Next Article in Journal
Delving into the Aftermath of a Disease-Associated Near-Extinction Event: A Five-Year Study of a Serpentovirus (Nidovirus) in a Critically Endangered Turtle Population
Previous Article in Journal
Association of the IFNG +874T/A Polymorphism with Symptomatic COVID-19 Susceptibility
Previous Article in Special Issue
Fluoxetine and Sertraline Potently Neutralize the Replication of Distinct SARS-CoV-2 Variants
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Why Certain Repurposed Drugs Are Unlikely to Be Effective Antivirals to Treat SARS-CoV-2 Infections

1
Center for ViroScience and Cure, Laboratory of Biochemical Pharmacology, Department of Pediatrics, Emory University School of Medicine and Children’s Healthcare of Atlanta, 1760 Haygood Drive, Atlanta, GA 30322, USA
2
Center for Cystic Fibrosis & Airways Disease Research, Division of Pulmonary, Allergy & Immunology, Cystic Fibrosis and Sleep, Emory University and Children’s Healthcare of Atlanta, 2015 Uppergate Drive, Atlanta, GA 30322, USA
3
Morningside Center for Innovative and Affordable Medicine, Departments of Medicine and Hematology and Oncology, Emory University School of Medicine, Atlanta, GA 30322, USA
*
Author to whom correspondence should be addressed.
Viruses 2024, 16(4), 651; https://doi.org/10.3390/v16040651
Submission received: 8 March 2024 / Revised: 10 April 2024 / Accepted: 16 April 2024 / Published: 22 April 2024

Abstract

:
Most repurposed drugs have proved ineffective for treating COVID-19. We evaluated median effective and toxic concentrations (EC50, CC50) of 49 drugs, mostly from previous clinical trials, in Vero cells. Ratios of reported unbound peak plasma concentrations, (Cmax)/EC50, were used to predict the potential in vivo efficacy. The 20 drugs with the highest ratios were retested in human Calu-3 and Caco-2 cells, and their CC50 was determined in an expanded panel of cell lines. Many of the 20 drugs with the highest ratios were inactive in human Calu-3 and Caco-2 cells. Antivirals effective in controlled clinical trials had unbound Cmax/EC50 ≥ 6.8 in Calu-3 or Caco-2 cells. EC50 of nucleoside analogs were cell dependent. This approach and earlier availability of more relevant cultures could have reduced the number of unwarranted clinical trials.

Graphical Abstract

1. Introduction

Since the start of the COVID-19 pandemic, caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), pharmaceutical companies and academic institutions have urgently searched for treatments and vaccines. Several specific and potent antiviral agents and vaccines have undergone various stages of clinical testing and regulatory approval. The three vaccines licensed in the USA include the Pfizer-BioNTech and Moderna COVID-19 mRNA vaccines, and the Novavax COVID-19 vaccine, which is a protein subunit vaccine [1]. They are considered safe and effective in reducing the number of individuals infected, and in decreasing the severity of the symptoms in vaccinated persons who experience breakthrough infections. However, vaccines have a limited impact for individuals with active SARS-CoV-2 infections and provide limited global protection against emerging virus strains [2,3]. Existing vaccines that target the spike protein of SARS-CoV-2 make cross resistance between vaccines theoretically possible, and the durability of protection is low, thus necessitating repeated vaccination and boosters.
Drugs developed specifically to block SARS-CoV-2 replication, as well as those considered for repurposing, are directed at viral targets other than the spike protein, making cross resistance with vaccines unlikely [4]. Remdesivir is an antiviral nucleoside analog (NA) produced by Gilead Sciences Inc which was originally developed for the treatment of hepatitis C virus (HCV), Ebola virus, and Marburg virus infections [5], but it was never approved by the Food and Drug Administration (FDA) for these indications. Intravenous (IV) remdesivir was shown to shorten the hospitalization duration of hospitalized patients with COVID-19 by 5 days (ACTT-1 trial) and improve their survival outcomes [6,7,8]. A more recent clinical study with IV remdesivir in high-risk non-hospitalized patients (PINETREE trial) demonstrated an 87% reduction in hospitalizations following a short 3-day course [9]. Remdesivir is now FDA approved for the treatment of COVID-19 in adults, and pediatric patients aged ≥ 28 days and weighing ≥ 3 kg, for both non-hospitalized patients with mild to moderate COVID-19 who are at high risk of progressing to severe disease, and in hospitalized patients [10,11]. Ritonavir-boosted nirmatrelvir (PaxlovidTM, Pfizer Inc New York City, NY, USA) was the first oral formulation to receive full FDA approval for home treatment in high-risk adults with co-morbidities, starting within a few days of their first symptoms [12,13,14]. Paxlovid™ was shown to reduce the risk of hospitalization or death by 89% (if administered within a few days of their first symptoms) in persons over 12 years of age [12,15]. The current FDA COVID-19 Treatment Guidelines recommend using Paxlovid for 5 days in non-hospitalized adult and pediatric patients, aged ≥12 years and weighing ≥ 40 kg, with mild to moderate COVID-19 who are at high risk of disease progression [16]. Treatment should be initiated as soon as possible and within 5 days of symptom onset [14]. Molnupiravir (LagevrioTM, Merck & Co., Inc., Rahway, NJ, USA) is an orally administered antiviral NA which received emergency use approval from the FDA for home treatment in high-risk adults with co-morbidities, starting within a few days of their first symptoms [17]. Molnupiravir is a 5′-ester prodrug of N4-hydroxycytidine, which was originally discovered as a broad-spectrum antiviral agent in our laboratories at Emory University and Pharmasset Inc., and was routinely used as a positive control in antiviral screening assays, including for HCV and other viruses [18,19,20,21]. It was repurposed as an ester prodrug by Drug Innovation Ventures at Emory (DRIVE LLC, Atlanta, GA, USA) for coronaviruses, and licensed to Ridgeback Biotherapeutic LP, and then to Merck & Co., Inc. [22]. Molnupiravir, like most antiviral NAs, does not require boosting and is less susceptible to drug–drug interactions (DDIs) than protease inhibitors. The drug’s mechanism of action involves inducing lethal mutations in the SARS-CoV-2 genome. Furthermore, molnupiravir can have mutagenic effects in mammalian cells and is not recommended in pregnant women [23,24,25,26]. Obeldesivir (GS-5245, Gilead Sciences Inc, San Francisco, CA, USA) is an orally administered 5′-isobutyl prodrug of GS-441524 (parent compound of remdesivir) being evaluated in two Phase 3 clinical trials (ClinicalTrials.gov Identifiers: NCT05603143 and NCT05715528) [27,28].
The drug-resistant lineages of SARS-CoV-2 have been isolated in infected cells serially passaged in increasing concentrations of remdesivir (or GS-441524) and nirmatrelvir [29,30]. Resistance to remdesivir (or GS-441524) is associated with mutations in the RNA-dependent RNA polymerase gene, resulting in a 2.5 to 10-fold increase in EC50 [29,31], while resistance to nirmatrelvir is associated with mutations in the 3-chymotrypsin-like cysteine protease (3CLpro) gene [30]. Documented clinical cases of resistance to remdesivir (or GS-441524) and nirmatrelvir are rare [32,33], likely due to the relatively short duration of treatment of most patients with COVID-19. The most common 3CLpro mutations selected for nirmatrelvir in the clinic, one year after the FDA approval of nirmatrelvir–ritonavir, were E166V and S144E, which reduce the inhibitory activity of nirmatrelvir by >100-fold and were present in <1 per million sequences [33]. Molnupiravir also has a high barrier to resistance in vitro [34]. However, molnupiravir’s signature mutations were detected in a study of newly infected persons in close contact with molnupiravir-treated individuals who had not cleared the virus [35]. Since resistance to these antivirals may occur, it is important to monitor for any increase in resistant SARS-CoV-2 strains in the population, especially with the emergence of new SARS-CoV-2 variants. Since drug resistance may occur, and mutant SARS-CoV-2 lineages continue to emerge, further drug development is needed to expand treatment options.
Antiviral agents developed for other diseases (e.g., HIV, viral hepatitis, and for veterinary viruses) have also been considered as potential agents to use against SARS-CoV-2, but none have been useful when rigorously evaluated in cell culture or in clinical trials [36,37,38,39,40,41]. Other medications approved for unrelated conditions (e.g., antimalarial drugs [42]; COX inhibitors [43,44]; glucocorticoids [45]; non-steroidal anti-inflammatory agents [46,47,48]; immunosuppressive monoclonal antibodies [49]; histamine blockers (H1 or H2 receptors) [50,51]; leukotriene inhibitors [52]; antioxidants [53,54]; mucolytics [55,56]; antibacterial antibiotics [57,58]; anti-fungal compounds [59]; anti-parasitic agents [60,61]; lipid lowering drugs [62]; oncology drugs [63,64]; selective serotonin uptake inhibitor antidepressants [65]; drugs used to treat diabetes [66]; lipid lowering agents [67]; dietary supplements [68,69]; and food additives [70]) have all been considered. However, most of these repurposed drugs have shown marginal or no benefit in clinical studies (discussed below).
In this study, we provide a side-by-side evaluation of the antiviral activity of 49 drugs previously considered for inclusion in regimens for treating individuals infected with SARS-CoV-2, using remdesivirTM as a positive control in a cell-based assay system. The in vitro potency measurements of all these drugs were compared with peak drug concentrations observed in the clinic to assess the potential clinical relevance of these compounds [71].

2. Materials and Methods

2.1. Drug Candidates

All drug candidates were purchased from Sigma-Aldrich (St. Louis, MO, USA) or a reliable source (e.g., extraction from prescription drugs) with at least 98% purity as determined by LC-MS-MS.

2.2. Cells and Viruses

Cell and virus culture techniques were adapted from previous projects in our laboratory [72,73]. Briefly: An African green monkey kidney epithelial cell line (Vero-CCL-81; CCL-81) [74], a human colorectal adenocarcinoma cell line (Caco-2; HTB-37™), a human lung cancer cell line (Calu-3; HTB-55)s [75,76], and a T-lymphoblastoid cell line (CCRF-CEM; CCL-119) were all obtained from the American Type Culture Collection (ATCC, Manassas, VA, USA). A hepatic cancer cell line (Huh-7; JCRB0403) was obtained from the Japanese Collection of Research Bioresources Cell Bank (JCRB, Osaka, Japan) [77]. Primary human peripheral blood mononuclear cells (PBMs) were isolated from blood pooled from multiple donors purchased from the New York Blood Center (NYBC; New York City, NY, USA) and stimulated with phytohemagglutinin [78]. Vero cells were cultured in minimum essential medium (MEM) supplemented with 10% fetal bovine serum (FBS). Huh-7 cells were cultured in Dulbecco’s modified eagle medium (DMEM) with 10% FBS. Caco-2 and Calu-3 cells were cultured in DMEM/F12 supplemented with 10% FBS. CEM and PBM cells were cultured in Roswell Park Memorial Institute (RPMI) 1640 Medium supplemented with 10 and 20% FBS, respectively. Media was supplemented with 2 µM L-glutamine and 100U Penicillin/Streptomycin. Media was purchased from Corning (Corning, NY, USA), and a single lot of heat-inactivated FBS (VWR, Radnor, PA, USA) was used. SARS-CoV-2, isolate hCoV-19/USA-WA1/2020, NR-52281, deposited by the Centers for Disease Control and Prevention and obtained through BEI Resources, NIAID, NIH: SARS-Related Coronavirus 2, was used.

2.3. In Vitro Evaluation of Compounds versus SARS-CoV-2

The anti-SARS-CoV-2 EC50 values and cytotoxicities (CC50) of drugs were initially evaluated in Vero cells. Drugs having the following properties were subsequently evaluated in human cells: (1) unbound Cmax/EC50 > 0.007 and EC50 and CC50 in Vero CCL-81 SARS-CoV-2 cells; (2) or having been tested or considered for clinical testing in individuals with COVID-19; (3) or reported active in an in vivo SARS-CoV-2 infection model; (4) or presumed active based on mechanistic studies.

2.3.1. Measurement of Half Maximal Cytotoxic Concentration (CC50)

The CC50 values of promising compounds were also measured in human PBM, Huh-7, CEM, Caco-2, and Calu-3. For the PBM, Huh-7, and CEM assays, cells were plated directly with the compound and allowed to grow under proliferative (10% FBS) conditions for 3 (Vero), 4 (Huh-7 and CEM), or 5 days (PBM). For Caco-2 and Calu-3 assay, cells were assayed under non-proliferative conditions, whereby they were grown to 90% confluence in media with 10% FBS, before replacing with fresh media containing 2% FBS and dissolved drug and incubated for 3 days. At the end of the incubation period, cells were tested for their ability to metabolize MTS (3-(4,5-dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium) as previously described [72]. In brief, 15 µL of MTS reagent (Promega©, Madison, WI, USA) was added per well and optical density reads were taken after 1–4 h at 490 nm absorbance on a multi-mode plate reader (Synergy, BioTek®, Winooski, VT, USA). After subtraction of the media-only controls, CC50 values were calculated using the Chou and Talalay method [79]. Drug binding in 2% FBS was expected to be negligible, so that CC50 measured in Calu-3 and Caco-2 cells was assumed equal to the CC50 of unbound drug. Unbound CC50 values in other cells were calculated by multiplying the measured CC50 values by the unbound fraction of the compound reported in human plasma, normalized to the FBS concentration in the media.

2.3.2. Measurement of In Vitro Anti-SARS-CoV-2 Activity (EC50 and EC90)

The drugs were initially evaluated for anti-SARS-CoV-2 activity in Vero CCL-81 cells, followed by an evaluation in Calu-3 and Caco-2 cells, using a virus yield inhibition assay with virus quantification by qRT-PCR, as described previously [80]. Briefly, a monolayer of each cell line was prepared in a 96-well plate using media containing 10% FBS. Once the cells reached confluency, they were then treated with 2-fold serial dilutions of each compound in triplicate and infected with SARS-CoV-2 (hCoV-19/USA/WA1/2020 strain) at a multiplicity of infection (MOI) of 0.1 (Vero) and 0.01 (Calu3 and Caco2). The cells were then incubated at 37 °C for 48 h (Vero) and 72 h (Calu3 and Caco2) in the presence of 5% CO2. The cell infections were performed in medium containing 2% of heat-inactivated FBS. After incubation, 100 µL of the supernatant was collected into 150 µL of RLT Buffer (Qiagen©, Germantown, MD, USA) for downstream RNA extraction (RNeasy 96 extraction kit, Qiagen©, Germantown, MD, USA) and subsequent qRT-PCR to detect the viral load using a 6-carboxyfluorescein (FAM)-labeled probe with primers against the SARS-CoV-2 non-structural protein 3 (nsp3). (SARS-CoV-2 FWD: AGA AGA TTG GTT AGA TGA TGA TAG T; SARS-CoV-2 REV:TTC CAT CTC TAA TTG AGG TTG AAC C; SARS-CoV-2 Probe: 56-FAM/TC CTC ACT GCC GTC TTG TTG ACC A/3BHQ_1).
The antiviral NAs and nucleoside base precursors, other than favipiravir, were evaluated up to 20 μM (>Cmax observed in the clinic) [81,82,83,84]. Favipiravir (AVIGANTM, Taisho Toyama Pharmaceutical Co., Ltd., Subsidiary of Fujifilm, Tokyo, Japan) was evaluated up to 300 μM, as it is administered at a considerably higher dose and has a higher Cmax than the other NAs tested (Table 1) [85,86]. Other drugs were tested up to 100 μM. Only drugs with measurable, low EC50 values were retested to ensure reproducibility.

2.4. Pharmacokinetics (PK) and Assessment of the Relative Clinical Potential of Compounds

Peak plasma concentrations (Cmax, μM = ng/mL/MW, where MW is the molecular weight of the compound) and the unbound fraction of most drugs were obtained from the literature, or from FDA-approved package inserts. The unbound fraction of apilimod (not found in the literature) was estimated based on chemical structure, using a PK computer program, which uses a random forest machine learning method (DruMAP ver. 1.5, Mizhguchi Laboratory, Tokyo, Japan, https://drumap.nibiohn.go.jp/, Accessed 5 April 2024) [87]. Drugs that exhibited binding to plasma proteins were assumed to equilibrate rapidly between the protein-bound and unbound fractions in the plasma, and that only the unbound fraction was free to interact with cells [88]. The unbound concentration of drug in the plasma at Cmax (unbound Cmax) was calculated by multiplying Cmax by the unbound fraction of drug.
Table 1. Anti-SARS-CoV-2 activity and Cmax/EC50 of drugs in Calu-3, Caco-2, and Vero CCL-81 cells.
Table 1. Anti-SARS-CoV-2 activity and Cmax/EC50 of drugs in Calu-3, Caco-2, and Vero CCL-81 cells.
Cmax (μM) Calu-3 SARS-CoV-2Caco-2 SARS-CoV-2Vero CCL-81 SARS-CoV-2Therapeutic CategoryRefs
Drug
Typical dose
(% CV or range)
{unbound}
FunEC50/EC90
(μM)
Unbound
Cmax/EC50
EC50/EC90
(μM)
Unbound
Cmax/EC50
EC50/EC90
(μM)
unbound
Cmax/EC50
NA
* VV116
300 mg bid
10.6
(25%)
{8.37}
------34---105---10.5NA[89]
** Molnupiravir (NHC)
800 mg bid
8.99
(37%)
{7.10}
10.4/1.8 221.7/12.25.30.6/1.44.58NA[17]
* Obeldesivir
350 mg bid
7.33
(32%)
{5.79}
------19---72---7.2NA[28]
GS-441524
750 mg/d
3.05
(23%)
{2.41}
0.790.3/2.348.030.08/1.4300.8/1.43.0NA[90,91]
Remdesivir
200 mg d1,
100 mg qd
3.70
(19%)
{0.44}
0.120.2/0.72.20.02/0.1223.2/4.70.14NA[10]
Nirmatrelvir
300 mg + 100 mg ritonavir bid.
6.86
(33%)
{2.13}
0.311.8/8.80.40.1/0.3 6.83.7/5.60.21NNAV[92]
Nitazoxanide1.4
(19%)
{0.014}
0.0122.8/42.90.0154.4/11.10.0781.4/3.90.25APA[93,94]
Ivermectin
390–470 μg/kg qd
0.30
(66%)
{0.28}
0.076.2/9.30.0034.5/8.90.0052.1/4.50.010APA[71,95,96,97]
Imatinib
400 mg qd
5.17
(30.3%)
{0.26}
0.0577.9/95.80.00337.9/880.0064.6/14.50.056ONC[98]
Apilimod mesylate
150 mg qd
0.54
{0.043}
0.08>100/NENE 85/>1000.00050.7/7.4* 0.062ILi[99,100]
Celecoxib
200 mg bid
1.85
(38%)
{0.056}
0.033.0/6.30.0190.7/1.20.087.2/14.60.008NSAID[101]
Zileuton
600 m qd
2.29
{0.16}
0.07>100/NENE>1000.0036.7/35.80.02LM[102]
Daclatasvir
60 mg qd
1.91
(13%)
{0.02}
0.01>100/NENE44.7/89.40.00050.7/3.50.03NNAV[103]
Fenofibrate
200 mg qd (micronized form)
28.36
(19%)
{0.28}
0.01>100/NENE>100/NENE42.7/88.70.007LL[104]
Ebselen
800 mg qd
0.256
(47%)
{0.013}
0.056.5/>100.002>100/NENE1.1/2.70.012AO[105]
Favipiravir411
(45%)
{193}
0.46>300/NENE256/>3000.7490.6/234.12.1NA[85]
Fluvoxamine (maleate)
100 mg qd
0.039
(0.02–0.06)
{0.003}
0.23>100/NE0.00517.9/47.50.0017>100/NENESSUi[106,107]
Honokiol
not approved
NE0.3669.2/94.069.2/949.6/80.4NE20.1/84NENP[108]
Iota-Carrageenan
(topical)
------0.5/1.9---59.2/93.3---0.7/1.6---DS
Mefenamic acid
500 mg day 1, 250 mg qid
15.83
(11–22)
{1.58}
0.1> 100/NENE>100/NENE87.6/>1000.018NSA[109,110]
Cmax = peak plasma concentration, Fun = unbound (presumed active) fraction of drugs in plasma; EC50 and EC90 are the in vitro concentrations of drug which inhibit viral replication by 50 and 90%, respectively. CC50 = drug concentration which inhibits cell growth by 50% (means, % CV, or ranges (when available) and mean unbound concentrations are shown). Unbound Cmax/EC50 ratio = Cmax × Fun/EC50. qd = once daily, bid = twice daily, tid = three times daily. NE = not estimated with existing data. Drug binding in 2% FBS was expected to be negligible, so that the EC50 measured in Calu-3, Caco-2 cells, and Vero cell was assumed equal to those of the unbound (presumed active) fraction of drug. * Oral administration of obeldesivir delivers GS-441524 into the plasma. VV116TM delivers the deuterated form of GS-441524 into plasma, which was assumed to have similar PK as the non-deuterated form. Therefore, the Cmax and cellular pharmacology of these drugs are reported in terms of GS-441524 (tabulated as ---). ** Oral administration of molnupiravir delivers NHC in the plasma, so the Cmax and cellular pharmacology is reported in terms of NHC (β-D-N4-hydroxycytidine). The drugs are listed in descending order of the ratio of unbound drug in plasma to antiviral potency SARS-CoV-2 measured in vitro (Cmax/EC50) in the order: Calu-3, Caco-2, and Vero cells. NE = not estimated. Therapeutic category; AO: antioxidant; APA: anti-parasitic agent; DS: dietary supplement; FA: food additive; AH: histamine 1 or 2 blocker; IL: inhibitor of IL-12 and IL-23 production; LL: lipid-lowering drug; LM: leukotriene modulator; NA: antiviral analog of a nucleoside or nucleoside base; NNAV: non-nucleoside and non-nucleoside-base analog antiviral agent; NSAID: non-steroidal anti-inflammatory drug.

3. Results

Table 1 summarizes the anti-SARS-CoV-2 activities in human Calu-3 and Caco-2 cells, and in the Vero CCL 81 cells. Drugs are tabulated in descending order of their mean unbound Cmax/EC50 ratios in rank order of Calu-3, Caco-2, and Vero CCL-81 cells, respectively. Higher ratios were assumed to correspond with a greater in vivo potential for efficacy. This ratio is considered a lenient indicator of potential efficacy of compound, as plasma concentrations typically decline shortly after oral delivery or after the completion of IV infusion [71].
The toxicities of the drugs in Calu-3, Caco-2, Vero CCL-81, PBM, CEM, and Huh-7 cells are listed in Table 2. Drugs are listed in the same order as in Table 1.
Table 3 summarizes the anti-SARS-CoV-2 activity, cellular toxicity, and Cmax at typical dose regimens of drugs tested only in Vero CCL-81. The following drugs were not tabulated due to a lack of antiviral activity in Vero CCL-81 cells: NAs and nucleoside-base precursors (emtricitabine, lamivudine, ribavirin, sofosbuvir, tenofovir alafenamide, tenofovir disoproxil fumarate, kinetin). Drugs which were inactive at concentrations up to 100 μM included: cetirizine (antihistamine), dexamethasone (glucocorticoid), L-ascorbic acid (antioxidant), fluvoxamine maleate (serotonin re-uptake inhibitor), metformin (used to treat diabetes), and plerixafor (used to treat malignancies).
The EC50, EC90, and CC50 values for NA and nucleoside-base analogs were especially celldependent (discussed below). Most drugs with anti-SARS-CoV-2 activity had unbound EC50 values similar to or below the CC50 values in the cells used for the assay, suggesting that their antiviral activity was not due to cytotoxicity per se. Except for certain antiviral NAs, the EC50 values in the Vero cells were lower than in the Caco-2 or Calu-3 cells (discussed below). The NAI developed as anti-SARS-CoV-2 agents had the highest protein adjusted Cmax/EC50 ratios in Calu-3 (2.2 and 34) and Caco-2 cells (2.2 to 34). The SARS-CoV-2 protease inhibitor, nirmatrelvir (active component of PaxlovidTM), had an adjusted Cmax/EC50 ratio = 0.42 and 6.8 in Calu-3 and Caco-2 cells, respectively. Drugs that were ineffective in the clinic had considerably lower Cmax/EC50 ratios (e.g., ivermectin had Cmax/EC50 ratios = 0.003 and 0.005 in Calu-3 and Caco-2 cells, respectively). The remaining drugs had Cmax/EC50 ratios that were considerably lower than that of ivermectin in these human cell lines.

4. Discussion

4.1. Relevance of Cell Lines and SARS-CoV-2 Variants

Since the start of the COVID-19 pandemic, many studies have investigated the possible repurposing of existing medications for the treatment of this disease [42,128,129]. In the present study, initial antiviral evaluations were performed in Vero CCL-81 cells, a kidney epithelial cell line derived from the African Green monkey. Vero cells are commonly used for in vitro evaluation as they are highly permissive to SARS-CoV-2, possibly related to a defective interferon response [42]. Certain drugs, such as chloroquine and hydroxychloroquine (not tested in this study), gained emergency approval from the US FDA during the start of the pandemic, but proved ineffective [42]. Initial optimism may have been fueled by in vitro potency reports on these drugs which relied mainly on experiments performed in Vero cells [130]. SARS-CoV-2 binds the cellular receptor ACE2 and must be activated by proteolysis with either a surface-expressed protease, like TMPRSS2, or by cathepsin L in the endosome [131]. In Vero cells, SARS-CoV-2 is primarily activated by cathepsin L in acidic endosomes, where it is uncoated before being released into the cytoplasm [132]. Chloroquine and hydroxychloroquine are di-protonic weak bases which raise the pH of the endosomes, thereby inhibiting virus uncoating and replication. However, in human lung and colon cells (primary sites of infection), SARS-CoV-2 is primarily activated by TMPRSS2, which does not require lysosomal activation and is not inhibited by chloroquine [131].
SARS-CoV-2 primarily infects the lung and GI tract in individuals with COVID-19 [132]. Therefore, the drugs that demonstrated potency in Vero cells were further evaluated in Calu-3 cells, a lung epithelial cell line originally derived from a 23-year-old man with a lung sarcoma, and Caco-2 cells, a cell line derived from a 72-year-old man with colorectal cancer. Notably, the EC50 and CC50 values of the drugs differed in the various cell systems.
Data from the Calu-3 cells are prioritized over that of the Caco-2 cells in Table 1, since lung involvement is the most common serious manifestation of COVID-19, ranging from asymptomatic disease or mild pneumonia to severe disease associated with hypoxia, and critical disease associated with shock, respiratory failure, and multiorgan failure or death. However, Caco-2 data are also considered important since the GI symptoms during a COVID-19 infections can be severe [132].
The SARS-CoV-2 isolate, hCoV-19/USA-WA1/2020, NR-52281, was used in this study as it is well studied in the literature, which allows our data to be compared with other studies. However, since the virus continues to mutate, further studies may be indicated should drug-resistant strains emerge [2,3].

4.2. Potent Antivirals: Unbound Cmax/EC50 Ratios > 1 in Calu-3, Caco-2, and Vero CCL-81 Cells

Of the six drugs in this category, five are NAs, including remdesivir, an FDA-approved IV-administered prodrug of GS-441524 (used as a positive control in the in vitro assays), and orally administered drugs, GS-441524, VV116 (prodrug of a deuterated version of GS-441524), Obeldesivir (a prodrug of GS-441524), and molnupiravir (prodrug of NHC). Nirmatrelvir, the SARS-CoV-2 protease inhibitor in PaxlovidTM, is also included in this category.
In general, the NAs enter cells rapidly, via the equilibrative nucleoside transporters 1 and 2 (Ent 1 and 2), and undergo intracellular phosphorylation to their nucleoside 5′-triphosphate (NTP) forms, which inhibit SARS-CoV-2 RNA polymerase, composed of the nonstructural protein 12 (nsp12) and the co-factors, nsp7/nsp8, a key enzymatic system responsible for the replication of the viral RNA [133,134]. Following the administration of NAs to patients, cellular concentrations of the accumulated NTP decline slower than NAI in the plasma. Therefore, the dosing frequency of the antiviral NAs are typically related to the respective cellular stabilities (t1/2) of the NA-TP in the tissues susceptible to infection [135,136].
TheIV-administered remdesivir had unbound Cmax/EC50 ratios of 2.2, 22, and 0.14 in Calu-2, Caco-2, and Vero CCL-81 cells, respectively. Remdesivir was nontoxic in Vero CCL-81, Calu-3, and Caco-2 cells (CC50 > 100 μM), but PBM cells, CEM, and Huh-7 cells were found to be sensitive (CC50 = 3.7, 10.1 and 1.9, μM, respectively). Remdesivir is a chiral phosphoramidate protide of GS-441524 which increases the lipophilicity of the drug and enhances its ability to penetrate lipid cell membranes via diffusion [137]. Intracellular esterase-1 (CES1), cathepsin-A (CTSA), and the phosphoramidase HNT1 then convert it to GS-441524 monophosphate (MP), thereby bypassing the rate limiting initial phosphorylation to GS-441524-MP. GS-441524-MP is rapidly phosphorylated by cellular kinases to the active 5′-triphosphate form (GS-443902, GS-441524-TP), which acts as a competitive non-obligate chain terminator of RdRp during viral replication.
The differing EC50 and CC50 values between cell lines may result from elevated cellular levels of carboxylesterase 1 (CES1) and cathepsin A (CatA) in various cell lines. Although remdesivir is rapidly cleared from the plasma following IV infusion (plasma t1/2 ~ 1 h), it is administered once per day due to the cellular stability of GS-441524-TP (e.g., the t1/2 of GS-441524-TP in human blood mononuclear cells following an IV infusion over 2 h, in vivo, is about 36–43 h [138]). In contrast, the NTP of NHC has a cellular t1/2 of 4–5 h in human astrocytes and human bronchial tracheal epithelial cells (hBTEC), and is significantly less (0.4–1.1 h) in primary hepatocytes [139]. Oral administration of remdesivir is not feasible for treating SARS-CoV-2 lung infections, as the phosphoramidate protide is de-esterified during the first-pass metabolism by CES1, forming GS-441524, which is phosphorylated in hepatocytes before reaching the systemic circulation [140,141,142]. Studies in African Green monkeys revealed that <1% of orally administered remdesivir enters the systemic circulation intact [143]. Jubilant Pharma Ltd. reported on an oral formulation of remdesivir designed “to sidestep hepatic metabolism”. Safety and PK studies in animals and healthy human volunteers in India produced promising safety and absorption data, but no efficacy data were reported [144].
Orally administered derivatives of the parent NA GS-441524 are in various stages of development [142,145]. The EC50 of GS-441524 versus SARS-CoV-2 in Calu-3, Caco-2, and Vero CCL-81 cells were 0.3, 0.08, and 0.8 μM, respectively (Table 1). GS-441524 enters cells via nucleoside transporters, and the initial phosphorylation step forming GS-441524-MP is rate limiting, resulting in a lower cellular anti-SARS-CoV-2 activity for GS-441524 in comparison with remdesivir in primary lung cells and other cells [91,137,146]. Notably, orally administered GS-441524 improved survival and reduced lung inflammation and injury in a mouse model of SARS-CoV-2 infection [147]. However, the oral bioavailability of GS-441524 varies with species and is about 3% and 89% in Cynomolgus monkeys and dogs, respectively [145]. A PK and efficacy study of GS-441524 in an African Green monkey infection model of SARS-CoV-2 reported an oral bioavailability of <10% [148].
The unbound Cmax/EC50 ratios of IV-administered remdesivir in the Calu-3 cells were 2.2, versus 22 for the orally administered molnupiravir (Table 1). Taken in isolation, this would incorrectly imply that orally administered molnupiravir has a greater efficacy than IV-administered remdesivir. However, the EC50 values used in the calculation were measured in vitro after hours-long incubation, when the ratio of NAs in the medium to their respective intracellular NA-TPs were at a steady state and does not reflect differences in the plasma t1/2 (e.g., 0.5 h for remdesivir versus 3.3 h for NHC, and 6–7 h for GS-441524). It also does not reflect differences in the accumulation dynamics of the NA-TP, which may be 16-fold slower for GS-441524 than remdesivir [143]. Nor does it reflect differences in the cellular decay half-lives of the various NA-TPs. The combined effect(s) of these factors on the dynamics of NA-TP accumulation during treatment could be assessed, if needed, using more comprehensive pharmacometrics models [149,150].
The SARS-CoV-2 protease inhibitor, nirmatrelvir (coadministered with the Cyp-3A4 inhibitor ritonavir in PaxlovidTM), is also considered a potent antiviral as it has unbound Cmax/EC50 ratios of 0.4, 6.8, and 0.21 in the Calu-3, Caco-2, and Vero CC-L81 cells, respectively. Paxlovid administered in the early stages of infection prevented serious COVID-19 illness and hospitalization by >95% [151,152].

4.3. Drugs with Moderate Unbound Cmax/EC50 Ratios in Calu-3 and/or Caco-2, and Vero CCL-81 Cells

Nitazoxanide is a broad-spectrum thiazolide antiparasitic agent, approved for the treatment of Cryptosporidium parvum and Giardia duodenalis infections in children aged ≥ 1 year and in adults [93]. It had an unbound Cmax/EC50 ratio of 0.25 in Vero CCL-81 cells, suggesting a weak potential for in vivo efficacy. However, the unbound Cmax/EC50 ratios of nitazoxanide in the human Calu-3 and Caco-2 cells were 0.015 and 0.078, respectively (Table 1), suggestive of a low potential for in vivo efficacy. The EC50 in the Caco-2 cells (4.4 μM) was less than 2-fold lower than the CC50 versus the Caco-2, Vero, and PBM cells (Table 2), suggestive of a potential toxicity at effective antiviral concentrations. Nitazoxanide is rapidly metabolized to its active metabolite, tizoxanide, and displays in vitro antiviral activity against a range of viruses, including the influenza viruses, hepatitis B and C viruses, norovirus, rotavirus, Ebola virus, Middle East respiratory syndrome coronavirus (MERS-CoV), and SARS-CoV-2. Nitazoxanide inhibits host enzymes, which impairs the post-translational processing of viral proteins. It also inhibits pro-inflammatory cytokines.
Due to its reported in vitro inhibition of SARS-CoV-2 replication and favorable PK of nitazoxanide, the COVER randomized trial, conducted in Johannesburg, South Africa, evaluated whether nitazoxanide or a combination of sofosbuvir plus daclatasvir could lower the risk of SARS-CoV-2 infection in 828 healthcare workers and others at high risk of infection. Neither nitazoxanide nor sofosbuvir plus daclatasvir significantly prevented SARS-CoV-2 infection in healthcare workers and others at high risk of infection [153]. The COVID-19 Treatment Guidelines Panel of the NIH recommends against the use of nitazoxanide for the treatment of COVID-19, except in a clinical trial [154]. Daclatasvir, an inhibitor of the HCV nonstructural protein, NS5A, had unbound Cmax/EC50 ratios of 0.03 and 0.0005 in Vero CCL-81 and Caco-2 cells, respectively. However, it was inactive in the Calu-3 cells (EC50 > 100 μM). In the same clinical trial, COVER, a combination of sofosbuvir with daclatasvir did not demonstrate protect against infection compared to a no-intervention control arm of the study [155].
Ivermectin had unbound Cmax/EC50 ratios equal to 0.003, 0.005, and 0.01 in Calu-3, Caco-2, and Vero CCL-81 cells, respectively, suggestive of a marginal antiviral effect at clinical doses. The unbound Cmax/EC50 ratio in the Vero CCL-81 cells is similar to the ratio reported by Peña-Silva et al. in Vero E6 cells. Their report concluded that ivermectin is unlikely to be clinically effective for the prevention or treatment of COVID-19 [72]. A multicenter open-label, randomized, controlled adaptive platform pharmacometrics trial of high-dose ivermectin in individuals with early COVID-19 (PLATCOV) showed that ivermectin does not increase the rate of SARS-CoV-2 clearance from the oral compartment [156]. The current NIH COVID-19 Treatment Guidelines do not recommend the use of ivermectin for the prevention or treatment of COVID-19 [157].
Favipiravir was inactive in Calu-3 cells (EC50 > 300 µM) but had activity in Caco-2 at high concentrations (EC50 = 256 µM, unbound Cmax/EC50 = 0.74) and in Vero CCL-81 cells (EC50 = 90.6 µM, unbound Cmax/EC50 = 2.1). It is a pyrazine carboxamide nucleosidase precursor, which is metabolized intracellularly to its active form, favipiravir-ribofuranosyl-5′-triphosphate, and may inhibit viral replication by inducing lethal mutations in the viral genome, in a mechanism similar to NHC [158]. Favipiravir (AVIGAN, T705 tablets, Taisho Toyama Pharmaceutical Co., Ltd, a Subsidiary of Fujifilm, Tokyo, Japan) was approved in Japan in 2014 for stockpiling for pandemic preparedness only, but not yet for the treatment of seasonal influenza. Moreover, it was marketed in China as a second-line treatment of novel or reemerging influenza outbreaks. However, favipiravir has poor antiviral activity in primary human respiratory cells, in agreement with the lack of activity observed in the human Calu-3 cells (Table 1) [157]. The drug is administered at high doses to treat influenza, as it has low to moderate antiviral efficacy at lower dosages [86]. Four separate mammalian species administered with favipiravir at doses equivalent to human regimens demonstrated delayed development or embryonic death in the first trimester, so this drug is not recommended for use in pregnant women or children [158]. A systematic review and meta-analysis of 12 clinical trials, involving 1636 patients with moderate to severe COVID-19, did not reveal significant differences in fatality and the requirement of mechanical ventilation between individuals treated with favipiravir and with standard of care [159].
Imatinib, a small molecule inhibitor targeting multiple tyrosine kinases, used to treat chronic myelogenous leukemia (CML) and acute lymphocytic leukemia (ALL), was weakly active in the Calu-3 (EC50 = 77.9 µM) and Caco-2 (37.9 µM) cells and more active in Vero CCL-81 cells (4.6 µM), with unbound Cmax/EC50 ratios of 0.003, 0.006, and 0.056, respectively [64,99]. A randomized, double-blind, placebo-controlled clinical trial conducted in 385 adult patients with severe COVID-19 infections did not meet its primary outcome of reducing the time needed to remain on a ventilator and the need for supplemental oxygen for more than 48 consecutive hours in patients requiring supplemental oxygen [64].
Celecoxib, a COX-2 enzyme inhibitor used for the treatment of rheumatoid arthritis, had unbound Cmax/EC50 ratios of 0.019, 0.08, and 0.008 in the Calu-3, Caco-2, and Vero CCL-81 cells, respectively, suggesting weak antiviral activity. However, since pro-inflammatory mediators, such as COX-2, p38 MAPK, IL-1b, IL-6, and TGF-β, play pivotal roles in coronavirus-related cell death, cytokine storm, and pulmonary interstitial fibrosis, blocking these mediators of inflammation with celecoxib could be beneficial in treating individuals with mild to moderate infections [43,44].
Apilimod is a specific inhibitor of PIKfyve kinase, a lipid enzyme which is also involved in the transport of filoviruses, including the Ebola virus and Marburg virus, into cells, and was originally developed to treat rheumatoid arthritis and Crohn’s disease [160,161]. Apilimod had unbound Cmax/EC50 ratios of 0.062 and 0.0005 in the Vero CCL-81 and Caco-2 cells, respectively, but was inactive in the Calu-3 cells. Apilimod inhibited SARS-CoV-2 replication in human pneumocyte-like cells derived from induced pluripotent stem cells, and in a primary human lung explant model, but worsened disease in a COVID-19 murine model [144,162,163]. The binding of apilimod in plasma was not previously reported; however, a computer analysis based on chemical structure (Section 2.4) estimated an unbound fraction of 0.062, which suggests that the discordance between in vitro potency and in vivo efficacy may result from binding to plasma proteins in vivo. AI Therapeutics sponsored a randomized, double-blind, placebo-controlled clinical trial in individuals to test the safety, tolerability, and efficacy of orally administered apilimod for the prevention of COVID-19 progression [164]. The trial was completed, but as of now the results have not been published. Baranov et al. raised a concern that since apilimod might prevent viral invasion by inhibiting host cell proteases, the same proteases that are critical for the antigen presentation leading to T-cell activation, and thus this could dampen antibody production. In addition, there is evidence from both in vitro studies and the clinic that apilimod blocks antiviral immune responses [165].
Zileuton had weak predicted efficacy in the Vero (unbound Cmax/EC50 = 0.02) and Caco-2 cells (Cmax/EC50 = 0.003) and was inactive in the Calu-3 cells (EC50 > 100 µM). Although leukotriene inhibitors are unlikely to inhibit SARS-CoV-2 replication, they may inhibit SARS-CoV-2-related inflammation [166]. A retrospective study reported that leukotriene inhibitors in combination with dexamethasone provided a mortality benefit in hospitalized patients with COVID-19 presenting with an oxygen saturation of below 50%. A cohort of patients in that study that received leukotriene inhibitors without dexamethasone had lower markers of inflammation and reduced cytokine storm [167].
Ebselen had an unbound Cmax/EC50 ratio of 0.012 in the Vero CCL-81 cells, 0.002 in the Calu-3 cells, and was inactive in the Caco-2 cells (EC50 > 100 µM). Ebselen is a synthetic organoselenium compound which mimics the action of glutathione peroxidase and peroxiredoxin enzymes [168]. Ebselen forms selenosulfide bonds with thiols, which results in antiviral, antibacterial, and anti-inflammatory effects [105,169]. It has been proposed that the main protease (Mpro) of SARS-CoV-2 is a potential drug target of ebselen [55]. A clinical trial has been enrolled with orally administered ebselen (SPI-1005), sponsored by Sound Pharmaceuticals Inc. [170].

4.4. Drugs with Measurable EC50 in Caco-2 Cells and EC50 > 100 μM in Calu-3 and Vero CCL-81 Cells

Fluvoxamine had an EC50 of >100 in the Vero CCL-81 and Calu-3 cells, and EC50 of 17.9 μM (unbound Cmax/EC50 ratio = 0.0005) in the Caco-2 cells. Fluvoxamine is a selective serotonin reuptake inhibitor [171] approved by the FDA for the treatment of obsessive-compulsive disorder and is widely used for other conditions, including depression, but is not approved for the treatment of any infection. Fluoxetine, a related drug, was reported to block SARS-CoV-2 replication efficiently ex vivo in human lung tissue [172]. An observational multicenter retrospective study with fluoxetine, including 7230 adults hospitalized for COVID-19, reported a significant association between antidepressant use and the reduced risk of intubation or death (p < 0.001), suggesting that anti-depressant drug use could be associated with lower risk of death or intubation in patients hospitalized for COVID-19 [171]. A retrospective cohort study using a database containing 83,584 patients with COVID-19 in 87 healthcare centers across the US, which included 3401 adult patients with COVID-19 who were prescribed either fluoxetine hydrochloride or fluvoxamine maleate (SSRIs), also reported a significant association between fluvoxamine or fluoxetine use and the reduced risk of death compared to a matched control, which was not observed in outpatients administered with other SSRIs [173]. Guo et al. performed a fixed effects and sensitivity meta-analysis on the combined data (2196 patients) from three clinical trials (STOP COVID 1 and 2, and the TOGETHER Trial) to assess the use of fluvoxamine during the early stage of a COVID-19 infection for reducing the risk of hospitalization, relative to the control group. The study concluded that although patients receiving fluvoxamine were 31% less likely to exhibit clinical deterioration or hospitalization compared with the placebo, more evidence from future trials is warranted to support this finding. The COVID-19 Treatment Guidelines Panel stated that there is presently insufficient evidence to recommend either for or against the use of fluvoxamine for the treatment of COVID-19 [174].

4.5. Drugs with Measurable Unbound Cmax/EC50 Ratios in Vero Cells Only

Fenofibrate is used to treat dyslipidemia [175]. Its lipid-modifying effects are mediated by the activation of peroxisome proliferator-activated receptor-α. It also reduces the levels of fibrinogen, C-reactive protein, and various pro-inflammatory markers [176]. The fenofibric acid metabolite was shown to destabilize the receptor binding domain of SARS-CoV-2 and to reduce virus replication in the Vero cells (EC50 = 7 μM) [177]. In our assays, fenofibrate was weakly active in the Vero CCL-81 cells (EC50 = 42.7 μM, Cmax/EC50 ratio = 0.007), and was inactive in the Calu-3 and Caco-2 cells. A multinational double-blinded trial was conducted in 701 inpatients and nonhospitalized patients with COVID-19 within 14 days of symptom onset, who were randomized to receive 145 mg of an oral fenofibrate formulation qd (n = 351) versus placebo (n = 350) for 10 days (NCT04517396) [67]. No significant differences were observed between the arms of the study using a “symptoms severity score” metric, which considered time to death, duration of mechanical ventilation, oxygenation, hospitalization, and symptom severity (primary endpoint). Also, no differences were observed in all-cause death across the arms’ secondary and exploratory endpoints.
Mefenamic acid had an unbound Cmax/EC50 ratio of 0.018 in the Vero CCL-81 cells and was inactive in the Caco-2 and Calu-3 cells (EC50 > 100 µM). A randomized double-blind placebo-controlled trial in 38 ambulatory individuals with COVID-19 reported that patients receiving mefenamic acid with standard medical care had an about 16-fold higher probability of achieving patient-acceptable symptoms on day 8 of treatment, compared with those receiving a placebo plus standard medical care, and had reduced symptomatology [178]. Since mefenamic acid is a non-steroidal anti-inflammatory agent, it was unclear whether the clinical benefit was due to modulation of inflammation or other mechanisms.

4.6. Drugs with EC50 > 100 μM in Calu-3, Caco-2, and Vero CCL-81 Cells

Kinetin (MB-905, N6-furfurylaminopurine) previously proved ineffective for the treatment of patients with familial dysautonomia [84,179]. Kinetin was reported to inhibit SARS-CoV-2 replication in vitro at sub-micromolar concentrations in human hepatic and pulmonary cell lines, and to reduce viral replication, IL-6, and TNF levels in infected monocytes [180]. Kinetin riboside 5′-TP) inhibits the SARS-CoV-2 RNA polymerase with an IC50 three-fold higher than that of GS-443902-TP (the active cellular metabolite of remdesivir). Kinetin was reported to produce a similar error prone catastrophic replication of the exonuclease of the viral RdRp as molnupiravir, but had negative Ames and micronucleus tests, suggesting a low potential for mutagenesis. The kinetics of cellular accumulation and stability of kinetin riboside 5′-TP were not reported in that study. Kinetin had a 50% oral bioavailability and exhibited satisfactory plasma PK in mice and rats and was ~80% bound to plasma proteins SARS-CoV-2-infected transgenic mice expressing human ACE2 treated with an oral dose of kinetin (140 mg/kg per day) exhibited a decrease in viral replication of the gamma variant, and reduced lung necrosis, hemorrhage and inflammation, and increased survival at plasma concentrations lower than those achieved and shown to be safe in clinical trials of kinetin in patients with familial dysautonomia [179,180]. Paradoxically, in our assays, kinetin did not inhibit SARS-CoV-2 and was not cytotoxic up to 100 µM. However, significant differences in the experimental design and use of the gamma variant by Souza et al. (we employed the Washington strain) could potentially explain these differences.
Metformin, a biguanide molecule, is the main first-line drug for the treatment of type 2 diabetes, particularly in overweight individuals [181]. Metformin was reported to inhibit SARS-CoV-2 replication in cell culture and in human lung ex vivo at physiologically relevant concentrations [182,183,184]. However, in our assays, we did not observe an EC50 or cytotoxicity at concentrations up to 100 μM (10-fold higher than the plasma Cmax (bound + unbound) = 10.2 μM) [185]. Retrospective studies reported an association between metformin use and less severe COVID-19 in patients already administering metformin [186,187,188]. The COVID-OUT trial was a Phase 3, randomized, placebo-controlled trial conducted in the United States from 21 May 2021 to 28 January 2022, which tested the effectiveness of metformin, ivermectin, and fluvoxamine in preventing serious SARS-CoV-2 infection in non-hospitalized adults. The study found no reduction in hypoxemia, emergency room visits, hospitalization, or death associated with any of the three drugs [189]. A secondary analysis of the data suggested that metformin may reduce a composite of emergency room visit, hospitalization, or death in overweight or obese individuals with COVID-19, and that further studies with metformin may be warranted in a similar population. The trial was lengthened to study the effects of these drugs on long COVID, a heterogeneous group, ranging from a single symptom to serious multi-organ involvement, and from mild and short-lived to chronically debilitating [190,191]. The study concluded that outpatient treatment with metformin reduced the incidence of long COVID by about 41%. The absolute reduction in long COVID incidence was 4.1%, compared with the placebo, suggesting only a marginal benefit [192]. The exact pathophysiology of long COVID is unknown, but it is likely to be multifactorial, including the inflammatory cascade during an acute infection and persistent viral replication [193].

4.7. Drugs with EC50 < 100 μM in Vero-CCL-81, Calu-3, and Caco-2 Cells but without Published Human PK Data

Honokiol is a natural polyphenolic compound extracted from the bark and leaves of the magnolia grandiflora, and is used in traditional medicine for treating a variety of ailments (e.g., malignancies, neurologic diseases, muscle spasms, depression, thromboses, microbial infections, and others) [194]. Tanikawa et al. reported that honokiol partially inhibited the furin-like enzymes responsible for cleaving a motif on the S1/S2 boundary of the spike (S) protein of SARS-CoV-2, and inhibited SARS-CoV-2 infectivity in Vero E6 cells with EC50 = 13 μM, and CC50 = 54 μM [195]. They concluded that honokiol and crude drugs which contain honokiol may benefit individuals with COVID-19. In our assays, the EC50 and unbound CC50 values of honokiol were as follows: 20.1 versus 37.8 μM in the Vero CCL-81 cells; 69.2 versus 27.5 μM in the Calu-3 cells; and 9.6 versus 18.4 μM in the Caco-2. The unbound CC50 values in human PBM, CEM and Huh7 cells were 26.5, 26.7, and 12.4 μM, respectively, indicative of toxicity at concentrations similar to those needed to inhibit SARS-CoV-2 replication. The PK of honokiol in humans remains unpublished. However, studies in rats and mice indicate that oral absorption is limited by poor solubility, intestinal metabolism during absorption, and first-pass hepatic metabolism. Magnolol, a similar compound (2,2-diol versus 2,4-diol) extracted from the bark and leaves of the magnolia grandiflora, has a 4% oral bioavailability in rats [196]. These data suggest that honokiol is unlikely to benefit individuals with COVID-19.

4.8. Compounds for Intranasal Administration to Protect against SARS-CoV-2 Infection

Carrageenans are naturally occurring, linear, sulfated polysaccharides extracted from red edible seaweeds, and they are widely used in the food industry as gelling, thickening, and stabilizing agents [197]. Iota-carrageenan was previously reported to inhibit SARS-CoV-2 replication in Vero cells [198]. In our assays, iota-carrageenan had an EC50 of 0.7, 0.5, and 59.2 μM in the Vero CCL-81, Calu-3, and Caco-2 cells, respectively. It had a CC50 of 95.9 μM in the Vero CCL-81 cells and >100 μM in the other cell systems tested. Previous animal experiments, which included repeated dose, local tolerance, and toxicity studies with intranasally applied 0.12% iota-carrageenan for 7 or 28 days in New Zealand White rabbits, as well as nebulized 0.12% iota-carrageenan administered to F344 rats for 7 days, revealed no penetration of iota-carrageenan across the nasal mucosa into the blood stream. The data do not provide any evidence for local intolerance or toxicity when carrageenan is applied intranasally or by inhalation. No signs for immunogenicity or immunotoxicity have been observed in the in vivo studies [199]. However, a pilot study of a topical nasal spray containing iota-carrageenan, in 300 hospital personnel dedicated to caring for patients with COVID-19 disease, demonstrated the following relative risk reduction: 79.8% (95% CI 5.3 to 95.4; p = 0.03). However, the absolute risk reduction was 4% (95% CI 0.6 to 7.4), suggesting a marginal benefit in this population [200].

5. Conclusions

There is a need for additional safe, effective, and inexpensive medications for the treatment or prevention of SARS-CoV-2 infection, and drug repurposing may be an option to identify them. We evaluated a wide range of drugs which were considered for repurposing. However, drugs that were not specifically developed as SARS-CoV-2 antiviral agents had the marginal ability to inhibit the replication of this virus at clinically relevant concentrations. Vero cells are highly permissive to this virus and thus widely used in the laboratory for antiviral drug evaluation. However, the data presented herein demonstrated that the potency and toxicities of some drugs, especially NA inhibitors were cell-system dependent. This was not unexpected since different cell types may have unequal distributions of drug metabolizing enzymes and membrane transporters [201], and certain antiviral agents (e.g., remdesivir and nirmatrelvir) are substrates of the Cytochrome P450 enzymes, P-glycoprotein, and other cell membrane transporters [202,203]. Also, NA phosphorylation may be cell-dependent [138,139]. It is informative to relate in vitro potency to reported in vivo plasma concentrations. The underlying assumption being that the concentration of the active (presumed unbound) form of the direct-acting antivirals (or the intracellular NTP) remains above the EC50 at the infection site for a sufficient fraction of the dose interval to inhibit viral replication. Many drugs that proved ineffective in the COVID-19 clinical trials exhibited an unbound peak plasma concentration (unbound Cmax) considerably less than the EC50 in the Calu-3 and Caco-2 lung-derived cells. Antiviral agents effective in controlled clinical trials had unbound Cmax/EC50 ≥ 6.8 in Calu-3 or Caco-2 cells. More comprehensive PK analysis and modeling is warranted for the drugs exhibiting higher unbound Cmax/EC50 ratios to assess their potential duration of efficacy. This could be achieved by superimposing a plot of simulated drug concentrations versus time profiles in plasma, generated using a PK model, with EC50 values in the physiologically relevant cells normalized for protein binding [204]. PK models may be expanded to include in vitro derived phosphorylation kinetics of NA (using a relevant cell line) to generate NTP versus time profiles, versus dose and time. The NTP concentration profiles can, in turn, be compared in vitro with the EC50, estimating the NTP concentration versus that of the viral polymerase [151,152]. Since SARS-CoV-2 infects other more shielded tissues (e.g., the central nervous system), drug exposures in other, more shielded organ sites (e.g., from animal studies) should be also considered [205].
The literature summary presented for many drugs suggests that they could be considered as possible adjunct therapies in combination with drugs known to inhibit SARS-CoV-2 replication to treat COVID-19-related symptoms in mild to moderate infections. Karim and Devranim commented in a recent editorial concerning the failure of the COVID-OUT trial on the use of oral metformin, ivermectin, or fluvoxamine to protect individuals with mild to moderate COVID-19 from severe disease (above) [206]. They concluded the following: “With respect to clinical decisions about COVID-19 treatment, some drug choices, especially those that have negative WHO recommendations, are clearly wrong. In keeping with evidence-based medical practice, patients with COVID-19 must be treated with efficacious medications; they deserve nothing less”. Similarly, an editorial by Dr. Abrescia, concerning a study which showed a lack of association between antiretroviral use and the acquisition or severe outcomes of SARS-CoV-2 infection in people with HIV in the Netherlands, recommended the following: “If SARS-CoV-2 infection does take place, drugs proven to be effective against SARS-CoV-2 (remdesivir or nirmatrelvir/ritonavir) should be started within five days of the onset of symptoms” [207,208]. Our calculations of the unbound Cmax/EC50 ratios and EC50 measurements in the cell systems relevant to in vivo SARS-CoV-2 infection lends further support to that conclusion. Many drugs may have potential in alleviating the symptoms of COVID-19 and/or in improving the outcomes of SARS-CoV-2 infection, despite having minimal or marginal direct antiviral effects. These should be tested as adjuncts during therapy with potent antiviral agents.

Author Contributions

Conceptualization, R.F.S., K.Z., S.J.H., R.M.T. and V.S.; Methodology, R.F.S., K.Z., S.J.H., J.C.L. and R.M.T.; Formal Analysis, S.J.H., K.Z. and R.F.S.; Investigation, K.Z., R.D., J.A.D.-B., S.L.G., T.M., M.B., B.S.D. and D.P.; Data Curation, S.J.H., F.A., J.J.K., F.A. and R.F.S.; Writing—Original Draft Preparation, S.J.H., J.J.K., F.A. and R.F.S.; Writing S.J.H., K.Z. and R.F.S.; Review and Editing, R.F.S., S.J.H., J.J.K., K.Z., F.A., R.M.T. and V.S.; Visualization, S.J.H.; Supervision, R.F.S. and T.M.; Project Administration, R.F.S.; Funding Acquisition, R.F.S., R.M.T. and V.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded in part by NIH grants 1RO1-AI-161570 NIH/NIAID (Kovari/Fitzpatrick/Schinazi), and RO1-MH-116695 NIH/NIAID (Schinazi/Tyor). We also thank the Woodruff Health Sciences 2020 COVID-19 CURE award (Tirouvanziam/Schinazi/Sukhatme), Imagine, Innovate, and Impact (I3) Team Award Emory University School of Medicine, and the Center for AIDS Research grant P30-AI-050409 NIH/NIAID (Emory University).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

All pertinent data are reported in the manuscript.

Acknowledgments

The Coronavirus 2, Isolate hCoV-19/USA-WA1/2020, NR-52281 was deposited by the Centers for Disease Control and Prevention and obtained through BEI Resources, NIAID, NIH: SARS-Related Coronavirus 2, Isolate hCoV-19/USA-WA1/2020, NR-52281.

Conflicts of Interest

R.F.S. has received royalties from Ely Lilly from the sales of baricitinib, used for the treatment of COVID-19. RFS received research funding from Pfizer for his group at Emory University. Emory University has reviewed and approved his conflict of interest. Other authors have no conflicts of interest to declare regarding this manuscript.

Abbreviations

Bidtwice per day
Cmaxpeak concentration of drug observed in plasma
CC50drug concentration that inhibits cell division by 50% in vitro
DMEMDulbecco’s Modified Eagle Medium
EC50/EC90median/90th percentile effective antiviral concentration measured in cell culture
FBSfetal bovine serum
FDAUS Federal Drug administration
Fununbound fraction of drug in plasma
HHour
LC-MS-MSchemical analysis which couples liquid chromatography with mass spectrometry
MEMMinimum Essential Media
MOImultiplicity of infection (MOI)
MTS3-(4,5-Dimethylthiazol-2-yl)-5-(3-carboxymethoxyphenyl)-2-(4-sulfophenyl)-2H-tetrazolium
NAnucleoside analog
NEnot estimated
NIHUS National Institutes of Health
PDPharmacodynamics
PKpharmacokinetics
Qdonce per day
t1/2elimination half-life
Tidthree times per day
qRT-PCR quantitative real-time PCR
RPMIRoswell Park Memorial Institute 1640 Media
unbound Cmax/EC50ratio of Cmax/EC50 normalized for protein binding.

References

  1. U.S. Centers for Disease Control and Prevention. Overview of COVID-19 Vaccines. Available online: https://www.cdc.gov/coronavirus/2019-ncov/vaccines/different-vaccines/overview-COVID-19-vaccines.html (accessed on 28 November 2023).
  2. Yap, C.; Ali, A.; Prabhakar, A.; Prabhakar, A.; Pal, A.; Lim, Y.Y.; Kakodkar, P. Comprehensive literature review on COVID-19 vaccines and role of SARS-CoV-2 variants in the pandemic. Ther. Adv. Vaccines Immunother. 2021, 9, 25151355211059791. [Google Scholar] [CrossRef] [PubMed]
  3. Fan, Y.J.; Chan, K.H.; Hung, I.F. Safety and efficacy of COVID-19 vaccines: A systematic review and meta-analysis of different vaccines at phase 3. Vaccines 2021, 9, 989. [Google Scholar] [CrossRef] [PubMed]
  4. Narayanan, A.; Narwal, M.; Majowicz, S.A.; Varricchio, C.; Toner, S.A.; Ballatore, C.; Brancale, A.; Murakami, K.S.; Jose, J. Identification of SARS-CoV-2 inhibitors targeting Mpro and PLpro using in-cell-protease assay. Commun. Biol. 2022, 5, 169. [Google Scholar] [CrossRef]
  5. Malin, J.J.; Suarez, I.; Priesner, V.; Fatkenheuer, G.; Rybniker, J. Remdesivir against COVID-19 and other viral diseases. Clin. Microbiol. Rev. 2020, 34, e00162-20. [Google Scholar] [CrossRef] [PubMed]
  6. 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. [Google Scholar] [CrossRef] [PubMed]
  7. Chokkalingam, A.P.; Hayden, J.; Goldman, J.D.; Li, H.; Asubonteng, J.; Mozaffari, E.; Bush, C.; Wang, J.R.; Kong, A.; Osinusi, A.O.; et al. Association of remdesivir treatment with mortality among hospitalized adults with COVID-19 in the United States. JAMA Netw. Open 2022, 5, e2244505. [Google Scholar] [CrossRef] [PubMed]
  8. Mozaffari, E.; Chandak, A.; Zhang, Z.; Liang, S.; Thrun, M.; Gottlieb, R.L.; Kuritzkes, D.R.; Sax, P.E.; Wohl, D.A.; Casciano, R.; et al. Remdesivir treatment in hospitalized patients with coronavirus disease 2019 (COVID-19): A comparative analysis of in-hospital all-cause mortality in a large multicenter observational cohort. Clin. Infect. Dis. 2022, 75, e450–e458. [Google Scholar] [CrossRef]
  9. 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. [Google Scholar] [CrossRef] [PubMed]
  10. Gilead Sciences, Inc. VEKLURY® (Remdesivir) for Injection, for Intravenous Use: Highlights of Prescribing Information. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2022/214787Orig1s015lbl.pdf (accessed on 18 December 2022).
  11. U.S. National Institutes of Health. COVID-19 Treatment Guidelines. Remdesivir. Available online: https://www.covid19treatmentguidelines.nih.gov/therapies/antivirals-including-antibody-products/remdesivir/ (accessed on 29 February 2024).
  12. Pfizer Inc. FDA Advisory Committee Votes in Support of Favorable Benefit-Risk Profile for Pfizer’s Paxlovid™. Available online: https://www.pfizer.com/news/press-release/press-release-detail/fda-advisory-committee-votes-support-favorable-benefit-risk (accessed on 28 March 2023).
  13. Parums, D.V. Editorial: Current status of oral antiviral drug treatments for SARS-CoV-2 infection in non-hospitalized patients. Med. Sci. Monit. 2022, 28, e935952. [Google Scholar] [CrossRef] [PubMed]
  14. National Institutes of Health. COVID-19 Treatment Guidelines—Ritonavir -Boosted Nirmatrelvir. Available online: https://www.covid19treatmentguidelines.nih.gov/therapies/antivirals-including-antibody-products/ritonavir-boosted-nirmatrelvir--paxlovid-/ (accessed on 5 April 2024).
  15. Pfizer Inc. Pfizer Announces Additional Phase 2/3 Study Results Confirming Robust Efficacy of Novel COVID-19 Oral Antiviral Treatment Candidate in Reducing Risk of Hospitalization or Death. Available online: https://www.pfizer.com/news/press-release/press-release-detail/pfizer-announces-additional-phase-23-study-results (accessed on 15 December 2021).
  16. U.S. Centers for Disease Control and Prevention. Underlying Medical Conditions Associated with Higher Risk for Severe COVID-19: Information for Healthcare Professionals. Available online: https://www.cdc.gov/coronavirus/2019-ncov/hcp/clinical-care/underlyingconditions.html (accessed on 28 November 2023).
  17. Merck & Co. Inc. Fact Sheet for Healthcare Providers: Emergency Use Authorization for Lagevrio™ (Molnupiravir) Capsules. Available online: https://www.merck.com/eua/molnupiravir-hcp-fact-sheet.pdf (accessed on 6 March 2023).
  18. Stuyver, L.J.; Whitaker, T.; McBrayer, T.R.; Hernandez-Santiago, B.I.; Lostia, S.; Tharnish, P.M.; Ramesh, M.; Chu, C.K.; Jordan, R.; Shi, J.; et al. Ribonucleoside analogue that blocks replication of bovine viral diarrhea and hepatitis C viruses in culture. Antimicrob. Agents Chemother. 2003, 47, 244–254. [Google Scholar] [CrossRef] [PubMed]
  19. Hollecker, L.; Choo, H.; Chong, Y.; Chu, C.K.; Lostia, S.; McBrayer, T.R.; Stuyver, L.J.; Mason, J.C.; Du, J.; Rachakonda, S.; et al. Synthesis of beta-enantiomers of N4-hydroxy-3′-deoxypyrimidine nucleosides and their evaluation against bovine viral diarrhoea virus and hepatitis C virus in cell culture. Antivir. Chem. Chemother. 2004, 15, 43–55. [Google Scholar] [CrossRef] [PubMed]
  20. Costantini, V.P.; Whitaker, T.; Barclay, L.; Lee, D.; McBrayer, T.R.; Schinazi, R.F.; Vinjé, J. Antiviral activity of nucleoside analogues against norovirus. Antivir. Ther. 2012, 17, 981–991. [Google Scholar] [CrossRef] [PubMed]
  21. Ehteshami, M.; Tao, S.; Zandi, K.; Hsiao, H.M.; Jiang, Y.; Hammond, E.; Amblard, F.; Russell, O.O.; Merits, A.; Schinazi, R.F. Characterization of β-d-N(4)-hydroxycytidine as a novel inhibitor of Chikungunya virus. Antimicrob. Agents Chemother. 2017, 61, e02395-16. [Google Scholar] [CrossRef] [PubMed]
  22. Painter, G.R.; Natchus, M.G.; Cohen, O.; Holman, W.; Painter, W.P. Developing a direct acting, orally available antiviral agent in a pandemic: The evolution of molnupiravir as a potential treatment for COVID-19. Curr. Opin. Virol. 2021, 50, 17–22. [Google Scholar] [CrossRef] [PubMed]
  23. Swanstrom, R.; Schinazi, R.F. Lethal mutagenesis as an antiviral strategy. Science 2022, 375, 497–498. [Google Scholar] [CrossRef] [PubMed]
  24. Zhou, S.; Hill, C.S.; Sarkar, S.; Tse, L.V.; Woodburn, B.M.D.; Schinazi, R.F.; Sheahan, T.P.; Baric, R.S.; Heise, M.T.; Swanstrom, R. β-d-N4-hydroxycytidine inhibits SARS-CoV-2 through lethal mutagenesis but is also mutagenic to mammalian cells. J. Infect. Dis. 2021, 224, 415–419. [Google Scholar] [CrossRef] [PubMed]
  25. Troth, S.; Butterton, J.; DeAnda, C.S.; Escobar, P.; Grobler, J.; Hazuda, D.; Painter, G. Letter to the editor in response to zhou et al. J. Infect. Dis. 2021, 224, 1442–1443. [Google Scholar] [CrossRef] [PubMed]
  26. U.S. National Institutes of Health. COVID-19 Treatment Guidelines—Molnupiravir. Available online: https://www.covid19treatmentguidelines.nih.gov/therapies/antivirals-including-antibody-products/molnupiravir/ (accessed on 26 June 2023).
  27. Mackman, R.L.; Kalla, R.; Babusis, D.; Pitts, J.; Barrett, K.T.; Chun, K.; Pont, V.D.; Rodriguez, L.; Moshiri, J.; Xu, Y.; et al. Discovery of GS-5245 (Obeldesivir), an oral prodrug of nucleoside GS-441524 that exhibits antiviral efficacy in SARS-CoV-2 infected african green monkeys. J. Med. Chem. 2023, 66, 11701–11717. [Google Scholar] [CrossRef] [PubMed]
  28. Anoshchenko, O.; Abdelghany, M.; Hyland, R.H.; Davies, S.; Mkay, C.; Shen, G.; Xiao, D.; Winter, H.; Llewellyn, J.; Humeniuk, R. Pharmacokinetics, safety, and tolerability of obeldesivir (OBV; GS-5245) in healthy participants. Poster # P260 Presented at the 33rd European Congress of Clinical Microbiology and Infectious Diseases (ECCMID), 15–18 April 2023. Available online: https://www.natap.org/2023/HIV/092523_03.htm (accessed on 5 April 2024).
  29. Stevens, L.J.; Pruijssers, A.J.; Lee, H.W.; Gordon, C.J.; Tchesnokov, E.P.; Gribble, J.; George, A.S.; Hughes, T.M.; Lu, X.; Li, J.; et al. Mutations in the SARS-CoV-2 RNA-dependent RNA polymerase confer resistance to remdesivir by distinct mechanisms. Sci. Transl. Med. 2022, 14, eabo0718. [Google Scholar] [CrossRef] [PubMed]
  30. Costacurta, F.; Dodaro, A.; Bante, D.; Schöppe, H.; Sprenger, B.; Moghadasi, S.A.; Fleischmann, J.; Pavan, M.; Bassani, D.; Menin, S.; et al. A comprehensive study of SARS-CoV-2 main protease (M(pro)) inhibitor-resistant mutants selected in a VSV-based system. bioRxiv 2023. [Google Scholar] [CrossRef]
  31. Gandhi, S.; Klein, J.; Robertson, A.J.; Peña-Hernández, M.A.; Lin, M.J.; Roychoudhury, P.; Lu, P.; Fournier, J.; Ferguson, D.; Mohamed Bakhash, S.A.K.; et al. De novo emergence of a remdesivir resistance mutation during treatment of persistent SARS-CoV-2 infection in an immunocompromised patient: A case report. Nat. Commun. 2022, 13, 1547. [Google Scholar] [CrossRef] [PubMed]
  32. Focosi, D.; Maggi, F.; McConnell, S.; Casadevall, A. Very low levels of remdesivir resistance in SARS-COV-2 genomes after 18 months of massive usage during the COVID19 pandemic: A GISAID exploratory analysis. Antivir. Res. 2022, 198, 105247. [Google Scholar] [CrossRef] [PubMed]
  33. Ip, J.D.; Wing-Ho Chu, A.; Chan, W.M.; Cheuk-Ying Leung, R.; Umer Abdullah, S.M.; Sun, Y.; Kai-Wang To, K. Global prevalence of SARS-CoV-2 3CL protease mutations associated with nirmatrelvir or ensitrelvir resistance. EBioMedicine 2023, 91, 104559. [Google Scholar] [CrossRef] [PubMed]
  34. Strizki, J.M.; Gaspar, J.M.; Howe, J.A.; Hutchins, B.; Mohri, H.; Nair, M.S.; Kinek, K.C.; McKenna, P.; Goh, S.L.; Murgolo, N. Molnupiravir maintains antiviral activity against SARS-CoV-2 variants and exhibits a high barrier to the development of resistance. Antimicrob. Agents Chemother. 2024, 68, e0095323. [Google Scholar] [CrossRef]
  35. Sanderson, T.; Hisner, R.; Donovan-Banfield, I.; Hartman, H.; Løchen, A.; Peacock, T.P.; Ruis, C. A molnupiravir-associated mutational signature in global SARS-CoV-2 genomes. Nature 2023, 623, 594–600. [Google Scholar] [CrossRef] [PubMed]
  36. Alavian, G.; Kolahdouzan, K.; Mortezazadeh, M.; Torabi, Z.S. Antiretrovirals for prophylaxis against COVID-19: A comprehensive literature review. J. Clin. Pharmacol. 2021, 61, 581–590. [Google Scholar] [CrossRef] [PubMed]
  37. Grieb, P.; Swiatkiewicz, M.; Prus, K.; Rejdak, K. Amantadine for COVID-19. J. Clin. Pharmacol. 2021, 61, 412–413. [Google Scholar] [CrossRef] [PubMed]
  38. Sharun, K.; Tiwari, R.; Dhama, K. Protease inhibitor GC376 for COVID-19: Lessons learned from feline infectious peritonitis. Ann. Med. Surg. 2021, 61, 122–125. [Google Scholar] [CrossRef] [PubMed]
  39. Hsu, C.K.; Chen, C.Y.; Chen, W.C.; Lai, C.C.; Hung, S.H.; Lin, W.T. The effect of sofosbuvir-based treatment on the clinical outcomes of patients with COVID-19: A systematic review and meta-analysis of randomized controlled trials. Int. J. Antimicrob. Agents 2022, 59, 106545. [Google Scholar] [CrossRef] [PubMed]
  40. Xu, Y.; Li, M.; Zhou, L.; Liu, D.; He, W.; Liang, W.; Sun, Q.; Sun, H.; Li, Y.; Liu, X. Ribavirin treatment for critically Ill COVID-19 patients: An observational study. Infect. Drug Resist. 2021, 14, 5287–5291. [Google Scholar] [CrossRef] [PubMed]
  41. RECOVERY Collaborative Group. Lopinavir-ritonavir in patients admitted to hospital with COVID-19 (RECOVERY): A randomised, controlled, open-label, platform trial. Lancet 2020, 396, 1345–1352. [Google Scholar] [CrossRef] [PubMed]
  42. Aherfi, S.; Pradines, B.; Devaux, C.; Honore, S.; Colson, P.; Scola, B.; Raoult, D. Drug repurposing against SARS-CoV-1, SARS-CoV-2 and MERS-CoV. Future Microbiol. 2021, 16, 1341–1370. [Google Scholar] [CrossRef] [PubMed]
  43. Baghaki, S.; Yalcin, C.E.; Baghaki, H.S.; Aydin, S.Y.; Daghan, B.; Yavuz, E. COX2 inhibition in the treatment of COVID-19: Review of literature to propose repositioning of celecoxib for randomized controlled studies. Int. J. Infect. Dis. 2020, 101, 29–32. [Google Scholar] [CrossRef]
  44. Ghaznavi, H.; Mohammadghasemipour, Z.; Shirvaliloo, M.; Momeni, M.K.; Metanat, M.; Gorgani, F.; Abedipour, F.; Mohammadi, M.; Sartipi, M.; Khorashad, A.R.S.; et al. Short-term celecoxib (celebrex) adjuvant therapy: A clinical trial study on COVID-19 patients. Inflammopharmacology 2022, 30, 1645–1657. [Google Scholar] [CrossRef] [PubMed]
  45. Cordeiro, L.P.; Linhares, E.; Nogueira, F.G.O.; Moreira-Silva, D.; Medeiros-Lima, D.J.M. Perspectives on glucocorticoid treatment for COVID-19: A systematic review. Pharmacol. Rep. 2021, 73, 728–735. [Google Scholar] [CrossRef] [PubMed]
  46. GómezMoore, N.; Bosco-Levy, P.; Thurin, N.; Blin, P.; Droz-Perroteau, C. NSAIDs and COVID-19: A Systematic Review and Meta-analysis. Drug Saf. 2021, 44, 929–938. [Google Scholar] [CrossRef]
  47. Kelleni, M.T. NSAIDs/nitazoxanide/azithromycin repurposed for COVID-19: Potential mitigation of the cytokine storm interleukin-6 amplifier via immunomodulatory effects. Expert. Rev. Anti Infect. Ther. 2022, 20, 17–21. [Google Scholar] [CrossRef] [PubMed]
  48. Vaja, R.; Chan, J.S.K.; Ferreira, P.; Harky, A.; Rogers, L.J.; Gashaw, H.H.; Kirkby, N.S.; Mitchell, J.A. The COVID-19 ibuprofen controversy: A systematic review of NSAIDs in adult acute lower respiratory tract infections. Br. J. Clin. Pharmacol. 2021, 87, 776–784. [Google Scholar] [CrossRef] [PubMed]
  49. U.S. Food & 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 (accessed on 24 February 2023).
  50. Chiu, L.; Shen, M.; Lo, C.H.; Chiu, N.; Chen, A.; Shin, H.J.; Prsic, E.H.; Hur, C.; Chow, R.; Lebwohl, B. Effect of famotidine on hospitalized patients with COVID-19: A systematic review and meta-analysis. PLoS ONE 2021, 16, e0259514. [Google Scholar] [CrossRef] [PubMed]
  51. Hogan Ii, R.B.; Hogan Iii, R.B.; Cannon, T.; Rappai, M.; Studdard, J.; Paul, D.; Dooley, T.P. Dual-histamine receptor blockade with cetirizine—Famotidine reduces pulmonary symptoms in COVID-19 patients. Pulm. Pharmacol. Ther. 2020, 63, 101942. [Google Scholar] [CrossRef]
  52. Kerget, B.; Kerget, F.; Aydin, M.; Karasahin, O. Effect of montelukast therapy on clinical course, pulmonary function, and mortality in patients with COVID-19. J. Med. Virol. 2021, 94, 1950–1958. [Google Scholar] [CrossRef] [PubMed]
  53. Thomas, S.; Patel, D.; Bittel, B.; Wolski, K.; Wang, Q.; Kumar, A.; Il’Giovine, Z.J.; Mehra, R.; McWilliams, C.; Nissen, S.E.; et al. Effect of High-Dose Zinc and Ascorbic Acid Supplementation vs Usual Care on Symptom Length and Reduction Among Ambulatory Patients With SARS-CoV-2 Infection: The COVID A to Z Randomized Clinical Trial. JAMA Netw. Open 2021, 4, e210369. [Google Scholar] [CrossRef] [PubMed]
  54. Sies, H.; Parnham, M.J. Potential therapeutic use of ebselen for COVID-19 and other respiratory viral infections. Free Radic. Biol. Med. 2020, 156, 107–112. [Google Scholar] [CrossRef] [PubMed]
  55. Tolouian, R.; Mulla, Z.D.; Jamaati, H.; Babamahmoodi, A.; Marjani, M.; Eskandari, R.; Dastan, F. Effect of bromhexine in hospitalized patients with COVID-19. J. Investig. Med. 2021, 71, 691–699. [Google Scholar] [CrossRef] [PubMed]
  56. Mikhaylov, E.N.; Lyubimtseva, T.A.; Vakhrushev, A.D.; Stepanov, D.; Lebedev, D.S.; Vasilieva, E.Y.; Konradi, A.O.; Shlyakhto, E.V. Bromhexine Hydrochloride Prophylaxis of COVID-19 for Medical Personnel: A Randomized Open-Label Study. Interdiscip. Perspect. Infect. Dis. 2022, 2022, 4693121. [Google Scholar] [CrossRef] [PubMed]
  57. Butler, C.C.; Yu, L.M.; Dorward, J.; Gbinigie, O.; Hayward, G.; Saville, B.R.; Van Hecke, O.; Berry, N.; Detry, M.A.; Saunders, C.; et al. Doxycycline for community treatment of suspected COVID-19 in people at high risk of adverse outcomes in the UK (PRINCIPLE): A randomised, controlled, open-label, adaptive platform trial. Lancet Respir. Med. 2021, 9, 1010–1020. [Google Scholar] [CrossRef] [PubMed]
  58. Egiz, A.; Gala, D. Clofazimine: Another potential magic bullet for the treatment of COVID-19? Postgrad. Med. J. 2021, 98, e124. [Google Scholar] [CrossRef] [PubMed]
  59. Blum, V.F.; Cimerman, S.; Hunter, J.R.; Tierno, P.; Lacerda, A.; Soeiro, A.; Cardoso, F.; Bellei, N.C.; Maricato, J.; Mantovani, N.; et al. Nitazoxanide superiority to placebo to treat moderate COVID-19—A Pilot prove of concept randomized double-blind clinical trial. EClinicalMedicine 2021, 37, 100981. [Google Scholar] [CrossRef] [PubMed]
  60. Hill, A.; Mirchandani, M.; Pilkington, V. Ivermectin for COVID-19: Addressing potential bias and medical fraud. Open Forum. Infect. Dis. 2022, 9, ofab645. [Google Scholar] [CrossRef] [PubMed]
  61. Izcovich, A.; Peiris, S.; Ragusa, M.; Tortosa, F.; Rada, G.; Aldighieri, S.; Reveiz, L. Bias as a source of inconsistency in ivermectin trials for COVID-19: A systematic review. Ivermectin’s suggested benefits are mainly based on potentially biased results. J. Clin. Epidemiol. 2021, 144, 43–55. [Google Scholar] [CrossRef] [PubMed]
  62. Feher, M.; Joy, M.; Munro, N.; Hinton, W.; Williams, J.; de Lusignan, S. Fenofibrate as a COVID-19 modifying drug: Laboratory success versus real-world reality. Atherosclerosis 2021, 339, 55–56. [Google Scholar] [CrossRef]
  63. Aman, J.; Duijvelaar, E.; Botros, L.; Kianzad, A.; Schippers, J.R.; Smeele, P.J.; Azhang, S.; Bartelink, I.H.; Bayoumy, A.A.; Bet, P.M.; et al. Imatinib in patients with severe COVID-19: A randomised, double-blind, placebo-controlled, clinical trial. Lancet Respir. Med. 2021, 9, 957–968. [Google Scholar] [CrossRef] [PubMed]
  64. Roschewski, M.; Lionakis, M.S.; Sharman, J.P.; Roswarski, J.; Goy, A.; Monticelli, M.A.; Roshon, M.; Wrzesinski, S.H.; Desai, J.V.; Zarakas, M.A.; et al. Inhibition of bruton tyrosine kinase in patients with severe COVID-19. Sci. Immunol. 2020, 5, eabd0110. [Google Scholar] [CrossRef] [PubMed]
  65. Mas, M.; García-Vicente, J.A.; Estrada-Gelonch, A.; Pérez-Mañá, C.; Papaseit, E.; Torrens, M.; Farré, M. Antidepressant drugs and COVID-19: A review of basic and clinical evidence. J. Clin. Med. 2022, 11, 4038. [Google Scholar] [CrossRef]
  66. Bramante, C.T.; Ingraham, N.E.; Murray, T.A.; Marmor, S.; Hovertsen, S.; Gronski, J.; McNeil, C.; Feng, R.; Guzman, G.; Abdelwahab, N.; et al. Metformin and risk of mortality in patients hospitalised with COVID-19: A retrospective cohort analysis. Lancet Healthy Longev. 2021, 2, e34–e41. [Google Scholar] [CrossRef]
  67. Chirinos, J.A.; Lopez-Jaramillo, P.; Giamarellos-Bourboulis, E.J.; Dávila-Del-Carpio, G.H.; Bizri, A.R.; Andrade-Villanueva, J.F.; Salman, O.; Cure-Cure, C.; Rosado-Santander, N.R.; Cornejo Giraldo, M.P.; et al. A randomized clinical trial of lipid metabolism modulation with fenofibrate for acute coronavirus disease 2019. Nat. Metab. 2022, 4, 1847–1857. [Google Scholar] [CrossRef] [PubMed]
  68. Lan, S.H.; Lee, H.Z.; Chao, C.M.; Chang, S.P.; Lu, L.C.; Lai, C.C. Efficacy of melatonin in the treatment of patients with COVID-19: A systematic review and meta-analysis of randomized controlled trials. J. Med. Virol. 2022, 94, 2102–2107. [Google Scholar] [CrossRef] [PubMed]
  69. Song, J.; Zhang, L.; Xu, Y.; Yang, D.; Zhang, L.; Yang, S.; Zhang, W.; Wang, J.; Tian, S.; Yang, S.; et al. The comprehensive study on the therapeutic effects of baicalein for the treatment of COVID-19 in vivo and in vitro. Biochem. Pharmacol. 2021, 183, 114302. [Google Scholar] [CrossRef] [PubMed]
  70. Hemila, H.; Chalker, E. Carrageenan nasal spray may double the rate of recovery from coronavirus and influenza virus infections: Re-analysis of randomized trial data. Pharmacol. Res. Perspect. 2021, 9, e00810. [Google Scholar] [CrossRef] [PubMed]
  71. Peña-Silva, R.; Duffull, S.B.; Steer, A.C.; Jaramillo-Rincon, S.X.; Gwee, A.; Zhu, X. Pharmacokinetic considerations on the repurposing of ivermectin for treatment of COVID-19. Br. J. Clin. Pharmacol. 2021, 87, 1589–1590. [Google Scholar] [CrossRef] [PubMed]
  72. LeCher, J.C.; Zandi, K.; Costa, V.V.; Amblard, F.; Tao, S.; Patel, D.; Lee, S.; da Silva Santos, F.R.; Goncalves, M.R.; Queroz-Junior, C.M.; et al. Discovery of a 2′-Fluoro,2′-bromouridine phosphoramidate prodrug exhibiting anti-yellow fever virus activity in culture and in mice. Microorganisms 2022, 10, 2098. [Google Scholar] [CrossRef] [PubMed]
  73. Zandi, K.; Musall, K.; Oo, A.; Cao, D.; Liang, B.; Hassandarvish, P.; Lan, S.; Slack, R.L.; Kirby, K.A.; Bassit, L.; et al. Baicalein and baicalin inhibit SARS-CoV-2 RNA-dependent-RNA polymerase. Microorganisms 2021, 9, 893. [Google Scholar] [CrossRef] [PubMed]
  74. American Type Culture Collection (ATCC). Vero CCL-81™. Available online: https://www.atcc.org/products/ccl-81 (accessed on 18 December 2022).
  75. American Type Culture Collection (ATCC). Caco-2 HTB-37™. Available online: https://www.atcc.org/products/htb-37 (accessed on 18 December 2022).
  76. American Type Culture Collection (ATCC). Calu-3 HTB-55™. Available online: https://www.atcc.org/products/htb-55 (accessed on 18 December 2022).
  77. JCRB Cell Bank. JCRB0403 HuH-7. Available online: https://cellbank.nibiohn.go.jp/~cellbank/en/search_res_det.cgi?RNO=jcrb0403 (accessed on 18 December 2022).
  78. Patil, S.; Kamath, S.; Sanchez, T.; Neamati, N.; Schinazi, R.F.; Buolamwini, J.K. Synthesis and biological evaluation of novel 5(H)-phenanthridin-6-ones, 5(H)-phenanthridin-6-one diketo acid, and polycyclic aromatic diketo acid analogs as new HIV-1 integrase inhibitors. Bioorg. Med. Chem. 2007, 15, 1212–1228. [Google Scholar] [CrossRef] [PubMed]
  79. Chou, T.C.; Talalay, P. Quantitative analysis of dose-effect relationships: The combined effects of multiple drugs or enzyme inhibitors. Adv. Enzym. Regul. 1984, 22, 27–55. [Google Scholar] [CrossRef] [PubMed]
  80. Zandi, K.; Amblard, F.; Musall, K.; Downs-Bowen, J.; Kleinbard, R.; Oo, A.; Cao, D.; Liang, B.; Russell, O.O.; McBrayer, T.; et al. Repurposi[Singh, 2011 #363]ng nucleoside analogs for human coronaviruses. Antimicrob. Agents Chemother. 2020, 65, e01652-20. [Google Scholar] [CrossRef] [PubMed]
  81. Gilead Sciences Inc. Sofosbuvir and Velpatasvir Tablets. Highlights of Prescribing Information. Available online: https://www.asegua.com/~/media/Files/pdfs/medicines/liver-disease/asegua/asegua_sof_vel_pi.pdf (accessed on 18 December 2022).
  82. Gilead Sciences. DESCOVY® (Emtricitabine and Tenofovir Alafenamide) Tablets. Highlights of Prescribing Information. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2019/208215s012lbl.pdf (accessed on 18 December 2022).
  83. Gilead Sciences. TRUVADA® (Emtricitabine and Tenofovir Disoproxil Fumarate). Highlights of Prescribing Information. Available online: https://www.gilead.com/~/media/files/pdfs/medicines/hiv/truvada/truvada_pi.pdf (accessed on 18 December 2022).
  84. Axelrod, F.B.; Liebes, L.; Gold-Von Simson, G.; Mendoza, S.; Mull, J.; Leyne, M.; Norcliffe-Kaufmann, L.; Kaufmann, H.; Slaugenhaupt, S.A. Kinetin improves IKBKAP mRNA splicing in patients with familial dysautonomia. Pediatr. Res. 2011, 70, 480–483. [Google Scholar] [CrossRef] [PubMed]
  85. Taisho Toyama Pharmaceutical Co., Ltd. AVIGAN Tablets (290 mg Favipiravir). Available online: https://www.cdc.gov.tw/File/Get/ht8jUiB_MI-aKnlwstwzvw (accessed on 18 December 2022).
  86. Hashemian, S.M.; Farhadi, T.; Velayati, A.A. A review on favipiravir: The properties, function, and usefulness to treat COVID-19. Expert. Rev. Anti Infect. Ther. 2021, 19, 1029–1037. [Google Scholar] [CrossRef] [PubMed]
  87. Watanabe, R.; Esaki, T.; Kawashima, H.; Natsume-Kitatani, Y.; Nagao, C.; Ohashi, R.; Mizuguchi, K. Predicting fraction unbound in human plasma from chemical structure: Improved accuracy in the low value ranges. Mol. Pharm. 2018, 15, 5302–5311. [Google Scholar] [CrossRef] [PubMed]
  88. Zeitlinger, M.A.; Derendorf, H.; Mouton, J.W.; Cars, O.; Craig, W.A.; Andes, D.; Theuretzbacher, U. Protein binding: Do we ever learn? Antimicrob. Agents Chemother. 2011, 55, 3067–3074. [Google Scholar] [CrossRef] [PubMed]
  89. Qian, H.J.; Wang, Y.; Zhang, M.Q.; Xie, Y.C.; Wu, Q.Q.; Liang, L.Y.; Cao, Y.; Duan, H.Q.; Tian, G.H.; Ma, J.; et al. Safety, tolerability, and pharmacokinetics of VV116, an oral nucleoside analog against SARS-CoV-2, in Chinese healthy subjects. Acta Pharmacol. Sin. 2022, 43, 3130–3138. [Google Scholar] [CrossRef] [PubMed]
  90. Yan, V. First-in-Woman safety, tolerability, and pharmacokinetics of orally administered GS-441524: A broad-spectrum antiviral treatment for COVID-19. In OSF Preprints; Center for Open Science: Charlottesville, VA, USA, 2021. [Google Scholar]
  91. Tao, S.; Zandi, K.; Bassit, L.; Ong, Y.T.; Verma, K.; Liu, P.; Downs-Bowen, J.A.; McBrayer, T.; LeCher, J.C.; Kohler, J.J.; et al. Comparison of anti-SARS-CoV-2 activity and intracellular metabolism of remdesivir and its parent nucleoside. Curr. Res. Pharmacol. Drug Discov. 2021, 2, 100045. [Google Scholar] [CrossRef] [PubMed]
  92. Pfizer Inc. Highlights of Prescribing Information Paxlovid™. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2023/217188s000lbl.pdf (accessed on 3 November 2023).
  93. Romark Laboratories, L.C. Alinia® (Nitazoxanide) Tablets. (Nitazoxanide) for Oral Suspension. Prescribing Information. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2005/021818lbl.pdf (accessed on 18 December 2022).
  94. Singh, N.; Narayan, S. Nitazoxanide: A broad spectrum antimicrobial. Med. J. Armed. Forces India 2011, 67, 67–68. [Google Scholar] [CrossRef] [PubMed]
  95. Guzzo, C.A.; Furtek, C.I.; Porras, A.G.; Chen, C.; Tipping, R.; Clineschmidt, C.M.; Sciberras, D.G.; Hsieh, J.Y.-K.; Lasseter, K.C. Safety, tolerability, and pharmacokinetics of escalating high doses of ivermectin in healthy adult subjects. J. Clin. Pharmacol. 2002, 42, 1122–1133. [Google Scholar] [CrossRef]
  96. Buonfrate, D.; Chesini, F.; Martini, D.; Roncaglioni, M.C.; Fernandez, M.L.O.; Alvisi, M.F.; De Simone, I.; Rulli, E.; Nobili, A.; Casalini, G.; et al. High-dose ivermectin for early treatment of COVID-19 (COVER study): A randomised, double-blind, multicentre, phase II, dose-finding, proof-of-concept clinical trial. Int. J. Antimicrob. Agents 2022, 59, 106516. [Google Scholar] [CrossRef] [PubMed]
  97. Muñoz, J.; Ballester, M.R.; Antonijoan, R.M.; Gich, I.; Rodríguez, M.; Colli, E.; Gold, S.; Krolewiecki, A.J. Safety and pharmacokinetic profile of fixed-dose ivermectin with an innovative 18mg tablet in healthy adult volunteers. PLoS Negl. Trop. Dis. 2018, 12, e0006020. [Google Scholar] [CrossRef] [PubMed]
  98. Peng, B.; Lloyd, P.; Schran, H. Clinical pharmacokinetics of imatinib. Clin. Pharmacokinet. 2005, 44, 879–894. [Google Scholar] [CrossRef] [PubMed]
  99. Harb, W.A.; Diefenbach, C.S.; Lakhani, N.; Rutherford, S.C.; Schreeder, M.T.; Ansell, S.M.; Sher, T.; Aboulafia, D.M.; Cohen, J.B.; Nix, D.; et al. Phase 1 clinical safety, pharmacokinetics (PK), and activity of apilimod dimesylate (LAM-002A), a first-in-class inhibitor of phosphatidylinositol-3-phosphate 5-kinase (PIKfyve), in ratients with relapsed or refractory B-cell malignancies. Blood 2017, 130, 4119. [Google Scholar]
  100. Ikonomov, O.C.; Sbrissa, D.; Shisheva, A. Small molecule PIKfyve inhibitors as cancer therapeutics: Translational promises and limitations. Toxicol. Appl. Pharmacol. 2019, 383, 114771. [Google Scholar] [CrossRef] [PubMed]
  101. G.D. Searle LLC. CELEBREX® (Celecoxib) Capsules Drug Safety Information. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2005/020998s017lbl.pdf (accessed on 18 December 2022).
  102. Abbott Laboratories, Inc. ZYFLO® (Zileuton Tablets) Drug Safety Data (Reference ID: 314232). Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2012/020471s017lbl.pdf (accessed on 18 December 2022).
  103. Gandhi, Y.; Eley, T.; Fura, A.; Li, W.; Bertz, R.J.; Garimella, T. Daclatasvir: A Review of Preclinical and Clinical Pharmacokinetics. Clin. Pharmacokinet. 2018, 57, 911–928. [Google Scholar] [CrossRef] [PubMed]
  104. Miller, D.B.; Spence, J.D. Clinical pharmacokinetics of fibric acid derivatives (fibrates). Clin. Pharmacokinet. 1998, 34, 155–162. [Google Scholar] [CrossRef] [PubMed]
  105. Kil, J.; Harruff, E.E.; Longenecker, R.J. Development of ebselen for the treatment of sensorineural hearing loss and tinnitus. Hear. Res. 2022, 413, 108209. [Google Scholar] [CrossRef] [PubMed]
  106. Jazz Pharmaceuticals Inc. LUVOX® (Fluvoxamine Maleate) Tablets 25 mg, 50 mg and 100 mg. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2007/021519lbl.pdf (accessed on 18 December 2022).
  107. Perucca, E.; Gatti, G.; Spina, E. Clinical pharmacokinetics of fluvoxamine. Clin. Pharmacokinet. 1994, 27, 175–190. [Google Scholar] [CrossRef] [PubMed]
  108. Wang, X.H.; Cai, L.L.; Zhang, X.Y.; Deng, L.Y.; Zheng, H.; Deng, C.Y.; Wen, J.; Zhao, X.; Wei, Y.; Chen, L. Improved solubility and pharmacokinetics of PEGylated liposomal honokiol and human plasma protein binding ability of honokiol. Int. J. Pharm. 2011, 410, 169–174. [Google Scholar] [CrossRef] [PubMed]
  109. Rawashdeh, N.M.; Najib, N.M.; Jalal, I.M. Comparative bioavailability of two capsule formulations of mefenamic acid. Int. J. Clin. Pharmacol. Ther. 1997, 35, 329–333. [Google Scholar] [PubMed]
  110. U.S. Food & Drug Administration. PONSTEL® (Mefenamic Acid Capsules, USP). Drug Safety Data. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2008/015034s040lbl.pdf (accessed on 18 December 2022).
  111. Warner Chilcott (US) Ltd. DORYX® (Doxycycline Hyclate) Delayed-Release Tablets, 75 mg, 100 mg and 150 mg for Oral Use. Highlights of Prescribing information. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2008/050795s005lbl.pdf (accessed on 14 November 2023).
  112. Alghamdi, W.A.; Al-Shaer, M.H.; Kipiani, M.; Barbakadze, K.; Mikiashvili, L.; Kempker, R.R.; Peloquin, C.A. Pharmacokinetics of bedaquiline, delamanid and clofazimine in patients with multidrug-resistant tuberculosis. J. Antimicrob. Chemother. 2021, 76, 1019–1024. [Google Scholar] [CrossRef]
  113. U.S. Food & Drug Administration. NDA Application 022519Orig1s000. Duexis (Ibuprofen/Famotidine). Available online: https://www.accessdata.fda.gov/drugsatfda_docs/nda/2011/022519Orig1s000ClinPharmR.pdf (accessed on 18 December 2022).
  114. Pathirana, S.; Jayawardena, S.; Meeves, S.; Thompson, G.A. Brompheniramine and chlorpheniramine pharmacokinetics following single-dose oral administration in children aged 2 to 17 years. J. Clin. Pharmacol. 2018, 58, 494–503. [Google Scholar] [CrossRef] [PubMed]
  115. Toutain, C.E.; Seewald, W.; Jung, M. The intravenous and oral pharmacokinetics of lotilaner in dogs. Parasit. Vectors 2017, 10, 522. [Google Scholar] [CrossRef] [PubMed]
  116. Nycomed GmbH. ALVESCO® (Ciclesonide) Inhalation Aerosol 80 mcg, Highlights of Prescribing Information. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2012/021658s006lbl.pdf (accessed on 18 December 2022).
  117. Merck & Co. Inc. SINGULAIR® (Montelukast Sodium) Drug Safety Data. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2009/020829s051_020830s052_021409s028lbl.pdf (accessed on 16 December 2022).
  118. Zhang, W.; Feng, F.; Wang, Y. Bioequivalence study of bromhexine by liquid chromatography-electrospray ionization-mass spectrometry after oral administration of bromhexine hydrochloride tablets. J. Pharm. Biomed. Anal. 2008, 48, 1206–1210. [Google Scholar] [CrossRef] [PubMed]
  119. Merck & Co. Inc. ZEPATIER® (Elbasvir and Grazoprevir) Tablets Highlighrs of Prescribing Information. Available online: https://www.merck.com/product/usa/pi_circulars/z/zepatier/zepatier_pi.pdf (accessed on 18 December 2022).
  120. deVries, T.; Dentiste, A.; Handiwala, L.; Jacobs, D. Bioavailability and pharmacokinetics of once-daily amantadine extended-release tablets in healthy volunteers: Results from three randomized, crossover, open-Label, Phase 1 studies. Neurol. Ther. 2019, 8, 449–460. [Google Scholar] [CrossRef] [PubMed]
  121. Kim, Y.; Liu, H.; Galasiti Kankanamalage, A.C.; Weerasekara, S.; Hua, D.H.; Groutas, W.C.; Chang, K.-O.; Pedersen, N.C. Reversal of the progression of fatal coronavirus infection in cats by a broad-spectrum coronavirus protease inhibitor. PLoS Pathog. 2016, 12, e1005531. [Google Scholar]
  122. Schwope, D.M.; Karschner, E.L.; Gorelick, D.A.; Huestis, M.A. Identification of recent cannabis use: Whole-blood and plasma free and glucuronidated cannabinoid pharmacokinetics following controlled smoked cannabis administration. Clin. Chem. 2011, 57, 1406–1414. [Google Scholar] [CrossRef] [PubMed]
  123. Harpsøe, N.G.; Andersen, L.P.; Gögenur, I.; Rosenberg, J. Clinical pharmacokinetics of melatonin: A systematic review. Eur. J. Clin. Pharmacol. 2015, 71, 901–909. [Google Scholar] [CrossRef] [PubMed]
  124. Chan, K.K.; Vyas, K.H.; Brandt, K.D. In vitro protein binding of diclofenac sodium in plasma and synovial fluid. J. Pharm. Sci 1987, 76, 105–108. [Google Scholar] [CrossRef] [PubMed]
  125. U.S. Food & Drug Administration. Diclofenac Potassium Powder for Oral Solution. Clinical Pharmacology Review. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/nda/2009/022165s000ClinPharmR.pdf (accessed on 14 November 2023).
  126. Ramaswamy, B.; Phelps, M.A.; Baiocchi, R.; Bekaii-Saab, T.; Ni, W.; Lai, J.P.; Wolfson, A.; Lustberg, M.E.; Wei, L.; Wilkins, D.; et al. A dose-finding, pharmacokinetic and pharmacodynamic study of a novel schedule of flavopiridol in patients with advanced solid tumors. Investig. New Drugs 2012, 30, 629–638. [Google Scholar] [CrossRef] [PubMed]
  127. AstraZeneca Pharmaceuticals LP. CALQUENCE® (Acalabrutinib) Capsules, for Oral Use. Highlights of Prescribing Information. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2017/210259s000lbl.pdf (accessed on 22 November 2023).
  128. Low, Z.Y.; Yip, A.J.W.; Lal, S.K. Repositioning ivermectin for COVID-19 treatment: Molecular mechanisms of action against SARS-CoV-2 replication. Biochim. Biophys. Acta Mol. Basis Dis. 2022, 1868, 166294. [Google Scholar] [CrossRef] [PubMed]
  129. Kim, T.; Lee, J.S.; Ju, Y.S. Experimental models for SARS-CoV-2 infection. Mol Cells 2021, 44, 377–383. [Google Scholar] [CrossRef] [PubMed]
  130. Hoffmann, M.; Mosbauer, K.; Hofmann-Winkler, H.; Kaul, A.; Kleine-Weber, H.; Kruger, N.; Gassen, N.C.; Muller, M.A.; Drosten, C.; Pohlmann, S. Chloroquine does not inhibit infection of human lung cells with SARS-CoV-2. Nature 2020, 585, 588–590. [Google Scholar] [CrossRef] [PubMed]
  131. Ou, T.; Mou, H.; Zhang, L.; Ojha, A.; Choe, H.; Farzan, M. Hydroxychloroquine-mediated inhibition of SARS-CoV-2 entry is attenuated by TMPRSS2. PLoS Pathog. 2021, 17, e1009212. [Google Scholar] [CrossRef] [PubMed]
  132. Gavriatopoulou, M.; Korompoki, E.; Fotiou, D.; Ntanasis-Stathopoulos, I.; Psaltopoulou, T.; Kastritis, E.; Terpos, E.; Dimopoulos, M.A. Organ-specific manifestations of COVID-19 infection. Clin. Exp. Med. 2020, 20, 493–506. [Google Scholar] [CrossRef]
  133. Jockusch, S.; Tao, C.; Li, X.; Chien, M.; Kumar, S.; Morozova, I.; Kalachikov, S.; Russo, J.J.; Ju, J. Sofosbuvir terminated RNA is more resistant to SARS-CoV-2 proofreader than RNA terminated by Remdesivir. Sci. Rep. 2020, 10, 16577. [Google Scholar] [CrossRef] [PubMed]
  134. Baddock, H.T.; Brolih, S.; Yosaatmadja, Y.; Ratnaweera, M.; Bielinski, M.; Swift, L.P.; Cruz-Migoni, A.; Fan, H.; Keown, J.R.; Walker, A.P.; et al. Characterization of the SARS-CoV-2 ExoN (nsp14ExoN-nsp10) complex: Implications for its role in viral genome stability and inhibitor identification. Nucleic Acids Res. 2022, 50, 1484–1500. [Google Scholar] [CrossRef] [PubMed]
  135. Hu, W.J.; Chang, L.; Yang, Y.; Wang, X.; Xie, Y.C.; Shen, J.S.; Tan, B.; Liu, J. Pharmacokinetics and tissue distribution of remdesivir and its metabolites nucleotide monophosphate, nucleotide triphosphate, and nucleoside in mice. Acta Pharmacol. Sin. 2021, 42, 1195–1200. [Google Scholar] [CrossRef]
  136. Pond, S.M.; Tozer, T.N. First-pass elimination. Basic concepts and clinical consequences. Clin. Pharmacokinet. 1984, 9, 1–25. [Google Scholar] [CrossRef] [PubMed]
  137. Hurwitz, S.J.; Schinazi, R.F. Practical considerations for developing nucleoside reverse transcriptase inhibitors. Drug Discov. Today Technol. 2012, 9, e183–e193. [Google Scholar] [CrossRef] [PubMed]
  138. Hurwitz, S.J.; Schinazi, R.F. Prodrug strategies for improved efficacy of nucleoside antiviral inhibitors. Curr. Opin HIV AIDS 2013, 8, 556–564. [Google Scholar] [CrossRef] [PubMed]
  139. Rasmussen, H.B.; Jürgens, G.; Thomsen, R.; Taboureau, O.; Zeth, K.; Hansen, P.E.; Hansen, P.R. Cellular uptake and Intracellular phosphorylation of GS-441524: Implications for Its effectiveness against COVID-19. Viruses 2021, 13, 1369. [Google Scholar] [CrossRef] [PubMed]
  140. Humeniuk, R.; Mathias, A.; Kirby, B.J.; Lutz, J.D.; Cao, H.; Osinusi, A.; Babusis, D.; Porter, D.; Wei, X.; Ling, J.; et al. Pharmacokinetic, pharmacodynamic, and drug-interaction profile of remdesivir, a SARS-CoV-2 replication inhibitor. Clin. Pharmacokinet. 2021, 60, 569–583. [Google Scholar] [CrossRef] [PubMed]
  141. European Medicines Agency. Assessment Report. Procedure under Article 5(3) of Regulation (EC) No 726/2004. Use of Molnupiravir for the Treatment of COVID-19. Available online: https://www.ema.europa.eu/en/documents/referral/lagevrio-also-known-molnupiravir-mk-4482-covid-19-article-53-procedure-assessment-report_en.pdf (accessed on 16 December 2022).
  142. Mackman, R.L.; Hui, H.C.; Perron, M.; Murakami, E.; Palmiotti, C.; Lee, G.; Stray, K.; Zhang, L.; Goyal, B.; Chun, K.; et al. Prodrugs of a 1′-CN-4-Aza-7,9-dideazaadenosine C-nucleoside leading to the discovery of remdesivir (gs-5734) as a potent inhibitor of respiratory syncytial virus with efficacy in the african green monkey model of RSV. J. Med. Chem. 2021, 64, 5001–5017. [Google Scholar] [CrossRef] [PubMed]
  143. Jubalent Pharma Ltd. Jubilant Reveals Promising Safety, Absorption Data for Oral Form of COVID-19 Drug Remdesivir. Available online: https://www.clinicaltrialsarena.com/news/jubilant-oral-form-remdesivir/ (accessed on 12 March 2023).
  144. Xie, J.; Wang, Z. Can remdesivir and its parent nucleoside GS-441524 be potential oral drugs? An in vitro and in vivo DMPK assessment. Acta Pharm. Sin. B 2021, 11, 1607–1616. [Google Scholar] [CrossRef] [PubMed]
  145. Rasmussen, H.B.; Thomsen, R.; Hansen, P.R. Nucleoside analog GS-441524: Pharmacokinetics in different species, safety, and potential effectiveness against COVID-19. Pharmacol. Res. Perspect. 2022, 10, e00945. [Google Scholar] [CrossRef] [PubMed]
  146. Pruijssers, A.J.; George, A.S.; Schäfer, A.; Leist, S.R.; Gralinksi, L.E.; Dinnon, K.H., 3rd; 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. [Google Scholar] [CrossRef] [PubMed]
  147. Li, Y.; Cao, L.; Li, G.; Cong, F.; Li, Y.; Sun, J.; Luo, Y.; Chen, G.; Li, G.; Wang, P.; et al. Remdesivir metabolite GS-441524 effectively Inhibits SARS-CoV-2 Infection in mouse models. J. Med. Chem. 2022, 65, 2785–2793. [Google Scholar] [CrossRef] [PubMed]
  148. Pitts, J.; Babusis, D.; Vermillion, M.S.; Subramanian, R.; Barrett, K.; Lye, D.; Ma, B.; Zhao, X.; Riola, N.; Xie, X.; et al. Intravenous delivery of GS-441524 is efficacious in the African green monkey model of SARS-CoV-2 infection. Antivir. Res. 2022, 203, 105329. [Google Scholar] [CrossRef] [PubMed]
  149. Hurwitz, S.J.; Asif, G.; Kivel, N.M.; Schinazi, R.F. Development of an optimized dose for coformulation of zidovudine with drugs that select for the K65R mutation using a population pharmacokinetic and enzyme kinetic simulation model. Antimicrob. Agents Chemother. 2008, 52, 4241–4250. [Google Scholar] [CrossRef] [PubMed]
  150. Schilling, W.H.K.; Jittamala, P.; Watson, J.A.; Ekkapongpisit, M.; Siripoon, T.; Ngamprasertchai, T.; Luvira, V.; Pongwilai, S.; Cruz, C.; Callery, J.J.; et al. Pharmacometrics of high-dose ivermectin in early COVID-19 from an open label, randomized, controlled adaptive platform trial (PLATCOV). Elife 2023, 12, e83201. [Google Scholar] [CrossRef]
  151. Hurwitz, S.J.; Schinazi, R.F. In silico study supports the efficacy of a reduced dose regimen for stavudine. Antivir. Res. 2011, 92, 372–377. [Google Scholar] [CrossRef]
  152. 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 M(pro) inhibitor clinical candidate for the treatment of COVID-19. Science 2021, 374, 1586–1593. [Google Scholar] [CrossRef]
  153. Najjar-Debbiny, R.; Gronich, N.; Weber, G.; Khoury, J.; Amar, M.; Stein, N.; Goldstein, L.H.; Saliba, W. Effectiveness of paxlovid in reducing severe COVID-19 and mortality in high risk patients. Clin. Infect. Dis. 2022, 76, e342–e349. [Google Scholar] [CrossRef] [PubMed]
  154. Sokhela, S.; Bosch, B.; Hill, A.; Simmons, B.; Woods, J.; Johnstone, H.; Akpomiemie, G.; Ellis, L.; Owen, A.; Casas, C.P.; et al. Randomized clinical trial of nitazoxanide or sofosbuvir/daclatasvir for the prevention of SARS-CoV-2 infection. J. Antimicrob. Chemother. 2022, 77, 2706–2712. [Google Scholar] [CrossRef] [PubMed]
  155. U.S. National Institutes of Health. COVID-19 Treatment Guidelines Nitazoxanide. Available online: https://www.covid19treatmentguidelines.nih.gov/ (accessed on 18 December 2022).
  156. U.S. National Institutes of Health. COVID-19 Treatment Guidelines Ivermectin. Available online: https://www.covid19treatmentguidelines.nih.gov/therapies/miscellaneous-drugs/ivermectin (accessed on 18 December 2022).
  157. Japan Pharmaceuticals and Medical Devices Agency. Evaluation and Licensing Division Report on the Deliberation Results—Avigan. Available online: https://www.pmda.go.jp/files/000210319.pdf (accessed on 12 November 2023).
  158. Logue, J.; Chakraborty, A.R.; Johnson, R.; Goyal, G.; Rodas, M.; Taylor, L.J.; Baracco, L.; McGrath, M.E.; Haupt, R.; Furlong, B.A.; et al. PIKfyve-specific inhibitors restrict replication of multiple coronaviruses in vitro but not in a murine model of COVID-19. Commun. Biol. 2022, 5, 808. [Google Scholar] [CrossRef] [PubMed]
  159. Yoon, J.J.; Toots, M.; Lee, S.; Lee, M.E.; Ludeke, B.; Luczo, J.M.; Ganti, K.; Cox, R.M.; Sticher, Z.M.; Edpuganti, V.; et al. Orally efficacious broad-spectrum ribonucleoside analog inhibitor of influenza and respiratory syncytial viruses. Antimicrob. Agents Chemother. 2018, 62, 10–1128. [Google Scholar] [CrossRef] [PubMed]
  160. Özlüşen, B.; Kozan, Ş.; Akcan, R.E.; Kalender, M.; Yaprak, D.; Peltek, İ.B.; Keske, Ş.; Gönen, M.; Ergönül, Ö. Effectiveness of favipiravir in COVID-19: A live systematic review. Eur. J. Clin. Microbiol. Infect. Dis. 2021, 40, 2575–2583. [Google Scholar] [CrossRef] [PubMed]
  161. Nelson, E.A.; Dyall, J.; Hoenen, T.; Barnes, A.B.; Zhou, H.; Liang, J.Y.; Michelotti, J.; Dewey, W.H.; DeWald, L.E.; Bennett, R.S.; et al. The phosphatidylinositol-3-phosphate 5-kinase inhibitor apilimod blocks filoviral entry and infection. PLoS Negl. Trop Dis. 2017, 11, e0005540. [Google Scholar] [CrossRef] [PubMed]
  162. Qiu, S.; Leung, A.; Bo, Y.; Kozak, R.A.; Anand, S.P.; Warkentin, C.; Salambanga, F.D.R.; Cui, J.; Kobinger, G.; Kobasa, D.; et al. Ebola virus requires phosphatidylinositol (3,5) bisphosphate production for efficient viral entry. Virology 2018, 513, 17–28. [Google Scholar] [CrossRef] [PubMed]
  163. Elkin, P.L.; Resendez, S.; Mullin, S.; Troen, B.R.; Mammen, M.J.; Chang, S.; Franklin, G.; McCray, W.; Brown, S.H. Leukotriene inhibitors with dexamethasone show promise in the prevention of death in COVID-19 patients with low oxygen saturations. J. Clin. Transl. Sci. 2022, 6, e74. [Google Scholar] [CrossRef] [PubMed]
  164. Sharpley, A.L.; Williams, C.; Holder, A.A.; Godlewska, B.R.; Singh, N.; Shanyinde, M.; MacDonald, O.; Cowen, P.J. A phase 2a randomised, double-blind, placebo-controlled, parallel-group, add-on clinical trial of ebselen (SPI-1005) as a novel treatment for mania or hypomania. Psychopharmacology 2020, 237, 3773–3782. [Google Scholar] [CrossRef] [PubMed]
  165. Riva, L.; Yuan, S.; Yin, X.; Martin-Sancho, L.; Matsunaga, N.; Pache, L.; Burgstaller-Muehlbacher, S.; De Jesus, P.D.; Teriete, P.; Hull, M.V.; et al. Discovery of SARS-CoV-2 antiviral drugs through large-scale compound repurposing. Nature 2020, 586, 113–119. [Google Scholar] [CrossRef] [PubMed]
  166. U.S. National Library of Medicine. A Study of LAM-002A for the Prevention of Progression of COVID-19. Available online: https://clinicaltrials.gov/ct2/show/NCT04446377 (accessed on 11 January 2023).
  167. Baranov, M.V.; Bianchi, F.; van den Bogaart, G. The PIKfyve inhibitor apilimod: A double-edged sword against COVID-19. Cells 2020, 10, 30. [Google Scholar] [CrossRef] [PubMed]
  168. Citron, F.; Perelli, L.; Deem, A.K.; Genovese, G.; Viale, A. Leukotrienes, a potential target for COVID-19. Prostaglandins Leukot Essent Fat. Acids 2020, 161, 102174. [Google Scholar] [CrossRef] [PubMed]
  169. Kil, J.; Pierce, C.; Tran, H.; Gu, R.; Lynch, E.D. Ebselen treatment reduces noise induced hearing loss via the mimicry and induction of glutathione peroxidase. Hear. Res. 2007, 226, 44–51. [Google Scholar] [CrossRef] [PubMed]
  170. U.S. National Library of Medicine. SPI-1005 Treatment in Moderate COVID-19 Patients. Available online: https://clinicaltrials.gov/ct2/show/NCT04484025 (accessed on 18 December 2022).
  171. Hoertel, N.; Sánchez-Rico, M.; Vernet, R.; Beeker, N.; Jannot, A.S.; Neuraz, A.; Salamanca, E.; Paris, N.; Daniel, C.; Gramfort, A.; et al. Association between antidepressant use and reduced risk of intubation or death in hospitalized patients with COVID-19: Results from an observational study. Mol. Psychiatry 2021, 26, 5199–5212. [Google Scholar] [CrossRef] [PubMed]
  172. Zimniak, M.; Kirschner, L.; Hilpert, H.; Geiger, N.; Danov, O.; Oberwinkler, H.; Steinke, M.; Sewald, K.; Seibel, J.; Bodem, J. The serotonin reuptake inhibitor Fluoxetine inhibits SARS-CoV-2 in human lung tissue. Sci. Rep. 2021, 11, 5890. [Google Scholar] [CrossRef] [PubMed]
  173. 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. [Google Scholar] [CrossRef] [PubMed]
  174. U.S. National Institutes of Health. COVID-19 Treatment Guidelines—Fluvoxamine. Available online: https://www.covid19treatmentguidelines.nih.gov/therapies/miscellaneous-drugs/fluvoxamine/ (accessed on 14 December 2022).
  175. McKeage, K.; Keating, G.M. Fenofibrate: A review of its use in dyslipidaemia. Drugs 2011, 71, 1917–1946. [Google Scholar] [CrossRef] [PubMed]
  176. Keating, G.M. Fenofibrate: A review of its lipid-modifying effects in dyslipidemia and its vascular effects in type 2 diabetes mellitus. Am. J. Cardiovasc. Drugs 2011, 11, 227–247. [Google Scholar] [CrossRef]
  177. Davies, S.P.; Mycroft-West, C.J.; Pagani, I.; Hill, H.J.; Chen, Y.H.; Karlsson, R.; Bagdonaite, I.; Guimond, S.E.; Stamataki, Z.; De Lima, M.A.; et al. The hyperlipidaemic drug fenofibrate significantly reduces infection by SARS-CoV-2 in cell culture models. Front. Pharmacol. 2021, 12, 660490. [Google Scholar] [CrossRef] [PubMed]
  178. Guzman-Esquivel, J.; Galvan-Salazar, H.R.; Guzman-Solorzano, H.P.; Cuevas-Velazquez, A.C.; Guzman-Solorzano, J.A.; Mokay-Ramirez, K.A.; Paz-Michel, B.A.; Murillo-Zamora, E.; Delgado-Enciso, J.; Melnikov, V.; et al. Efficacy of the use of mefenamic acid combined with standard medical care vs. standard medical care alone for the treatment of COVID-19: A randomized double-blind placebo-controlled trial. Int. J. Mol. Med. 2022, 49, 1–9. [Google Scholar] [CrossRef]
  179. Gold-von Simson, G.; Goldberg, J.D.; Rolnitzky, L.M.; Mull, J.; Leyne, M.; Voustianiouk, A.; Slaugenhaupt, S.A.; Axelrod, F.B. Kinetin in familial dysautonomia carriers: Implications for a new therapeutic strategy targeting mRNA splicing. Pediatr. Res. 2009, 65, 341–346. [Google Scholar] [CrossRef] [PubMed]
  180. Souza, T.M.L.; Pinho, V.D.; Setim, C.F.; Sacramento, C.Q.; Marcon, R.; Fintelman-Rodrigues, N.; Chaves, O.A.; Heller, M.; Temerozo, J.R.; Ferreira, A.C.; et al. Preclinical development of kinetin as a safe error-prone SARS-CoV-2 antiviral able to attenuate virus-induced inflammation. Nat. Commun. 2023, 14, 199. [Google Scholar] [CrossRef]
  181. Rydén, L.; Grant, P.J.; Anker, S.D.; Berne, C.; Cosentino, F.; Danchin, N.; Deaton, C.; Escaned, J.; Hammes, H.P.; Huikuri, H.; et al. ESC guidelines on diabetes, pre-diabetes, and cardiovascular diseases developed in collaboration with the EASD—Summary. Diab. Vasc. Dis. Res. 2014, 11, 133–173. [Google Scholar] [CrossRef] [PubMed]
  182. Ventura-López, C.; Cervantes-Luevano, K.; Aguirre-Sánchez, J.S.; Flores-Caballero, J.C.; Alvarez-Delgado, C.; Bernaldez-Sarabia, J.; Sánchez-Campos, N.; Lugo-Sánchez, L.A.; Rodríguez-Vázquez, I.C.; Sander-Padilla, J.G.; et al. Treatment with metformin glycinate reduces SARS-CoV-2 viral load: An in vitro model and randomized, double-blind, Phase IIb clinical trial. Biomed. Pharmacother. 2022, 152, 113223. [Google Scholar] [CrossRef] [PubMed]
  183. Parthasarathy, H.; Tandel, D.; Siddiqui, A.H.; Harshan, K.H. Metformin suppresses SARS-CoV-2 in cell culture. Virus. Res. 2022, 323, 199010. [Google Scholar] [CrossRef] [PubMed]
  184. Schaller, M.A.; Sharma, Y.; Dupee, Z.; Nguyen, D.; Urueña, J.; Smolchek, R.; Loeb, J.C.; Machuca, T.N.; Lednicky, J.A.; Odde, D.J.; et al. Ex vivo SARS-CoV-2 infection of human lung reveals heterogeneous host defense and therapeutic responses. JCI Insight 2021, 6, e148003. [Google Scholar] [CrossRef] [PubMed]
  185. Bristol-Myers Squibb Co. GLUCOPHAGE® (Metformin Hydrochloride) Tablets. Available online: https://www.accessdata.fda.gov/drugsatfda_docs/label/2017/020357s037s039,021202s021s023lbl.pdf (accessed on 26 July 2023).
  186. Bramante, C.T.; Buse, J.; Tamaritz, L.; Palacio, A.; Cohen, K.; Vojta, D.; Liebovitz, D.; Mitchell, N.; Nicklas, J.; Lingvay, I.; et al. Outpatient metformin use is associated with reduced severity of COVID-19 disease in adults with overweight or obesity. J. Med. Virol. 2021, 93, 4273–4279. [Google Scholar] [CrossRef] [PubMed]
  187. Li, Y.; Yang, X.; Yan, P.; Sun, T.; Zeng, Z.; Li, S. Metformin in patients with COVID-19: A systematic review and meta-analysis. Front. Med. 2021, 8, 704666. [Google Scholar] [CrossRef] [PubMed]
  188. Krausz, S.; Boumans, M.J.; Gerlag, D.M.; Lufkin, J.; van Kuijk, A.W.; Bakker, A.; de Boer, M.; Lodde, B.M.; Reedquist, K.A.; Jacobson, E.W.; et al. Brief report: A phase IIa, randomized, double-blind, placebo-controlled trial of apilimod mesylate, an interleukin-12/interleukin-23 inhibitor, in patients with rheumatoid arthritis. Arthritis. Rheum. 2012, 64, 1750–1755. [Google Scholar] [CrossRef] [PubMed]
  189. Bramante, C.T.; Huling, J.D.; Tignanelli, C.J.; Buse, J.B.; Liebovitz, D.M.; Nicklas, J.M.; Cohen, K.; Puskarich, M.A.; Belani, H.K.; Proper, J.L.; et al. Randomized trial of metformin, ivermectin, and fluvoxamine for COVID-19. N. Engl. J. Med. 2022, 387, 599–610. [Google Scholar] [CrossRef] [PubMed]
  190. Yang, C.; Zhao, H.; Tebbutt, S.J. A glimpse into long COVID and symptoms. Lancet Respir. Med. 2022, 10, e81. [Google Scholar] [CrossRef] [PubMed]
  191. U.S. Centers for Disease Control and Prevention. Post-COVID Conditions: Information for Healthcare Providers. Available online: https://www.cdc.gov/coronavirus/2019-ncov/hcp/clinical-care/post-covid-conditions.html (accessed on 26 July 2023).
  192. Bramante, C.T.; Buse, J.B.; Liebovitz, D.M.; Nicklas, J.M.; Puskarich, M.A.; Cohen, K.; Belani, H.K.; Anderson, B.J.; Huling, J.D.; Tignanelli, C.J.; et al. Outpatient treatment of COVID-19 and incidence of post-COVID-19 condition over 10 months (COVID-OUT): A multicentre, randomised, quadruple-blind, parallel-group, phase 3 trial. Lancet Infect. Dis. 2023, 23, 1119–1129. [Google Scholar] [CrossRef] [PubMed]
  193. Mantovani, A.; Morrone, M.C.; Patrono, C.; Santoro, M.G.; Schiaffino, S.; Remuzzi, G.; Bussolati, G. Long Covid: Where we stand and challenges ahead. Cell Death Differ. 2022, 29, 1891–1900. [Google Scholar] [CrossRef] [PubMed]
  194. Rauf, A.; Olatunde, A.; Imran, M.; Alhumaydhi, F.A.; Aljohani, A.S.M.; Khan, S.A.; Uddin, M.S.; Mitra, S.; Emran, T.B.; Khayrullin, M.; et al. Honokiol: A review of its pharmacological potential and therapeutic insights. Phytomedicine 2021, 90, 153647. [Google Scholar] [CrossRef] [PubMed]
  195. Tanikawa, T.; Hayashi, T.; Suzuki, R.; Kitamura, M.; Inoue, Y. Inhibitory effect of honokiol on furin-like activity and SARS-CoV-2 infection. J. Tradit. Complement. Med. 2022, 12, 69–72. [Google Scholar] [CrossRef] [PubMed]
  196. Lin, S.P.; Tsai, S.Y.; Lee Chao, P.D.; Chen, Y.C.; Hou, Y.C. Pharmacokinetics, bioavailability, and tissue distribution of magnolol following single and repeated dosing of magnolol to rats. Planta Med. 2011, 77, 1800–1805. [Google Scholar] [CrossRef] [PubMed]
  197. Pacheco-Quito, E.M.; Ruiz-Caro, R.; Veiga, M.D. Carrageenan: Drug delivery systems and other biomedical applications. Mar. Drugs 2020, 18, 583. [Google Scholar] [CrossRef] [PubMed]
  198. Bansal, S.; Jonsson, C.B.; Taylor, S.L.; Figueroa, J.M.; Dugour, A.V.; Palacios, C.; Vega, J.C. Iota-carrageenan and xylitol inhibit SARS-CoV-2 in Vero cell culture. PLoS ONE 2021, 16, e0259943. [Google Scholar] [CrossRef]
  199. Hebar, A.; Koller, C.; Seifert, J.M.; Chabicovsky, M.; Bodenteich, A.; Bernkop-Schnürch, A.; Grassauer, A.; Prieschl-Grassauer, E. Non-clinical safety evaluation of intranasal iota-carrageenan. PLoS ONE 2015, 10, e0122911. [Google Scholar] [CrossRef] [PubMed]
  200. Figueroa, J.M.; Lombardo, M.E.; Dogliotti, A.; Flynn, L.P.; Giugliano, R.; Simonelli, G.; Valentini, R.; Ramos, A.; Romano, P.; Marcote, M.; et al. Efficacy of a nasal spray containing iota-carrageenan in the postexposure prophylaxis of COVID-19 in hospital personnel dedicated to patients care with COVID-19 disease. Int. J. Gen. Med. 2021, 14, 6277–6286. [Google Scholar] [CrossRef] [PubMed]
  201. Nwabufo, C.K.; Hoque, M.T.; Yip, L.; Khara, M.; Mubareka, S.; Pollanen, M.S.; Bendayan, R. SARS-CoV-2 infection dysregulates the expression of clinically relevant drug metabolizing enzymes in Vero E6 cells and membrane transporters in human lung tissues. Front. Pharmacol. 2023, 14, 1124693. [Google Scholar] [CrossRef]
  202. Blair, H.A. Remdesivir: A Review in COVID-19. Drugs 2023, 83, 1215–1237. [Google Scholar] [CrossRef] [PubMed]
  203. Gerhart, J.; Cox, D.S.; Singh, R.S.P.; Chan, P.L.S.; Rao, R.; Allen, R.; Shi, H.; Masters, J.C.; Damle, B. A Comprehensive review of the clinical pharmacokinetics, pharmacodynamics, and drug interactions of nirmatrelvir/ritonavir. Clin. Pharmacokinet. 2024, 63, 27–42. [Google Scholar] [CrossRef] [PubMed]
  204. Wattanakul, T.; Chotsiri, P.; Scandale, I.; Hoglund, R.M.; Tarning, J. A pharmacometric approach to evaluate drugs for potential repurposing as COVID-19 therapeutics. Expert. Rev. Clin. Pharmacol. 2022, 15, 945–958. [Google Scholar] [CrossRef] [PubMed]
  205. Delorey, T.M.; Ziegler, C.G.K.; Heimberg, G.; Normand, R.; Yang, Y.; Segerstolpe, Å.; Abbondanza, D.; Fleming, S.J.; Subramanian, A.; Montoro, D.T.; et al. COVID-19 tissue atlases reveal SARS-CoV-2 pathology and cellular targets. Nature 2021, 595, 107–113. [Google Scholar] [CrossRef] [PubMed]
  206. Salim, A.K.; Ingraham, N.E.; Murray, T.A.; Marmor, S.; Hovertsen, S.; Gronski, J.; McNeil, C.; Feng, R.; Guzman, G.; Abdelwahab, N.; et al. Time to Stop Using Ineffective COVID-19 Drugs. N. Engl. J. Med. 2022, 387, 654–655. [Google Scholar]
  207. Abrescia, N. Preventing SARS-CoV-2 infection and its severe outcomes in HIV-infected people. Aids 2023, 37, 1473–1475. [Google Scholar] [CrossRef] [PubMed]
  208. Verburgh, M.L.; van der Valk, M.; Rijnders, B.J.A.; Reiss, P.; Wit, F. No association between use of tenofovir disoproxil fumarate, etravirine, or integrase-strand transfer inhibitors and acquisition or severe outcomes of SARS-CoV-2 infection in people with HIV in the Netherlands. Aids 2023, 37, 1481–1486. [Google Scholar] [CrossRef] [PubMed]
Table 2. Toxicities of the drugs in the Calu-3, Caco-2, Vero CCL-81, PBM, CEM, and Huh-7 cells.
Table 2. Toxicities of the drugs in the Calu-3, Caco-2, Vero CCL-81, PBM, CEM, and Huh-7 cells.
Calu-3Caco-2Vero PBMCEMHuh7
DrugCC50
(μM)
CC50
(μM)
{Unbound}
CC50
(μM)
{Unbound}
CC50
(μM)
{Unbound}
CC50
(μM)
{Unbound}
CC50
(μM)
{Unbound}
* NHC >100>10018.8
{18.8}
59.8
{59.8}
2.4
{2.4}
59.1
{59.1}
GS-441524>100>100>100>100>100>100
Remdesivir >100>100>1004.5
{3.7}
11
{10.1}
2.1
{1.9}
Nirmatrelvir >100>100>100>100>100>100
Nitazoxanide49.87.28.7
{7.8}
8.3
{6.7}
15.5
{14}
54.3
{48.9}
Ivermectin17.55.835.2
{31.9}
29.2
{23.8}
12
{9.5}
5.6
{5.1}
Imatinib906612.3
{11.1}
13.9
{11.3}
14.1
{12.8}
11.9
{10.8}
Apilimod mesylate >100>100>100>10068.5
{62.2}
60.3
{54.8}
Celecoxib70.559.150
{45.2}
>10077.1
{69.6}
31.1
{28.1}}
Zileuton46.7>100>100>100>100>100
Daclatasvir>10067.650.6
{45.6}
36.4
{29.2}
26.2
{23.6}
53.6
{48.3}
Fenofibrate≥100>100>10038.9
{31.2}
28.5>100
Ebselen>100>10022.4
{20.3}
93.2
{75.5}
46.7
{42.3}
84.3
{76.3}
Favipiravir>300>300>100>100>100>100
Fluvoxamine (maleate) >100>10035.5
{32.8}
35.4
{29.9}
20.0
{18.5}
36.7
{33.9}
Honokiol 27.518.440.4
{37.8}
30.4
{26.5}
28.6
{26.8}
13.2
{12.4}
Iota-Carrageenan>100>100>100>10095.9
{NE}
>100
Mefenamic acid>100>100>100>100>100>100
The unbound CC50 values were calculated by multiplying the measured CC50 values by the unbound fraction of drug reported in human plasma normalized to the FBS concentration in the media. * NHC (β-D-N4-hydroxycytidine).
Table 3. The anti-SARS-CoV-2 activity, toxicity, and typical Cmax of drugs tested only in Vero CCL-81 cells.
Table 3. The anti-SARS-CoV-2 activity, toxicity, and typical Cmax of drugs tested only in Vero CCL-81 cells.
Vero CCL-81 CellsCmax (μM)
Drug
Typical Dose
EC50/EC90 (μM)CC50
(μM)
(% CV or Range)
{Unbound}
FunUnbound
Cmax/EC50
Therapeutic CategoryRefs
Doxycycline
100 mg bid
then 100 mg/d
69.3/102≥1003.67 (50%)
{0.55}
0.150.008ABA[111]
Clofazimine
200 mg/d
0.5/334.31.67
(2.14–2.60) {0.002}
0.0010.003ABA[112]
Famotidine
40 mg bid
160/NE>1000.2 (29%)
{0.16}
0.80.001AH[113]
Chlorpheniramine
maleate
8 mg q 4–6 h
31/104>1000.027 (41%)
{0.008}
0.280.0002AH[114]
Lotilaner
20 mg/kg, once
(dog or cat)
1.9/4.944.36.72 (25%)
{0.007}
0.0010.0035APA[115]
Ciclesonide
Inhale 80 mg bid
4.1/7.833.90.001
{0.0000/1}
0.011.9 × 10−6GC[116]
Montelukast sodium
10 mg/d
4.9/7.235.90.58
(0.3–0.90)
{0.009}
0.010.002LM[117]
Bromhexine
16 mg bid
0.8/6.2>1000.093 (24%
{0.005})
0.050.006ML[118]
Elbasivir
50 mg/d
3.4/12.5>1000.26 (47%)
{0.003}
0.010.001NNAV[119]
Amantadine
129 mg bid
138/288>1002.17 (9%)
{0.002}
0.0010.0002NNAV[120]
Velpatasvir
100 mg/d
19.1/79.6>1000.29 (54%)
{0.001}
0.0057.8 × 10−5NNAV[81]
GC376
10 mg/kg
(in cats)
0.2/0.6>1001.7
{NE}
NENENNAV[121]
Cannabinol
cigarette (0.79 g) with 6.8% TH
2.7/8.737.30.032
(0.012–0.051) {0.013}
0.10.0011NP[122]
Melatonin
2–6 mg/d
22/NE>1000.0215 (0.009–0.034)
{0.005}
0.220.0002NP[123]
Diclofenac sodium
50 mg tid
97.6/141>1002.41
{0.012}
0.0050.0001NSAID[124,125]
Flavopiridol
30 mg/m2 + 60 mg/m2 infusion
0.1/1.10.0311.92 (55.7%)
{0.096}
0.050.99ONC[126]
Acalabrutinib
100 mg bid
11.1/70.9>1000.69
{0.017}
0.0250.002ONC[127]
Fun = 1 – (fraction bound reported in the literature). Fun of apilimod was computed in silico, and that of lotilaner was approximated as 0.01, as it was described as highly bound. Within each therapeutic category, compounds are listed in descending order of adjusted Cmax/EC50 in Vero CCL-81 cells. qd = once daily, bid = twice daily, tid = three times daily. NE = not estimated. Therapeutic category: ABA: antibacterial antibiotic; AF: antifungal; AH: histamine 1 or 2 blocker; AP: anti-parasitic agent; GC: glucocorticoid; APA: anti-parasitic agent; NP: natural product; ML: mucolytic; NSAID: non-steroidal anti-inflammatory drug; ONC: oncology drug; NNAV: non-nucleoside and non-nucleoside-base analog antiviral drug.
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

Hurwitz, S.J.; De, R.; LeCher, J.C.; Downs-Bowen, J.A.; Goh, S.L.; Zandi, K.; McBrayer, T.; Amblard, F.; Patel, D.; Kohler, J.J.; et al. Why Certain Repurposed Drugs Are Unlikely to Be Effective Antivirals to Treat SARS-CoV-2 Infections. Viruses 2024, 16, 651. https://doi.org/10.3390/v16040651

AMA Style

Hurwitz SJ, De R, LeCher JC, Downs-Bowen JA, Goh SL, Zandi K, McBrayer T, Amblard F, Patel D, Kohler JJ, et al. Why Certain Repurposed Drugs Are Unlikely to Be Effective Antivirals to Treat SARS-CoV-2 Infections. Viruses. 2024; 16(4):651. https://doi.org/10.3390/v16040651

Chicago/Turabian Style

Hurwitz, Selwyn J., Ramyani De, Julia C. LeCher, Jessica A. Downs-Bowen, Shu Ling Goh, Keivan Zandi, Tamara McBrayer, Franck Amblard, Dharmeshkumar Patel, James J. Kohler, and et al. 2024. "Why Certain Repurposed Drugs Are Unlikely to Be Effective Antivirals to Treat SARS-CoV-2 Infections" Viruses 16, no. 4: 651. https://doi.org/10.3390/v16040651

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