Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 4213 2024-01-08 08:26:50 |
2 format correct -17 word(s) 4196 2024-01-09 09:01:18 | |
3 format correct + 1 word(s) 4197 2024-01-09 09:02:34 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Arman, B.Y.; Brun, J.; Hill, M.L.; Zitzmann, N.; Von Delft, A. SARS-CoV-2 Clinical Trial Results. Encyclopedia. Available online: https://encyclopedia.pub/entry/53529 (accessed on 02 May 2024).
Arman BY, Brun J, Hill ML, Zitzmann N, Von Delft A. SARS-CoV-2 Clinical Trial Results. Encyclopedia. Available at: https://encyclopedia.pub/entry/53529. Accessed May 02, 2024.
Arman, Benediktus Yohan, Juliane Brun, Michelle L. Hill, Nicole Zitzmann, Annette Von Delft. "SARS-CoV-2 Clinical Trial Results" Encyclopedia, https://encyclopedia.pub/entry/53529 (accessed May 02, 2024).
Arman, B.Y., Brun, J., Hill, M.L., Zitzmann, N., & Von Delft, A. (2024, January 08). SARS-CoV-2 Clinical Trial Results. In Encyclopedia. https://encyclopedia.pub/entry/53529
Arman, Benediktus Yohan, et al. "SARS-CoV-2 Clinical Trial Results." Encyclopedia. Web. 08 January, 2024.
SARS-CoV-2 Clinical Trial Results
Edit

The coronavirus disease 2019 (COVID-19) pandemic has claimed over 7 million lives worldwide, providing a stark reminder of the importance of pandemic preparedness. Due to the lack of approved antiviral drugs effective against coronaviruses at the start of the pandemic, the world largely relied on repurposed efforts. Here, the researchers summarise results from randomised controlled trials to date, as well as selected in vitro data of directly acting antivirals, host-targeting antivirals, and immunomodulatory drugs. Overall, repurposing efforts evaluating directly acting antivirals targeting other viral families were largely unsuccessful, whereas several immunomodulatory drugs led to clinical improvement in hospitalised patients with severe disease. In addition, accelerated drug discovery efforts during the pandemic progressed to multiple novel directly acting antivirals with clinical efficacy, including small molecule inhibitors and monoclonal antibodies.

COVID-19 SARS-CoV-2 coronavirus therapeutics clinical trial drug discovery

1. Directly Acting Antivirals (DAA)

Directly acting antivirals (DAA) that include small molecules and some antibodies, target virally encoded proteins directly and efficiently suppress viral replication in vivo [1] (reviewed in [2][3]). DAA commonly targets viral entry and fusion, or key enzymes like viral polymerases and proteases such as the SARS-CoV2 3-chymotrypsin-like protease (3CLpro), also known as main protease (Mpro), and papain-like protease (PLpro) [1][3]. By targeting viral replication directly, DAA are generally thought to have fewer side effects than HTA [4]. However, by targeting the virus directly, DAA are known to cause mutations that can lead to reduced drug efficacy and drug resistance [5][6].
Rapid SARS-CoV-2 small molecule drug discovery efforts were carried out during the 2019 SARS-CoV-2 pandemic, targeting the viral polymerase (remdesivir, Gilead; molnupiravir, Merck, and others) and main protease (nirmatrelvir, Pfizer; ensitrelvir, Shionogi; and others). The pandemic also gave rise to innovative and collaborative approaches such as collaborative alliances across biopharmaceutical companies creating public–private partnerships such as the Innovative Medicines Initiative Corona Accelerated R&D in Europe (https://www.imi-care.eu/, accessed on 8 November 2023), close-knit consortia of companies, and even completely open science consortia with company participation such as the COVID Moonshot [7][8].

1.1. Protease Inhibitors

SARS-CoV-2 Mpro has been a drug development target from early in the pandemic, with numerous novel SARS-CoV-2 inhibitors in the clinical and preclinical pipeline. Two SARS-CoV-2 MPro inhibitors are now under EUA, including nirmatrelvir/ritonavir (brand name Paxlovid) and ensitrelvir (brand name Xocova).
The covalent SARS-CoV-2 Mpro inhibitor nirmatrelvir (PF-07321332) developed by Pfizer is available in combination with ritonavir (available as Paxlovid) and has been evaluated in multiple RCTs to date. Nirmatrelvir showed potent antiviral activity in in vitro assays, with a half-maximum effective concentration (EC50) of 231 nM in Vero E6 cells [9]. In RCTs, when nirmatrelvir/ritonavir was given within three days of symptoms onset, nirmatrelvir-treated individuals had a significantly reduced risk of severe COVID-19 in comparison to placebo-treated controls, alongside reduced viral loads at day five of treatment [10]. A recent phase 2 RCT study result revealed that nirmatrelvir/ritonavir was more effective than the other leading oral antiviral drug for patients with COVID-19 [11]. A drawback of the combination with the pharmacokinetic (PK) enhancer ritonavir are significant drug-drug interactions, preventing the use in patients with severe renal or hepatic impairment [12][13][14]. A second-generation broad-spectrum protease inhibitor, PF-0781883 has been recently disclosed by Pfizer and is currently in Phase 2 clinical trials, with an initial human PK profile suggesting no further need to co-administer with ritonavir [15].
Ensitrelvir (S-217622) was developed by Shionogi as an oral noncovalent nonpeptidic Mpro inhibitor of SARS-CoV-2 using virtual and biological screening of compound libraries and a structural-based drug design strategy of the hit compounds [16][17]. In comparison to nirmatrelvir/ritonavir, ensitrelvir shows an optimized PK profile [17] and displays potent in vitro antiviral activity against SARS-CoV-2 variant of concern (VOC) including Delta (EC50 = 34.8 nM) and Omicron BA.1 (EC50 = 23.9 nM) variants, as well as other coronaviruses: SARS-CoV-1 (EC50 = 0.21 μM), MERS-CoV (EC50 = 1.4 μM), and HCoV OC43 (EC90 = 0.074 μM) [17][18]. Ensitrelvir is progressing to Phase 2/3 clinical trials. Results disclosed from the Phase 2a trial showed a significant reduction in viral titer and RNA on day four, a median time of negative RT-PCR conversion of two days, and acceptable adverse events [19]. This drug recently received emergency regulatory approval from the Ministry of Health, Labour and Welfare of Japan, with ongoing Phase 3 clinical trials and plans to extend the approval for worldwide use.
Other oral main protease inhibitor candidates in clinical trials include PBI-0451 from Pardes Biosciences and EDP-235 from Enanta Pharmaceuticals [20]. In addition, novel initiatives like the crowdsourced open-science effort COVID Moonshot brought together academic and industrial partners from across the world to develop SARS-CoV-2 Mpro inhibitors [8][21]. More than 2400 molecules were synthesised and rapidly shared to create a rich intellectual property-free dataset. The search for a potent covalent COVID-19 Mpro inhibitor has also been aided by a computational pipeline to efficiently identify irreversible inhibitors [21]. Due to its open access and free data-sharing nature, the Moonshot fragment dataset may have aided in the early development of the Shionogi Mpro compound.
Due to its saliency as a target, repurposing efforts early in the SARS-CoV-2 pandemic focused on the Mpro. Repurposed protease inhibitors evaluated in RCTs include lopinavir (with and without ritonavir), darunavir, and danoprevir; however, none of the tested inhibitors showed efficacy against SARS-CoV-2.
Lopinavir is a protease inhibitor that has been developed for the treatment of human immunodeficiency virus (HIV) and is used in combination with the PK enhancer ritonavir that blocks the enzyme cytochrome P450 3A [22]. Lopinavir showed some in vitro activity against both SARS-CoV-1 and SARS-CoV-2 with half maximum inhibitory concentrations (IC50) of 50 and 26 μM, respectively [22]. Lopinavir/ritonavir was assessed in multiple COVID-19 RCTs, however, it failed to show any clinical benefit and is therefore not recommended as COVID-19 standard care for hospitalised patients [23][24].
Darunavir is another inhibitor of the HIV protease that is in clinical use in combination with the PK enhancer cobicistat [25]. Darunavir was not active against SARS-CoV-2 in vitro at clinically relevant concentrations [26] and in line with these results, did not show clinical efficacy in COVID-19 RCTs.
Danoprevir, a hepatitis C virus (HCV) protease (NS3/4A) inhibitor that is used in combination with ritonavir [27] showed cellular activity against SARS-CoV-2 with an EC50 of 87 µM in Vero E6 cells [28]. In hospitalised patients infected with SARS-CoV-2, danoprevir with ritonavir shows a good safety profile [27] with shorter times to PCR negativity and shorter hospital stays compared to the lopinavir/ritonavir group [29].

1.2. RNA-Dependent RNA Polymerase (RdRp) Inhibitor

The RNA-dependent RNA polymerase (RdRp) is essential for viral replication and is highly conserved among coronaviruses and positive-strand RNA viruses, and a clinically validated target across many viruses [30][31]. Several existing RdRp inhibitors developed against other RNA viruses, including remdesivir [32][33], molnupiravir [34], and favipiravir [35], were in late-stage development or clinical use at the start of the pandemic and showed promising in vitro activity against the SARS-CoV-2. Subsequently, numerous clinical trials were conducted to explore their potential against SARS-CoV-2 infection.
Remdesivir is a nucleotide analogue inhibitor of the RdRp that was originally developed for the Ebola virus (EBOV) [36]. It shows potent broad-spectrum antiviral activity against pathogenic animal and human coronaviruses in vitro [32][33][37]. However, remdesivir’s intravenous administration limits its widespread use in the community. Based on multiple clinical trials, remdesivir was the first antiviral drug approved for the treatment of COVID-19 in adults and paediatric patients (≥28 days old and weighing ≥ 3 kg) with a positive SARS-CoV-2 test, for both hospitalised patients, and non-hospitalised individuals with mild-to-moderate COVID-19 that are at high risk for progression to severe COVID-19 [38]. One randomised, double-blind, placebo-controlled clinical trial (ACTT-1 study) showed a shorter median time to recovery for remdesivir-treated patients compared to placebo [39]. Two additional trials sponsored by Gilead Sciences informed the approval. An open-label clinical trial of hospitalised adults with moderate COVID-19 showed an improvement in symptoms in patients receiving a five-day course of remdesivir, however, no improvement was demonstrated in those receiving a 10-day course of remdesivir [40]. A third randomised, open-label trial in hospitalised adults with severe COVID-19 showed no statistically significant differences in recovery or mortality rates [41]. However, it is important to note that with DAA, initiating treatment early whilst viral loads are high is crucial for compound efficacy, and variable treatment windows may explain conflicting clinical trial results [42]. An oral pro-drug of remdesivir, obeldesivir, shows excellent cross-reactivity against multiple coronaviruses including MERS, in vitro [43]. A recent Phase 3 RCT to evaluate the efficacy and safety for the treatment of COVID-19 in non-hospitalised patients with a high risk for disease progression was discontinued due to lower than expected incidence rates and related hospitalisations or all-cause death, whilst a Phase 3 RCT in hospitalised patients is still ongoing [44].
Molnupiravir is an orally available RdRp inhibitor with broad-spectrum antiviral activity [34] and was initially developed as an oral antiviral against Venezuelan equine encephalitis virus (VEEV) [30]. It acts as a nucleoside analogue in the RNA elongation process where it incorporates mutation errors that accumulate into an “error catastrophe” and subsequent virus replication failure [30][34][45]. Molnupiravir is authorised for emergency use for SARS-CoV-2 infection by the FDA in adults with confirmed mild-to-moderate SARS-CoV-2 infection, including those who are at high risk for progression to severe disease and those for whom alternative treatment options are not accessible or clinically appropriate. The FDA authorisation was based on the randomised, double-blind, placebo-controlled MOVe-OUT trial, demonstrating a reduction in hospitalization and mortality in the molnupiravir-treated group compared to the placebo [46]. However, the European Medicines Agency refused market authorisation for molnupiravir [47], arguing that it was not possible to conclude that molnupiravir reduces the risk of hospitalisation or death in adults at risk of severe disease.
Several other RdRp inhibitors that have been developed for other viral infections have been investigated for clinical efficacy in COVID-19 as part of intensive repurposing efforts. Favipiravir is a guanine analogue that selectively inhibits the RdRp and has been developed as a novel antiviral compound against influenza by Toyama Chemical Co. Favipiravir has been evaluated against SARS-CoV-2 infection in multiple clinical trials, with no significant effects on hospitalisation and mortality even when administered within 5 days of symptoms developing [48]. Sofosbuvir, a potent inhibitor developed against the RdRp of HCV, has and is in clinical use in combination for the treatment of chronic HCV infection in combination with ribavirin, ledipasvir, and daclatasvir for different genotypes. In multiple RCTs conducted with sofosbuvir in SARS-CoV-2 infection, no effect on endpoints was observed, apart from a small open-label study demonstrating an effect on median hospital stay [49]. Azvudine, a nucleoside analogue RdRp inhibitor originally developed against HIV infection [50][51] has been approved for the treatment of COVID-19 by the Chinese regulatory agency in 2022, citing a phase 3 clinical trial showing “improved clinical symptoms”, compared to a placebo [52][53]. However, detailed clinical trial data has not been published to date. Publicly available data is limited, with studies available showing an impact on SARS-CoV-2 viral replication and a shortened time to viral clearance in patients with mild COVID-19, compared to the standard antiviral treatment [54], as well as an impact on disease progression outcome in retrospective studies.

1.3. Monoclonal Antibodies

Neutralising monoclonal antibodies, mostly targeting the receptor binding domain (RBD) of the SARS-CoV-2 spike protein, has been tested in several clinical trials [3] and are reviewed elsewhere [55]. They offer a highly effective therapy against viral infections. However, their use is limited in clinical practice due to (i) the need for intravenous application which limits their use to hospital settings [56][57]; (ii) high costs that restrict access for large parts of the global population [58]; (iii) the reduced efficacy of some antibodies against VoCs [59]. Specifically, some antibodies with good efficacy against earlier SARS-CoV-2 variants were less potent against Omicron and its sub-variants [60]. Antibodies are effective in mild to moderate disease if given early. Further developments in the use and applicability of monoclonal antibodies for preventing COVID-19 and for early therapy are underway, including alternative applications such as nasal sprays [61].
Examples of monoclonal antibodies developed against SARS-CoV-2 include the antibody cocktail bamlanivimab with etesevimab from Eli Lilly and Company [62], Regeneron’s REGN-COV2 casirivimab and imdevimab cocktail [56][63], and the monotherapy sotrovimab by Vir Biotechnology in collaboration with GlaxoSmithKline [64].
The combination of bamlanivimab and etesevimab lowers COVID-19-related hospitalisation and mortality in mild to moderate COVID-19 cases and reduces SARS-CoV-2 viral load [65]. The use of antibody cocktails that target different neutralisation epitopes may be one way to maintain their therapeutic efficacy [66]. The non-competing monoclonal antibodies casirivimab and imdevimab bind to two different sites on the RBD of the SARS-CoV-2 spike protein resulting in the blockage of viral entry into host cells [56]. This antibody combination therapy reduced mortality at 28 days in seronegative patients but not in seropositive patients at baseline [56].
The neutralising antibody sotrovimab, developed by GlaxoSmithKline and Vir Biotechnology neutralises SARS-CoV-2 by targeting a highly conserved epitope in the RBD of the spike protein [57]. Treatment with sotrovimab reduced the risk of COVID-19 disease progression to hospitalisation or death in mild to moderate high-risk patients [57]. Sotrovimab monotherapy retained activity against SARS-CoV-2 VOCs including Omicron [64], however, has been associated with the rapid development of spike gene mutations in vitro [59] and in vivo [64].
Bebtelovimab, developed by Eli Lilly, is a neutralising IgG1 monoclonal antibody that binds to an epitope within the RBD of the SARS-CoV-2 S protein with broad neutralising activity to all SARS-CoV-2 VOCs, including the Omicron variant [67]. Bebtelovimab has been studied in the BLAZE-4 clinical trial and is associated with greater viral clearance. It is effective for the treatment of mild to moderate COVID-19 in adults and children above 12 years old who are at risk of disease progression and hospitalisation [68].
The long-lasting monoclonal antibody combination of tixagevimab and cilgavimab (brand name Evusheld from AstraZeneca) was derived from antibodies isolated from B cells of patients infected with SARS-CoV-2 [69][70].
The FDA removed its EUA for several monoclonal antibodies including bebtelovimab, tixagevimab-cilgavimab, sotrovimab, bamlanivimab-etesevimab, and casirivimab-imdevimab [71], due to their reduced efficacy against Omicron and its subvariants that are now circulating at high frequency [59][60].

2. Host-Targeting Antivirals (HTA)

Host-targeting antivirals (HTA) are drugs that modify host cell pathways required for viral replication [72]. They often modulate virus-host interactions by targeting human proteins used by viruses. As they may target host cell pathways used by several viral families, HTAs may be suitable as a first-line antiviral drug when a novel virus emerges, provided the novel virus relies on the same host pathway. Conceptually, HTAs could be given early in a pandemic, prior to emerging virweuses being characterised at the molecular level, and may carry a lower risk of drug resistance [5][6]

2.1. Inhibitors of Viral Cell Entry

SARS-CoV-2 enters host cells by attaching to the cell membrane and subsequently fusing in the endosome. Entry requires the spike glycoprotein, which is cleaved by the host protease furin into S1 and S2 subunits during viral release from an infected cell. The subunits remain non-covalently associated and are present on mature virions as trimeric spikes. The S1 subunits bind the obligate SARS-CoV-2 receptor ACE2 with their RBDs, while the S2 subunits anchor the spike protein to the membrane and contain the fusion peptide. When S1 binds ACE2, an S2′ site within S2 is exposed and triggers its cleavage by the transmembrane serine protease 2 (TMPRSS2) at the cell surface, or by cathepsin L in the endosome after receptor-mediated endocytosis [73]. Every step in this intricate entry process is important and presents a potential DAA or HTA target. 
Camostat mesylate, an inhibitor of TMPRSS2, has been reported as a potential entry inhibitor [74]. Although considered safe and well-tolerated, no clinical benefit was observed from RCTs results in terms of reduction of the disease progression or mortality.
Umifenovir (arbidol) was developed for the treatment of influenza virus as a hemagglutinin inhibitor preventing virus-mediated fusion with the cell membrane, blocking viral entry into target cells [75][76]. Unsurprisingly, as SARS-CoV-2 viral entry does not depend on hemagglutinin, RCT results evaluating its activity against COVID-19 showed no favourable effect and limited efficacy of arbidol in treating COVID-19 [77][78].
Some inhibitors work by interacting with dynamin, a GTPase responsible for clathrin-dependent endocytosis, which is essential for coronaviruses to enter the cell [79]. Chlorpromazine (CPZ), an antipsychotic medication, has been repurposed as a COVID-19 therapeutic based on this mechanism and reported for its antiviral activity against MERS-CoV and SARS-CoV-1 [80]. These findings led to a clinical trial that observed a lower incidence of symptomatic COVID-19 among patients after treatment with CPZ [81].
Disulfiram is a hepatic aldehyde dehydrogenase inhibitor that is used to treat chronic alcoholism [82][83]. It reduced the incidence and severity of COVID-19 in a retrospective observational study [82], although no impact on viral load was shown [83]. Further large-scale clinical trials are needed to assess the findings.
Fluvoxamine is a psychotropic medication that belongs to a group of functional inhibitors of acid sphingomyelinase (FIASMA) which were evaluated against COVID-19 [84]. The rationale for using FIASMA is linked to the role of lipid rafts in viral entry. Sphingomyelinase activity is triggered by SARS-CoV-2 binding which in turn leads to the formation of ceramide-enriched membrane domains that help viral entry by clustering ACE2 [85]. Treatment with fluvoxamine inhibits the sphingomyelinase activity formation of these domains in vitro [85][86]. Fluvoxamine and other FIASMA drugs were associated with reduced mortality in COVID-19 patients were tested in a large cohort study [87]. Although three RCT studies showed favourable clinical benefits of fluvoxamine, clinical evidence was deemed insufficient to issue a treatment recommendation in the National Institutes of Health (NIH) treatment guidelines [88].

2.2. Inhibitor of Viral Glycoprotein Folding

Enveloped viruses require the glycoprotein-folding machinery in the host endoplasmic reticulum (ER) to correctly fold their glycosylated proteins [5]. Pivotal players in this ER quality control (ERQC) pathway are the ER alpha glucosidases, which can be targeted for example by iminosugars [89][90][91]. Partially inhibiting the ERQC prevents the proper folding and incorporation of viral glycoproteins into budding viruses, as shown previously for other enveloped viruses such as HIV [92], human papillomavirus [93], dengue [91][94][95], influenza [91][96], hepatitis B virus [97], HCV [98], Zika virus [99], Marburg virus [100] and EBOV [101], and could lead to a potential broad-spectrum antiviral drug also against coronaviruses. The extensively glycosylated SARS-CoV-2 spike protein is essential for viral entry, and inhibition of its proper glycosylation leads to antiviral effects. The monocyclic UV-4 (N-(9-methoxynonyl)-1-deoxynojirimycin) or MON-DNJ prevented SARS-CoV-2-induced Vero cell death and reduced viral replication in vitro after 24 h of treatment [5][102]. The results are encouraging and need to be further tested in vivo and in clinical trials.

2.3. Host-Targeting Antivirals with Unknown Mechanism

Some repurposed drugs work by targeting the host, although the exact antiviral mechanisms are unknown. Ivermectin, an FDA-approved anti-parasitic drug was reported to have an in vitro antiviral activity to SARS-CoV-2 [103]. However, several adequately powered RCTs in Brazil [104], the US [105][106], and Malaysia [107] failed to report a clinical benefit from the use of ivermectin in COVID-19 outpatients and it is not approved or authorised by the FDA for the treatment of COVID-19.
Similar attention was given to the anti-malaria drug chloroquine and its derivatives [108]. Chloroquine was reported to have efficacy and acceptable safety against COVID-19-associated pneumonia in multi-centre clinical trials conducted in China [109] and its use attracted disproportionate attention during the coronavirus pandemic, spurred by preliminary studies and endorsement from political leaders [110]. The chloroquine derivative hydroxychloroquine was tested in RCTs with limited to no clinical benefit for COVID-19 [111]. Based on sufficiently powered randomised trials [112][113] however, NIH treatment guidelines recommend against the use of both chloroquine and hydroxychloroquine for the treatment of COVID-19 [88].
The anti-protozoal drug nitazoxanide also showed in vitro activity against SARS-CoV-2 [114], and has been tested in multiple RCTs. A possible mechanism of action may be linked to inhibition of SARS-CoV-2 spike-induced syncytia and its binding to TMEM16 [115]. Nitazoxanide did not show efficacy in a number of RCTs when used at approved doses, and NIH treatment guidelines recommend against the use of nitazoxanide for the treatment of COVID-19 [88]. However, nitazoxanide did not show serious adverse events when evaluated in a Phase 1 study at higher doses of 1500 mg twice daily, at which it may provide antiviral efficacy according to the pharmacokinetic modelling [116].

3. Immunomodulatory Drugs

Immunomodulatory drugs are commonly used in autoimmune disease and include both monoclonal antibodies and small molecule inhibitors. They can either target cytokines directly, such as monoclonal antibodies against interleukins (ILs) or tumour necrosis factor (TNF)-alpha, inhibit proteins involved in inflammatory signalling pathways such as the janus kinase (JAK) inhibitors, or interfere with the hormonal regulation of inflammation such as corticosteroids [117][118]. Multiple monoclonal antibodies and immunomodulatory compounds were evaluated in COVID-19 infection, driven by the aim to impact on COVID-19 associated systemic inflammation that can be associated with heightened cytokine release, as indicated by elevated blood levels of IL-6, C-reactive protein (CRP), D-dimer and ferritin [119][120][121][122]. Crucial for the assessment of the effects of immunomodulatory drugs were large-scale platform trials initiated early in the pandemic.

3.1. Corticosteroids

Corticosteroids bind to the glucocorticoid receptor and inhibit the synthesis of multiple inflammatory proteins through the suppression of genes that encode them, as well as promoting anti-inflammatory signals [122][123]. Coordinated from Oxford, the open-label RECOVERY trial recruited over 43,000 hospitalised patients with COVID-19 participants worldwide and randomly assigned patients to treatment groups that included dexamethasone, hydroxychloroquine, lopinavir-ritonavir, or azithromycin, and compared them to usual care, with the primary endpoint mortality at 28 days [124]. Dexamethasone treatment significantly decreased mortality in patients who were receiving either invasive mechanical ventilation or oxygen alone at randomization [125]. Another inhaled corticosteroid, budesonide, was assessed in the PRINCIPLE study in non-hospitalised patients with COVID-19, and improved recovery time with the potential to reduce hospital admissions or deaths [126].

3.2. Host-Targeting Monoclonal Antibodies

Multiple immunomodulatory monoclonal antibodies, directed against cytokines such as against tumour necrosis factor alpha (TNF-α), Interleukin (IL)-1, and IL-6, were assessed in COVID-19 patients [122]. Tocilizumab is a recombinant humanised monoclonal antibody that binds to interleukin-6 receptors, thereby blocking the activity of the pro-inflammatory cytokine. IL-6 is produced by a variety of cell types including lymphocytes, monocytes, and fibroblasts, and has been shown to be induced by SARS-CoV-2 infection in bronchial epithelial cells. The IL-6 inhibitor tocilizumab is in clinical use for several inflammatory diseases such as rheumatoid arthritis, giant cell arteritis, polyarticular juvenile idiopathic arthritis, and systemic juvenile idiopathic arthritis. As of 24 June 2021, the FDA has authorised the use of tocilizumab under EUA for the treatment of COVID-19 in hospitalised adults who are receiving systemic corticosteroids and require supplemental oxygen, mechanical ventilation, or extracorporeal membrane oxygenation (ECMO) [127]. The authorisation was based on the results of 4 clinical trials: RECOVERY, EMPACTA, COVACTA, and REMDACTA, with the decision to grant EUA mainly based on the positive results of RECOVERY and EMPACTA that demonstrated an impact on mortality and a composite readout of mechanical ventilation and mortality [128][129].
IL-1 inhibitors include the IL-1 receptor antagonist Kineret (brand name anakinra), the IL-1 “Trap” rilonacept, and the neutralising monoclonal antibody to IL-1β canakinumab [122][130]. Anakinra was considered a safe and efficient treatment for severe forms of COVID-19 with a significant survival benefit in critically ill patients and features of macrophage activation-like syndrome [131][132], with most RCTs supporting the clinical benefit of this drug for COVID-19.
TNF-inhibitors have been used in severe cases of autoimmune inflammatory diseases such as rheumatoid arthritis, inflammatory bowel disease, or ankylosing spondylitis [133]. Several formulations of TNF inhibitors are currently available, including adalimumab, etanercept, and infliximab [122]. Elevated serum levels of TNF-α and soluble TNF-Receptor 1 have been detected in COVID-19 patients with severe infection, providing a rationale for the use of TNF-α inhibitors in SARS-CoV-2 infection [134]. However, concerns regarding the potential suppression of antiviral immune responses have been raised by an observational study showing a negative impact on SARS-CoV-2 antibody levels following natural infection in patients with inflammatory bowel disease treated with infliximab [135][136]. In the CATALYST open-label phase 2 trial, infliximab showed no impact on CRP levels as a measure for inflammation in SARS-CoV-2 infection [137]. Recently reported results of the ACTIV-1 trial, a placebo-controlled, masked, RCT among patients hospitalised for COVID-19, reported a significant benefit of infliximab on mortality [138], but no impact on the primary study endpoint length of pneumonia [139].
The latest addition to monoclonal antibodies under EUA is vilobelimab, a monoclonal antibody that specifically binds to the soluble human complement C5a, a product of complement activation. C5a activates the innate immune response, including the local release of histamines, contributing to inflammation and local tissue damage. In vivo studies demonstrated that an anti-C5a monoclonal antibody inhibited acute lung injury in a human C5a receptor knock-in mouse model [140]. The phase 3 PANAMO RCT study recorded an improvement in invasive mechanically ventilated patients’ survival that led to a decrease in mortality with the use of vilobelimab [141].

3.3. Janus Kinase (JAK) Inhibitors

The primary mechanism of action of Janus kinases (JAK) is the phosphorylating signal transducer and activator of transcription (STAT), a key player in signalling pathways involved signalling, growth, survival, inflammation, and immune activation. Inhibiting JAK prevents the phosphorylation of key signalling proteins involved in inflammation pathways, thereby blocking cytokine signalling [122][142]. Tofacitinib and baricitinib are orally available small-molecule JAK inhibitors approved for the treatment of rheumatoid arthritis [142] and have been evaluated in multiple RCTs in patients with SARS-CoV-2 infection, leading to a treatment recommendation in hospitalised adults that require respiratory support [88]. In brief, baricitinib had an impact on mortality [143][144], as well as reducing intensive care unit admissions, lowering the requirement for invasive mechanical ventilation, and improving patients’ oxygenation index [145][146], effects that were maintained in a meta-analysis [147]. Furthermore, baricitinib in combination with remdesivir showed superior results compared to remdesivir alone in reducing patients’ recovery time and improving their clinical status [148].
In summary, for immunomodulatory drugs, the NIH issued guidelines based on the existing evidence, recommending dexamethasone, tocilizumab, and baricitinib for hospitalised patients with COVID-19, whilst the evidence for anakinra, inhaled corticosteroids, and vilobelimab was deemed inconclusive, despite all compounds receiving EUA from the FDA.

References

  1. Groaz, E.; De Clercq, E.; Herdewijn, P. Anno 2021: Which Antivirals for the Coming Decade? Annu. Rep. Med. Chem. 2021, 57, 49–107.
  2. Teoh, S.L.; Lim, Y.H.; Lai, N.M.; Lee, S.W.H. Directly Acting Antivirals for COVID-19: Where Do We Stand? Front. Microbiol. 2020, 11, 1857.
  3. Mei, M.; Tan, X. Current Strategies of Antiviral Drug Discovery for COVID-19. Front. Mol. Biosci. 2021, 8, 671263.
  4. Jonsson, C.B.; Golden, J.E.; Meibohm, B. Time to “Mind the Gap” in Novel Small Molecule Drug Discovery for Direct-Acting Antivirals for SARS-CoV-2. Curr. Opin. Virol. 2021, 50, 1–7.
  5. Dwek, R.A.; Bell, J.I.; Feldmann, M.; Zitzmann, N. Host-Targeting Oral Antiviral Drugs to Prevent Pandemics. Lancet Lond. Engl. 2022, 399, 1381–1382.
  6. Robinson, P.C.; Liew, D.F.L.; Tanner, H.L.; Grainger, J.R.; Dwek, R.A.; Reisler, R.B.; Steinman, L.; Feldmann, M.; Ho, L.-P.; Hussell, T.; et al. COVID-19 Therapeutics: Challenges and Directions for the Future. Proc. Natl. Acad. Sci. USA 2022, 119, e2119893119.
  7. Von Delft, A.; Hall, M.D.; Kwong, A.D.; Purcell, L.A.; Saikatendu, K.S.; Schmitz, U.; Tallarico, J.A.; Lee, A.A. Accelerating Antiviral Drug Discovery: Lessons from COVID-19. Nat. Rev. Drug Discov. 2023, 22, 585–603.
  8. Boby, M.L.; Fearon, D.; Ferla, M.; Filep, M.; Koekemoer, L.; Robinson, M.C.; The COVID Moonshot Consortium; Chodera, J.D.; Lee, A.A.; London, N.; et al. Open Science Discovery of Potent Noncovalent SARS-CoV-2 Main Protease Inhibitors. Science 2023, 382, eabo7201.
  9. Owen, D.R.; Allerton, C.M.N.; Anderson, A.S.; Aschenbrenner, L.; Avery, M.; Berritt, S.; Boras, B.; Cardin, R.D.; Carlo, A.; Coffman, K.J.; et al. An Oral SARS-CoV-2 Mpro Inhibitor Clinical Candidate for the Treatment of COVID-19. Science 2021, 374, 1586–1593.
  10. Hammond, J.; Leister-Tebbe, H.; Gardner, A.; Abreu, P.; Bao, W.; Wisemandle, W.; Baniecki, M.; Hendrick, V.M.; Damle, B.; Simón-Campos, A.; et al. Oral Nirmatrelvir for High-Risk, Nonhospitalized Adults with COVID-19. N. Engl. J. Med. 2022, 386, 1397–1408.
  11. Schilling, W.H.K.; Jittamala, P.; Watson, J.A.; Boyd, S.; Luvira, V.; Siripoon, T.; Ngamprasertchai, T.; Batty, E.M.; Cruz, C.; Callery, J.J.; et al. Antiviral Efficacy of Molnupiravir versus Ritonavir-Boosted Nirmatrelvir in Patients with Early Symptomatic COVID-19 (PLATCOV): An Open-Label, Phase 2, Randomised, Controlled, Adaptive Trial. Lancet Infect. Dis. 2023, 24, 36–45.
  12. Heskin, J.; Pallett, S.J.C.; Mughal, N.; Davies, G.W.; Moore, L.S.P.; Rayment, M.; Jones, R. Caution Required with Use of Ritonavir-Boosted PF-07321332 in COVID-19 Management. Lancet Lond. Engl. 2022, 399, 21–22.
  13. DTB Drug Review. Two New Oral Antivirals for COVID-19: ▼molnupiravir and ▼nirmatrelvir plus Ritonavir. Drug Ther. Bull. 2022, 60, 73.
  14. Lim, S.; Tignanelli, C.J.; Hoertel, N.; Boulware, D.R.; Usher, M.G. Prevalence of Medical Contraindications to Nirmatrelvir/Ritonavir in a Cohort of Hospitalized and Nonhospitalized Patients with COVID-19. Open Forum Infect. Dis. 2022, 9, ofac389.
  15. Tuttle, J. Discovery of PF-07817883: A next Generation Oral Protease Inhibitor for the Treatment of COVID-19. Available online: https://acs.digitellinc.com/sessions/584379/view (accessed on 7 September 2023).
  16. Tyndall, J.D.A. S-217622, a 3CL Protease Inhibitor and Clinical Candidate for SARS-CoV-2. J. Med. Chem. 2022, 65, 6496–6498.
  17. Unoh, Y.; Uehara, S.; Nakahara, K.; Nobori, H.; Yamatsu, Y.; Yamamoto, S.; Maruyama, Y.; Taoda, Y.; Kasamatsu, K.; Suto, T.; et al. Discovery of S-217622, a Noncovalent Oral SARS-CoV-2 3CL Protease Inhibitor Clinical Candidate for Treating COVID-19. J. Med. Chem. 2022, 65, 6499–6512.
  18. Sasaki, M.; Tabata, K.; Kishimoto, M.; Itakura, Y.; Kobayashi, H.; Ariizumi, T.; Uemura, K.; Toba, S.; Kusakabe, S.; Maruyama, Y.; et al. S-217622, a SARS-CoV-2 Main Protease Inhibitor, Decreases Viral Load and Ameliorates COVID-19 Severity in Hamsters. Sci. Transl. Med. 2023, 15, eabq4064.
  19. Mukae, H.; Yotsuyanagi, H.; Ohmagari, N.; Doi, Y.; Imamura, T.; Sonoyama, T.; Fukuhara, T.; Ichihashi, G.; Sanaki, T.; Baba, K.; et al. A Randomized Phase 2/3 Study of Ensitrelvir, a Novel Oral SARS-CoV-2 3C-Like Protease Inhibitor, in Japanese Patients with Mild-to-Moderate COVID-19 or Asymptomatic SARS-CoV-2 Infection: Results of the Phase 2a Part. Antimicrob. Agents Chemother. 2022, 66, e0069722.
  20. Antonopoulou, I.; Sapountzaki, E.; Rova, U.; Christakopoulos, P. Inhibition of the Main Protease of SARS-CoV-2 (Mpro) by Repurposing/Designing Drug-like Substances and Utilizing Nature’s Toolbox of Bioactive Compounds. Comput. Struct. Biotechnol. J. 2022, 20, 1306–1344.
  21. Zaidman, D.; Gehrtz, P.; Filep, M.; Fearon, D.; Gabizon, R.; Douangamath, A.; Prilusky, J.; Duberstein, S.; Cohen, G.; Owen, C.D.; et al. An Automatic Pipeline for the Design of Irreversible Derivatives Identifies a Potent SARS-CoV-2 Mpro Inhibitor. Cell Chem. Biol. 2021, 28, 1795–1806.
  22. Cattaneo, D.; Cattaneo, D.; Gervasoni, C.; Corbellino, M.; Galli, M.; Riva, A.; Gervasoni, C.; Clementi, E.; Clementi, E. Does Lopinavir Really Inhibit SARS-CoV-2? Pharmacol. Res. 2020, 158, 104898.
  23. Horby, P.W.; Mafham, M.; Bell, J.L.; Linsell, L.; Staplin, N.; Emberson, J.; Palfreeman, A.; Raw, J.; Elmahi, E.; Prudon, B.; et al. Lopinavir–Ritonavir in Patients Admitted to Hospital with COVID-19 (RECOVERY): A Randomised, Controlled, Open-Label, Platform Trial. Lancet 2020, 396, 1345–1352.
  24. Patel, T.K.; Patel, P.B.; Barvaliya, M.; Saurabh, M.K.; Bhalla, H.L.; Khosla, P.P. Efficacy and Safety of Lopinavir-Ritonavir in COVID-19: A Systematic Review of Randomized Controlled Trials. J. Infect. Public Health 2021, 14, 740–748.
  25. Tashima, K.; Crofoot, G.; Tomaka, F.L.; Kakuda, T.N.; Brochot, A.; Van de Casteele, T.; Opsomer, M.; Garner, W.; Margot, N.; Custodio, J.M.; et al. Cobicistat-Boosted Darunavir in HIV-1-Infected Adults: Week 48 Results of a Phase IIIb, Open-Label Single-Arm Trial. AIDS Res. Ther. 2014, 11, 39.
  26. De Meyer, S.; Bojkova, D.; Cinatl, J.; Van Damme, E.; Buyck, C.; Van Loock, M.; Woodfall, B.; Ciesek, S. Lack of Antiviral Activity of Darunavir against SARS-CoV-2. Int. J. Infect. Dis. 2020, 97, 7–10.
  27. Chen, H.; Zhang, Z.; Wang, L.; Huang, Z.; Gong, F.; Li, X.; Chen, Y.; Wu, J.J. First Clinical Study Using HCV Protease Inhibitor Danoprevir to Treat COVID-19 Patients. Medicine (Baltimore) 2020, 99, e23357.
  28. Gammeltoft, K.A.; Zhou, Y.; Duarte Hernandez, C.R.; Galli, A.; Offersgaard, A.; Costa, R.; Pham, L.V.; Fahnøe, U.; Feng, S.; Scheel, T.K.H.; et al. Hepatitis C Virus Protease Inhibitors Show Differential Efficacy and Interactions with Remdesivir for Treatment of SARS-CoV-2 In Vitro. Antimicrob. Agents Chemother. 2021, 65, e0268020.
  29. Zhang, Z.; Wang, S.; Tu, X.; Peng, X.; Huang, Y.; Wang, L.; Ju, W.; Rao, J.; Li, X.; Zhu, D.; et al. A Comparative Study on the Time to Achieve Negative Nucleic Acid Testing and Hospital Stays between Danoprevir and Lopinavir/Ritonavir in the Treatment of Patients with COVID-19. J. Med. Virol. 2020, 92, 2631–2636.
  30. Cully, M. A Tale of Two Antiviral Targets—And the COVID-19 Drugs That Bind Them. Nat. Rev. Drug Discov. 2021, 21, 3–5.
  31. Vicenti, I.; Zazzi, M.; Saladini, F. SARS-CoV-2 RNA-Dependent RNA Polymerase as a Therapeutic Target for COVID-19. Expert Opin. Ther. Pat. 2021, 31, 325–337.
  32. Pruijssers, A.J.; George, A.S.; Schäfer, A.; Leist, S.R.; Gralinksi, L.E.; Dinnon, K.H.; Yount, B.L.; Agostini, M.L.; Stevens, L.J.; Chappell, J.D.; et al. Remdesivir Inhibits SARS-CoV-2 in Human Lung Cells and Chimeric SARS-CoV Expressing the SARS-CoV-2 RNA Polymerase in Mice. Cell Rep. 2020, 32, 107940.
  33. Wang, M.; Cao, R.; Zhang, L.; Yang, X.; Liu, J.; Xu, M.; Shi, Z.; Hu, Z.; Zhong, W.; Xiao, G. Remdesivir and Chloroquine Effectively Inhibit the Recently Emerged Novel Coronavirus (2019-nCoV) in Vitro. Cell Res. 2020, 30, 269–271.
  34. Sheahan, T.P.; Sims, A.C.; Zhou, S.; Graham, R.L.; Pruijssers, A.J.; Agostini, M.L.; Leist, S.R.; Schäfer, A.; Dinnon, K.H.; Stevens, L.J.; et al. An Orally Bioavailable Broad-Spectrum Antiviral Inhibits SARS-CoV-2 in Human Airway Epithelial Cell Cultures and Multiple Coronaviruses in Mice. Sci. Transl. Med. 2020, 12, eabb5883.
  35. Shannon, A.; Selisko, B.; Le, N.-T.-T.; Huchting, J.; Touret, F.; Piorkowski, G.; Fattorini, V.; Ferron, F.; Decroly, E.; Meier, C.; et al. Rapid Incorporation of Favipiravir by the Fast and Permissive Viral RNA Polymerase Complex Results in SARS-CoV-2 Lethal Mutagenesis. Nat. Commun. 2020, 11, 4682.
  36. Warren, T.K.; Jordan, R.; Lo, M.K.; Ray, A.S.; Mackman, R.L.; Soloveva, V.; Siegel, D.; Perron, M.; Bannister, R.; Hui, H.C.; et al. Therapeutic Efficacy of the Small Molecule GS-5734 against Ebola Virus in Rhesus Monkeys. Nature 2016, 531, 381–385.
  37. Sheahan, T.P.; Sims, A.C.; Graham, R.L.; Menachery, V.D.; Gralinski, L.E.; Case, J.B.; Leist, S.R.; Pyrc, K.; Feng, J.Y.; Trantcheva, I.; et al. Broad-Spectrum Antiviral GS-5734 Inhibits Both Epidemic and Zoonotic Coronaviruses. Sci. Transl. Med. 2017, 9, eaal3653.
  38. Food and Drug Administration FDA Approves First Treatment for COVID-19. Available online: https://www.fda.gov/news-events/press-announcements/fda-approves-first-treatment-covid-19 (accessed on 4 July 2023).
  39. Beigel, J.H.; Tomashek, K.M.; Dodd, L.E.; Mehta, A.K.; Zingman, B.S.; Kalil, A.C.; Hohmann, E.; Chu, H.Y.; Luetkemeyer, A.; Kline, S.; et al. Remdesivir for the Treatment of COVID-19—Final Report. N. Engl. J. Med. 2020, 383, 1813–1826.
  40. Spinner, C.D.; Gottlieb, R.L.; Criner, G.J.; Arribas López, J.R.; Cattelan, A.M.; Soriano Viladomiu, A.; Ogbuagu, O.; Malhotra, P.; Mullane, K.M.; Castagna, A.; et al. Effect of Remdesivir vs Standard Care on Clinical Status at 11 Days in Patients with Moderate COVID-19: A Randomized Clinical Trial. JAMA 2020, 324, 1048–1057.
  41. Goldman, J.D.; Lye, D.C.B.; Hui, D.S.; Marks, K.M.; Bruno, R.; Montejano, R.; Spinner, C.D.; Galli, M.; Ahn, M.-Y.; Nahass, R.G.; et al. Remdesivir for 5 or 10 Days in Patients with Severe COVID-19. N. Engl. J. Med. 2020, 383, 1827–1837.
  42. Gottlieb, R.L.; Vaca, C.E.; Paredes, R.; Mera, J.; Webb, B.J.; Perez, G.; Oguchi, G.; Ryan, P.; Nielsen, B.U.; Brown, M.; et al. Early Remdesivir to Prevent Progression to Severe COVID-19 in Outpatients. N. Engl. J. Med. 2022, 386, 305–315.
  43. Martinez, D.R.; Moreira, F.R.; Zweigart, M.R.; Gully, K.L.; la Cruz, G.D.; Brown, A.J.; Adams, L.E.; Catanzaro, N.; Yount, B.; Baric, T.J.; et al. Efficacy of the Oral Nucleoside Prodrug GS-5245 (Obeldesivir) against SARS-CoV-2 and Coronaviruses with Pandemic Potential. bioRxiv 2023.
  44. Gilead Sciences Gilead Sciences Statement on Phase 3 Obeldesivir Clinical Trials in COVID-19: BIRCH Study to Stop Enrollment While OAKTREE Study Nears Full Enrollment. Available online: https://www.gilead.com/news-and-press/company-statements/gilead-sciences-statement-on-phase-3-obeldesivir-clinical-trials-in-covid-19-birch-study-to-stop-enrollment-while-oaktree-study-nears-full-enrollment (accessed on 14 November 2023).
  45. Jayk Bernal, A.; Gomes da Silva, M.M.; Musungaie, D.B.; Kovalchuk, E.; Gonzalez, A.; Delos Reyes, V.; Martín-Quirós, A.; Caraco, Y.; Williams-Diaz, A.; Brown, M.L.; et al. Molnupiravir for Oral Treatment of COVID-19 in Nonhospitalized Patients. N. Engl. J. Med. 2022, 386, 509–520.
  46. Johnson, M.G.; Strizki, J.M.; Brown, M.L.; Wan, H.; Shamsuddin, H.H.; Ramgopal, M.; Florescu, D.F.; Delobel, P.; Khaertynova, I.; Flores, J.F.; et al. Molnupiravir for the Treatment of COVID-19 in Immunocompromised Participants: Efficacy, Safety, and Virology Results from the Phase 3 Randomized, Placebo-Controlled MOVe-OUT Trial. Infection 2023, 51, 1273–1284.
  47. European Medicines Agency Refusal of the Marketing Authorisation for Lagevrio (Molnupiravir). Available online: https://www.ema.europa.eu/en/documents/smop-initial/questions-answers-refusal-marketing-authorisation-lagevrio-molnupiravir_en.pdf (accessed on 4 July 2023).
  48. McMahon, J.H.; Lau, J.S.Y.; Coldham, A.; Roney, J.; Hagenauer, M.; Price, S.; Bryant, M.; Garlick, J.; Paterson, A.; Lee, S.J.; et al. Favipiravir in Early Symptomatic COVID-19, a Randomised Placebo-Controlled Trial. EClinicalMedicine 2022, 54, 101703.
  49. Sadeghi, A.; Ali Asgari, A.; Norouzi, A.; Kheiri, Z.; Anushirvani, A.; Montazeri, M.; Hosamirudsai, H.; Afhami, S.; Akbarpour, E.; Aliannejad, R.; et al. Sofosbuvir and Daclatasvir Compared with Standard of Care in the Treatment of Patients Admitted to Hospital with Moderate or Severe Coronavirus Infection (COVID-19): A Randomized Controlled Trial. J. Antimicrob. Chemother. 2020, 75, 3379–3385.
  50. Wang, R.-R.; Yang, Q.-H.; Luo, R.-H.; Peng, Y.-M.; Dai, S.-X.; Zhang, X.-J.; Chen, H.; Cui, X.-Q.; Liu, Y.-J.; Huang, J.-F.; et al. Azvudine, a Novel Nucleoside Reverse Transcriptase Inhibitor Showed Good Drug Combination Features and Better Inhibition on Drug-Resistant Strains than Lamivudine in Vitro. PLoS ONE 2014, 9, e105617.
  51. Zhang, J.-L.; Li, Y.-H.; Wang, L.-L.; Liu, H.-Q.; Lu, S.-Y.; Liu, Y.; Li, K.; Liu, B.; Li, S.-Y.; Shao, F.-M.; et al. Azvudine Is a Thymus-Homing Anti-SARS-CoV-2 Drug Effective in Treating COVID-19 Patients. Signal Transduct. Target. Ther. 2021, 6, 414.
  52. Ye, Y. China Approves First Homegrown COVID Antiviral. Nature 2022.
  53. Yu, B.; Chang, J. The First Chinese Oral Anti-COVID-19 Drug Azvudine Launched. Innov. Camb. Mass 2022, 3, 100321.
  54. Ren, Z.; Luo, H.; Yu, Z.; Song, J.; Liang, L.; Wang, L.; Wang, H.; Cui, G.; Liu, Y.; Wang, J.; et al. A Randomized, Open-Label, Controlled Clinical Trial of Azvudine Tablets in the Treatment of Mild and Common COVID-19, A Pilot Study. Adv. Sci. Weinh. Baden-Wurtt. Ger. 2020, 7, 2001435.
  55. Hwang, Y.-C.; Lu, R.-M.; Su, S.-C.; Chiang, P.-Y.; Ko, S.-H.; Ke, F.-Y.; Liang, K.-H.; Hsieh, T.-Y.; Wu, H.-C. Monoclonal Antibodies for COVID-19 Therapy and SARS-CoV-2 Detection. J. Biomed. Sci. 2022, 29, 1.
  56. Abani, O.; Abbas, A.; Abbas, F.; Abbas, M.; Abbasi, S.; Abbass, H.; Abbott, A.; Abdallah, N.; Abdelaziz, A.; Abdelfattah, M.; et al. Casirivimab and Imdevimab in Patients Admitted to Hospital with COVID-19 (RECOVERY): A Randomised, Controlled, Open-Label, Platform Trial. Lancet 2022, 399, 665–676.
  57. Gupta, A.; Gonzalez-Rojas, Y.; Juarez, E.; Crespo Casal, M.; Moya, J.; Falci, D.R.; Sarkis, E.; Solis, J.; Zheng, H.; Scott, N.; et al. Early Treatment for COVID-19 with SARS-CoV-2 Neutralizing Antibody Sotrovimab. N. Engl. J. Med. 2021, 385, 1941–1950.
  58. Flaxman, A.D.; Issema, R.; Barnabas, R.V.; Ross, J.M. Estimated Health Outcomes and Costs of COVID-19 Prophylaxis with Monoclonal Antibodies Among Unvaccinated Household Contacts in the US. JAMA Netw. Open 2022, 5, e228632.
  59. Rockett, R.; Basile, K.; Maddocks, S.; Fong, W.; Agius, J.E.; Johnson-Mackinnon, J.; Arnott, A.; Chandra, S.; Gall, M.; Draper, J.; et al. Resistance Mutations in SARS-CoV-2 Delta Variant after Sotrovimab Use. N. Engl. J. Med. 2022, 386, 1477–1479.
  60. Lim, S.P. Targeting SARS-CoV-2 and Host Cell Receptor Interactions. Antiviral Res. 2022, 210, 105514.
  61. Ku, Z.; Xie, X.; Hinton, P.R.; Liu, X.; Ye, X.; Muruato, A.E.; Ng, D.C.; Biswas, S.; Zou, J.; Liu, Y.; et al. Nasal Delivery of an IgM Offers Broad Protection from SARS-CoV-2 Variants. Nature 2021, 595, 718–723.
  62. Jones, B.E.; Brown-Augsburger, P.L.; Corbett, K.S.; Westendorf, K.; Davies, J.; Cujec, T.P.; Wiethoff, C.M.; Blackbourne, J.L.; Heinz, B.A.; Foster, D.; et al. The Neutralizing Antibody, LY-CoV555, Protects against SARS-CoV-2 Infection in Nonhuman Primates. Sci. Transl. Med. 2021, 13, eabf1906.
  63. Weinreich, D.M.; Sivapalasingam, S.; Norton, T.; Ali, S.; Gao, H.; Bhore, R.; Xiao, J.; Hooper, A.T.; Hamilton, J.D.; Musser, B.J.; et al. REGEN-COV Antibody Combination and Outcomes in Outpatients with COVID-19. N. Engl. J. Med. 2021, 385, e81.
  64. Destras, G.; Bal, A.; Simon, B.; Lina, B.; Josset, L. Sotrovimab Drives SARS-CoV-2 Omicron Variant Evolution in Immunocompromised Patients. Lancet Microbe 2022, 3, e559.
  65. Dougan, M.; Nirula, A.; Azizad, M.; Mocherla, B.; Gottlieb, R.L.; Chen, P.; Hebert, C.; Perry, R.; Boscia, J.; Heller, B.; et al. Bamlanivimab plus Etesevimab in Mild or Moderate COVID-19. N. Engl. J. Med. 2021, 385, 1382–1392.
  66. Weisblum, Y.; Schmidt, F.; Zhang, F.; DaSilva, J.; Poston, D.; Lorenzi, J.C.; Muecksch, F.; Rutkowska, M.; Hoffmann, H.-H.; Michailidis, E.; et al. Escape from Neutralizing Antibodies by SARS-CoV-2 Spike Protein Variants. eLife 2020, 9, e61312.
  67. Westendorf, K.; Žentelis, S.; Wang, L.; Foster, D.; Vaillancourt, P.; Wiggin, M.; Lovett, E.; van der Lee, R.; Hendle, J.; Pustilnik, A.; et al. LY-CoV1404 (Bebtelovimab) Potently Neutralizes SARS-CoV-2 Variants. Cell Rep. 2022, 39, 110812.
  68. Dougan, M.; Azizad, M.; Chen, P.; Feldman, B.; Frieman, M.; Igbinadolor, A.; Kumar, P.; Morris, J.; Potts, J.; Baracco, L.; et al. Bebtelovimab, Alone or Together with Bamlanivimab and Etesevimab, as a Broadly Neutralizing Monoclonal Antibody Treatment for Mild to Moderate, Ambulatory COVID-19. MedRxiv 2022.
  69. Abramowicz, M. Tixagevimab and Cilgavimab (Evusheld) for Pre-Exposure Prophylaxis of COVID-19. JAMA 2022, 327, 384–385.
  70. Levin, M.J.; Ustianowski, A.; De Wit, S.; Launay, O.; Avila, M.; Templeton, A.; Yuan, Y.; Seegobin, S.; Ellery, A.; Levinson, D.J.; et al. Intramuscular AZD7442 (Tixagevimab–Cilgavimab) for Prevention of COVID-19. N. Engl. J. Med. 2022, 386, 2188–2200.
  71. Food and Drug Administration Emergency Use Authorizations for Drugs and Non-Vaccine Biological Products. Available online: https://www.fda.gov/drugs/emergency-preparedness-drugs/emergency-use-authorizations-drugs-and-non-vaccine-biological-products (accessed on 7 May 2023).
  72. Singh, M.; de Wit, E. Antiviral Agents for the Treatment of COVID-19: Progress and Challenges. Cell Rep. Med. 2022, 3, 100549.
  73. Jackson, C.B.; Farzan, M.; Chen, B.; Choe, H. Mechanisms of SARS-CoV-2 Entry into Cells. Nat. Rev. Mol. Cell Biol. 2022, 23, 3–20.
  74. Hoffmann, M.; Kleine-Weber, H.; Schroeder, S.; Krüger, N.; Herrler, T.; Erichsen, S.; Schiergens, T.S.; Herrler, G.; Wu, N.-H.; Nitsche, A.; et al. SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and Is Blocked by a Clinically Proven Protease Inhibitor. Cell 2020, 181, 271–280.e8.
  75. Boriskin, Y.S.; Leneva, I.A.; Pécheur, E.-I.; Polyak, S.J. Arbidol: A Broad-Spectrum Antiviral Compound That Blocks Viral Fusion. Curr. Med. Chem. 2008, 15, 997–1005.
  76. Zeng, L.-Y.; Yang, J.; Liu, S. Investigational Hemagglutinin-Targeted Influenza Virus Inhibitors. Expert Opin. Investig. Drugs 2017, 26, 63–73.
  77. Li, Y.; Xie, Z.; Lin, W.; Cai, W.; Wen, C.; Guan, Y.; Mo, X.; Wang, J.; Wang, Y.; Peng, P.; et al. Efficacy and Safety of Lopinavir/Ritonavir or Arbidol in Adult Patients with Mild/Moderate COVID-19: An Exploratory Randomized Controlled Trial. Med. N. Y. N 2020, 1, 105–113.e4.
  78. Deng, L.; Li, C.; Zeng, Q.; Liu, X.; Li, X.; Zhang, H.; Hong, Z.; Xia, J. Arbidol Combined with LPV/r versus LPV/r Alone against Corona Virus Disease 2019: A Retrospective Cohort Study. J. Infect. 2020, 81, e1–e5.
  79. Burkard, C.; Verheije, M.H.; Wicht, O.; van Kasteren, S.I.; van Kuppeveld, F.J.; Haagmans, B.L.; Pelkmans, L.; Rottier, P.J.M.; Bosch, B.J.; de Haan, C.A.M. Coronavirus Cell Entry Occurs through the Endo-/Lysosomal Pathway in a Proteolysis-Dependent Manner. PLoS Pathog. 2014, 10, e1004502.
  80. De Wilde, A.H.; Jochmans, D.; Posthuma, C.C.; Zevenhoven-Dobbe, J.C.; van Nieuwkoop, S.; Bestebroer, T.M.; van den Hoogen, B.G.; Neyts, J.; Snijder, E.J. Screening of an FDA-Approved Compound Library Identifies Four Small-Molecule Inhibitors of Middle East Respiratory Syndrome Coronavirus Replication in Cell Culture. Antimicrob. Agents Chemother. 2014, 58, 4875–4884.
  81. Plaze, M.; Attali, D.; Petit, A.-C.; Blatzer, M.; Simon-Loriere, E.; Vinckier, F.; Cachia, A.; Chrétien, F.; Gaillard, R. Repurposing Chlorpromazine to Treat COVID-19: The reCoVery Study. L’Encephale 2020, 46, 169–172.
  82. Fillmore, N.; Bell, S.; Shen, C.; Nguyen, V.; La, J.; Dubreuil, M.; Strymish, J.; Brophy, M.; Mehta, G.; Wu, H.; et al. Disulfiram Use Is Associated with Lower Risk of COVID-19: A Retrospective Cohort Study. PLoS ONE 2021, 16, e0259061.
  83. Tamburin, S.; Mantovani, E.; De Bernardis, E.; Zipeto, D.; Lugoboni, F.; Agostoni, C.; Almasio, R.; Avveduti, P.; Baisini, O.; Ballerio, M.; et al. COVID-19 and Related Symptoms in Patients under Disulfiram for Alcohol Use Disorder. Intern. Emerg. Med. 2021, 16, 1729–1731.
  84. Hoertel, N.; Sánchez-Rico, M.; Gulbins, E.; Kornhuber, J.; Carpinteiro, A.; Abellán, M.; de la Muela, P.; Vernet, R.; Beeker, N.; Neuraz, A.; et al. Association between FIASMA Psychotropic Medications and Reduced Risk of Intubation or Death in Individuals with Psychiatric Disorders Hospitalized for Severe COVID-19: An Observational Multicenter Study. Transl. Psychiatry 2022, 12, 90.
  85. Carpinteiro, A.; Edwards, M.J.; Hoffmann, M.; Kochs, G.; Gripp, B.; Weigang, S.; Adams, C.; Carpinteiro, E.; Gulbins, A.; Keitsch, S.; et al. Pharmacological Inhibition of Acid Sphingomyelinase Prevents Uptake of SARS-CoV-2 by Epithelial Cells. Cell Rep. Med. 2020, 1, 100142.
  86. Kornhuber, J.; Tripal, P.; Reichel, M.; Mühle, C.; Rhein, C.; Muehlbacher, M.; Groemer, T.W.; Gulbins, E. Functional Inhibitors of Acid Sphingomyelinase (FIASMAs): A Novel Pharmacological Group of Drugs with Broad Clinical Applications. Cell Physiol. Biochem. Int. J. Exp. Cell Physiol. Biochem. Pharmacol. 2010, 26, 9–20.
  87. Oskotsky, T.; Maric, I.; Tang, A.; Oskotsky, B.; Wong, R.J.; Aghaeepour, N.; Sirota, M.; Stevenson, D.K. Mortality Risk Among Patients with COVID-19 Prescribed Selective Serotonin Reuptake Inhibitor Antidepressants. JAMA Netw. Open 2021, 4, e2133090.
  88. COVID-19 Treatment Guidelines Panel Coronavirus Disease 2019 (COVID-19) Treatment Guidelines. National Institutes of Health. Available online: https://www.covid19treatmentguidelines.nih.gov/ (accessed on 14 July 2023).
  89. Alonzi, D.S.; Scott, K.A.; Dwek, R.A.; Zitzmann, N. Iminosugar Antivirals: The Therapeutic Sweet Spot. Biochem. Soc. Trans. 2017, 45, 571–582.
  90. Evans DeWald, L.; Starr, C.; Butters, T.; Treston, A.; Warfield, K.L. Iminosugars: A Host-Targeted Approach to Combat Flaviviridae Infections. Antiviral Res. 2020, 184, 104881.
  91. Warfield, K.L.; Alonzi, D.S.; Hill, J.C.; Caputo, A.T.; Roversi, P.; Kiappes, J.L.; Sheets, N.; Duchars, M.; Dwek, R.A.; Biggins, J.; et al. Targeting Endoplasmic Reticulum α-Glucosidase I with a Single-Dose Iminosugar Treatment Protects against Lethal Influenza and Dengue Virus Infections. J. Med. Chem. 2020, 63, 4205–4214.
  92. Fowler, P.A.; Haines, A.H.; Taylor, R.J.K.; Chrystal, E.J.T.; Gravestock, M.B. Synthesis and Biological Activity of Acyclic Anlogues of Nojirimycin. J. Chem. Soc. Perkin 1 1994, 2229–2235.
  93. Wetherill, L.F.; Wasson, C.W.; Swinscoe, G.; Kealy, D.; Foster, R.; Griffin, S.; Macdonald, A. Alkyl-Imino Sugars Inhibit the pro-Oncogenic Ion Channel Function of Human Papillomavirus (HPV) E5. Antiviral Res. 2018, 158, 113–121.
  94. Miller, J.L.; Tyrrell, B.E.; Zitzmann, N. Mechanisms of Antiviral Activity of Iminosugars Against Dengue Virus. Adv. Exp. Med. Biol. 2018, 1062, 277–301.
  95. Perera, N.; Brun, J.; Alonzi, D.S.; Tyrrell, B.E.; Miller, J.L.; Zitzmann, N. Antiviral Effects of Deoxynojirimycin (DNJ)-Based Iminosugars in Dengue Virus-Infected Primary Dendritic Cells. Antiviral Res. 2022, 199, 105269.
  96. Tyrrell, B.E.; Sayce, A.C.; Warfield, K.L.; Miller, J.L.; Zitzmann, N. Iminosugars: Promising Therapeutics for Influenza Infection. Crit. Rev. Microbiol. 2017, 43, 521–545.
  97. Durantel, D.; Alotte, C.; Zoulim, F. Glucosidase Inhibitors as Antiviral Agents for Hepatitis B and C. Curr. Opin. Investig. Drugs Lond. Engl. 2000 2007, 8, 125–129.
  98. Durantel, D. Celgosivir, an Alpha-Glucosidase I Inhibitor for the Potential Treatment of HCV Infection. Curr. Opin. Investig. Drugs Lond. Engl. 2000 2009, 10, 860–870.
  99. Bhushan, G.; Lim, L.; Bird, I.; Chothe, S.K.; Nissly, R.H.; Kuchipudi, S.V. Iminosugars with Endoplasmic Reticulum α-Glucosidase Inhibitor Activity Inhibit ZIKV Replication and Reverse Cytopathogenicity in Vitro. Front. Microbiol. 2020, 11, 531.
  100. Warfield, K.L.; Warren, T.K.; Qiu, X.; Wells, J.; Mire, C.E.; Geisbert, J.B.; Stuthman, K.S.; Garza, N.L.; Van Tongeren, S.A.; Shurtleff, A.C.; et al. Assessment of the Potential for Host-Targeted Iminosugars UV-4 and UV-5 Activity against Filovirus Infections In Vitro and In Vivo. Antiviral Res. 2017, 138, 22–31.
  101. Miller, J.L.; Spiro, S.G.; Dowall, S.D.; Taylor, I.; Rule, A.; Alonzi, D.S.; Sayce, A.C.; Wright, E.; Bentley, E.M.; Thom, R.; et al. Minimal In Vivo Efficacy of Iminosugars in a Lethal Ebola Virus Guinea Pig Model. PLoS ONE 2016, 11, e0167018.
  102. Clarke, E.C.; Nofchissey, R.A.; Ye, C.; Bradfute, S.B. The Iminosugars Celgosivir, Castanospermine and UV-4 Inhibit SARS-CoV-2 Replication. Glycobiology 2021, 31, 378–384.
  103. Caly, L.; Druce, J.D.; Catton, M.G.; Jans, D.A.; Wagstaff, K.M. The FDA-Approved Drug Ivermectin Inhibits the Replication of SARS-CoV-2 in Vitro. Antiviral Res. 2020, 178, 104787.
  104. Reis, G.; Silva, E.A.S.M.; Silva, D.C.M.; Thabane, L.; Milagres, A.C.; Ferreira, T.S.; dos Santos, C.V.Q.; Campos, V.H.S.; Nogueira, A.M.R.; de Almeida, A.P.F.G.; et al. Effect of Early Treatment with Ivermectin among Patients with COVID-19. N. Engl. J. Med. 2022, 386, 1721–1731.
  105. Naggie, S.; Boulware, D.R.; Lindsell, C.J.; Stewart, T.G.; Gentile, N.; Collins, S.; McCarthy, M.W.; Jayaweera, D.; Castro, M.; Sulkowski, M.; et al. Effect of Ivermectin vs Placebo on Time to Sustained Recovery in Outpatients with Mild to Moderate COVID-19: A Randomized Clinical Trial. JAMA 2022, 328, 1595–1603.
  106. Naggie, S.; Boulware, D.R.; Lindsell, C.J.; Stewart, T.G.; Slandzicki, A.J.; Lim, S.C.; Cohen, J.; Kavtaradze, D.; Amon, A.P.; Gabriel, A.; et al. Effect of Higher-Dose Ivermectin for 6 Days vs Placebo on Time to Sustained Recovery in Outpatients with COVID-19: A Randomized Clinical Trial. JAMA 2023, 329, 888–897.
  107. Lim, S.C.L.; Hor, C.P.; Tay, K.H.; Mat Jelani, A.; Tan, W.H.; Ker, H.B.; Chow, T.S.; Zaid, M.; Cheah, W.K.; Lim, H.H.; et al. Efficacy of Ivermectin Treatment on Disease Progression Among Adults with Mild to Moderate COVID-19 and Comorbidities: The I-TECH Randomized Clinical Trial. JAMA Intern. Med. 2022, 182, 426–435.
  108. Touret, F.; de Lamballerie, X. Of Chloroquine and COVID-19. Antiviral Res. 2020, 177, 104762.
  109. Gao, J.; Tian, Z.; Yang, X. Breakthrough: Chloroquine Phosphate Has Shown Apparent Efficacy in Treatment of COVID-19 Associated Pneumonia in Clinical Studies. Biosci. Trends 2020, 14, 72–73.
  110. Ledford, H. Chloroquine Hype Is Derailing the Search for Coronavirus Treatments. Nature 2020, 580, 573.
  111. RECOVERY Collaborative Group; Horby, P.; Mafham, M.; Linsell, L.; Bell, J.L.; Staplin, N.; Emberson, J.R.; Wiselka, M.; Ustianowski, A.; Elmahi, E.; et al. Effect of Hydroxychloroquine in Hospitalized Patients with COVID-19. N. Engl. J. Med. 2020, 383, 2030–2040.
  112. Abella, B.S.; Jolkovsky, E.L.; Biney, B.T.; Uspal, J.E.; Hyman, M.C.; Frank, I.; Hensley, S.E.; Gill, S.; Vogl, D.T.; Maillard, I.; et al. Efficacy and Safety of Hydroxychloroquine vs Placebo for Pre-Exposure SARS-CoV-2 Prophylaxis Among Health Care Workers: A Randomized Clinical Trial. JAMA Intern. Med. 2021, 181, 195–202.
  113. Rajasingham, R.; Bangdiwala, A.S.; Nicol, M.R.; Skipper, C.P.; Pastick, K.A.; Axelrod, M.L.; Pullen, M.F.; Nascene, A.A.; Williams, D.A.; Engen, N.W.; et al. Hydroxychloroquine as Pre-Exposure Prophylaxis for Coronavirus Disease 2019 (COVID-19) in Healthcare Workers: A Randomized Trial. Clin. Infect. Dis. 2021, 72, e835–e843.
  114. Lokhande, A.S.; Devarajan, P.V. A Review on Possible Mechanistic Insights of Nitazoxanide for Repurposing in COVID-19. Eur. J. Pharmacol. 2021, 891, 173748.
  115. Braga, L.; Ali, H.; Secco, I.; Chiavacci, E.; Neves, G.; Goldhill, D.; Penn, R.; Jimenez-Guardeño, J.M.; Ortega-Prieto, A.M.; Bussani, R.; et al. Drugs That Inhibit TMEM16 Proteins Block SARS-CoV-2 Spike-Induced Syncytia. Nature 2021, 594, 88–93.
  116. Walker, L.E.; FitzGerald, R.; Saunders, G.; Lyon, R.; Fisher, M.; Martin, K.; Eberhart, I.; Woods, C.; Ewings, S.; Hale, C.; et al. An Open Label, Adaptive, Phase 1 Trial of High-Dose Oral Nitazoxanide in Healthy Volunteers: An Antiviral Candidate for SARS-CoV-2. Clin. Pharmacol. Ther. 2022, 111, 585–594.
  117. Ferner, R.E.; Aronson, J.K. Chloroquine and Hydroxychloroquine in COVID-19. BMJ 2020, 369, m1432.
  118. Misra, D.P.; Agarwal, V.; Gasparyan, A.Y.; Zimba, O. Rheumatologists’ Perspective on Coronavirus Disease 19 (COVID-19) and Potential Therapeutic Targets. Clin. Rheumatol. 2020, 39, 2055–2062.
  119. Zhou, F.; Yu, T.; Du, R.; Fan, G.; Liu, Y.; Liu, Z.; Xiang, J.; Wang, Y.; Song, B.; Gu, X.; et al. Clinical Course and Risk Factors for Mortality of Adult Inpatients with COVID-19 in Wuhan, China: A Retrospective Cohort Study. Lancet Lond. Engl. 2020, 395, 1054–1062.
  120. Huang, C.; Wang, Y.; Li, X.; Ren, L.; Zhao, J.; Hu, Y.; Zhang, L.; Fan, G.; Xu, J.; Gu, X.; et al. Clinical Features of Patients Infected with 2019 Novel Coronavirus in Wuhan, China. Lancet Lond. Engl. 2020, 395, 497–506.
  121. Wang, Z.; Yang, B.; Li, Q.; Wen, L.; Zhang, R. Clinical Features of 69 Cases with Coronavirus Disease 2019 in Wuhan, China. Clin. Infect. Dis. 2020, 71, 769–777.
  122. Burrage, D.R.; Koushesh, S.; Sofat, N. Immunomodulatory Drugs in the Management of SARS-CoV-2. Front. Immunol. 2020, 11, 1844.
  123. Ramamoorthy, S.; Cidlowski, J.A. Corticosteroids: Mechanisms of Action in Health and Disease. Rheum. Dis. Clin. N. Am. 2016, 42, 15–31.
  124. Normand, S.-L.T. The RECOVERY Platform. N. Engl. J. Med. 2021, 384, 757–758.
  125. RECOVERY Collaborative Group; Horby, P.; Lim, W.S.; Emberson, J.R.; Mafham, M.; Bell, J.L.; Linsell, L.; Staplin, N.; Brightling, C.; Ustianowski, A.; et al. Dexamethasone in Hospitalized Patients with COVID-19. N. Engl. J. Med. 2021, 384, 693–704.
  126. Yu, L.-M.; Bafadhel, M.; Dorward, J.; Hayward, G.; Saville, B.R.; Gbinigie, O.; Hecke, O.V.; Ogburn, E.; Evans, P.H.; Thomas, N.P.B.; et al. Inhaled Budesonide for COVID-19 in People at High Risk of Complications in the Community in the UK (PRINCIPLE): A Randomised, Controlled, Open-Label, Adaptive Platform Trial. Lancet 2021, 398, 843–855.
  127. Food and Drug Administration Coronavirus (COVID-19) Update: FDA Authorizes Drug for Treatment of COVID-19. Available online: https://www.fda.gov/news-events/press-announcements/coronavirus-covid-19-update-fda-authorizes-drug-treatment-covid-19#:~:text=Today%2C%20the%20U.S.%20Food%20and,oxygen%2C%20non%2Dinvasive%20or%20invasive (accessed on 25 July 2023).
  128. Abani, O.; Abbas, A.; Abbas, F.; Abbas, M.; Abbasi, S.; Abbass, H.; Abbott, A.; Abdallah, N.; Abdelaziz, A.; Abdelfattah, M.; et al. Tocilizumab in Patients Admitted to Hospital with COVID-19 (RECOVERY): A Randomised, Controlled, Open-Label, Platform Trial. Lancet 2021, 397, 1637–1645.
  129. Salama, C.; Han, J.; Yau, L.; Reiss, W.G.; Kramer, B.; Neidhart, J.D.; Criner, G.J.; Kaplan-Lewis, E.; Baden, R.; Pandit, L.; et al. Tocilizumab in Patients Hospitalized with COVID-19 Pneumonia. N. Engl. J. Med. 2021, 384, 20–30.
  130. Langer-Gould, A.; Smith, J.B.; Gonzales, E.G.; Castillo, R.D.; Figueroa, J.G.; Ramanathan, A.; Li, B.H.; Gould, M.K. Early Identification of COVID-19 Cytokine Storm and Treatment with Anakinra or Tocilizumab. Int. J. Infect. Dis. 2020, 99, 291–297.
  131. Geng, J.; Wang, F.; Huang, Z.; Chen, X.; Wang, Y. Perspectives on Anti-IL-1 Inhibitors as Potential Therapeutic Interventions for Severe COVID-19. Cytokine 2021, 143, 155544.
  132. Karakike, E.; Dalekos, G.N.; Koutsodimitropoulos, I.; Saridaki, M.; Pourzitaki, C.; Papathanakos, G.; Kotsaki, A.; Chalvatzis, S.; Dimakopoulou, V.; Vechlidis, N.; et al. ESCAPE: An Open-Label Trial of Personalized Immunotherapy in Critically Lll COVID-19 Patients. J. Innate Immun. 2021, 14, 218–228.
  133. Feldmann, M.; Maini, R.N.; Woody, J.N.; Holgate, S.T.; Winter, G.; Rowland, M.; Richards, D.; Hussell, T. Trials of Anti-Tumour Necrosis Factor Therapy for COVID-19 Are Urgently Needed. Lancet Lond. Engl. 2020, 395, 1407–1409.
  134. Mortaz, E.; Tabarsi, P.; Jamaati, H.; Dalil Roofchayee, N.; Dezfuli, N.K.; Hashemian, S.M.; Moniri, A.; Marjani, M.; Malekmohammad, M.; Mansouri, D.; et al. Increased Serum Levels of Soluble TNF-α Receptor Is Associated with ICU Mortality in COVID-19 Patients. Front. Immunol. 2021, 12, 592727.
  135. Kennedy, N.A.; Goodhand, J.R.; Bewshea, C.; Nice, R.; Chee, D.; Lin, S.; Chanchlani, N.; Butterworth, J.; Cooney, R.; Croft, N.M.; et al. Anti-SARS-CoV-2 Antibody Responses Are Attenuated in Patients with IBD Treated with Infliximab. Gut 2021, 70, 865–875.
  136. Chanchlani, N.; Lin, S.; Chee, D.; Hamilton, B.; Nice, R.; Zehra, A.; Bewshea, C.; Cipriano, B.; Derikx, L.A.A.P.; Dunlop, A.; et al. Adalimumab and Infliximab Impair SARS-CoV-2 Antibody Responses: Results from a Therapeutic Drug Monitoring Study in 11422 Biologic-Treated Patients. J. Crohns Colitis 2021, 16, jjab153.
  137. Fisher, B.A.; Veenith, T.; Slade, D.; Gaskell, C.; Rowland, M.; Whitehouse, T.; Scriven, J.; Parekh, D.; Balasubramaniam, M.S.; Cooke, G.; et al. Namilumab or Infliximab Compared with Standard of Care in Hospitalised Patients with COVID-19 (CATALYST): A Randomised, Multicentre, Multi-Arm, Multistage, Open-Label, Adaptive, Phase 2, Proof-of-Concept Trial. Lancet Respir. Med. 2022, 10, 255–266.
  138. Velez, M.P.; McCarthy, M.W. Infliximab as a Potential Treatment for COVID-19. Expert Rev. Anti Infect. Ther. 2023, 21, 1–5.
  139. O’Halloran, J.A.; Ko, E.R.; Anstrom, K.J.; Kedar, E.; McCarthy, M.W.; Panettieri, R.A., Jr.; Maillo, M.; Nunez, P.S.; Lachiewicz, A.M.; Gonzalez, C.; et al. Abatacept, Cenicriviroc, or Infliximab for Treatment of Adults Hospitalized with COVID-19 Pneumonia: A Randomized Clinical Trial. JAMA 2023, 330, 328–339.
  140. Carvelli, J.; Demaria, O.; Vély, F.; Batista, L.; Benmansour, N.C.; Fares, J.; Carpentier, S.; Thibult, M.-L.; Morel, A.; Remark, R.; et al. Association of COVID-19 Inflammation with Activation of the C5a-C5aR1 Axis. Nature 2020, 588, 146–150.
  141. Vlaar, A.P.J.; Witzenrath, M.; van Paassen, P.; Heunks, L.M.A.; Mourvillier, B.; de Bruin, S.; Lim, E.H.T.; Brouwer, M.C.; Tuinman, P.R.; Saraiva, J.F.K.; et al. Anti-C5a Antibody (Vilobelimab) Therapy for Critically Ill, Invasively Mechanically Ventilated Patients with COVID-19 (PANAMO): A Multicentre, Double-Blind, Randomised, Placebo-Controlled, Phase 3 Trial. Lancet Respir. Med. 2022, 10, 1137–1146.
  142. Taylor, P.C. Clinical Efficacy of Launched JAK Inhibitors in Rheumatoid Arthritis. Rheumatol. Oxf. Engl. 2019, 58, i17–i26.
  143. Marconi, V.C.; Ramanan, A.V.; de Bono, S.; Kartman, C.E.; Krishnan, V.; Liao, R.; Piruzeli, M.L.B.; Goldman, J.D.; Alatorre-Alexander, J.; de Cassia Pellegrini, R.; et al. Efficacy and Safety of Baricitinib for the Treatment of Hospitalised Adults with COVID-19 (COV-BARRIER): A Randomised, Double-Blind, Parallel-Group, Placebo-Controlled Phase 3 Trial. Lancet Respir. Med. 2021, 9, 1407–1418.
  144. RECOVERY Collaborative Group Baricitinib in Patients Admitted to Hospital with COVID-19 (RECOVERY): A Randomised, Controlled, Open-Label, Platform Trial and Updated Meta-Analysis. Lancet Lond. Engl. 2022, 400, 359–368.
  145. Lin, Z.; Niu, J.; Xu, Y.; Qin, L.; Ding, J.; Zhou, L. Clinical Efficacy and Adverse Events of Baricitinib Treatment for Coronavirus Disease-2019 (COVID-19): A Systematic Review and Meta-Analysis. J. Med. Virol. 2021, 94, 1523–1534.
  146. Wolfe, C.R.; Tomashek, K.M.; Patterson, T.F.; Gomez, C.A.; Marconi, V.C.; Jain, M.K.; Yang, O.O.; Paules, C.I.; Palacios, G.M.R.; Grossberg, R.; et al. Baricitinib versus Dexamethasone for Adults Hospitalised with COVID-19 (ACTT-4): A Randomised, Double-Blind, Double Placebo-Controlled Trial. Lancet Respir. Med. 2022, 10, 888–899.
  147. Patoulias, D.; Doumas, M.; Papadopoulos, C.; Karagiannis, A. Janus Kinase Inhibitors and Major COVID-19 Outcomes: Time to Forget the Two Faces of Janus! A Meta-Analysis of Randomized Controlled Trials. Clin. Rheumatol. 2021, 40, 4671–4674.
  148. Kalil, A.C.; Patterson, T.F.; Mehta, A.K.; Tomashek, K.M.; Wolfe, C.R.; Ghazaryan, V.; Marconi, V.C.; Ruiz-Palacios, G.M.; Hsieh, L.; Kline, S.; et al. Baricitinib plus Remdesivir for Hospitalized Adults with COVID-19. N. Engl. J. Med. 2021, 384, 795–807.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , ,
View Times: 202
Entry Collection: COVID-19
Revisions: 3 times (View History)
Update Date: 09 Jan 2024
1000/1000