COBICISTAT FOR PREVENTION AND/OR TREATMENT OF CORONAVIRUS INFECTIONS

Abstract
The present invention relates to cobicistat and its derivatives or prodrugs for use in the prophylaxis and/or treatment of severe acute respiratory syndrome coronavirus type 2 (SARS-CoV-2) infection, severe acute respiratory syndrome coronavirus (SARS-CoV) infection and/or Middle East respiratory syndrome coronavirus (MERS-CoV) infection. The present invention further relates to methods of prevention and/or treatment of SARS-CoV-2 infection.
Description

The present invention relates to cobicistat and its derivatives or prodrugs for use in the prophylaxis and/or treatment of severe acute respiratory syndrome coronavirus type 2 (SARS-CoV-2) infection, severe acute respiratory syndrome coronavirus (SARS-CoV) infection and/or Middle East respiratory syndrome coronavirus (MERS-CoV) infection. The present invention further relates to methods of prevention and/or treatment of SARS-CoV-2 infection.


BACKGROUND OF THE INVENTION

The announcement of the outbreak of Severe Acute Respiratory Syndrome Coronavirus type 2 (SARS-CoV-2) in December 2019 was followed by quick and pandemic spread of the infection, leading to a medical, economic and social crisis. One of the most challenging health emergencies of the past hundred years, the SARS-CoV-2 pandemic is highlighting the danger posed by RNA viruses, also in countries where they were absent or considered eradicated. The pathogenic effects of SARS CoV-2 can lead to the coronavirus disease 2019 (COVID-19), characterized by severe pneumonia with a high fatality rate, reaching a 40% among hospitalized old patients. Other coronaviruses associated with severe disease and high mortality are MERS-CoV, which can lead to the Middle East Respiratory Syndrome (MERS) and SARS-CoV, which bears a close genetic similarity with SARS-CoV-2 and was the causative agent of the Severe Acute Respiratory Syndrome (SARS).


SARS-CoV-2 belongs to the enveloped positive-sense RNA coronaviruses (Pal et al., 2020), and its genome has a length of 29.9 kb with 12 functional open reading frames (ORFs), along with a set of 9 sub-genomic mRNAs which are carriers of a conserved leader sequence, 9 transcription-regulatory sequences, and 2 terminal untranslated regions (Fehr et al., 2015). The genome encodes a total of 9,860 amino acids and, in particular, four main structural proteins: the spike (S)-glycoprotein, the small envelope/E glycoprotein, the membrane/M glycoprotein and the nucleocapsid/N protein (Pal et al., 2020, Jiang et al., 2020). Additionally, the SARS-CoV-2 genome encodes 16 non-structural proteins (NSPs), which encompass the two viral cysteine proteases, i.e. NSP3/papain-like protease and NSP5/3C-like protease (3CLpro, also known as main protease). Apart from the proteases, the viral NSPs encompass other key viral enzymes such as NSP12/RNA-dependent RNA polymerase (RdRP) and NSP13/helicase, which are essential for the transcription and replication of the virus (Pal et al., 2020).


The ongoing pandemic of SARS-CoV-2 poses the challenge of quick development of antiviral therapies. SARS-CoV-2 is an enveloped, positive sense, RNA virus of the Coronaviridae family, which includes other human-infecting pathogens such as SARS-CoV and MERS-CoV (V'kovski et al. 2020). Currently, there are no widely approved antivirals to treat infection with Coronaviruses. Substantial effort has been devoted to identifying inhibitors of SARS-CoV-2 replication through repurposing of compounds approved for treating other clinical indications. Repositioned drugs offer the advantage of a well-known safety profile and the possibility of faster clinical testing, which is essential during a sudden epidemic outbreak (Pushpakom et al. 2019). Large scale clinical trials have identified immune modulating agents (e.g. dexamethasone (Johnson and Vinetz 2020; RECOVERY Collaborative Group, Horby, Lim, et al. 2020)) as potential treatments for Coronavirus disease 2019 (COVID-19). However, direct acting antiviral agents have shown limited clinical benefits so far. In particular, a set of antiviral drugs initially identified as effective in vitro (remdesivir, chloroquine/hydroxychloroquine) has been unable to reproducibly decrease mortality in placebo-controlled trials (M. Wang et al. 2020; Beigel et al. 2020; Y. Wang et al. 2020; RECOVERY Collaborative Group, Horby, Mafham, et al. 2020).


Complete inhibition of SARS-CoV-2 replication will likely require combinations of antivirals, in line with previous evidence on other RNA viruses (Pawlotsky et al. 2015; Gulick and Flexner 2019). Candidate inhibitors have been proposed to target several critical steps of SARS-CoV-2 replication, including viral entry, polyprotein cleavage by viral proteases, transcription and viral RNA replication (Guy et al. 2020). SARS-CoV-2 entry is mediated by the spike glycoprotein (S-glycoprotein), which binds through its 51 subunit to the cellular receptor Angiotensin-converting enzyme 2 (ACE2). Upon binding, the viral entry requires a proteolytic activation of the S2 subunit leading to the fusion of the viral envelope with the host cell membrane (Hoffmann et al. 2020). The study of candidate inhibitors of SARS-CoV-2 entry has mainly focused on monoclonal antibodies and small molecules to target the association of the receptor binding domain (RBD) of the S-glycoprotein to ACE-2 (Xiu et al. 2020). Interestingly, the intensively studied antimalarials chloroquine and hydroxychloroquine have been suggested to impair SARS-CoV-2 entry in vitro by both decreasing the binding of the RBD to ACE2 and by decreasing endosomal acidification (Liu et al. 2020).


Upon viral membrane fusion, the viral RNA is released to the cytosol and translated into two large polyproteins that are cleaved into non-structural proteins (nsp) by two viral proteases, the main protease (3CLpro) and the papain-like protease (PLpro). A large body of work to identify antivirals against SARS-CoV-2 has focused on research on these viral proteases. Initial drug repurposing efforts focused on inhibitors of the HIV-1 protease, such as lopinavir and darunavir, alone or in combination with pharmacological boosters. These inhibitors, however, proved poorly effective in inhibiting 3CLpro activity in vitro (Mandi et al. 2020) and did not offer reproducible clinical benefit (Cao et al. 2020; Chen et al. 2020; E. J. Kim et al. 2020). Larger drug screenings have so far relied on a combination of in-silico and in vitro tools (Jin et al. 2020). In particular, libraries of compounds have been screened through molecular docking and many candidate drugs have shown favorable binding properties to the SARS-CoV-2 proteases when analyzed by molecular dynamics (Razzaghi-Asl et al. 2020). Overall, however, repurposed inhibitors of SARS-CoV-2 proteases have generally shown half-maximal inhibitory concentration (IC50) values that were incompatible with dosages achievable in vivo.


The nsps generated by polyprotein cleavage by the viral proteases support the transcription and replication of the viral genome, which is catalyzed by the activity of the RNA-dependent RNA polymerase (RdRP). Owing to its crucial role and high evolutionary conservation, this viral enzyme represents a very attractive therapeutic target, which has so far been exploited by repurposing the anti-Ebola virus drug remdesivir (Mulangu et al. 2019; Beigel et al. 2020; Y. Wang et al. 2020). Other potential RdRP inhibitors, repurposed from treatment of HCV, HIV-1 and influenza virus have been proposed as well (Jácome et al. 2020; Chien et al. 2020; Jockusch et al. 2020). Among them, Favipiravir and Molnupiravir (MK-4482) have shown in vivo therapeutic potential by decreasing viral burden and transmission in hamster and ferret models of the infection, respectively (Cox, Wolf, and Plemper 2021; Kaptein et al. 2020). Viral transcripts generated by the RdRP are used for assembly of new virions by budding into the lumen of the ER-Golgi intermediate compartment (ERGIC) (Klein et al. 2020). The assembly is driven by the structural proteins M and E which are responsible for the incorporation of the N protein forming ribonucleoprotein complexes containing the viral genome. After the budding is completed, viruses are released from the cell either by exocytosis or through lysosomal organelle trafficking (V′kovski et al. 2020). So far, drug candidates proposed to target viral assembly/budding have not advanced beyond in-silico predictions (Gupta et al. 2020).


A major limitation hampering the development of combined antiviral strategies against SARS-CoV-2 is the lack of data on drug interactions. Initial guidelines have cautioned against the combined use of potentially effective compounds, such as remdesivir and chloroquine/hydroxychloroquine, on the basis of the possible interference of the latter with remdesivir metabolism through the efflux pump P-glycoprotein (P-gp) [(Gilead. Summary on compassionate use) (Leegwater et al. 2020; Arribas et al. 2020)]. On the other hand, extensive first pass metabolism by the liver is known to limit bioavailability of remdesivir forcing its intravenous administration, limiting both its scalability and, likely, antiviral efficacy (Siegel et al. 2017).


It is an objective of the present invention to provide means for antiviral therapies of SARS.


SUMMARY OF THE INVENTION

According to the present invention this object is solved by providing cobicistat for use in the prophylaxis and/or treatment of severe acute respiratory syndrome coronavirus type 2 (SARS-CoV-2) infection, severe acute respiratory syndrome coronavirus (SARS-CoV) infection and/or Middle East respiratory syndrome coronavirus (MERS-CoV) infection, wherein cobicistat or a derivative or prodrug thereof is used, said derivative or prodrug is ritonavir or desoxy-ritonavir.


According to the present invention this object is solved by a method of prevention and/or treatment of severe acute respiratory syndrome coronavirus type 2 (SARS-CoV-2) infection, severe acute respiratory syndrome coronavirus (SARS-CoV) infection and/or Middle East respiratory syndrome coronavirus (MERS-CoV) infection,


comprising the step of

    • administering to a subject in need thereof a therapeutically amount of cobicistat or a derivative or prodrug thereof is used, as defined in any one of claims 1 to 10,
    • wherein said derivative or prodrug is ritonavir or desoxy-ritonavir.


DESCRIPTION OF THE PREFERRED EMBODIMENTS OF THE INVENTION

Before the present invention is described in more detail below, it is to be understood that this invention is not limited to the particular methodology, protocols and reagents described herein as these may vary. It is also to be understood that the terminology used herein is for the purpose of describing particular embodiments only, and is not intended to limit the scope of the present invention which will be limited only by the appended claims. Unless defined otherwise, all technical and scientific terms used herein have the same meanings as commonly understood by one of ordinary skill in the art. For the purpose of the present invention, all references cited herein are incorporated by reference in their entireties.


Concentrations, amounts, and other numerical data may be expressed or presented herein in a range format. It is to be understood that such a range format is used merely for convenience and brevity and thus should be interpreted flexibly to include not only the numerical values explicitly recited as the limits of the range, but also to include all the individual numerical values or sub-ranges encompassed within that range as if each numerical value and sub-range is explicitly recited. As an illustration, a numerical range of “1 to 20” should be interpreted to include not only the explicitly recited values of 1 to 20, but also include individual values and sub-ranges within the indicated range. Thus, included in this numerical range are individual values such as 1, 2, 3, 4, 5 . . . 17, 18, 19, 20 and sub-ranges such as from 2 to 10, 8 to 15, etc. This same principle applies to ranges reciting only one numerical value, such as “higher than 150 mg per day”. Furthermore, such an interpretation should apply regardless of the breadth of the range or the characteristics being described.


Cobicistat for Use in the Prophylaxis and Treatment of SARS-CoV-2

As outlined above, the present invention provides cobicistat for use in the prophylaxis and/or treatment of coronavirus infection.


As outlined above, the present invention provides cobicistat for use in the prophylaxis and/or treatment of severe acute respiratory syndrome coronavirus type 2 (SARS-CoV-2) infection, severe acute respiratory syndrome coronavirus (SARS-CoV) infection and/or Middle East respiratory syndrome coronavirus (MERS-CoV) infection,


According to the invention, cobicistat or a derivative or prodrug thereof is used. When throughout this application reference is made to “cobicistat” respective compounds which are derivatives and prodrugs are meant to be included under the proviso that said compounds show similar anti-SARS-CoV-2 activity than cobicistat.


A derivative or prodrug of cobicistat is ritonavir or desoxy-ritonavir.


Name: Cobicistat
Other IDs: GS 9350/GS-9350/GS9350

DrugBank ID: DB09065 (https://www.drugbank.ca/drugs/DB09065)


ATC code: V03AX03 (WHO) (https://www.whocc.no/atc_ddd_index/?code=V03 AX03)


Chemical structure: 1,3-thiazol-5-ylmethyl [(2R,5R)-5-{[(2S)-2-({[(2-isopropyl-1,3-thiazol-4-yl)methyl](methyl)carbamoyl}amino)-4-(morpholin-4-yl)butanoyl]amino}-1,6-diphenylhexan-2-yl]carbamate,


with a molecular formula of C40H53N7O5S2 and a molecular weight of 776.0 g/mol (information retrieved from: National Center for Biotechnology Information. PubChem Database. Cobicistat, CID=25151504, https://pubchem.ncbi.nlm.nih.gov/compound/Cobicistat; accessed on Jun. 16, 2020).




embedded image


Cobicistat, marketed under the trade name of Tybost®, is an approved therapy for treating HIV-1 infection. As of yet, cobicistat has not been used as a direct antretroviral, but rather exerts its effect as a pharmacokinetic enhancer (booster) for other antiretrovirals, such as the integrase inhibitor elvitegravir and some protease inhibitors (e.g. darunavir).


At a molecular level, cobicistat can selectively inhibit cytochrome P450 3A isoforms (CYP3A) and block P-glycoprotein efflux transporters, thus increasing the systemic exposure of co-administered agents, such as antiretroviral drugs, which are metabolized by CYP3A enzymes (von Hentig et al., 2015).


Despite the existence of clinical guidelines and observations conducted by experts in the field, so far nobody hypothesized that cobicistat could be a direct-acting antiviral agent against SARS-CoV-2, as it was only described and utilized as a pharmacological booster of HIV-1 protease inhibitor. This lack of attention was likely due to the fact that it was known that Cobicistat is devoid of any direct antiviral activity against HIV-1, displaying an EC50 against the HIV-1 protease of more than 30 μM (https://www.selleckchem.com/products/cobicistat-gs-9350.html). Moreover, despite being approved by FDA since 2014, cobicistat was never tested during the MERS-CoV outbreak, for which, as in the case of SARS-CoV-2, there are no effective treatments.


Preferably, cobicistat is provided for use in the prophylaxis and/or treatment of Coronavirus disease 2019 (COVID-19).


Thereby, any stage of COVID-19 is comprised.


Preferably, prophylaxis or prevention comprises pre- and post-exposure prophylaxis to or prevention of SARS-CoV-2 infection.


In a preferred embodiment, cobicistat is used in combination with one or more further drug.


Preferably, the one or more further drug is selected from

    • an antiviral agent,
    • a protease inhibitor, preferably an HIV protease inhibitor,
    • a substrate of cytochrome P450-3As (CYP3A) and/or P-glycoprotein (P-gp),
    • an anti-inflammatory glucocorticoid,
    • a januskinase (JAK) inhibitor, and/or
    • a palmitoyl protein thioesterase 1 (PPT1) inhibitor,
    • a monoclonal antibody targeting viral replication or host inflammation.


Preferred antiviral agents are remdesivir, chloroquine or hydroxychloroquine, molnupiravir or favipiravir.


Preferred HIV protease inhibitors are tipranavir, nelfinavir, lopinavir and atazanavir.


A preferred substrates of cytochrome P450-3As (CYP3A) and/or P-glycoprotein (P-gp) is plitidepsin.


Plitidepsin (aplidin) is metabolised through cytochrome P450-3A; the cellular target of cobicistat. It inhibits translation elongation factor eEF1A (White et al., 2021).


Preferred anti-inflammatory glucocorticoids are dexamethasone, prednisone, methylprednisolone and hydrocortisone.


Preferred januskinase (JAK) inhibitors are baricitinib, ruxolitinib, and upadacitinib.


A preferred palmitoyl protein thioesterase 1 (PPT1) inhibitor is GNS561.


A preferred monoclonal antibody targeting host inflammation is tocilizumab.


In a preferred embodiment the further drug is remdesivir, i.a. a combination of cobicistat with remdesivir is preferred.


In one embodiment, the one or more further drug is remdesivir, tipranavir, chloroquine, hydroxychloroquine, molnupiravir, favipiravir, nelfinavir, lopinavir, atazanavir, plitidepsin, dexamethasone, baricitinib and/or GNS561.


In a preferred embodiment cobicistat is used in combination with remdesivir in further combination with one or more further drug, preferably with an HIV protease inhibitor, more preferably tipranavir, nelfinavir, lopinavir or atazanavir.


In one embodiment, cobicistat is used in combination with chloroquine in further combination with one or more further drug, preferably with an HIV protease inhibitor, more preferably tipranavir, nelfinavir, lopinavir or atazanavir.


Route of Administration and Therapeutically Amount


A “therapeutically amount” or “therapeutically effective amount”, both of which terms are used herein interchangeably, of cobicistat according to the present invention is the amount which results in the desired therapeutic result.


In a preferred embodiment, cobicistat is administered in a therapeutically amount, which is higher than the dosage used for HIV-1 treatment.


The dosage of cobicistat for HIV-1 treatment is 150 mg per day. Thus, cobicistat is preferably administered in a therapeutically amount higher than 150 mg per day.


In HIV-1 treatment, Cobicistat is administered orally in combination with the HIV-1 protease inhibitors atazanavir (trade name of the combination: Evotaz®) or darunavir (trade names of the combination: Prezcobix® in the US and Rezolsta® in the EU) or with a combination of several antiretrovirals (trade names Stribild®, Genvoya®, Symtuza®). In all these fixed-dose combinations Cobicistat is administered in a tablet at 150 mg/day.


Cobicistat can be administered via systemic delivery, oral, intranasal, via inhalation, intravenous, or any combination thereof.


Preferred routes of administration of cobicistat are:

    • oral
    • intranasal
    • via inhalation


      or combinations thereof.


In one embodiment, cobicistat is administered orally.


The daily dosage for oral administration is preferably in the range from 10 mg to 1,200 mg, more preferably, 300 to 1,000 mg.


In one embodiment, cobicistat is administered intranasally and/or via inhalation.


An inhaled form of cobicistat could overcome its rapid turnover and allow its delivery at micromolar concentrations in the lung.


In one embodiment, the administration is preferably via a dry powder inhaler, and cobicistat is preferably in solid form.


In one embodiment, the administration is preferably via a nebulizer or a soft mist spray dispenser, and cobicistat is preferably resuspended in an aqueous medium.


The amount administered is preferably at micromolar concentrations, such as in the range from about 2 to 30 μM per day, such as 2 to 15 μM per day.


For example, cobicistat is administered intranasally, through inhalation, at an equivalent concentration of 2-15 μM.


Methods of Prevention and/or Treatment


As outlined above, the present invention provides a method of prevention and/or treatment of severe acute respiratory syndrome coronavirus type 2 (SARS-CoV-2) infection, severe acute respiratory syndrome coronavirus (SARS-CoV) infection and/or Middle East respiratory syndrome coronavirus (MERS-CoV) infection.


Said method comprises the step of

    • administering to a subject in need thereof a therapeutically amount of cobicistat or a derivative or prodrug thereof is used.


As discussed above, said derivative or prodrug is ritonavir or desoxy-ritonavir.


Preferably, the method according to the invention is a method for the prophylaxis and/or treatment of Coronavirus disease 2019 (COVID-19).


Preferably, prophylaxis or prevention comprises pre- and post-exposure prophylaxis to or prevention of SARS-CoV-2 infection.


In a preferred embodiment, cobicistat is administered in combination with one or more further drug.


Preferably, the one or more further drug is selected from

    • an antiviral agent,
    • a protease inhibitor, preferably an HIV protease inhibitor,
    • a substrate of cytochrome P450-3As (CYP3A) and/or P-glycoprotein (P-gp),
    • an anti-inflammatory glucocorticoid,
    • a januskinase (JAK) inhibitor, and/or
    • a palmitoyl protein thioesterase 1 (PPT1) inhibitor,
    • a monoclonal antibody targeting viral replication or host inflammation.


Preferred antiviral agents are remdesivir, chloroquine or hydroxychloroquine, molnupiravir or favipiravir.


Preferred HIV protease inhibitors are tipranavir, nelfinavir, lopinavir and atazanavir.


A preferred substrate of cytochrome P450-3As (CYP3A) and/or P-glycoprotein (P-gp) is plitidepsin.


Preferred anti-inflammatory glucocorticoids are dexamethasone, prednisone, methylprednisolone and hydrocortisone.


Preferred januskinase (JAK) inhibitors are baricitinib, ruxolitinib and upadacitinib.


A preferred palmitoyl protein thioesterase 1 (PPT1) inhibitor is GNS561.


A preferred monoclonal antibody targeting host inflammation is tocilizumab.


In a preferred embodiment the further drug is remdesivir, i.a. a combination of cobicistat with remdesivir is administered.


In one embodiment, the one or more further drug is remdesivir, tipranavir, chloroquine, hydroxychloroquine, molnupiravir, favipiravir, nelfinavir, lopinavir, atazanavir, plitidepsin, dexamethasone, baricitinib and/or GNS561.


In a preferred embodiment cobicistat is administered in combination with remdesivir in further combination with one or more further drug, preferably with an HIV protease inhibitor, more preferably tipranavir, nelfinavir, lopinavir or atazanavir.


In one embodiment, cobicistat is administered in combination with chloroquine in further combination with one or more further drug, preferably with an HIV protease inhibitor, more preferably tipranavir, nelfinavir, lopinavir or atazanavir.


Preferably the administration is oral, intranasal and/or via inhalation.


The therapeutically amount is preferably higher than the dosage used for HIV-1 treatment, i.e. higher than 150 mg per day.


Cobicistat can be administered via systemic delivery, oral, intranasal, via inhalation, intravenous, or any combination thereof.


Preferred routes of administration of cobicistat are:

    • oral
    • intranasal
    • via inhalation


      or combinations thereof.


In one embodiment, the administration of cobicistat is oral, preferably at a daily dosage in the range from 10 mg to 1,200 mg, more preferably, 300 to 1,000 mg.


In one embodiment, the administration of cobicistat is intranasal and/or via inhalation,


wherein, preferably, the administration is via a dry powder inhaler, and cobicistat is preferably in solid form,


or via a nebulizer or a soft mist spray dispenser, and cobicistat is preferably resuspended in an aqueous medium.


Preferably, the amount administered is at micromolar concentrations, such as in the range from about 2 to 30 μM per day, such as 2 to 15 μM per day.


Further Description of Preferred Embodiments

Here, we demonstrate that the FDA-approved CYP3A inhibitor cobicistat, typically used as a booster of HIV-1 protease inhibitors (Sherman et al. 2015), can block SARS-CoV-2 replication in vitro in cell lines of lung and gut origin. While cobicistat was identified through in-silico screening of 3CLpro inhibitors, our data point towards an effect on the S-protein, which in the presence of cobicistat showed decreased ability to form syncytia in cells overexpressing the S-protein. The antiviral concentrations of cobicistat, while well tolerated in vitro, are clearly above those used for HIV-1 treatment, but compatible with plasma levels previously reached at higher doses in mice as well as in humans. In combination with remdesivir, cobicistat exhibits a synergistic effect in rescuing cell viability and abrogating viral replication in both cell lines and in a primary colon organoid. Overall, our data show that cobicistat has a dual activity both as antiviral drug and as pharmacoenhancer, thus potentially providing a basis for combined therapies aimed at complete suppression SARS-CoV-2 replication.


Abstract


Combinations of direct-acting antivirals are needed to minimize drug-resistance mutations and stably suppress replication of RNA viruses. Currently, there are limited therapeutic options against the Severe Acute Respiratory Syndrome Corona Virus 2 (SARS-CoV-2) and testing of a number of drug regimens has led to conflicting results. Here we show for the first time that cobicistat, which is an-FDA approved drug-booster that blocks the activity of the drug metabolizing proteins Cytochrome P450-3As (CYP3As) and P-glycoprotein (P-gp), can have antiviral activity and inhibit SARS-CoV-2 replication. This was unexpected as cobicistat was specifically developed to be “inert” against the HIV-1 protease and to exert solely a booster effect (Xu et al., 2010). Our cell-to-cell membrane fusion assays indicated that the antiviral effect of cobicistat is exerted through inhibition of spike protein-mediated membrane fusion. Incubation with low micromolar concentrations of cobicistat decreased viral replication in three different cell lines including cells of lung and gut origin. These concentrations of cobicistat were previously deemed unnecessary as the inhibitory activity of the drug on CYP3A requires only low nanomolar concentrations (Xu et al., 2010). Indeed, clinical trials testing drug regimens including cobicistat had only considered standard dosing of cobicistat and had not postulated any antiviral effect of this drug and were aimed solely at testing the antiviral activity of HIV-1 protease inhibitors (Chen et al. 2020). When cobicistat was used in combination with the putative CYP3A target and nucleoside analog remdesivir, a synergistic effect on the inhibition of viral replication was observed in cell lines and in a primary human colon organoid.


The cobicistat/remdesivir combination was able to potently abate viral replication to levels comparable to mock-infected cells leading to an almost complete rescue of infected cell viability. These data highlight cobicistat as a therapeutic for treating SARS-CoV-2 infection and as a building block of combination therapies for COVID-19.


Results


In-Silico and In Vitro Analyses Identify Cobicistat as a Candidate Inhibitor of SARS-CoV-2 Replication

To identify potential inhibitors of SARS-CoV-2 replication we performed a structure-based virtual screening of the Drugbank library of compounds approved for clinical use. Candidate drugs were ranked based on their docking score to the substrate-binding site of 3CLpro, i.e. the site essential for the proteolytic function. Our results highlighted seventeen top candidate inhibitors, including compounds used to treat parasitic as well as viral infections. Among the latter, the HIV-1 protease inhibitor nelfinavir, which was one of the top scoring compounds in our analysis, was previously shown to decrease SARS-CoV and SARS-CoV-2 replication in vitro (Yamamoto et al. 2004, n.d.) (Table 1). Two additional drugs used for treatment of HIV-1 displayed top docking scores, i.e. the protease inhibitor tipranavir and, unexpectedly, the CYP3A inhibitor cobicistat, which was previously designed as a molecule devoid of antiviral activity (Xu et al., 2010). The latter was a particularly interesting candidate, given its activity as a booster for HIV-1 protease inhibitors (Sherman et al. 2015), which renders it a promising candidate for combination therapies. Additional in-silico investigation of the binding poses and stability of cobicistat to the 3CLpro of SARS-CoV-2 corroborated a high predicted affinity for the target (FIGS. 1A and B). These in-silico results prompted us to test the effect of cobicistat on SARS-CoV-2 replication in vitro. For this purpose, we conducted a time course analysis of the effect of different concentrations of cobicistat on intracellular viral RNA replication and release of virus into the culture supernatant of Calu-3 cells (FIG. 1C to E). Analysis of virus RNA amounts by qPCR showed a dose dependent inhibitory effect of low micromolar concentrations of cobicistat (FIGS. 1D and E). This effect was visible in both supernatants and cellular extracts, and was reproducible when samples were assayed with two different sets of primers [i.e. N1 and N2 primer sets recommended by the Center of Disease Control (FIGS. 1D and E with the N1 primer set; data for N2 primer set not shown)]. Of note, pre-incubation or treatment upon infection with cobicistat did not increase the antiviral effects as compared to adding the drug two or four hours post-infection, potentially suggesting an effect on late stages of the viral life cycle.


Taken together, these data show that cobicistat has a direct antiviral effect on SARS-CoV-2 replication in vitro.













TABLE 1









Docking


DRUGBANK_ID
Drug groups
Generic name
Main indication
score



















DB01362
approved
Iohexol
Contrast agent
−11.72


DB09134
approved
Ioversol
Contrast agent
−11.03


DB12407
approved;
Iobitridol
Contrast agent
−10.22



investigational


DB12615
approved;
Plazomicin
Antibiotic for urinary tract
−9.43



investigational

infections


DB00932
approved;
Tipranavir
HIV protease inhibitor
−8.06



investigational


DB00220
approved
Nelfinavir
HIV protease inhibitor
−7.91


DB08909
approved
Glycerol
Nitrogen-binding agent for
−7.86




phenylbutyrate
management of urea





cycle disorders


DB00905
approved;
Bimatoprost
Analog of prostaglandin
−7.67



investigational

F2α for treatment of





glaucoma


DB08889
approved;
Carfilzomib
Proteasome Inhibitor (anti-
−7.54



investigational

cancer)


DB09065
approved
Cobicistat
CYP3A inhibitor for
−7.12





boosting HIV-1 protease





inhibitors


DB04868
approved;
Nilotinib
Tyrosine kinase inhibitor
−7.05



investigational

for treatment of chronic





myelogenous leukemia


DB01288
approved;
Fenoterol
Beta adrenergic agonist for
−7.05



investigational

asthma treatment


DB00482
approved;
Celecoxib
Nonsteroidal anti-
−6.80



investigational

inflammatory drug


DB13931
approved
Netarsudil
Rho kinase inhibitor for
−6.75





treatment of glaucoma


DB11611
approved
Lifitegrast
Anti-Inflammatory for
−6.45





treatment of





keratoconjunctivitis sicca


DB11979
approved;
Elagolix
gonadotropin-releasing
−5.72



investigational

hormone antagonist for





treatment of endometriosis





pain


DB01116
approved;
Trimethaphan
nicotinic antagonist used to
−5.70



investigational

counteract hypertension









The Antiviral Concentration Range of Cobicistat is Well Tolerated In Vitro, but Above Plasma Levels Equivalents Achievable Through Standard Dosing of the Drug

We next analyzed more thoroughly the antiviral effects of cobicistat using three cell lines of different origin, i.e. Calu-3 cells (human lung), Vero E6 cells (african green monkey kidney) and T84 cells (human gut), to reflect various known or putative tissue compartments of SARS-CoV-2 replication. Each cell line was infected using two different multiplicities of infection [(MOI) 0.05 and 0.5] and left untreated or treated with various concentrations of cobicistat 2 h post-infection. In all cell lines, cobicistat showed a dose dependent effect in decreasing viral RNA release in supernatant (FIG. 2A). In line with this, the higher concentrations of cobicistat tested (5-10 μM) were able to partially rescue viability of infected cells, as shown by both MTT and crystal violet assay (FIGS. 2A and B), while being well tolerated by uninfected cells (FIGS. 2A and D). Overall, the range of IC50 concentrations of cobicistat (0.58-8.76 μM) was dependent on the MOI of the infection and on the cell type, but always far below the half cytotoxic concentration (CC50) range of the drug on the same cell types (38.66-53 μM). We then compared our in vitro results with previously known pharmacokinetic properties of cobicistat in humans and mice, namely the peak plasma levels detectable in mice (Pharmacology Review of Cobicistat—Application number: 203-094) and in humans (Mathias et al. 2010; Kakuda et al. 2014).


Interestingly, maximum plasma concentrations achievable through standard dosing of cobicistat (150 mg/day as a booster for HIV-1 protease inhibitors) (Deeks 2014) were well below (≈1 μM) most IC50 values obtained in our experiments (FIG. 2C). In line with this, clinical testing of the HIV-1 protease darunavir, boosted by standard concentrations of cobicistat, did not yield clinical benefit to SARS-CoV-2 infected patients (Chen et al. 2020). On the other hand, plasma levels achievable through a higher dosage of cobicistat, that was tested in tolerability studies of this drug (400 mg/day) (Mathias et al. 2010), were above IC50 values calculated when cells were infected using a 0.05 MOI (FIG. 2C). Moreover, plasma levels achievable in mice through a higher cobicistat dosage shown to be safe in this animal model (50 mg/Kg) were clearly above all IC50 values calculated in our experiments, while remaining below the CC50 concentrations (FIG. 2C).


Overall, our data show that non-toxic concentrations of cobicistat can consistently decrease SARS-CoV-2 replication in various cellular infection models. Moreover, these data prove that plasma concentrations obtained through standard HIV-1 dosing of cobicistat are below those required to highlight the antiviral effect of cobicistat.


Cobicistat Decreases S-Glycoprotein Content and Syncytia Formation In Vitro

To characterize the mechanism of the antiviral effects of cobicistat, we analyzed the catalytic activity of 3CLpro using a previously described FRET assay (Zhang et al. 2020). Apart from cobicistat, compounds tested included HIV-1 protease inhibitors highlighted by our molecular docking [nelfinavir, tipranavir] or previously administered in clinical trials as SARS-CoV-2 therapeutics [darunavir (Chen et al. 2020), lopinavir (Cao et al. 2020)], as well as two positive controls known to inhibit 3CLpro activity [MG132 and GC376 (Ma et al. 2020)].


While treatment with the known inhibitors of 3CLpro, such as GC376 and MG-132, potently reduced the catalytic activity of the enzyme, cobicistat was surprisingly inactive (FIGS. 3A and B). Among the top scoring compounds in our docking analysis (Table 1), only tipranavir proved able to partially inhibit 3CLpro activity, although at relatively high concentrations (EC50 47 μM; FIG. 3A,B).


In light of the lack of effect of cobicistat on 3CLpro, we proceeded to analyze the possible impact of cobicistat on other key viral proteins. To reduce the bias of the analysis, while retaining a representative model of the infection, we performed western blot analysis of Vero E6 cell lysates using previously validated patient sera to detect viral proteins (Pape et al. 2020). The results showed the reduction of a high molecular weight band (≈250 kDa) when infected cells were incubated with low micromolar concentrations of cobicistat (data not shown). Based on the known molecular weights of SARS-CoV-2 proteins, we postulated that the patterns detected with patient sera corresponded to dimers/trimers of the S-protein (Ou et al. 2020; Algaissi et al. 2020) and to the nucleoprotein (N-protein) (Algaissi et al. 2020) of the virus. To confirm this hypothesis, we performed western blot analysis using monoclonal antibodies against the S and N proteins (FIG. 3C). The results confirmed the observation that S-protein levels are decreased by cobicistat (FIG. 3C). Moreover, the data indicated a high inhibition at the level of the S2 subunit (≈100 KDa) of the S-protein, i.e. the subunit responsible for the fusion to the host cell and subsequent viral entry (FIG. 3C). To isolate the possible effect of cobicistat on S-protein-mediated fusion, we used a cellular assay measuring syncytia formation in Vero E6 cells transfected with the S-protein. The results showed decreased syncytia formation when cells were incubated with cobicistat or when sera from SARS-CoV-2 patients were used as controls to block S-protein fusion (FIGS. 3D and E). Of note, both western blot analysis and the syncytia assay indicated an effect of cobicistat in the 5-10 μM range, which corresponds to the range of most IC50 values calculated on the basis of viral RNA levels in supernatants (FIG. 2A, C).


Overall, these data show that the antiviral effect of cobicistat is not mediated by inhibition of 3CLpro activity, but is rather exerted, at least partially, through impairment of S-protein-mediated fusion.


Cobicistat Potently Enhances the Antiviral Effect of Remdesivir in Cell Lines and a Primary Colon Organoid

We then tested the potential of cobicistat to exert a double activity as direct inhibitor of SARS-CoV-2 replication and as pharmacoenhancer of other antivirals. To this aim, we evaluated remdesivir as a candidate compound to synergize with cobicistat. The choice of remdesivir was motivated by its known activity as an inhibitor of SARS-CoV-2 RdRP, as well as by its postulated susceptibility to extensive first pass liver metabolism, potentially mediated by the cellular targets of cobicistat CYP3A and P-gp [E.M.A., Human Medicines Division. Summary on compassionate use Remdesivir Gilead (Siegel et al. 2017)]. We thus examined the in silico predicted affinity of remdesivir for the main members of the CYP3A family (CYP3A4 and 5), as well as for P-gp. The SwissADME server (Daina, Michielin, and Zoete 2017) predicted remdesivir to be both a CYP3A4 and P-gp substrate by using machine learning models with 79% and 88% accuracy, respectively. Similarly, the pkCSM (Pires, Blundell, and Ascher 2015) and CYPreact (Tian et al. 2018) servers also predicted remdesivir to be both a P-gp and CYP3A4 substrate, but not an inhibitor. Finally, remdesivir displayed high docking scores to the active sites of CYP3A4, CYP3A5 and P-gp (data not shown), which were comparable to those of ritonavir and cobicistat, i.e. known inhibitors with well characterized binding (data not shown).


To identify the most suitable in vitro model for testing the combination of remdesivir and cobicistat, we first examined the relative expression levels of CYP3A4, CYP3A5 and P-gp in different human tissues and cell lines susceptible to SARS-CoV-2 infection (FIG. 4). Both transcriptomic analysis (data not shown) and qPCR analysis highlighted liver, gut and kidney as major compartments of CYP3A4/5 and P-gp expression (FIG. 4A), as previously described (von Richter et al. 2004; Wessler et al. 2013). On the other hand, primary lung tissues were characterized by lower CYP3A4/5 and P-gp expression, while the cell line Calu-3 showed intermediate characteristics, with low CYP3A4/5 and high P-gp expression, in line with upregulation of the latter marker in cancer cells (Bradley and Ling 1994). Of note, SARS-CoV-2 infection was associated with altered expression of these genes. Overall, infected cells and primary gut organoids showed a trend towards upregulation of P-gp, CYP3A4 and CYP3A5 expression (FIG. 4B). Of note, cell lines of gut origin and Vero E6 cells displayed a peculiar trend showing opposite expression patterns of CYP3A4 and CYP3A5 upon infection (FIG. 4B). Given their divergent response to the infection, we decided to use both Vero E6 and T84 cells as models for testing cobicistat and remdesivir, to obtain data on the efficacy of the drug combination and on its possible reliance on increased expression of either CYP3A4 or CYP3A5.


While treatment with remdesivir-only displayed antiviral activity at previously described levels (data not shown), the combined use of cobicistat and remdesivir was able to significantly enhance the effect of each drug alone, in both cell lines (FIGS. 5A-G, and FIGS. 6A-D). Suprisingly, this synergistic effect was most visible when cobicistat was administered at low micromolar concentrations, thus proving that the synergism is not merely driven by a booster effect of cobicistat, but also by its direct antiviral activity which had never before been postulated by people skilled in the art. In particular, the drug combination was able to almost completely abrogate viral infection/replication, as measured by IF (FIG. 5A, B; FIG. 6B), and qPCR (FIG. 5C-E; FIG. 6C), analysis. In line with this potent antiviral activity, the cobicistat/remdesivir combination also displayed a synergistic effect in inhibiting the cytopathic effects of SARS-CoV-2, thus restoring viability of infected cells to levels comparable to mock-infected controls (FIG. 5F; FIG. 6A,B,D). We then tested the effect of the drug combination on a primary human colon organoid (FIG. 5G), which is susceptible to SARS-CoV-2 infection, as previously described (Stanifer et al. 2020). Also in this case, the addition of cobicistat enhanced the antiviral effect of remdesivir. Finally, we tested the ability of a combination of cobicistat and chloroquine in rescuing cytopathic effects of SARS-CoV-2 infection in two cell lines (FIGS. 7A and B). While the effects of this combination were lower than those observed when treating with cobicistat and remdesivir, the use of chloroquine with cobicistat was synergistic in one of the two cell lines considered (i.e. Calu-3 cells) (FIG. 7A).


Overall, our data prove that the combination of cobicistat and remdesivir can suppress viral replication in different cellular models of SARS-CoV-2 infection and show that cobicistat can exert a double activity as direct antiviral agent and as pharmacoenhancer.


Discussion


The data herein presented demonstrate the antiviral activity of the FDA-approved drug cobicistat and support its role for combined antiviral therapies against SARS-CoV-2. The use of drug combinations targeting different steps of the viral life cycle is a well-established paradigm for treating RNA-virus infections (Bartlett et al. 2001; Naggie and Muir 2017). Translating this concept to SARS-CoV-2 drug development has, however, proven challenging due to the paucity of effective drug candidates available. In particular, compounds showing promise in initial studies, have failed to reproducibly decrease the mortality and morbidity of the infection (M. Wang et al. 2020; Beigel et al. 2020; Y. Wang et al. 2020; RECOVERY Collaborative Group, Horby, Mafham, et al. 2020). Similarly disappointing results were observed in the early stages of HIV-1 drug discovery, and might be partially explained by the inability of candidate antivirals to reach in vivo concentrations sufficient to completely block viral replication. The use of pharmacoenhancers such as cobicistat (Sherman et al. 2015) could help to overcome these limitations.


While the present study exclusively focused on the combination of cobicistat and remdesivir, more than 30% of all drugs are metabolized by the main cellular targets of cobicistat (i.e. CYP3A4/5) (van Waterschoot et al. 2010). For example, the recently described SARS-CoV-2 inhibitor plitidepsin (White et al. 2021) is mainly metabolized by CYP3A4 in vitro (Brandon et al. 2007). Therefore, it is conceivable that a synergistic effect similar to that described for remdesivir can be obtained by coupling cobicistat to other antiviral agents. In particular, other compounds tested in clinical trials of SARS-CoV-2 patients, such as chloroquine/hydroxychloroquine (K.-A. Kim et al. 2003) and lopinavir (van Waterschoot et al. 2010), are well known substrates of CYP3A. The booster effect of cobicistat would be further complemented by the own antiviral activity of this drug, which was proven herein in vitro on several models of SARS-CoV-2 infection. In line with this, we observed the strongest synergistic effect with remdesivir, when cobicistat was used at concentrations above its IC50 levels, suggesting that the hitherto unknown antiviral effects of cobicistat contribute to the observed synergism. Of note, the concentration range in which cobicistat could inhibit SARS-CoV-2 replication was higher than that achievable through standard dosages (i.e. 150 mg/day) approved for treatment of HIV-1 infection (Deeks 2014). Therefore, the antiviral effect of cobicistat requires administration of the drug at higher dosages (e.g. 400 mg/day) which result in plasma levels compatible with the antiviral concentrations described in our study (Mathias et al. 2010). These observations can also explain the lack of success of an early trial testing the HIV-1 protease inhibitor darunavir, boosted by a standard cobicistat dose (Chen et al. 2020). It is important to note that the authors of said study never considered the addition of cobicistat to have any antiviral potential and concluded from the study the lack of antiviral activity of darunavir, without hypothesizing the possibility to increase the dosage of cobicistat.


Another possible limitation of candidate antivirals for SARS-CoV-2 treatment is the inability to reach specific tissue reservoirs of the infection. Remdesivir is case in point, due to its quick metabolization and poor intestinal absorption (Hu et al. 2020). Of note, previous experience with HIV-1 protease inhibitors suggests that cobicistat might overcome this limitation (Lepist et al. 2012), in line with the synergistic effect that we observed when treating primary colon organoid and T84 colon adenocarcinoma cells with the combination of cobicistat and remdesivir. Intriguingly, the tissue penetration and activity of cobicistat in the main sites of CYP3A expression (i.e. gut and liver) can be relevant also for the route of administration of remdesivir. Currently, remdesivir requires intravenous administration due to its extensive first pass metabolism (Jorgensen, Kebriaei, and Dresser 2020), but its coupling with cobicistat can improve its absorption, perhaps allowing oral formulation of the drug. Increasing the scalability of remdesivir might per se improve its therapeutic potential, as an early treatment of the infection might prevent hospitalization and development of severe COVID-19, a stage where the efficacy of remdesivir could not be firmly established (Y. Wang et al. 2020).


Overall, our study introduces cobicistat as an agent for inhibiting SARS-CoV-2 replication and for combination therapies aimed at blocking or reversing the onset of COVID-19.


The following examples and drawings illustrate the present invention without, however, limiting the same thereto.





BRIEF DESCRIPTION OF THE DRAWINGS


FIG. 1 Cobicistat is an inhibitor of SARS-CoV-2 replication.


A-C) In-silico docking (A,B) and molecular dynamics (C-E) analysis of the putative mode and energy of binding of cobicistat to SARS-CoV-2 3CLpro.


A) Docking pose showing the ligand interaction of cobicistat to the active site of 3CLpro and the formation of hydrogen bonds to ASN142, GLY143 and GLN189 of 3CLpro.


B) Overlay of crystal structures of SARS-Cov-2 3CLpro showing the amino acids important for the binding of cobicistat to the active site of the enzyme. Residues of the catalytic dyad (Cys145 and His41) of 3CLpro were among the highest contributors to non covalent binding to cobicistat. The source and list of structures used are detailed in Example 1.


C) Schematic representation of time course experiments evaluating in vitro inhibition of SARS-CoV-2 replication by cobicistat.


D,E) Effect of various concentrations of cobicistat, added according to the scheme of (D) on intracellular and supernatant SARS-CoV-2 RNA content in Calu-3 cells. Viral RNA content was measured by qPCR using the 2019-nCoV_N1 primer set (Center of Disease Control). Fold change values in intracellular RNA (D) were calculated by the delta-delta CT method, using the Tata-binding protein (TBP) gene as housekeeper control. Expression levels in supernatant (E) were quantified using an in vitro transcribed standard curve generated as described in Example 1. Data are expressed as mean with SD and were analyzed by two-way ANOVA followed by Dunnet's post-test (N=3 independent experiments). *P<0.05; **P<0.01; ***P<0.001.



FIG. 2. Cobicistat decreases replication of SARS-CoV-2 and rescues viability of infected cells in multiple in vitro models.


A,B) Effect of serial dilutions of cobicistat on SARS-CoV-2 RNA concentration in supernatants (A) and on the viability of infected and uninfected cell lines of lung (Calu-3), gut (T84) and kidney (Vero E6) origin (A,B). Cells were infected with SARS-CoV-2 at two different MOIs (0.05 and 0.5) and left untreated or treated with cobicistat two hours post-infection. Forty-eight hours post-infection supernatants were collected and viral RNA was assayed by qPCR while cellular viability was measured by MTT assay (A) or by crystal violet staining (B). Inhibition of viral replication was calculated as described in Example 1 while viability data were normalized to the uninfected or to the untreated control. Half maximal inhibitory (IC50) concentration values were calculated by nonlinear regression. Each point in panel A represents a mean of 3 independent experiments. Pictures in panel B are derived from infections at MOI 0.5 (Calu-3 and T84 cells) or MOI 0.05 (Vero E6 cells).


C) Comparison between the IC50 and CC50 values of cobicistat determined in vitro and the peak plasma levels detectable in mice and in after administration of a single dose of the drug. Determination of in vitro CC50 values is based on the data shown in (D).


D) Uninfected cell lines of lung (Calu-3), gut (T84) and kidney (Vero E6) origin were left untreated or treated with serial dilutions of cobicistat. Forty-eight hours post-treatment cellular viability was measured by MTT assay. Data, expressed as mean±SD of three independent experiments, were normalized to the untreated control and CC50 values were calculated by nonlinear regression.



FIG. 3. Cobicistat decreases SARS-CoV-2 S-protein content and syncytia formation.


A,B). Screening of putative inhibitors of the enzymatic activity of 3CLpro. The activity of 3 CLpro was measured by FRET assay and normalized over the untreated condition (A). Apart from cobicistat, compounds tested included HIV-1 protease inhibitors [nelfinavir, tipranavir] and compounds previously administered in clinical trials as SARS-CoV-2 therapeutics [darunavir, lopinavir], as well as two positive controls known to inhibit 3CLpro activity [MG132 and GC376]. EC50 values were calculated by nonlinear regression (B).


C) Effect of cobicistat on the expression of S- and N-proteins in SARS-CoV-2 infected Vero E6 cells. Cells were infected at 0.5 MOI and left untreated or treated, two hours post-infection, with various concentrations of cobicistat, of the RdRP inhibitor remdesivir, or the 3CLpro inhibitor GC376. Cells were harvested 24 hours post-treatment and subjected to protein extraction and subsequent analysis by Western Blot. Expression of S- and N-proteins, and expression of the housekeeping protein actin-(3, were detected using primary monoclonal antibodies followed by incubation with fluorescent-conjugated secondary antibodies and detection on a LI-COR Odyssey® CLx instrument. Data are representative of three independent experiments.


D,E) Effect of cobicistat on S-protein-mediated syncytia formation. Vero E6 cells were transfected with the SARS-CoV-2 S-protein and left untreated or treated with various concentrations of cobicistat or with sera isolated from convalescent SARS-CoV-2 patients (1:100 dilution). Syncytia formation was examined 24 hours post-transfection by immunofluorescence (IF) staining for DAPI and S-protein (D) and quantified as the number of cells forming syncytia (E). Data were analyzed using the nonparametric Kruskal-Wallis test followed by Dunn's post-test. Horizontal lines represent mean values. **P<0.01; ****P<0.0001. Scale bar=50 μM.



FIG. 4. Expression of the metabolic targets of cobicistat in uninfected and SARS-CoV-2 infected cell lines and primary human organoids.


A,B) The relative expression of CYP3A4/5 and P-gp was analyzed by qPCR in uninfected (A) and SARS-CoV-2 infected or mock infected (B) cells. Infections were carried out at MOI 0.5 for 48 hours. Raw data were used to calculate delta CT values (A), by using the TBP gene as housekeeping control. Fold changes, in infected over mock infected cells, were then calculated using the delta-delta CT method. Data in (B) are expressed as mean±SD (N=3).



FIG. 5. The combination of cobicistat and remdesivir synergistically inhibits SARS-CoV-2 activity.


A-F) Synergistic activity of cobicistat and remdesivir in inhibiting replication and cytopathic effects of SARS-CoV-2 in Vero E6 cells. Cells were infected at 0.5 MOI and left untreated or treated with the drugs at the indicated concentrations two hours-post infection. Forty Eight hours post-treatment: cells were fixed for immunofluorescence (IF) staining (A,B), supernatants were collected for qPCR (C-E) or cellular viability was analyzed (F). For IF detection, cells were stained with sera of SARS-CoV-2 patients and with the J2 antibody, which binds to double stranded RNA (Pape et al. 2020). The percentage of infected cells was determined by automatic acquisition of nine images per well (A), as described in Example 1. Scale bar=100 μM. Viral RNA in supernatants was detected by qPCR using an in vitro transcribed standard curve for absolute quantification (C-E) and data, expressed as mean±SD, were transformed as Logic) to restore normality and analyzed by one-way ANOVA, followed by Holm-Sidak's post-test (C). Cellular viability was measured by MTT assay (F).


Isobologram analysis of synergism (D) (Chou 2010) was performed using the IC90 values for SARS-CoV-2 replication of cobicistat, remdesivir, or their combination, calculated by non-linear regression. Synergism analyses of the inhibition of viral replication (E) or cytopathic effects (F) were performed with the SynergyFinder web-tool using the Zero Interaction Potency (ZIP) model based on inhibition values calculated as described in Example 1.


G) Effect of the combination of cobicistat and remdesivir on SARS-CoV-2 RNA expression in supernatants of a primary human colon organoid. Treatment with cobicistat/remdesivir was performed two hours post-infection and supernatants were collected forty-eight hours post-treatment. Viral RNA was quantified as described for panel (C).


For all panels N=3 independent experiments, except for panel E (N=2 independent experiments) and panel G (N=2 replicates from one colon organoid donor). ***P<0.001; **P<0.01; *P<0.05.



FIG. 6. Synergistic antiviral effect of cobicistat and remdesivir in the Vero E6 and T84 cell lines.


Effect of combined treatment of cobicistat and remdesivir on the viability of SARS-CoV-2 infected Vero E6 cells (A) and on viral replication (B,C) and inhibition of cytopathic effects (D) in T84 cells. Cells were infected at 0.5 MOI and left untreated or treated with the drugs at the indicated concentrations two hours-post infection. Forty Eight hours post-treatment: cells were fixed for crystal violet (A) or immunofluorescence (IF) (B) staining, supernatants were collected for qPCR (C), or cellular viability was analyzed (D). For IF detection, cells were stained with sera of SARS-CoV-2 patients (B). Viral RNA in supernatants was detected by qPCR using an in vitro transcribed standard curve for absolute quantification and data, expressed as mean±SD, were analyzed by non-parametric Friedman test, followed by Dunn's post-test (C). Scale bar=100 μM. Cellular viability was measured by MTT assay and synergism analysis of the inhibition cytopathic effects was performed with the SynergyFinder web-tool using the Zero Interaction Potency (ZIP) model based on inhibition values calculated as described in the Methods section. For panels C,D N=3 independent experiments. *P<0.05.



FIG. 7 Synergistic antiviral effect of cobicistat and chloroquine in the Calu-3 and T84 cell lines.


Effect of combined treatment of cobicistat and remdesivir on the viability of SARS-CoV-2 infected Calu-3 (A) and T84 (B) cells. Cells were infected at 0.5 MOI and left untreated or treated with the drugs at the indicated concentrations two hours-post infection. Forty Eight hours post-treatment cellular viability was analyzed by MTT assay. Synergism analysis of the inhibition cytopathic effects was performed with the SynergyFinder web-tool using the Zero Interaction Potency (ZIP) model.





EXAMPLES
Example 1 Material and Methods
1. Virtual Screening and Molecular Docking

Identification of potentially active SARS-CoV-2 inhibitors with desirable Absorption, Distribution, Metabolism, Excretion and Toxicity (ADME-Tox) properties, was performed by structure-based virtual screening (SBVS) of Drugbank V. 5.1.5(72) compounds targeting the three-dimensional structure of SARS-CoV-2 3CLpro. The analysis was focused on the substrate-binding site, which is located between domain I and II of 3CLpro. The binding site was identified using the publicly available 3D crystal structure [Protein Data Bank (PDB) ID: 6W63]. Structures of the previously described non-covalent protease inhibitor X77 (Andrianov et al., 2020), natively co-crystallized with 3CLpro were used as a reference for the identification of binding-site coordinates and dimensions for the virtual screening workflow, as well as for the docking validation of positions generated from the screening.


Protein structure analysis and preparation for docking were performed using the Schrödinger protein preparation wizard (Schrödinger Inc). Missing hydrogen atoms were added, bond orders were corrected and unknown atom types were assigned. Protein side-chain amides were fixed using program default parameters and missing protein side chains were filled-in using the prime tool. All non-amino acid residues, including water molecules, were removed. Further, unrelated ligand molecules were removed and active ligand structures were extracted and isolated in separate files. Finally, the minimization of protein strain energy was achieved through restrained minimization options with default parameters. The centroids of extracted ligands were then used to identify the binding site with coordinates and dimensions extended for 20 Å stored as Glide grid file. Drug screening was performed using the Glide software (Friesner et al., 2004). High throughput virtual screening (HTVS) was performed with the fastest search configurations. After post-docking minimization, the top-scoring tenth percentile of the output docked structures were subjected to the standard precision docking stage (SP). Then, active ligand structures were extracted and isolated in separate files. Finally, the top 10% scoring compounds were selected and retained only if their good scoring states were confirmed by Extra precision docking.


Remdesivir docking to CYP3A4, CYP3A5 and P-gp structures was performed to assess its capacity as a substrate/inhibitor for these proteins. CYP3A4, CYP3A5 and P-gp structures were collected from Protein Data Bank (PDB), IDs: 5VC0, 5VEU and 6QEE, respectively, and were subjected to the same preparation steps described above. Native inhibitors were used for identification of binding sites; the centroid of the known inhibitor Zosuquidar was used to identify the drug binding pocket of the P-gp protein structure. Further, co-crystallized Ritonavir was used for identification of the drug binding pocket in both CYP3A4/5. Receptor grids were generated for protein structures, for both CYP3A4 and CYP3A5. The heme iron of the Protoporphyrin ring was added as metal coordination constraint, allowing metal-ligand interaction in the subsequent docking steps. Docking was performed using flexible ligand conformer sampling allowing ring sampling with a 2.5 kcal/mol window. Retained poses for the initial docking phase were set to 5000 poses and only 800 best poses per ligand were selected for energy minimization. Finally, post-docking minimization was carried out for 10 poses per ligand with a 0.5 kcal/mol threshold for rejecting minimized poses.


2. Cell Lines and Primary Human Organoids

The following cell lines were used for infection and/or relative quantification of gene expression: Calu-3 (ATCC HTB-55), Caco-2 (ATCC HTB-37), T84 (ATCC CCL-24) and VeroE6 (ATCC CRL-1586). Primary organoids derived from human colon and ileum were seeded in 2D as described in (Stanifer et al. 2020). Culture conditions and susceptibility to SARS-CoV-2 infection have been previously described (Cortese et al. 2020; Stanifer et al. 2020).


3. Virus Stock Production and Infection

Viral stocks used for infections were produced by passaging the BavPatl/2020 SARS-CoV-2 strain in Vero E6 cells and the infectious titer was estimated by plaque assay, as previously described. Infection experiments were conducted using 25,000 or 250,000 cells per well in 96 and 12 well plates, respectively. Cell lines were infected at 0.05 or 0.5 MOI in medium with low FCS content (2%). Colon organoids were infected in a 24-well plate using 60000 plaque forming units (PFU) per well. Two hours post-infection cells were washed twice in PBS and resuspended in complete medium.


4. Drug Treatments

The following compounds were tested to determine their effects on 3CLpro activity, cytotoxicity or inhibition of SARS-CoV-2 replication: cobicistat (#sc-500831; Santa Cruz Biotechnology), remdesivir (#S78932; Selleckchem Chemicals), tipranavir (#sc-220260; Santa Cruz Biotechnology), nelfinavir mesylate hydrate (#PZ0013, Sigma-Aldrich), darunavir, lopinavir (both obtained through the AIDS Research and Reference Reagent Program, Division of AIDS, NIAID), MG-132 (#M8699; Sigma-Aldrich), GC376 (BPS Bioscience), Chloroquine (#C 6628, Sigma Aldrich).


5. RNA Isolation and cDNA Retrotranscription


RNA extraction was performed on cell lysates or supernatants using the NucleoSpin RNA, Mini kit for RNA purification (Macherey-Nagel, Duren, Germany) according to the manufacturer's instructions. The concentration of RNA extracted from cell lysates was measured using a P-class P 300 NanoPhotometer (Implen GmbH, Munich, Germany).


Retrotranscription to cDNA was performed with 500 ng of intracellular RNA or 10 μL of RNA from supernatants, using High-Capacity cDNA Reverse Transcription Kit (Applied Biosystems, Foster City, Calif., USA) following the manufacturer's instructions.


6. SARS-CoV-2 RNA Standard

For the preparation of a viral RNA standard to use in qPCR for quantification of viral copies in supernatants, SARS-CoV-2 N sequence was reverse transcribed from total RNA isolated from cells infected with the SARS-CoV-2 BavPatl stain using Superscript 3 and specific primers (TTAGGCCTGAGTTGAGTCA, SEQ ID NO. 1). The resulting cDNA was amplified and cloned into the pJET1.2 plasmid. Ten μg of plasmid DNA was linearized by Adel restriction enzyme digestion and DNA was purified using the NucleoSpin Gel and PCR Clean-up kit (Macherey-Nagel, Düren, Germany). For in vitro transcription T7 RNA polymerase was used as previously described (Fischl and Bartenschlager 2013). In vitro transcripts were purified by phenol-chloroform extraction and resuspended in RNase-free water. RNA integrity was confirmed by agarose gel electrophoresis.


7. qPCR Analysis


Gene and/or viral expression were analyzed by SYBR green qPCR using, for each reaction, 10 μL of SsoFast™ EvaGreen® Supermix (Bio-Rad Laboratories, Hercules, Calif., USA), 500 nM of forward and reverse primer (0.1 μL each from 100 μM stock), 8.8 μL water and 1 μL cDNA. The primers used are listed in Table 2. The qPCR reaction was performed on a CFX96/C1000 Touch qPCR system (Bio-Rad Laboratories, Hercules, Calif., USA) using the following PCR program: polymerase activation/DNA denaturation 98° C. for 3 min, followed by 45 cycles of denaturation at 98° C. for 10 s; annealing/extension at 60° C. for 40 s and a final extension step at the end of the program at 65° C. for 30 s. Gene expression data were normalized using the delta-delta CT method [2(−ΔΔC(T)) method] (Livak and Schmittgen 2001), using the Tata-binding protein (TBP) gene as housekeeper control.









TABLE 2







List of qPCR primers used in the study













SEQ


Name
Sequence
Source
ID NO.





2019-nCoV_N1-
GAC CCC AAA ATC
(1)
 2


Forward
AGC GAA AT




2019-nCoV_N1-
TCT GGT TAG TGC

 3


Reverse
CAG TTG AAT CTG







2019-nCoV_N2-
TTA CAA ACA TTG
(2)
 4


Forward
GCC GCA AA




2019-nCoV_N2-
GCG CGA CAT TCC

 5


Reverse
GAA GAA







Hum Cyp3A4-
TGA TGG CTC TCA

 6


Forward
TCC CAG AC







Cyp3A4-
AGC CCC ACA CTT

 7


Reverse
TTC CAT AC







AGM Cyp3A4-
TGA TGG ACC TCA

 8


Forward
TCC CAG AC







Hum Cyp3A5-
CGA CAA ACA AAA

 9


Forward
GCA CCG AC







Hum Cyp3A5-
TTA TTG ACT GGG

10


Reverse
CTG CGA G







AGM Cyp3A5-
CGA CAA ACA AAA

11


Forward
GCA CCG AG







AGM Cyp3A5-
TAA TTG ATT GGG

12


Reverse
CCA CGA G







P-gp(MDR1)-F
CCC ATC ATT GCA
Gao et
13



ATA GCA GG
al., 2015



P-gp(MDR1)-R
TGT TCA AAC TTC

14



TGC TCC TGA







TBP-F
CCA CTC ACA GAC
Stanifer
15



TCT CAC AAC
et al.,



TBP-R
CTG CGG TAC AAT
2020
16



CCC AGA ACT





Hum = human


AGM = african green monkey


(1) https://www.cdc.gov/coronavirus/2019-ncov/lab/rt-pcr-panel-primer-probes.html


(2) https://www.cdc.gov/coronavirus/2019-ncov/lab/rt-pcr-panel-primer-probes.html






8. Western Blot

For Western blot experiments 0.5×106 cells were lysed in a buffer (20 mM Tris-HCl, pH 7.4, 1 mM EDTA, 150 mM NaCl, 0.5% Nonidet P-40, 0.1% SDS, and 0.5% sodium deoxycholate supplemented with protease and phosphatase inhibitors (Sigma-Aldrich, Saint Louis, Mich., USA). Lysates were boiled at 95° C. for 10 min and sonicated for 5 min using a Bioruptor® Plus sonication device (Diagenode, Liege, Belgium). Protein lysates were then run on a precast NuPAGEBis-Tris 4-12% (Thermo Fisher Scientific, Waltham, Mass., USA) SDS-PAGE at 100-120 V and transferred onto a nitrocellulose membrane (GE Healthcare, Little Chalfont, UK) for 2.5 h at 25 V using a Trans-Blot device for semi-dry transfer (Bio-Rad Laboratories, Hercules, Calif., USA). Membranes were blocked using the LI-COR Intercept (PBS) Blocking Buffer (LI-COR Biosciences, Lincoln, Nebr., USA) for 1 h at RT and incubated overnight at 4° C. with the following primary antibodies in blocking buffer with 0.2% Tween 20: α-β-actin (1:10,000), (Sigma-Aldrich, Saint Louis, Mich., USA), a-SARS-CoV-2 spike protein [(rabbit; 1:1000) ab252690 Abcam], α-SARS-CoV-2 nucleocapsid [(mouse; 1:1000) AB_2827977, Sino Biological)], sera of SARS-CoV-2 positive individuals (1:200). Sera were collected as described in (Pape et al. 2020), following signing of informed consent by the donors, as well as ethical approval by Heidelberg University Hospital. After primary antibody incubation, membranes were washed three times with 0.1% PBS-Tween and incubated for 1 h with the following fluorescence-conjugated secondary antibodies: IRDye® 800CW Goat anti-Human IgG, IRDye® 800CW anti rabbit, IRDye® 700CW anti mouse (LI-COR Biosciences, Lincoln, Nebr., USA). All secondary antibodies were diluted 1:15000 in blocking buffer+0.2% Tween. After three washes with 0.1% PBS-Tween and one wash in PBS, fluorescence signals were acquired using a LI-COR Odyssey® CLx instrument.


9. Re-Processing of Microarray and RNA-Seq Data

Microarray gene expression data for CYP3A4/5 and P-gp in different anatomical tissues or cell lines were retrieved from Homo Sapiens Affymetrix Human Genome U133 Plus 2.0 Array dataset. Data were filtered by applying the criteria “Healthy sample status” and “No experimental treatment”. From the initial list, tissues with sample size<25 were filtered out. The anatomy search tool was used to plot Log 2 expression ratios of the tested genes. Gene expression data in cell lines were retrieved, apart from the aforementioned microarray dataset, from the RNAseq “mRNA Gene Level Homo sapiens (ref: Ensembl 75)” dataset. The cell line condition filter was used to refine the analysis and include exclusively cell lines susceptible to SARS-CoV-2 infection (i.e. T84, Caco2, Calu-3 and A-549).


10. Cell Viability

Cell viability was evaluated by (3-[4,5-dimethylthiazol-2-yl]-2,5 diphenyl tetrazolium bromide) (MTT) assay and by crystal violet staining as previously described (Shytaj et al. 2020; Feoktistova, Geserick, and Leverkus 2016). Briefly, the MTT assay was conducted using the CellTiter 96® Non-Radioactive Cell Proliferation Assay (MTT) (Promega; Madison, Wis., USA). Cells were plated in a 96-well plate at a concentration of 3×106 cells/mL in 100 μl of medium. The MTT solution (15 μl) was added to each well and, after 2-4 h, the reaction was stopped by the addition of 100 μl of 10% SDS. Absorbance values were acquired using an Infinite 200 PRO (Tecan, Männedorf, Switzerland) multimode plate reader at 570 nm wavelength.


For the crystal violet staining, cells were fixed in 6% formaldehyde and incubated with 0.1% crystal violet for 15 mins. Unbound staining was then washed with H2O and cells were imaged using a Nikon Eclipse Ts2-FL microscope.


11. 3CLpro FRET Assay

The activity of 3CLpro was measured by FRET assay (BPS Bioscience, San Diego, Calif., USA) according to the manufacturer's instructions and as previously described (Zhang et al. 2020). Briefly, serial dilutions of test compounds and known 3CLpro were incubated in a 384 well plate with the 3CLpro and its appropriate buffer, containing 0.5 M DTT. Wells without drugs or without 3CLpro were used as positive control of 3CLpro activity and blank control, respectively. After a 30 min incubation, the 3CLpro substrate was added to each well and the plate was stored for 4 hours in the dark. The fluorescence signal was acquired on an Infinite 200 PRO (Tecan, Männedorf, Switzerland) using an excitation wavelength of 360 nm and a detection wavelength of 460 nm. All Three separate experiments were conducted, with each experiment performed in duplicate. Relative 3CLpro was expressed as percentage of the positive control after subtraction of the blank.


12. Immunofluorescence and Syncytia Formation Assay

Cells were seeded on iBIDI glass bottom 96 well plate and infected with SARS-CoV-2 strain BavPatl/2020 for 24-48 h at MOI 0.5. Cells were rinsed in PBS and fixed with 6% PFA, followed by permeabilization with 0.5% Triton X100 (Sigma) in PBS for 15 minutes. Cells were then subjected to a standard immunofluorescence staining protocol. Briefly, cells were blocked in 2% milk (Roth) in PBS and incubated with primary antibodies in PBS (anti ds-RNA mouse monoclonal J2 antibody (Scicons) 1:2000 and patient serum 1:250). Cells were washed twice in PBS 0.02% tween and incubated with secondary antibody in PBS (1:1000 anti-mouse 568, Goat anti-human IgG-AlexaFluor 488 (Invitrogen, Thermofisher Scientific) for immunoglobulins detection in human serum and goat anti-mouse IgG-AlexaFluor 568 (Invitrogen, Thermofisher Scientific) for dsRNA detection). Nuclei were counterstained with Hoechst 33342 (Thermofisher Scientific, 0.002 μg/ml in PBS) for 5 minutes, washed twice with PBS and stored at +4° C. until imaging.


For syncytia formation assay, Vero E6 cells (0.2×106 cells/well) were seeded on cover slips in a 12 well plate 24 h prior transfection. Cells were transfected using TransIT-2020 or TransIT-LT1 (Mirus) with 0.75 μg of pCDNA3.1(+)-SARS-CoV-2-S and 100 μl Opti-MEM per well. 2 h post transfection, cells were treated with cobicistat (final concentration of 1 μM, 5 μM and 10 μM), serum of patients (1:500 or 1:100) or DMSO (same concentration as in 10 μM cobicistat). 24 h post transfection, cells were washed twice with PBS and fixed in 4% PFA for 20 min at room temperature. After another washing step, cells were permeabilized in 0.5% Triton for 5 min at room temperature, washed and blocked in 3% lipid-free BSA in PBS-0.1% Tween-20 for 1 h at room temperature. After washing, cells were stained with the primary rabbit polyclonal anti-SARS-CoV-2 spike glycoprotein antibody (1:1000, Abcam) for 1 h at room temperature or overnight at 4° C. After washing, cells were incubated with the secondary Alexa Fluor 488 goat anti-rabbit IgG antibody (1:500, Life Technologies) for 1 h at room temperature. After washing, cells were incubated with DAPI (1:1000, Sigma-Aldrich) for 1 min followed by washing with PBS and deionized water. Images were acquired with Nikon Eclipse Ts2-FL Inverted Microscope. Syncytia with three or more nuclei surrounded by the antibody staining were used for the quantification. The edges of the antibody staining were overdrawn with the polygon selection tool in ImageJ.


13. Microscopy and Image Analysis

Cells were imaged using motorized Nikon Ti2 widefield microscope or with Nikon/Andor (CSU W1) spinning disc using a Plan Apo lambda 20×/0.75 air objective and a back-illuminated EM-CCD camera (Andor iXon DU-888). JOBS module was used for automatic acquisition of 9 images per well. Images were acquired in 3 channels using the following excitation/emission settings: Ex 377/50, Em 447/60 (Hoechst); Ex 482/35, Em 536/40 (AlexaFluor 488); Ex 562/40, Em 624/40 (AlexaFluor 568). When spinning disc was used the excitation was performed with 405 nm, 488 nm and 561 nm lasers.


Quantification of infected cells (expressed as percentage of total cells imaged per well) was performed using a custom-made macro in ImageJ. After camera offset subtraction and local background subtraction using the rolling ball algorithm, nuclei were segmented using automated local thresholding based on the Niblack method. Region of interest (represented by the ring (5 pixel wide) around the nucleus) was determined for each individual cell. Median signal intensity was measured in the region of interest in Alexa488 (serum) and Alexa568 (dsRNA) channels. Threshold for calling infected cells was manually determined for each individual experiment using the data from mock transfected cells. The same image analysis procedure and threshold was used for all wells within one experiment.


14. Statistical Analysis

Data normality assumptions were tested by D'Agostino & Pearson normality test (for >3). Multiple group comparisons were conducted by non-parametric Kruksal Wallis test, followed by Dunn's post-test, or by Two-Way ANOVA followed by Dunnet's post-test. Half maximal inhibitory (IC50) and cytotoxic (CC50) concentrations of the compounds tested were estimated by nonlinear regression using relative inhibition values calculated according to the formula: % inhibition=100*(1−(X−mock infected)/(infected untreated−mock infected)), where X is each given treatment condition. Data analysis was conducted using GraphPad Prism v6 (GraphPad Software, San Diego, Calif., USA). Synergy scores were calculated using the SynergyFinder web-tool (Ianevski et al. 2020) using the Zero Interaction Potency (ZIP) model (Yadav et al. 2015).


The features disclosed in the foregoing description, in the claims and/or in the accompanying drawings may, both separately and in any combination thereof, be material for realizing the invention in diverse forms thereof.


REFERENCES



  • Algaissi, Abdullah, Mohamed A. Alfaleh, Sharif Hala, Turki S. Abujamel, Sawsan S. Alamri, Sarah A. Almahboub, Khalid A. Alluhaybi, et al. 2020. “SARS-CoV-2 S1 and N-Based Serological Assays Reveal Rapid Seroconversion and Induction of Specific Antibody Response in COVID-19 Patients.” Scientific Reports 10 (1): 16561.

  • Andrianov, Alexander M., Yuri V. Kornoushenko, Anna D. Karpenko, Ivan P. Bosko, and Alexander V. Tuzikov. 2020. “Computational Discovery of Small Drug-like Compounds as Potential Inhibitors of SARS-CoV-2 Main Protease.” Journal of Biomolecular Structure & Dynamics, July, 1-13

  • Arribas, Jose Ramon, Jose Ramon Arribas, Arun J. Sanyal, Alex Soriano, Bum Sik Chin, Bum Sik Chin, Shirin Kalimuddin, et al. 2020. “557. Impact of Concomitant Hydroxychloroquine Use on Safety and Efficacy of Remdesivir in Moderate COVID-19 Patients.” Open Forum Infectious Diseases. https://doi.org/10.1093/ofid/ofaa439.751.

  • Bartlett, J. A., R. DeMasi, J. Quinn, C. Moxham, and F. Rousseau. 2001. “Overview of the Effectiveness of Triple Combination Therapy in Antiretroviral-Naive HIV-1 Infected Adults.” AIDS 15 (11): 1369-77.

  • Beigel, John H., Kay M. Tomashek, Lori E. Dodd, Aneesh K. Mehta, Barry S. Zingman, Andre C. Kalil, Elizabeth Hohmann, et al. 2020. “Remdesivir for the Treatment of Covid-19—Final Report.” The New England Journal of Medicine 383 (19): 1813-26.

  • Bradley, G., and V. Ling. 1994. “P-Glycoprotein, Multidrug Resistance and Tumor Progression.” Cancer Metastasis Reviews 13 (2): 223-33.

  • Brandon, Esther F. A., Rolf W. Sparidans, Ronald D. van Ooijen, Irma Meijerman, Luis Lopez Lazaro, Ignacio Manzanares, Jos H. Beijnen, and Jan H. M. Schellens. 2007. “In Vitro Characterization of the Human Biotransformation Pathways of Aplidine, a Novel Marine Anti-Cancer Drug.” Investigational New Drugs 25 (1): 9-19.

  • Cao, Bin, Yeming Wang, Danning Wen, Wen Liu, Jingli Wang, Guohui Fan, Lianguo Ruan, et al. 2020. “A Trial of Lopinavir-Ritonavir in Adults Hospitalized with Severe Covid-19.” The New England Journal ofMedicine 382 (19): 1787-99.

  • Chen, Jun, Lu Xia, Li Liu, Qingnian Xu, Yun Ling, Dan Huang, Wei Huang, et al. 2020. “Antiviral Activity and Safety of Darunavir/Cobicistat for the Treatment of COVID-19.” Open Forum Infectious Diseases 7 (7): ofaa241.

  • Chien, Minchen, Thomas K. Anderson, Steffen Jockusch, Chuanjuan Tao, Xiaoxu Li, Shiv Kumar, James J. Russo, Robert N. Kirchdoerfer, and Jingyue Ju. 2020. “Nucleotide Analogues as Inhibitors of SARS-CoV-2 Polymerase, a Key Drug Target for COVID-19.” Journal of Proteome Research 19 (11): 4690-97.

  • Chou, Ting-Chao. 2010. “Drug Combination Studies and Their Synergy Quantification Using the Chou-Talalay Method.” Cancer Research 70 (2): 440-46.

  • Cortese, Mirko, Ji-Young Lee, Berati Cerikan, Christopher J. Neufeldt, Viola M. J. Oorschot, Sebastian Köhrer, Julian Hennies, et al. 2020. “Integrative Imaging Reveals SARS-CoV-2-Induced Reshaping of Subcellular Morphologies.” Cell Host & Microbe 28 (6): 853-66.e5.

  • Cox, Robert M., Josef D. Wolf, and Richard K. Plemper. 2021. “Therapeutically Administered Ribonucleoside Analogue MK-4482/EIDD-2801 Blocks SARS-CoV-2 Transmission in Ferrets.” Nature Microbiology 6 (1): 11-18.

  • Daina, Antoine, Olivier Michielin, and Vincent Zoete. 2017. “SwissADME: A Free Web Tool to Evaluate Pharmacokinetics, Drug-Likeness and Medicinal Chemistry Friendliness of Small Molecules.” Scientific Reports 7 (March): 42717.

  • Deeks, Emma D. 2014. “Cobicistat: A Review of Its Use as a Pharmacokinetic Enhancer of Atazanavir and Darunavir in Patients with HIV-1 Infection.” Drugs 74 (2): 195-206.

  • Fehr A R, Perlman S. Coronaviruses: An overview of their replication and pathogenesis. Coronaviruses: Methods and Protocols. 2015; 1-23.

  • Feoktistova, Maria, Peter Geserick, and Martin Leverkus. 2016. “Crystal Violet Assay for Determining Viability of Cultured Cells.” Cold Spring Harbor Protocols 2016 (4): db.prot087379.

  • Fischl, Wolfgang, and Ralf Bartenschlager. 2013. “High-Throughput Screening Using Dengue Virus Reporter Genomes.” Methods in Molecular Biology 1030: 205-19.

  • Friesner, Richard A., Jay L. Banks, Robert B. Murphy, Thomas A. Halgren, Jasna J. Klicic, Daniel T.

  • Mainz, Matthew P. Repasky, et al. 2004. “Glide: A New Approach for Rapid, Accurate Docking and Scoring. 1. Method and Assessment of Docking Accuracy.” Journal of Medicinal Chemistry 47 (7): 1739-49.

  • Gao, B., Yang, F. M., Yu, Z. T., Li, R., Xie, F., Chen, J., Luo, H. J., & Zhang, J. C. (2015). Relationship between the expression of MDR1 in hepatocellular cancer and its biological behaviors. International journal of clinical and experimental pathology, 8(6), 6995-7001

  • Gulick, Roy M., and Charles Flexner. 2019. “Long-Acting HIV Drugs for Treatment and Prevention.” Annual Review of Medicine. https://doi.org/10.1146/annurev-med-041217-013717.

  • Gupta, Manoj Kumar, Sarojamma Vemula, Ravindra Donde, Gayatri Gouda, Lambodar Behera, and Ramakrishna Vadde. 2020. “Approaches to Detect Inhibitors of the Human Severe Acute Respiratory Syndrome Coronavirus Envelope Protein Ion Channel” Journal of Biomolecular Structure & Dynamics, April, 1-11.

  • Guy, R. Kiplin, R. Kiplin Guy, Robert S. DiPaola, Frank Romanelli, and Rebecca E. Dutch. 2020. “Rapid Repurposing of Drugs for COVID-19.” Science. https://doi.org/10.1126/science.abb9332.

  • Hoffmann, Markus, Hannah Kleine-Weber, and Stefan Pöhlmann. 2020. “A Multibasic Cleavage Site in the Spike Protein of SARS-CoV-2 Is Essential for Infection of Human Lung Cells.”Molecular Cell 78 (4): 779-84.e5.

  • Hoffmann, Markus, Hannah Kleine-Weber, Simon Schroeder, Nadine Kruger, Tanja Herrler, Sandra Erichsen, Tobias S. Schiergens, et al. 2020. “SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and Is Blocked by a Clinically Proven Protease Inhibitor.” Cell 181 (2): 271-80.e8.

  • Hu, Wen-Juan, Lu Chang, Ying Yang, Xin Wang, Yuan-Chao Xie, Jing-Shan Shen, Bo Tan, and Jia Liu. 2020. “Pharmacokinetics and Tissue Distribution of Remdesivir and Its Metabolites Nucleotide Monophosphate, Nucleotide Triphosphate, and Nucleoside in Mice.” Acta Pharmacologica Sinica, October. https://doi.org/10.1038/s41401-020-00537-9.

  • Ianevski, Aleksandr, Liye He, Tero Aittokallio, and Jing Tang. 2020. “SynergyFinder: A Web Application for Analyzing Drug Combination Dose-Response Matrix Data.” Bioinformatics 36 (8): 2645.

  • Jácome, Rodrigo, Jose Alberto Campillo-Balderas, Samuel Ponce de Leon, Arturo Becerra, and Antonio Lazcano. 2020. “Sofosbuvir as a Potential Alternative to Treat the SARS-CoV-2 Epidemic.” Scientific Reports 10 (1): 9294.

  • Jiang S, Hillyer C, Du L. Neutralizing Antibodies against SARS-CoV-2 and Other Human Coronaviruses. Trends in Immunology. 2020; July-August; 14(4): 407-412.

  • Jin, Zhenming, Xiaoyu Du, Yechun Xu, Yongqiang Deng, Meiqin Liu, Yao Zhao, Bing Zhang, et al. 2020. “Structure of M pro from SARS-CoV-2 and Discovery of Its Inhibitors.” Nature 582 (7811): 289-93.

  • Jockusch, Steffen, Chuanjuan Tao, Xiaoxu Li, Thomas K. Anderson, Minchen Chien, Shiv Kumar, James J. Russo, Robert N. Kirchdoerfer, and Jingyue Ju. 2020. “A Library of Nucleotide Analogues Terminate RNA Synthesis Catalyzed by Polymerases of Coronaviruses That Cause SARS and COVID-19.” Antiviral Research 180 (August): 104857.

  • Johnson, Raymond M., and Joseph M. Vinetz. 2020. “Dexamethasone in the Management of Covid-19.” BMJ. https://doi.org/10.1136/bmj.m2648.

  • Jorgensen, Sarah C. J., Razieh Kebriaei, and Linda D. Dresser. 2020. “Remdesivir: Review of Pharmacology, Pre-Clinical Data, and Emerging Clinical Experience for COVID-19.” Pharmacotherapy 40 (7): 659-71.

  • Josephson, F. 2010. “Drug-Drug Interactions in the Treatment of HIV Infection: Focus on Pharmacokinetic Enhancement through CYP3A Inhibition.” Journal of Internal Medicine 268 (6): 530-39.

  • Kakuda, Thomas N., Tom Van De Casteele, Romana Petrovic, Mark Neujens, Hiba Salih, Magda Opsomer, and Richard Mw Hoetelmans. 2014. “Bioequivalence of a Darunavir/cobicistat Fixed-Dose Combination Tablet versus Single Agents and Food Effect in Healthy Volunteers.” Antiviral Therapy 19 (6): 597-606.

  • Kaptein, Suzanne J. F., Sofie Jacobs, Lana Langendries, Laura Seldeslachts, Sebastiaan ter Horst, Laurens Liesenborghs, Bart Hens, et al. 2020. “Favipiravir at High Doses Has Potent Antiviral Activity in SARS-CoV-2-infected Hamsters, Whereas Hydroxychloroquine Lacks Activity.” Proceedings of the National Academy of Sciences. https://doi.org/10.1073/pnas.2014441117.

  • Kim, Kyoung-Ah, Ji-Young Park, Ji-Suk Lee, and Sabina Lim. 2003. “Cytochrome P450 2C8 and CYP3A4/5 Are Involved in Chloroquine Metabolism in Human Liver Microsomes.” Archives of Pharmacal Research 26 (8): 631-37.

  • Klein, Steffen, Mirko Cortese, Sophie L. Winter, Moritz Wachsmuth-Melm, Christopher J. Neufeldt, Berati Cerikan, Megan L. Stanifer, Steeve Boulant, Ralf Bartenschlager, and Petr Chlanda. 2020. “SARS-CoV-2 Structure and Replication Characterized by in Situ Cryo-Electron Tomography.” Nature Communications 11 (1): 5885.

  • Leegwater, Emiel, Anne Strik, Erik B. Wilms, Liesbeth B. E. Bosma, David M. Burger, Thomas H. Ottens, and Cees van Nieuwkoop. 2020. “Drug-Induced Liver Injury in a Patient With Coronavirus Disease 2019: Potential Interaction of Remdesivir With P-Glycoprotein Inhibitors.” Clinical Infectious Diseases. https://doi.org/10.1093/cid/ciaa883.

  • Lepist, Eve-Irene, Truc K. Phan, Anupma Roy, Leah Tong, Kelly Maclennan, Bernard Murray, and Adrian S. Ray. 2012. “Cobicistat Boosts the Intestinal Absorption of Transport Substrates, Including HIV Protease Inhibitors and GS-7340, in Vitro.” Antimicrobial Agents and Chemotherapy 56 (10): 5409-13.

  • Liu, Jia, Ruiyuan Cao, Mingyue Xu, Xi Wang, Huanyu Zhang, Hengrui Hu, Yufeng Li, Zhihong Hu, Wu Zhong, and Manli Wang. 2020. “Hydroxychloroquine, a Less Toxic Derivative of Chloroquine, Is Effective in Inhibiting SARS-CoV-2 Infection in Vitro.” Cell Discovery 6 (March): 16.

  • Livak, K. J., and T. D. Schmittgen. 2001. “Analysis of Relative Gene Expression Data Using Real-Time Quantitative PCR and the 2(-Delta Delta C(T)) Method.” Methods 25 (4): 402-8.

  • Ma, Chunlong, Michael Dominic Sacco, Brett Hurst, Julia Alma Townsend, Yanmei Hu, Tommy Szeto, Xiujun Zhang, et al. 2020. “Boceprevir, GC-376, and Calpain Inhibitors II, XII Inhibit SARS-CoV-2 Viral Replication by Targeting the Viral Main Protease.” Cell Research 30 (8): 678-92.

  • Mandi, Mohamed, János András Mótyán, Zsófia Ilona Szojka, Mária Golda, Márió Miczi, and József Tözsér. 2020. “Analysis of the Efficacy of HIV Protease Inhibitors against SARS-CoV-2's Main Protease.” Virology Journal 17 (1): 190.

  • Maier, James A., Carmenza Martinez, Koushik Kasavajhala, Lauren Wickstrom, Kevin E. Hauser, and Carlos Simmerling. 2015. “ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB.” Journal of Chemical Theory and Computation 11 (8): 3696-3713.

  • Mathias, A. A., P. German, B. P. Murray, L. Wei, A. Jain, S. West, D. Warren, J. Hui, and B. P. Kearney. 2010. “Pharmacokinetics and Pharmacodynamics of GS-9350: A Novel Pharmacokinetic Enhancer without Anti-HIV Activity.” Clinical Pharmacology and Therapeutics 87 (3): 322-29.

  • Miller, Bill R., 3rd, T. Dwight McGee Jr, Jason M. Swails, Nadine Homeyer, Holger Gohlke, and Adrian E. Roitberg. 2012. “MMPBSA.py: An Efficient Program for End-State Free Energy Calculations.” Journal of Chemical Theory and Computation 8 (9): 3314-21.

  • Mulangu, Sabue, Lori E. Dodd, Richard T. Davey Jr, Olivier Tshiani Mbaya, Michael Proschan, Daniel Mukadi, Mariano Lusakibanza Manzo, et al. 2019. “A Randomized, Controlled Trial of Ebola Virus Disease Therapeutics.” The New England Journal of Medicine 381 (24): 2293-2303.

  • Naggie, Susanna, and Andrew J. Muir. 2017. “Oral Combination Therapies for Hepatitis C Virus Infection: Successes, Challenges, and Unmet Needs.” Annual Review of Medicine 68 (January): 345-58.

  • Ou, Xiuyuan, Yan Liu, Xiaobo Lei, Pei Li, Dan Mi, Lili Ren, Li Guo, et al. 2020. “Characterization of Spike Glycoprotein of SARS-CoV-2 on Virus Entry and Its Immune Cross-Reactivity with SARS-CoV.” Nature Communications 11 (1): 1620.

  • Pal M, Berhanu G, Desalegn C, Kandi V. Severe Acute Respiratory Syndrome Coronavirus-2 (SARS-CoV-2): An Update. Cureus. 2020; March; 12(3): e7423.

  • Pape, Constantin, Roman Remme, Adrian Wolny, Sylvia Olberg, Steffen Wolf, Lorenzo Cerrone,

  • Mirko Cortese, et al. 2020. “Microscopy-based Assay for Semi-quantitative Detection of SARS-CoV-2 Specific Antibodies in Human Sera.” BioEssays. https://doi.org/10.1002/bies.202000257.

  • Pawlotsky, Jean-Michel, Jordan J. Feld, Stefan Zeuzem, and Jay H. Hoofnagle. 2015. “From Non-A, Non-B Hepatitis to Hepatitis C Virus Cure.” Journal of Hepatology 62 (1 Suppl): S87-99.

  • Pires, Douglas E. V., Tom L. Blundell, and David B. Ascher. 2015. “pkCSM: Predicting Small-Molecule Pharmacokinetic and Toxicity Properties Using Graph-Based Signatures.” Journal of Medicinal Chemistry 58 (9): 4066-72.

  • Pushpakom, Sudeep, Francesco Iorio, Patrick A. Eyers, K. Jane Escott, Shirley Hopper, Andrew Wells, Andrew Doig, et al. 2019. “Drug Repurposing: Progress, Challenges and Recommendations.” Nature Reviews. Drug Discovery 18 (1): 41-58.

  • Razzaghi-Asl, Nima, Ahmad Ebadi, Sara Shahabipour, and Danial Gholamin. 2020. “Identification of a Potential SARS-CoV2 Inhibitor via Molecular Dynamics Simulations and Amino Acid Decomposition Analysis.” Journal of Biomolecular Structure & Dynamics, July, 1-16.

  • RECOVERY Collaborative Group, Peter Horby, Wei Shen Lim, Jonathan R. Emberson, Marion

  • Mafham, Jennifer L. Bell, Louise Linsell, et al. 2020. “Dexamethasone in Hospitalized Patients with Covid-19—Preliminary Report.” The New England Journal of Medicine, July. https://doi.org/10.1056/NEJMoa2021436.

  • RECOVERY Collaborative Group, Peter Horby, Marion Mafham, Louise Linsell, Jennifer L. Bell, Natalie Staplin, Jonathan R. Emberson, et al. 2020. “Effect of Hydroxychloroquine in Hospitalized Patients with Covid-19.” The New England Journal ofMedicine 383 (21): 2030-40.

  • Richter, Oliver von, Oliver Burk, Martin F. Fromm, Klaus P. Thon, Michel Eichelbaum, and Kari T. Kivistö. 2004. “Cytochrome P450 3A4 and P-Glycoprotein Expression in Human Small Intestinal Enterocytes and Hepatocytes: A Comparative Analysis in Paired Tissue Specimens.” Clinical Pharmacology and Therapeutics 75 (3): 172-83.

  • Roe, Daniel R., and Thomas E. Cheatham 3rd. 2013. “PTRAJ and CPPTRAJ: Software for Processing and Analysis of Molecular Dynamics Trajectory Data.” Journal of Chemical Theory and Computation 9 (7): 3084-95.

  • Sherman, Elizabeth M., Marylee V. Worley, Nathan R. Unger, Timothy P. Gauthier, and Jason J. Schafer. 2015. “Cobicistat: Review of a Pharmacokinetic Enhancer for HIV Infection.” Clinical Therapeutics 37 (9): 1876-93.

  • Shytaj, Tart Luca, Bojana Lucic, Mattia Forcato, Carlotta Penzo, James Billingsley, Vibor Laketa, Steven Bosinger, et al. 2020. “Alterations of Redox and Iron Metabolism Accompany the Development of HIV Latency.” The EMBO Journal 39 (9): e102209.

  • Siegel, Dustin, Hon C. Hui, Edward Doerffler, Michael O. Clarke, Kwon Chun, Lijun Zhang, Sean Neville, et al. 2017. “Discovery and Synthesis of a Phosphoramidate Prodrug of a Pyrrolo[2,1-F][triazin-4-Amino] Adenine C-Nucleoside (GS-5734) for the Treatment of Ebola and Emerging Viruses.” Journal of Medicinal Chemistry. https://doi.org/10.102 Vacs medchem.6b01594.

  • Stanifer, Megan L., Carmon Kee, Mirko Cortese, Camila Metz Zumaran, Sergio Triana, Markus Mukenhirn, Hans-Georg Kraeusslich, Theodore Alexandrov, Ralf Bartenschlager, and Steeve Boulant. 2020. “Critical Role of Type III Interferon in Controlling SARS-CoV-2 Infection in Human Intestinal Epithelial Cells.” Cell Reports 32 (1): 107863.

  • Tian, Siyang, Yannick Djoumbou-Feunang, Russell Greiner, and David S. Wishart. 2018. “CypReact: A Software Tool for in Silico Reactant Prediction for Human Cytochrome P450 Enzymes.” Journal of Chemical Information and Modeling 58 (6): 1282-91.

  • V'kovski, Philip, Annika Kratzel, Silvio Steiner, Hanspeter Stalder, and Volker Thiel. 2020. “Coronavirus Biology and Replication: Implications for SARS-CoV-2.” Nature Reviews. Microbiology, October. https://doi.org/10.1038/s41579-020-00468-6.

  • von Hentig N. Clinical use of cobicistat as a pharmacoenhancer of human immunodeficiency virus therapy. HIV/AIDS—Research and Palliative Care. 2015; 8:1-16.

  • Wang, Junmei, Romain M. Wolf, James W. Caldwell, Peter A. Kollman, and David A. Case. 2004. “Development and Testing of a General Amber Force Field.” Journal of Computational Chemistry 25 (9): 1157-74.

  • Wang, Manli, Ruiyuan Cao, Leike Zhang, Xinglou Yang, Jia Liu, Mingyue Xu, Zhengli Shi, Zhihong Hu, Wu Zhong, and Gengfu Xiao. 2020. “Remdesivir and Chloroquine Effectively Inhibit the Recently Emerged Novel Coronavirus (2019-nCoV) in Vitro.” Cell Research 30 (3): 269-71.

  • Wang, Yeming, Dingyu Zhang, Guanhua Du, Ronghui Du, Jianping Zhao, Yang Jin, Shouzhi Fu, et al. 2020. “Remdesivir in Adults with Severe COVID-19: A Randomised, Double-Blind, Placebo-Controlled, Multicentre Trial.” The Lancet 395 (10236): 1569-78.

  • Waterschoot, R. A. B. van, R. ter Heine, E. Wagenaar, C. M. M. van der Kruijssen, R. W. Rooswinkel, A. D. R. Huitema, J. H. Beijnen, and A. H. Schinkel. 2010. “Effects of Cytochrome P450 3A (CYP3A) and the Drug Transporters P-Glycoprotein (MDR1/ABCB1) and MRP2 (ABCC2) on the Pharmacokinetics of Lopinavir.” British Journal of Pharmacology 160 (5): 1224-33.

  • Wessler, Jeffrey D., Laura T. Grip, Jeanne Mendell, and Robert P. Giugliano. 2013. “The P-Glycoprotein Transport System and Cardiovascular Drugs.” Journal of the American College of Cardiology 61 (25): 2495-2502.

  • White, Kris M., Romel Rosales, Soner Yildiz, Thomas Kehrer, Lisa Miorin, Elena Moreno, Sonia Jangra, et al. 2021. “Plitidepsin Has Potent Preclinical Efficacy against SARS-CoV-2 by Targeting the Host Protein eEF1A.” Science, January. https://doi.org/10.1126/science.abf4058.

  • Xiu, Siyu, Alexej Dick, Han Ju, Sako Mirzaie, Fatemeh Abdi, Simon Cocklin, Peng Zhan, and Xinyong Liu. 2020. “Inhibitors of SARS-CoV-2 Entry: Current and Future Opportunities.” Journal of Medicinal Chemistry 63 (21): 12256-74.

  • Xu L, Liu H, Murray B P, Callebaut C, Lee M S, Hong A, Strickley R G, Tsai L K, Stray K M, Wang Y, Rhodes G R, Desai M C. Cobicistat (GS-9350): A Potent and Selective Inhibitor of Human CYP3A as a Novel Pharmacoenhancer. ACS Med Chem Lett. 2010 Aug. 12; 1(5): 209-213.

  • Yadav, Bhagwan, Krister Wennerberg, Tero Aittokallio, and Jing Tang. 2015. “Searching for Drug Synergy in Complex Dose-Response Landscapes Using an Interaction Potency Model.” Computational and Structural Biotechnology Journal 13 (September): 504-13.

  • Yamamoto, Norio, Shutoku Matsuyama, Tyuji Hoshino, and Naoki Yamamoto. n.d. “Nelfinavir Inhibits Replication of Severe Acute Respiratory Syndrome Coronavirus 2 in Vitro.” https://doi.org/10.1101/2020.04.06.026476.

  • Yamamoto, Norio, Rongge Yang, Yoshiyuki Yoshinaka, Shinji Amari, Tatsuya Nakano, Jindrich Cinatl, Holger Rabenau, et al. 2004. “HIV Protease Inhibitor Nelfinavir Inhibits Replication of SARS-Associated Coronavirus.” Biochemical and Biophysical Research Communications 318 (3): 719-25.

  • Yang, Katherine. 2020. “What Do We Know About Remdesivir Drug Interactions?” Clinical and Translational Science 13 (5): 842-44.

  • Zanger, Ulrich M., and Matthias Schwab. 2013. “Cytochrome P450 Enzymes in Drug Metabolism: Regulation of Gene Expression, Enzyme Activities, and Impact of Genetic Variation.” Pharmacology & Therapeutics 138 (1): 103-41.

  • Zhang, Linlin, Daizong Lin, Xinyuanyuan Sun, Ute Curth, Christian Drosten, Lucie Sauerhering, Stephan Becker, Katharina Rox, and Rolf Hilgenfeld. 2020. “Crystal Structure of SARS-CoV-2 Main Protease Provides a Basis for Design of Improved a-Ketoamide Inhibitors.” Science 368 (6489): 409-12.


Claims
  • 1. A composition comprising cobicistat, or a derivative or prodrug thereof, for use in the prophylaxis and/or treatment of severe acute respiratory syndrome coronavirus type 2 (SARS-CoV-2) infection, severe acute respiratory syndrome coronavirus (SARS-CoV) infection and/or Middle East respiratory syndrome coronavirus (MERS-CoV) infection, wherein said derivative or prodrug is ritonavir or desoxy-ritonavir, and wherein said composition comprises a further drug selected from remdesivir, chloroquine, hydroxychloroquine, molnupiravir, tipranavir, nelfinavir, lopinavir, atazanavir, plitidepsin, favipiravir, an anti-inflammatory glucocorticoid, januskinase (JAK) inhibitor, a palmitoyl protein thioesterase 1 (PPT1) inhibitor, and a monoclonal antibody targeting viral replication or host inflammation.
  • 2-5. (canceled)
  • 6. The composition according to claim 1, wherein said further drug is selected from dexamethasone, prednisone, methylprednisolone, hydrocortisone, baricitinib, ruxolitinib, upadacitinib, GNS561, and tocilizumab.
  • 7. The composition according to claim 1, wherein the further drug is remdesivir.
  • 8. The composition according to claim 1, wherein the further drug is at least one of remdesivir, tipranavir, chloroquine, hydroxychloroquine, molnupiravir, favipiravir, nelfinavir, lopinavir, atazanavir, plitidepsin, dexamethasone, baricitinib, and GNS561.
  • 9. The composition according to claim 6, further comprising one or more further drug, selected from tipranavir, nelfinavir, lopinavir, and atazanavir.
  • 10. The composition according to claim 1, comprising cobicistat, or a derivative or prodrug thereof, and chloroquine in combination with one or more further drugs selected from tipranavir, nelfinavir, lopinavir, and atazanavir.
  • 11. The composition according to claim 1, wherein cobicistat is present in a therapeutically effective amount, which is higher than the dosage used for HIV-1 treatment.
  • 12-16. (canceled)
  • 17. A method of prevention and/or treatment of severe acute respiratory syndrome coronavirus type 2 (SARS-CoV-2) infection, severe acute respiratory syndrome coronavirus (SARS-CoV) infection and/or Middle East respiratory syndrome coronavirus (MERS-CoV) infection, comprising the step of: administering, to a subject in need of such prevention and/or treatment, a therapeutically effective amount of cobicistat or a derivative or prodrug thereofwherein said derivative or prodrug is ritonavir or desoxy-ritonavir.
  • 18. The method of claim 17, wherein the therapeutically amount is higher than the dosage used for HIV-1 treatment.
  • 19. The method of claim 17, wherein administration of cobicistat is oral, at a daily dosage in the range from 300 to 1,000 mg, or wherein administration of cobicistat is intranasal and/or via inhalation and the administration is via a dry powder inhaler, or via a nebulizer or a soft mist spray dispenser.
  • 20. (canceled)
  • 21. The method of claim 19, wherein administration is intranasal and/or inhalation and the amount administered is in the range from 2 to 15 μM per day.
  • 22. The method according to claim 17, wherein prophylaxis comprises pre- and post-exposure prophylaxis to SARS-CoV-2 infection.
  • 23. The method according to claim 17, wherein said cobicistat, or derivative or prodrug thereof, is administered in combination with one or more further drug.
  • 24. The method according to claim 23, wherein the one or more further drug is selected from remdesivir, chloroquine, hydroxychloroquine, molnupiravir, and favipiravir.
  • 25. The method according to claim 23, wherein the one or more further drug is selected from tipranavir, nelfinavir, lopinavir, and atazanavir.
  • 26. The method according to claim 23, wherein the one or more further drug is selected from plitidepsin, dexamethasone, prednisone, methylprednisolone, hydrocortisone, baricitinib, ruxolitinib, upadacitinib, GNS561, and tocilizumab.
  • 27. The method according to claim 23, wherein the further drug is remdesivir.
  • 28. The method according to claim 23, wherein the one or more further drug is selected from remdesivir, tipranavir, chloroquine, hydroxychloroquine, molnupiravir, favipiravir, nelfinavir, lopinavir, atazanavir, plitidepsin, dexamethasone, baricitinib, and GNS561.
  • 29. The method according to claim 23, wherein cobicistat is administered in combination with remdesivir and further in combination with one or more further drug, selected from tipranavir, nelfinavir, lopinavir, and atazanavir.
  • 30. The method according to claim 23, wherein cobicistat is administered in combination with chloroquine in combination with one or more further drug selected from tipranavir, nelfinavir, lopinavir, and atazanavir.
PCT Information
Filing Document Filing Date Country Kind
PCT/EP2021/055621 3/5/2021 WO
Provisional Applications (1)
Number Date Country
63040758 Jun 2020 US