like protease. Bioorganic Med. Chem. 2010, 18, 5504–5509. [Google Scholar] [CrossRef]
  91. Böttcher, E.; Matrosovich, T.; Beyerle, M.; Klenk, H.D.; Garten, W.; Matrosovich, M. Proteolytic Activation of Influenza Viruses by Serine Proteases TMPRSS2 and HAT from Human Airway Epithelium. J. Virol. 2006, 80, 9896–9898. [Google Scholar] [CrossRef]
  92. Bertram, S.; Dijkman, R.; Habjan, M.; Heurich, A.; Gierer, S.; Glowacka, I.; Welsch, K.; Winkler, M.; Schneider, H.; Hofmann-Winkler, H.; 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, 6150–6160. [Google Scholar] [CrossRef]
  93. Bertram, S.; Glowacka, I.; Müller, M.A.; Lavender, H.; Gnirss, K.; Nehlmeier, I.; Niemeyer, D.; He, Y.; Simmons, G.; Drosten, C.; et al. Cleavage and Activation of the Severe Acute Respiratory Syndrome Coronavirus Spike Protein by Human Airway Trypsin-Like Protease. J. Virol. 2011, 85, 13363–13372. [Google Scholar] [CrossRef]
  94. Beaulieu, A.; Gravel, É.; Cloutier, A.; Marois, I.; Colombo, É.; Désilets, A.; Verreault, C.; Leduc, R.; Marsault, É.; Richter, M.V. Matriptase Proteolytically Activates Influenza Virus and Promotes Multicycle Replication in the Human Airway Epithelium. J. Virol. 2013, 87, 4237–4251. [Google Scholar] [CrossRef]
  95. Whittaker, G.R.; Straus, M.R. Human matriptase/ST 14 proteolytically cleaves H7N9 hemagglutinin and facilitates the activation of influenza A/Shanghai/2/2013 virus in cell culture. Influenza Other Respir. Viruses 2020, 14, 189–195. [Google Scholar] [CrossRef]
  96. Zmora, P.; Hoffmann, M.; Kollmus, H.; Moldenhauer, A.S.; Danov, O.; Braun, A.; Winkler, M.; Schughart, K.; Pöhlmann, S. TMPRSS11A activates the influenza A virus hemagglutinin and the MERS coronavirus spike protein and is insensitive against blockade by HAI-1. J. Biol. Chem. 2018, 293, 13863–13873. [Google Scholar] [CrossRef]
  97. Kido, H.; Okumura, Y. MSPL/TMPRSS13. Front. Biosci. 2008, 13, 754–758. [Google Scholar] [CrossRef]
  98. Kishimoto, M.; Uemura, K.; Sanaki, T.; Sato, A.; Hall, W.W.; Kariwa, H.; Orba, Y.; Sawa, H.; Sasaki, M. TMPRSS11D and TMPRSS13 Activate the SARS-CoV-2 Spike Protein. Viruses 2021, 13, 384. [Google Scholar] [CrossRef]
  99. Yun Kim, S.; Park, D.; Oh, M.; Sellamuthu, S.; Park, W.J. Detection of site-specific proteolysis in secretory pathways. Biochem. Biophys. Res. Commun. 2002, 296, 419–424. [Google Scholar] [CrossRef]
  100. Mahmoud, I.S.; Jarrar, Y.B.; Alshaer, W.; Ismail, S. SARS-CoV-2 entry in host cells-multiple targets for treatment and prevention. Biochimie 2020, 175, 93–98. [Google Scholar] [CrossRef]
  101. Shirogane, Y.; Takeda, M.; Iwasaki, M.; Ishiguro, N.; Takeuchi, H.; Nakatsu, Y.; Tahara, M.; Kikuta, H.; Yanagi, Y. Efficient Multiplication of Human Metapneumovirus in Vero Cells Expressing the Transmembrane Serine Protease TMPRSS2. J. Virol. 2008, 82, 8942–8946. [Google Scholar] [CrossRef]
  102. 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, e01815-18. [Google Scholar] [CrossRef]
  103. Abe, M.; Tahara, M.; Sakai, K.; Yamaguchi, H.; Kanou, K.; Shirato, K.; Kawase, M.; Noda, M.; Kimura, H.; Matsuyama, S.; et al. TMPRSS2 Is an Activating Protease for Respiratory Parainfluenza Viruses. J. Virol. 2013, 87, 11930–11935. [Google Scholar] [CrossRef]
  104. Esram, P.; Arumugam, P. Development and Validation of an Enzymatic Assay for TMPRSS4: Evaluation of Molecular Inhibitors. J. Adv. Zool. 2023, 44, 309–321. [Google Scholar] [CrossRef]
  105. Kühn, N.; Bergmann, S.; Kösterke, N.; Lambertz, R.L.O.; Keppner, A.; Van den Brand, J.M.A.; Pöhlmann, S.; Weiß, S.; Hummler, E.; Hatesuer, B.; et al. The Proteolytic Activation of (H3N2) Influenza A Virus Hemagglutinin Is Facilitated by Different Type II Transmembrane Serine Proteases. J. Virol. 2016, 90, 4298–4307. [Google Scholar] [CrossRef]
  106. Zang, R.; Gomez Castro, M.F.; McCune, B.T.; Zeng, Q.; Rothlauf, P.W.; Sonnek, N.M.; Liu, Z.; Brulois, K.F.; Wang, X.; Greenberg, H.B.; et al. TMPRSS2 and TMPRSS4 promote SARS-CoV-2 infection of human small intestinal enterocytes. Sci. Immunol. 2020, 5, eabc3582. [Google Scholar] [CrossRef]
  107. 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]
  108. Murza, A.; Dion, S.P.; Boudreault, P.L.; Désilets, A.; Leduc, R.; Marsault, É. Inhibitors of type II transmembrane serine proteases in the treatment of diseases of the respiratory tract – A review of patent literature. Expert Opin. Ther. Pat. 2020, 30, 807–824. [Google Scholar] [CrossRef]
  109. 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]
  110. Beaudoin, C.A.; Pandurangan, A.P.; Kim, S.Y.; Hamaia, S.W.; Huang, C.L.H.; Blundell, T.L.; Chaitanya Vedithis, S.; Jackson, A.P. In silico analysis of mutations near S1/S2 cleavage site in SARS-CoV-2 spike protein reveals increased propensity of glycosylation in Omicron strain. J. Med. Virol. 2022, 94, 4181–4192. [Google Scholar] [CrossRef]
  111. Yao, Z.; Zhang, L.; Duan, Y.; Tang, X.; Lu, J. Molecular insights into the adaptive evolution of SARS-CoV-2 spike protein. J. Infect. 2024, 88, 106121. [Google Scholar] [CrossRef]
  112. Kirschke, H. Chapter 410—Cathepsin L. In Handbook of Proteolytic Enzymes, 3rd ed.; Rawlings, N.D., Salvesen, G., Eds.; Academic Press: Cambridge, MA, USA, 2013; pp. 1808–1817. Available online: https://www.sciencedirect.com/science/article/pii/B9780123822192004105 (accessed on 13 July 2023).
  113. 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]
  114. Du, X.; Tang, H.; Gao, L.; Wu, Z.; Meng, F.; Yan, R.; Qiao, S.; An, J.; Wang, C.; Qin, F.X.F. Omicron adopts a different strategy from Delta and other variants to adapt to host. Sig. Transduct. Target. Ther. 2022, 7, 45. [Google Scholar] [CrossRef]
  115. Willett, B.J.; Grove, J.; MacLean, O.A.; Wilkie, C.; De Lorenzo, G.; Furnon, W.; Cantoni, D.; Scott, S.; Logan, N.; Ashraf, S.; et al. SARS-CoV-2 Omicron is an immune escape variant with an altered cell entry pathway. Nat. Microbiol. 2022, 7, 1161–1179. [Google Scholar] [CrossRef]
  116. Mykytyn, A.Z.; Breugem, T.I.; Geurts, M.H.; Beumer, J.; Schipper, D.; Van Acker, R.; Van den Doel, P.B.; Van Royen, M.E.; Zhang, J.; Clevers, H.; et al. SARS-CoV-2 Omicron entry is type II transmembrane serine protease-mediated in human airway and intestinal organoid models. J. Virol. 2023, 97, e00851-23. [Google Scholar] [CrossRef]
  117. Scarcella, M.; d’Angelo, D.; Ciampa, M.; Tafuri, S.; Avallone, L.; Pavone, L.M.; De Pasquale, V. The Key Role of Lysosomal Protease Cathepsins in Viral Infections. Int. J. Mol. Sci. 2022, 23, 9089. [Google Scholar] [CrossRef]
  118. Mainou, B.A. The Orchestra of Reovirus Cell Entry. Curr. Clin. Micro. Rpt. 2017, 4, 142–149. [Google Scholar] [CrossRef]
  119. Baer, G.S.; Dermody, T.S. Mutations in reovirus outer-capsid protein sigma3 selected during persistent infections of L cells confer resistance to protease inhibitor E64. J. Virol. 1997, 71, 4921–4928. [Google Scholar] [CrossRef]
  120. Wilson, G.J.; Nason, E.L.; Hardy, C.S.; Ebert, D.H.; Wetzel, J.D.; Venkataram Prasad, B.V.; Dermody, T.S. A Single Mutation in the Carboxy Terminus of Reovirus Outer-Capsid Protein σ3 Confers Enhanced Kinetics of σ3 Proteolysis, Resistance to Inhibitors of Viral Disassembly, and Alterations in σ3 Structure. J. Virol. 2002, 76, 9832–9843. [Google Scholar] [CrossRef]
  121. Doyle, J.D.; Danthi, P.; Kendall, E.A.; Ooms, L.S.; Wetzel, J.D.; Dermody, T.S. Molecular Determinants of Proteolytic Disassembly of the Reovirus Outer Capsid. J. Biol. Chem. 2012, 287, 8029–8038. [Google Scholar] [CrossRef]
  122. Mainou, B.A.; Dermody, T.S. In Search of Cathepsins: How Reovirus Enters Host Cells. DNA Cell Biol. 2012, 31, 1646–1649. [Google Scholar] [CrossRef]
  123. Schornberg, K.; Matsuyama, S.; Kabsch, K.; Delos, S.; Bouton, A.; White, J. Role of Endosomal Cathepsins in Entry Mediated by the Ebola Virus Glycoprotein. J. Virol. 2006, 80, 4174–4178. [Google Scholar] [CrossRef]
  124. Diederich, S.; Sauerhering, L.; Weis, M.; Altmeppen, H.; Schaschke, N.; Reinheckel, T.; Erbar, S.; Maisner, A. Activation of the Nipah virus fusion protein in MDCK cells is mediated by cathepsin B within the endosome-recycling compartment. J. Virol. 2012, 86, 3736–3745. [Google Scholar] [CrossRef]
  125. Vogt, C.; Eickmann, M.; Diederich, S.; Moll, M.; Maisner, A. Endocytosis of the Nipah Virus Glycoproteins. J. Virol. 2005, 79, 3865–3872. [Google Scholar] [CrossRef]
  126. Corry, J.; Johnson, S.M.; Cornwell, J.; Peeples, M.E. Preventing Cleavage of the Respiratory Syncytial Virus Attachment Protein in Vero Cells Rescues the Infectivity of Progeny Virus for Primary Human Airway Cultures. J. Virol. 2016, 90, 1311–1320. [Google Scholar] [CrossRef]
  127. Biniossek, M.L.; Nägler, D.K.; Becker-Pauly, C.; Schilling, O. Proteomic Identification of Protease Cleavage Sites Characterizes Prime and Non-prime Specificity of Cysteine Cathepsins B., L., and S. J. Proteome Res. 2011, 10, 5363–5373. [Google Scholar] [CrossRef]
  128. Simmons, G.; Gosalia, D.N.; Rennekamp, A.J.; Reeves, J.D.; Diamond, S.L.; Bates, P. Inhibitors of cathepsin L prevent severe acute respiratory syndrome coronavirus entry. Proc. Natl. Acad. Sci. USA 2005, 102, 11876–11881. [Google Scholar] [CrossRef]
  129. 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, 6537–6545. [Google Scholar] [CrossRef]
  130. Belouzard, S.; Chu, V.C.; Whittaker, G.R. Activation of the SARS coronavirus spike protein via sequential proteolytic cleavage at two distinct sites. Proc. Natl. Acad. Sci. USA 2009, 106, 5871–5876. [Google Scholar] [CrossRef]
  131. 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.; et al. 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. Sig. Transduct. Target Ther. 2021, 6, 134. [Google Scholar] [CrossRef]
  132. Ebert, D.H.; Deussing, J.; Peters, C.; Dermody, T.S. Cathepsin L and Cathepsin B Mediate Reovirus Disassembly in Murine Fibroblast Cells. J. Biol. Chem. 2002, 277, 24609–24617. [Google Scholar] [CrossRef]
  133. Kleine-Weber, H.; Elzayat, M.T.; Hoffmann, M.; Pöhlmann, S. Functional analysis of potential cleavage sites in the MERS-coronavirus spike protein. Sci. Rep. 2018, 8, 16597. [Google Scholar] [CrossRef]
  134. Pager, C.T.; Craft, W.W.; Patch, J.; Dutch, R.E. A mature and fusogenic form of the Nipah virus fusion protein requires proteolytic processing by cathepsin L. Virology 2006, 346, 251–257. [Google Scholar] [CrossRef]
  135. Coleman, M.D.; Ha, S.D.; Haeryfar, S.M.M.; Barr, S.D.; Kim, S.O. Cathepsin B plays a key role in optimal production of the influenza A virus. J. Virol. Antivir. Res. 2018, 7, 1–20. [Google Scholar] [CrossRef]
  136. Bollavaram, K.; Leeman, T.H.; Lee, M.W.; Kulkarni, A.; Upshaw, S.G.; Yang, J.; Song, H.; Platt, M.O. Multiple sites on SARS-CoV-2 spike protein are susceptible to proteolysis by cathepsins B, K, L, S, and V. Protein Sci. 2021, 30, 1131–1143. [Google Scholar] [CrossRef]
  137. Golden, J.W.; Bahe, J.A.; Lucas, W.T.; Nibert, M.L.; Schiff, L.A. Cathepsin S Supports Acid-independent Infection by Some Reoviruses. J. Biol. Chem. 2004, 279, 8547–8557. [Google Scholar] [CrossRef]
  138. Brinkworth, R.I.; Tort, J.F.; Brindley, P.J.; Dalton, J.P. Phylogenetic relationships and theoretical model of human cathepsin W (lymphopain), a cysteine proteinase from cytotoxic T lymphocytes. Int. J. Biochem. Cell Biol. 2000, 32, 373–384. [Google Scholar] [CrossRef]
  139. Edinger, T.O.; Pohl, M.O.; Yángüez, E.; Stertz, S. Cathepsin W Is Required for Escape of Influenza A Virus from Late Endosomes. mBio 2015, 6, e00297. [Google Scholar] [CrossRef]
  140. Brömme, D. Chapter 409—Cathepsin K. In Handbook of Proteolytic Enzymes, 3rd ed.; Rawlings, N.D., Salvesen, G., Eds.; Academic Press: Cambridge, MA, USA, 2013; pp. 1801–1807. Available online: https://www.sciencedirect.com/science/article/pii/B9780123822192004099 (accessed on 4 May 2024).
  141. Brömme, D. Chapter 414—Cathepsin V. In Handbook of Proteolytic Enzymes, 3rd ed.; Rawlings, N.D., Salvesen, G., Eds.; Academic Press: Cambridge, MA, USA, 2013; pp. 1831–1834. Available online: https://www.sciencedirect.com/science/article/pii/B9780123822192004130 (accessed on 4 May 2024).
  142. Cabral-Pacheco, G.A.; Garza-Veloz, I.; Castruita-De la Rosa, C.; Ramirez-Acuña, J.M.; Perez-Romero, B.A.; Guerrero-Rodriguez, J.F.; Martinez-Avila, N.; Martinez-Fierro, M.L. The Roles of Matrix Metalloproteinases and Their Inhibitors in Human Diseases. Int. J. Mol. Sci. 2020, 21, 9739. [Google Scholar] [CrossRef]
  143. Yeo, S.J.; Yun, Y.J.; Lyu, M.A.; Woo, S.Y.; Woo, E.R.; Kim, S.J.; Lee, H.J.; Park, H.K.; Kook, Y.H. Respiratory syncytial virus infection induces matrix metalloproteinase-9 expression in epithelial cells. Arch. Virol. 2002, 147, 229–242. [Google Scholar] [CrossRef]
  144. Rojas-Quintero, J.; Wang, X.; Tipper, J.; Burkett, P.R.; Zuñiga, J.; Ashtekar, A.R.; Polverino, F.; Rout, A.; Yambayev, I.; Hernández, C.; et al. Matrix metalloproteinase-9 deficiency protects mice from severe influenza A viral infection. JCI Insight 2018, 3, e99022. [Google Scholar] [CrossRef]
  145. Tacon, C.E.; Wiehler, S.; Holden, N.S.; Newton, R.; Proud, D.; Leigh, R. Human Rhinovirus Infection Up-Regulates MMP-9 Production in Airway Epithelial Cells via NF-κB. Am. J. Respir. Cell Mol. Biol. 2010, 43, 201–209. [Google Scholar] [CrossRef]
  146. Elliott, M.B.; Welliver, R.C.; Laughlin, T.S.; Pryharski, K.S.; LaPierre, N.A.; Chen, T.; Souza, V.; Terio, N.B.; Hancock, G.E. Matrix metalloproteinase-9 and tissue inhibitor of matrix metalloproteinase-1 in the respiratory tracts of human infants following paramyxovirus infection. J. Med. Virol. 2007, 79, 447–456. [Google Scholar] [CrossRef]
  147. Phillips, J.M.; Gallagher, T.; Weiss, S.R. Neurovirulent Murine Coronavirus JHM.SD Uses Cellular Zinc Metalloproteases for Virus Entry and Cell-Cell Fusion. J. Virol. 2017, 91, e01564-16. [Google Scholar] [CrossRef]
  148. Yamamoto, M.; Gohda, J.; Kobayashi, A.; Tomita, K.; Hirayama, Y.; Koshikawa, N.; Seiki, M.; Semba, K.; Akiyama, T.; Kawaguchi, Y.; et al. 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, e0051922. [Google Scholar] [CrossRef]
  149. Chan, J.F.W.; Huang, X.; Hu, B.; Chai, Y.; Shi, H.; Zhu, T.; Yuen, T.T.T.; Liu, Y.; Liu, H.; Shi, J.; et al. Altered host protease determinants for SARS-CoV-2 Omicron. Sci. Adv. 2023, 9, eadd3867. [Google Scholar] [CrossRef]
  150. Fernandez-Patron, C.; Hardy, E. Matrix Metalloproteinases in Health and Disease in the Times of COVID-19. Biomolecules 2022, 12, 692. [Google Scholar] [CrossRef]
  151. Syed, F.; Li, W.; Relich, R.F.; Russell, P.M.; Zhang, S.; Zimmerman, M.K.; Yu, Q. Excessive Matrix Metalloproteinase-1 and Hyperactivation of Endothelial Cells Occurred in COVID-19 Patients and Were Associated With the Severity of COVID-19. J. Infect. Dis. 2021, 224, 60–69. [Google Scholar] [CrossRef]
  152. Safont, B.; Tarraso, J.; Rodriguez-Borja, E.; Fernández-Fabrellas, E.; Sancho-Chust, J.N.; Molina, V.; Lopez-Ramirez, C.; Lopez-Martinez, A.; Cabanes, L.; Andreu, A.L.; et al. Lung Function, Radiological Findings and Biomarkers of Fibrogenesis in a Cohort of COVID-19 Patients Six Months After Hospital Discharge. Arch. Bronconeumol. 2022, 58, 142–149. [Google Scholar] [CrossRef]
  153. Blascke de Mello, M.M.; Parente, J.M.; Schulz, R.; Castro, M.M. Matrix metalloproteinase (MMP)-2 activation by oxidative stress decreases aortic calponin-1 levels during hypertrophic remodeling in early hypertension. Vasc. Pharmacol. 2019, 116, 36–44. [Google Scholar] [CrossRef]
  154. D`Avila-Mesquita, C.; Couto, A.E.S.; Campos, L.C.B.; Vasconcelos, T.F.; Michelon-Barbosa, J.; Corsi, C.A.C.; Mestriner, F.; Petroski-Moraes, B.C.; Garbellinini-Diab, M.J.; Cuoto, D.M.S.; et al. MMP-2 and MMP-9 levels in plasma are altered and associated with mortality in COVID-19 patients. Biomed. Pharmacother. 2021, 142, 112067. [Google Scholar] [CrossRef]
  155. Salomão, R.; Assis, V.; De Sousa Neto, I.V.; Petriz, B.; Babault, N.; Durigan, J.L.Q.; De Cássia Marqueti, R. Involvement of Matrix Metalloproteinases in COVID-19: Molecular Targets, Mechanisms, and Insights for Therapeutic Interventions. Biology 2023, 12, 843. [Google Scholar] [CrossRef]
  156. 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]
  157. 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]
  158. Zhang, H.; Lv, P.; Jiang, J.; Liu, Y.; Yan, R.; Shu, S.; Hu, B.; Xiao, H.; Cai, K.; Yuan, S.; et al. Advances in developing ACE2 derivatives against SARS-CoV-2. Lancet Microbe 2023, 4, e369-78. [Google Scholar] [CrossRef]
  159. Heindl, M.R.; Rupp, A.L.; Schwerdtner, M.; Bestle, D.; Harbig, A.; De Rocher, A.; Schmacke, L.C.; Staker, B.; Steinmetzer, T.; Stein, D.A.; et al. ACE2 acts as a novel regulator of TMPRSS2-catalyzed proteolytic activation of influenza A virus in airway cells. J. Virol. 2024, 98, e00102-24. [Google Scholar] [CrossRef]
  160. Tai, W.; He, L.; Zhang, X.; Pu, J.; Voronin, D.; Jiang, S.; Zhou, S.; Gu, 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]
  161. Li, W.; Moore, M.J.; Vasilieva, N.; Sui, J.; Wong, S.K.; Berne, M.A.; Somasundaran, M.; Sullivan, J.L.; Luzuriaga, K.; Greenough, T.C.; et al. Angiotensin-converting enzyme 2 is a functional receptor for the SARS coronavirus. Nature 2003, 426, 450–454. [Google Scholar] [CrossRef]
  162. Hofmann, H.; Pyrc, K.; Van der Hoek, L.; Geier, M.; Berkhout, B.; Pöhlmann, S. Human coronavirus NL63 employs the severe acute respiratory syndrome coronavirus receptor for cellular entry. Proc. Natl. Acad. Sci. USA 2005, 102, 7988–7993. [Google Scholar] [CrossRef]
  163. Funk, C.J.; Wang, J.; Ito, Y.; Travanty, E.A.; Voelker, D.R.; Holmes, K.V.; Mason, R.J. Infection of human alveolar macrophages by human coronavirus strain 229E. J. Gen. Virol. 2012, 93 Pt 3, 494–503. [Google Scholar] [CrossRef]
  164. Hajizadeh, M.; Moosavi-Movahedi, Z.; Sheibani, N.; Moosavi-Movahedi, A.A. An outlook on suicide enzyme inhibition and drug design. J. Iran. Chem. Soc. 2022, 19, 1575–1592. [Google Scholar] [CrossRef]
  165. Ben-Tal, A.; Nir, K. Introduction to Proteins: Structure, Function, and Motion, 2nd ed.; Chapman and Hall/CRC: New York, NY, USA, 2018; 988p. [Google Scholar]
  166. Rajendran, L.; Knölker, H.J.; Simons, K. Subcellular targeting strategies for drug design and delivery. Nat. Rev. Drug Discov. 2010, 9, 29–42. [Google Scholar] [CrossRef]
  167. Donahue, N.D.; Acar, H.; Wilhelm, S. Concepts of nanoparticle cellular uptake, intracellular trafficking, and kinetics in nanomedicine. Adv. Drug Deliv. Rev. 2019, 143, 68–96. [Google Scholar] [CrossRef]
  168. Lotke, R.; Petersen, M.; Sauter, D. Restriction of Viral Glycoprotein Maturation by Cellular Protease Inhibitors. Viruses 2024, 16, 332. [Google Scholar] [CrossRef]
  169. Stein, P.; Chothia, C. Serpin tertiary structure transformation. J. Mol. Biol. 1991, 221, 615–621. [Google Scholar] [CrossRef]
  170. Nichols, D.B.; De Martini, W.; Cottrell, J. Poxviruses Utilize Multiple Strategies to Inhibit Apoptosis. Viruses 2017, 9, 215. [Google Scholar] [CrossRef]
  171. Varkoly, K.; Beladi, R.; Hamada, M.; McFadden, G.; Irving, J.; Lucas, A.R. Viral SERPINS—A Family of Highly Potent Immune-Modulating Therapeutic Proteins. Biomolecules 2023, 13, 1393. [Google Scholar] [CrossRef]
  172. Viswanathan, K.; Liu, L.; Vaziri, S.; Dai, E.; Richardson, J.; Togonu-Bickersteth, B.; Vatsya, P.; Christov, A.; Lucas, A.R. Myxoma viral serpin, Serp-1, a unique interceptor of coagulation and innate immune pathways. Thromb. Haemost. 2006, 95, 499–510. [Google Scholar] [CrossRef]
  173. Lomas, D.A.; Evans, D.L.; Upton, C.; McFadden, G.; Carrell, R.W. Inhibition of plasmin, urokinase, tissue plasminogen activator, and C1S by a myxoma virus serine proteinase inhibitor. J. Biol. Chem. 1993, 268, 516–521. [Google Scholar] [CrossRef]
  174. Guo, Q.; Yaron, J.R.; Wallen, J.W.; Browder, K.F.; Boyd, R.; Olson, T.L.; Burgin, M.; Ulrich, P.; Aliskevich, E.; Schutz, L.N.; et al. PEGylated Serp-1 Markedly Reduces Pristane-Induced Experimental Diffuse Alveolar Hemorrhage, Altering uPAR Distribution, and Macrophage Invasion. Front. Cardiovasc. Med. 2021, 8, 633212. [Google Scholar] [CrossRef]
  175. Tarighi, P.; Eftekhari, S.; Chizari, M.; Sabernavaei, M.; Jafari, D.; Mirzabeigi, P. A review of potential suggested drugs for coronavirus disease (COVID-19) treatment. Eur. J. Pharmacol. 2021, 895, 173890. [Google Scholar] [CrossRef]
  176. Wyde, P.R.; Chetty, S.N.; Jewell, A.M.; Boivin, G.; Piedra, P.A. Comparison of the inhibition of human metapneumovirus and respiratory syncytial virus by ribavirin and immune serum globulin in vitro. Antivir. Res. 2003, 60, 51–59. [Google Scholar] [CrossRef]
  177. Thomas, G.; Couture, F.; Kwiatkowska, A. The Path to Therapeutic Furin Inhibitors: From Yeast Pheromones to SARS-CoV-2. Int. J. Mol. Sci. 2022, 23, 3435. [Google Scholar] [CrossRef]
  178. Essalmani, R.; Jain, J.; Susan-Resiga, D.; Andréo, U.; Evagelidis, A.; Derbali, R.M.; Huynh, D.N.; Dallaire, F.; Laporte, M.; Delpal, A.; et al. Distinctive Roles of Furin and TMPRSS2 in SARS-CoV-2 Infectivity. J. Virol. 2022, 96, e00128-22. [Google Scholar] [CrossRef]
  179. Someya, A.; Tanaka, N.; Okuyama, A. Inhibition of influenza virus AWSN replication by a trypsin inhibitor, 6-amidino-2-naphthyl p-guanidinobenzoate. Biochem. Biophys. Res. Commun. 1990, 169, 148–152. [Google Scholar] [CrossRef]
  180. Hosoya, M.; Matsuyama, S.; Baba, M.; Suzuki, H.; Shigeta, S. Effects of protease inhibitors on replication of various myxoviruses. Antimicrob. Agents Chemother. 1992, 36, 1432–1436. [Google Scholar] [CrossRef]
  181. Hosoya, M.; Shigeta, S.; Ishii, T.; Suzuki, H.; Clercq, E.D. Comparative Inhibitory Effects of Various Nucleoside and Nonnucleoside Analogues on Replication of Influenza Virus Types A and B In Vitro and In Ovo. J. Infect. Dis. 1993, 168, 641–646. [Google Scholar] [CrossRef]
  182. Gunst, J.D.; Staerke, N.B.; Pahus, M.H.; Kristensen, L.H.; Bodilsen, J.; Lohse, N.; Dalgaard, L.S.; Brønnum, D.; Fröbert, O.; Hønge, B.; et al. Efficacy of the TMPRSS2 inhibitor camostat mesilate in patients hospitalized with Covid-19-a double-blind randomized controlled trial. EClinicalMedicine 2021, 35, 100849. [Google Scholar] [CrossRef]
  183. Zhirnov, O.P.; Klenk, H.D.; Wright, P.F. Aprotinin and similar protease inhibitors as drugs against influenza. Antivir. Res. 2011, 92, 27–36. [Google Scholar] [CrossRef]
  184. Grocott, H.P.; Sheng, H.; Miura, Y.; Sarraf-Yazdi, S.; Mackensen, G.B.; Pearlstein, R.D.; Warner, D.S. The effects of aprotinin on outcome from cerebral ischemia in the rat. Anesth. Analg. 1999, 88, 1–7. [Google Scholar] [CrossRef]
  185. Mangano, D.T.; Tudor, I.C.; Dietzel, C. The risk associated with aprotinin in cardiac surgery. N. Engl. J. Med. 2006, 354, 353–365. [Google Scholar] [CrossRef]
  186. Hamilton, B.S.; Chung, C.; Cyphers, S.Y.; Rinaldi, V.D.; Marcano, V.C.; Whittaker, G.R. Inhibition of influenza virus infection and hemagglutinin cleavage by the protease inhibitor HAI-2. Biochem. Biophys. Res. Commun. 2014, 450, 1070–1075. [Google Scholar] [CrossRef]
  187. Straus, M.R.; Kinder, J.T.; Segall, M.; Dutch, R.E.; Whittaker, G.R. SPINT2 inhibits proteases involved in activation of both influenza viruses and metapneumoviruses. Virology 2020, 543, 43–53. [Google Scholar] [CrossRef]
  188. Shapira, T.; Monreal, I.A.; Dion, S.P.; Buchholz, D.W.; Imbiakha, B.; Olmstead, A.D.; Jager, M.; Désilets, A.; Gao, G.; Martins, M.; et al. A TMPRSS2 inhibitor acts as a pan-SARS-CoV-2 prophylactic and therapeutic. Nature 2022, 605, 340–348. [Google Scholar] [CrossRef]
Figure 1. Substrate (left) cleaved by protease (right) with the scissile bond marked by an arrow.
Figure 1. Substrate (left) cleaved by protease (right) with the scissile bond marked by an arrow.
Viruses 16 00984 g001
Figure 2. Furin cleavage of influenza HA and SARS-CoV-2 S. HA has one cleavage site (CS) while S has two (S1/S2 and S2′). The location of the fusion peptide of each glycoprotein is indicated by FP.
Figure 2. Furin cleavage of influenza HA and SARS-CoV-2 S. HA has one cleavage site (CS) while S has two (S1/S2 and S2′). The location of the fusion peptide of each glycoprotein is indicated by FP.
Viruses 16 00984 g002
Figure 3. Artistic representation of one monomeric unit of SARS-CoV-2 S being cleaved by furin and then entering into the “up” conformation with the receptor binding domain exposed.
Figure 3. Artistic representation of one monomeric unit of SARS-CoV-2 S being cleaved by furin and then entering into the “up” conformation with the receptor binding domain exposed.
Viruses 16 00984 g003
Figure 4. Illustration of the different types of inhibition.
Figure 4. Illustration of the different types of inhibition.
Viruses 16 00984 g004
Table 1. Characteristics of well-known serine proteases outside of the TTSP family.
Table 1. Characteristics of well-known serine proteases outside of the TTSP family.
Serine ProteasesCleavage PreferencesExploiting Respiratory Viruses
ChymaseAromatic amino acids at P1, aliphatic amino acids from P2 to P4, S at P1′, E/D at P2′, A/V/G at P3′ [17]Mammalian Orthoreovirus [18]
Factor XaPreference for PXG/AR↓XXD [19]SARS-CoV-1 [20], SARS-CoV-2 [21]
FurinR-X-X-R↓S [22]Influenza virus [23], SARS-CoV-2 [24], MERS-CoV [25], RSV [26], HPIV [27], measles virus (MeV) [28], mumps virus [29]
Kallikrein-related peptidase 13 (KLK13]V/Y-R/L/F/M-R↓ [30]HKU-1 [31]
KLK1R/Y↓S/R [32]Influenza virus [33]
KLK5X(aliphatic/aromatic)-R/K -X(polar/aliphatic)- R↓ [30]Influenza virus [33]
Neutrophil elastaseA/V/I/T↓ [34]Mammalian Orthoreovirus [35], SARS-CoV-1 [36,37]
PC5/6R-X-R/K-R↓ [38]Influenza virus [39], RSV [39]
PlasminR/L↓ [40], preference for aromatic hydrophobic residue at P2 [41]Influenza virus [42], RSV [43], SARS-CoV-1 [44], SARS-CoV-2 [45]
ThrombinL-X-P-R↓S/A/G/T-X(aromatic)-R [46]RSV [43], SARS-CoV-2 [21]
TrypsinK/R↓ [34]SARS-CoV-1 [47,48], HMPV [49], RSV [43], HPIV [50]
Cathepsin GPreference for F, Y, W, or L at P1 [51], S at P6, negatively charged amino acid in P2′ positionMammalian Orthoreovirus [18], RSV [52]
Table 2. Characteristics of select Type II Transmembrane Serine Proteases.
Table 2. Characteristics of select Type II Transmembrane Serine Proteases.
Type II Transmembrane Serine Proteases (TTSPs)Cleavage PreferencesExploiting Respiratory Viruses
DESC1R-R/A/L-L-A↓ [87]Influenza virus [88], MERS-CoV [88], SARS-CoV-1 [88]
Human Airway Trypsin-like ProteaseR/K↓ [89,90]Influenza virus [91], HCoV-229E [92], SARS-CoV-1 [93], Mammalian Orthoreovirus [18]
MatriptaseMinimum: R/K [34]↓
Preferred: R-X(non-basic)-S-R↓ [87]
Influenza A virus [94,95]
TMPRSS11aUnconfirmed, putative R/K↓SARS-CoV-1 [44], influenza virus [96], MERS-CoV [96]
TMPRSS13/MSPLR/K↓, preference for dibasic P2-P1 [97] Influenza virus [88], SARS-CoV-1 [88], MERS-CoV [88], SARS-CoV-2 [98]
TMPRSS2R/K↓ [99]Influenza A + B virus [91], SARS-CoV-2 [100], HMPV [101], HCoV-229E [92], MERS-CoV [102], SARS-CoV [102], HPIV [103], Mammalian Orthoreovirus [18]
TMPRSS4Unconfirmed, putative R/K↓ [104]Influenza A virus [105], SARS-CoV-2 [106]
Table 3. Characteristics of select cathepsins.
Table 3. Characteristics of select cathepsins.
Cysteine ProteasesCleavage DeterminantsExploiting Respiratory Viruses
Cathepsin LPrefers aromatic or aliphatic residues in P2 [127]HCoV-229E [92], SARS-CoV-1 [128,129,130], SARS-CoV-2 [131], Mammalian Orthoreovirus [132], MERS-CoV [133], Nipah [134], RSV [126], Hendra [134]
Cathepsin BPrefers an aromatic or aliphatic residue and tolerates a basic P2, an aromatic residue in P1′ and a P3′ G [127] Influenza A virus [135], Mammalian Orthoreovirus [132], Nipah [124], SARS-CoV-2 [136]
Cathepsin SPrefers aliphatic residues in P2, G/E in P1 [127]Mammalian Orthoreovirus [137], SARS-CoV-2, SARS-CoV-1, RSV [52]
Cathepsin WW/F–L/V–G/A/R↓V–D/N/E/Q (suggested [138])Influenza A virus [139], RSV [52]
Cathepsin KPrefers non-aromatic hydrophobic residues in P2 [140]SARS-CoV-2 [136]
Cathepsin VPrefers hydrophobic residues in P2, P in P3 [141]SARS-CoV-2 [136]
Table 4. Characteristics of select metalloproteases.
Table 4. Characteristics of select metalloproteases.
MetalloproteaseExploiting Respiratory VirusesRole
MT-MMPSARS-CoV-2 [149]Proteolytic cleavage
ADAMSARS-CoV-2 [149]Proteolytic cleavage
ACE2SARS-CoV-2 [160], SARS-CoV [161], HcoV NL63 [162]|Influenza A virus [159], MERS-CoV [159]Receptor|Regulation of TMPRSS2
APNHCoV-229E [163]Receptor
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

Lubinski, B.; Whittaker, G.R. Host Cell Proteases Involved in Human Respiratory Viral Infections and Their Inhibitors: A Review. Viruses 2024, 16, 984. https://doi.org/10.3390/v16060984

AMA Style

Lubinski B, Whittaker GR. Host Cell Proteases Involved in Human Respiratory Viral Infections and Their Inhibitors: A Review. Viruses. 2024; 16(6):984. https://doi.org/10.3390/v16060984

Chicago/Turabian Style

Lubinski, Bailey, and Gary R. Whittaker. 2024. "Host Cell Proteases Involved in Human Respiratory Viral Infections and Their Inhibitors: A Review" Viruses 16, no. 6: 984. https://doi.org/10.3390/v16060984

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