Skip to main content
Springer Nature - PMC COVID-19 Collection logoLink to Springer Nature - PMC COVID-19 Collection
. 2016 Feb 12;15(5):327–347. doi: 10.1038/nrd.2015.37

Coronaviruses — drug discovery and therapeutic options

Alimuddin Zumla 1,#, Jasper F W Chan 2,#, Esam I Azhar 3,#, David S C Hui 4,#, Kwok-Yung Yuen 2,✉,#
PMCID: PMC7097181  PMID: 26868298

Key Points

  • Severe acute respiratory syndrome (SARS) and Middle East respiratory syndrome (MERS) are examples of emerging zoonotic coronavirus infections capable of person-to-person transmission that result in large-scale epidemics with substantial effects on patient health and socioeconomic factors. Unlike patients with mild illnesses that are caused by other human-pathogenic coronaviruses, patients with SARS or MERS coronavirus infections may develop severe acute respiratory disease with multi-organ failure. The case–fatality rates of SARS and MERS are approximately 10% and 35%, respectively.

  • Both SARS and MERS pose major clinical management challenges because there is no specific antiviral treatment that has been proven to be effective in randomized clinical trials for either infection. Substantial efforts are underway to discover new therapeutic agents for coronavirus infections.

  • Virus-based therapies include monoclonal antibodies and antiviral peptides that target the viral spike glycoprotein, viral enzyme inhibitors, viral nucleic acid synthesis inhibitors and inhibitors of other viral structural and accessory proteins.

  • Host-based therapies include agents that potentiate the interferon response or affect either host signalling pathways involved in viral replication or host factors utilized by coronaviruses for viral replication.

  • The major challenges in the clinical development of novel anti-coronavirus drugs include the limited number of suitable animal models for the evaluation of potential treatments for SARS and MERS, the current absence of new SARS cases, the limited number of MERS cases — which are also predominantly geographically confined to the Middle East — as well as the lack of industrial incentives to develop antivirals for mild infections caused by other, less pathogenic coronaviruses.

  • The continuing threat of MERS-CoV to global health 3 years after its discovery presents a golden opportunity to tackle current obstacles in the development of new anti-coronavirus drugs. A well-organized, multidisciplinary, international collaborative network consisting of clinicians, virologists and drug developers, coupled to political commitment, should be formed to carry out clinical trials using anti-coronavirus drugs that have already been shown to be safe and effective in vitro and/or in animal models, particularly lopinavir–ritonavir, interferon beta-1b and monoclonal antibodies and antiviral peptides targeting the viral spike glycoprotein.

Supplementary information

The online version of this article (doi:10.1038/nrd.2015.37) contains supplementary material, which is available to authorized users.

Subject terms: SARS virus, Vaccines, Drug discovery, Infectious diseases


Severe acute respiratory syndrome (SARS) and Middle East respiratory syndrome (MERS), which are caused by coronaviruses, have attracted substantial attention owing to their high mortality rates and potential to cause epidemics. Yuen and colleagues discuss progress with treatment options for these syndromes, including virus- and host-targeted drugs, and the challenges that need to be overcome in their further development.

Supplementary information

The online version of this article (doi:10.1038/nrd.2015.37) contains supplementary material, which is available to authorized users.

Abstract

In humans, infections with the human coronavirus (HCoV) strains HCoV-229E, HCoV-OC43, HCoV-NL63 and HCoV-HKU1 usually result in mild, self-limiting upper respiratory tract infections, such as the common cold. By contrast, the CoVs responsible for severe acute respiratory syndrome (SARS) and Middle East respiratory syndrome (MERS), which were discovered in Hong Kong, China, in 2003, and in Saudi Arabia in 2012, respectively, have received global attention over the past 12 years owing to their ability to cause community and health-care-associated outbreaks of severe infections in human populations. These two viruses pose major challenges to clinical management because there are no specific antiviral drugs available. In this Review, we summarize the epidemiology, virology, clinical features and current treatment strategies of SARS and MERS, and discuss the discovery and development of new virus-based and host-based therapeutic options for CoV infections.

Supplementary information

The online version of this article (doi:10.1038/nrd.2015.37) contains supplementary material, which is available to authorized users.

Main

Coronaviruses (CoVs; subfamily Coronavirinae, family Coronaviridae, order Nidovirales) are a group of highly diverse, enveloped, positive-sense, single-stranded RNA viruses that cause respiratory, enteric, hepatic and neurological diseases of varying severity in a broad range of animal species, including humans1,2,3. CoVs are subdivided into four genera: Alphacoronavirus, Betacoronavirus (βCoV), Gammacoronavirus and Deltacoronavirus2,3,4,5,6,7. Over the past 12 years, two novel βCoVs, severe acute respiratory syndrome CoV (SARS-CoV) and Middle East respiratory syndrome CoV (MERS-CoV), have emerged, and these viruses can cause severe human diseases8,9. The lack of effective drug treatment and associated high morbidity and mortality rates of these two CoVs as well as their potential to cause epidemics highlight the need for novel drug discovery for the treatment of CoV infections.

Epidemiology of SARS and MERS

SARS. SARS-CoV emerged first in southern China and rapidly spread around the globe in 2002–2003 (Refs 8,10,11). In November 2002, an unusual epidemic of atypical pneumonia with a high rate of nosocomial transmission to health-care workers occurred in Foshan, Guangdong, China12,13. In March 2003, a novel CoV was confirmed to be the causative agent for SARS, and was thus named SARS-CoV14,15,16,17. A 64-year-old nephrologist who travelled from southern China to Hong Kong on 21 February 2003 became the index case of subsequent large community and health-care-associated outbreaks of SARS in Hong Kong and other regions10,11,18,19,20,21. The high infectivity of SARS was highlighted by the super-spreading event at a major teaching hospital in Hong Kong in which 138 people, including many previously healthy health-care workers, were infected within 2 weeks of exposure to an index patient who was being managed in a general medical ward for community-acquired pneumonia10,22. Through international air travel, SARS-CoV was spread to 29 countries and regions with a total of 8,098 cases and 774 fatalities (9.6% of cases) by the end of the epidemic in July 2003 (Ref. 23) (see Supplementary information S1 (figure, parts a,b)).

A retrospective serological survey suggested that cross-species transmission of SARS-CoV or its variants from animal species to humans might have occurred frequently in the wet market, where high seroprevalence was detected among asymptomatic animal handlers24. A close variant of SARS-CoV was isolated in palm civets in Dongmen market, Shenzhen, China, in 2003 (Ref. 25). During the small-scale SARS outbreaks in late 2003 and early 2004, three of the four patients had direct or indirect contact with palm civets26,27. However, viral genetic sequence analysis demonstrated that the SARS-CoV-like virus had not been circulating among masked palm civets in markets for a long time, and a serological study showed that only caged market civets and not wild civets were infected with the SARS-CoV-like virus28. CoVs that are highly similar to SARS-CoV have been isolated from Chinese horseshoe bats since 2005 (Refs 29,30,31,32). These SARS-like CoVs from bats share 88–95% sequence homology with human or civet CoV isolates, which suggests that bats were probably the natural reservoir of a close ancestor of SARS-CoV4,33,34.

MERS. The isolation of a novel βCoV from a patient in Jeddah, Saudi Arabia, who died of severe pneumonia and multi-organ failure in June 2012, was first reported in September 2012 (Ref. 35). Initially named 'human coronavirus Erasmus Medical Center', the virus was later renamed MERS-CoV by international consensus, and the disease was called Middle East respiratory syndrome (MERS)36. Retrospective analysis of a cluster of nosocomial cases in April 2012 in Jordan confirmed that MERS-CoV was also responsible for that outbreak37. Over the past 3 years, MERS-CoV has continued to spread within and beyond the Middle East, and there are ongoing reports of sporadic cases and community and health-care-associated clusters of infected individuals in the Middle East, especially in Saudi Arabia and the United Arab Emirates9,38. Travel-related cases and clusters have also been increasingly reported on other continents9. As of 9 October 2015, 1,593 laboratory-confirmed cases of MERS, including 568 deaths, have been reported to the World Health Organization39 (see Supplementary information S1 (figure, parts c,d)).

MERS-CoV is considered primarily to be a zoonotic virus that has the capability of non-sustained person-to-person spread9. Serological and virological studies have shown that camels and bats are the most likely animal reservoirs of MERS-CoV9,40,41,42,43,44,45,46,47. Although not all primary cases of MERS were individuals who had direct contact with camels, such exposure is considered to be an important factor for the spread of MERS-CoV, as evidenced by the substantially increased seroprevalence of anti-MERS-CoV antibodies among individuals with occupational exposure to camels, such as camel shepherds and slaughterhouse workers, relative to the general population in Saudi Arabia48,49. Person-to-person transmission of MERS-CoV has occurred in health-care facilities and family clusters50,51,52,53. The recent, large health-care-associated outbreaks in Jeddah and South Korea have been attributed to poor compliance with infection control measures54,55. Further studies are needed to fully understand the exact mode of transmission and other potential sources of MERS-CoV for optimization of treatment and prevention strategies for MERS56,57.

Clinical features of SARS and MERS. Patients with SARS or MERS present with various clinical features, ranging from asymptomatic or mild respiratory illness to fulminant severe acute respiratory disease with extrapulmonary manifestations8,9. Both diseases have predominantly respiratory manifestations, but extrapulmonary features may occur in severe cases56 (see Supplementary information S2 (table)). Notably, early treatment is especially important for patients with severe MERS because this disease progresses to respiratory distress, renal failure and death much more rapidly than SARS does. The three- to four-fold higher case-fatality rate of MERS relative to SARS may be related to the higher median age and prevalence of comorbidities in patients with MERS as well as the different pathogenesis of the two diseases9,58,59,60,61. Comorbidities associated with severe MERS include obesity, diabetes mellitus, systemic immunocompromising conditions and chronic cardiac and pulmonary diseases9,60,62,63. Although the rate of secondary transmission among household contacts of index MERS patients (which is approximately 4%) and the estimated pandemic potential of MERS-CoV are lower than those of SARS-CoV, the rapidly progressive clinical course and high fatality of MERS continues to pose a major threat to at-risk populations64,65,66,67,68,69,70,71 (see Supplementary information S2 (table)).

Current management strategies for SARS and MERS. Supportive care — including organ support and prevention of complications, especially acute respiratory distress syndrome, organ failure and secondary nosocomial infections — remains the most important management strategy for SARS and MERS, as there is currently no specific antiviral treatment that has been proven to be effective in randomized controlled trials8,9,56,72,73,74,75. Numerous compounds have been found to inhibit the entry and/or replication of SARS-CoV and MERS-CoV in cell culture or in animal models, but activity in vitro and even in animal experiments does not necessarily translate into efficacy in humans8,9. Owing to the high morbidity and mortality rates of SARS and MERS, some of these antiviral drugs and immunomodulators have been used empirically or evaluated in uncontrolled trials8,10,21,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90 (Table 1). Substantial efforts are underway to discover new therapeutic agents for CoV infections and these investigations are based on our understanding of the basic virology of CoVs. Importantly, treatment with these investigational therapies requires application of standard research treatment protocols and systematic clinical and virological data collection in controlled research trials, with the approval of the local ethics committee.

Table 1. Therapeutic interventions used in patients with SARS and MERS.

Type of intervention Therapeutic intervention Treatment effects Refs
Treatments used for SARS patients
Antivirals Ribavirin No significant effect on clinical outcome 10,21
Ribavirin, lopinavir–ritonavir + corticosteroids Patients who received ribavirin, lopinavir–ritonavir and a corticosteroid had lower 21-day ARDS and death rates than those who received ribavirin and a corticosteroid 76,77
Interferon combination Interferon alfa-1 + corticosteroid Associated with improved oxygen saturation and more rapid resolution of radiographic lung opacities than systemic corticosteroid alone (uncontrolled study) 78
Corticosteroids Pulsed methylprednisolone Associated with an increased 30-day mortality rate (adjusted OR = 26.0, 95% CI = 4.4–154.8). Disseminated fungal infection and avascular osteonecrosis occurred following prolonged systemic corticosteroid therapy 79,80,81
A randomized, placebo-controlled study showed that plasma SARS-CoV RNA levels in weeks 2–3 of the illness were higher in patients given hydrocortisone (n = 10) than those given normal saline (n = 7) in the early phase of the illness, suggesting that early use of pulsed methylprednisolone might prolong viraemia 82
Convalescent-phase plasma Convalescent-phase plasma therapy Has been used for severe respiratory tract infections including SARS and influenza. A systematic review and exploratory meta-analysis of patients with SARS or influenza treated with convalescent-phase plasma showed a reduction in mortality, but the treatment success was determined by its availability and timely administration 85,272,273
Among 80 non-randomized SARS patients who were given convalescent-phase plasma, the discharge rate at day 22 was 58.3% for patients (n = 48) treated within 14 days of illness onset versus 15.6% for those (n = 32) treated beyond 14 days 83,84
Treatments used for MERS patients
Combination of antivirals and interferons Ribavirin + interferon alfa-2a or interferon alfa-2b No significant effect on clinical outcome; case–control study showed significantly improved survival (14 out of 20 and 7 out of 24 in the treated and control groups, respectively; P = 0.004) at 14 days, but not at 28 days 86,87,88,89
Ribavirin + interferon beta-1a Retrospective analyses showed no significant effect on clinical outcome 89
Ribavirin, lopinavir–ritonavir + interferon alfa-2a Viraemia resolved 2 days after commencement of treatment in a patient with severe MERS 90
Corticosteroids Pulsed methylprednisolone Patients with severe MERS who were treated with systemic corticosteroid with or without antivirals and interferons had no favourable response 87,88,274

ARDS, acute respiratory distress syndrome; CI, confidence interval; CoV, coronavirus; MERS, Middle East respiratory syndrome; OR, odds ratio; SARS, severe acute respiratory syndrome.

Development of anti-CoV therapeutics

Key CoV targets for new drug development. Despite their high species diversity, CoVs share key genomic elements that are essential for the design of therapeutic targets (Fig. 1). The large replicase polyprotein 1a (pp1a) and pp1ab, which are encoded by the 5′-terminal open reading frame 1a/b (ORF1a/b), are cleaved by two viral proteases, the papain-like protease (PLpro) and the 3C-like protease (3CLpro), to produce non-structural proteins (NSPs) such as RNA-dependent RNA polymerase (RdRp) and helicase, which are involved in the transcription and replication of the virus9,91 (Fig. 2). Numerous enzyme inhibitors targeting these proteins have shown anti-CoV activity in vitro.

Figure 1. Genomes and structures of SARS-CoV and MERS-CoV.

Figure 1

The typical coronavirus (CoV) genome is a single-stranded, non-segmented RNA genome, which is approximately 26–32 kb. It contains 5′-methylated caps and 3′-polyadenylated tails and is arranged in the order of 5′, replicase genes, genes encoding structural proteins (spike glycoprotein (S), envelope protein (E), membrane protein (M) and nucleocapsid protein (N)), polyadenylated tail and then the 3′ end. The partially overlapping 5′-terminal open reading frame 1a/b (ORF1a/b) is within the 5′ two-thirds of the CoV genome and encodes the large replicase polyprotein 1a (pp1a) and pp1ab. These polyproteins are cleaved by papain-like cysteine protease (PLpro) and 3C-like serine protease (3CLpro) to produce non-structural proteins, including RNA-dependent RNA polymerase (RdRp) and helicase (Hel), which are important enzymes involved in the transcription and replication of CoVs. The 3′ one-third of the CoV genome encodes the structural proteins (S, E, M and N), which are essential for virus–cell-receptor binding and virion assembly, and other non-structural proteins and accessory proteins that may have immunomodulatory effects297. Particle image from Ref. 296, Nature Publishing Group. MERS, Middle East respiratory syndrome; SARS, severe acute respiratory syndrome.

PowerPoint slide

Figure 2. Virus-based and host-based treatment options targeting the coronavirus replication cycle.

Figure 2

Binding between the receptor-binding domain on the S1 subunit of spike glycoprotein (S) and the host receptor triggers conformational changes in the S2 subunit of S. This leads to fusion of the viral and cell membranes. Coronaviruses (CoVs) enter the host cell using the endosomal pathway and/or the cell surface non-endosomal pathway. Endosomal cell entry of CoVs is facilitated by low pH and the pH-dependent endosomal cysteine protease cathepsins. S is activated and cleaved into the S1 and S2 subunits by other host proteases, such as transmembrane protease serine 2 (TMPRSS2) and TMPRSS11D, which enables cell surface non-endosomal virus entry at the plasma membrane. Middle East respiratory syndrome (MERS)-CoV S is additionally activated by the serine endoprotease furin. CoVs then dissemble intracellularly to release the nucleocapsid and viral RNA into the cytoplasm for the translation of ORF1a/b into the large replicase polyprotein 1a (pp1a) and pp1ab and for the replication of genomic RNA. pp1a and pp1ab are cleaved by papain-like protease (PLpro) and 3C-like protease (3CLpro) to produce non-structural proteins (NSPs), including RNA-dependent RNA polymerase (RdRp) and helicase, which are involved in the transcription and replication of the virus. The NSPs produced by the cleavage of pp1a and pp1ab form the replication–transcription complex. Attachment of the hydrophobic domains of the CoV replication–transcription complex to the limiting membrane derived from the endoplasmic reticulum (ER) produces typical CoV replication structures including double-membrane vesicles and convoluted membranes. The full-length positive-strand genomic RNA is transcribed to form a full-length negative-strand template for synthesis of new genomic RNAs and overlapping subgenomic negative-strand templates. Subgenomic mRNAs are then synthesized and translated to produce the structural and accessory proteins. The helical nucleocapsid formed by the assembly of nucleocapsid protein (N) and genomic RNA interacts with the other structural proteins to form the assembled virion, which is then released by exocytosis into the extracellular compartment. Virus- and host-based treatment options are highlighted in red and blue, respectively. +, positive-strand RNA; −, negative-strand RNA; AP, accessory protein; CYP, cyclophilin; dec-RVKR-CMK, decanoyl-Arg-Val-Lys-Arg-chloromethylketone; DRACO, double-stranded RNA-activated caspase oligomerizer; E, envelope protein; ER, endoplasmic reticulum; ERGIC, endoplasmic reticulum Golgi intermediate compartment; ERK, extracellular signal-regulated kinase; M, membrane; mAb, monoclonal antibody; MAPK, mitogen-activated protein kinase; MPA, mycophenolic acid; mTOR, mammalian target of rapamycin; N, nucleocapsid protein; NAAE, N-(2-aminoethyl)-1-aziridine-ethanamine; NFAT, nuclear factor of activated T cells; ORF, open reading frame; PI3K, phosphoinositide 3-kinase; poly(I:C), polyinosinic:polycytidylic acid; RdRp, RNA-dependent RNA polymerase; S, spike glycoprotein; SARS-CoV, severe acute respiratory syndrome coronavirus; siRNA, small interfering RNA. *Only siRNAs that have been evaluated in published reports are included. siRNAs directed against other parts of the CoV genome would also be expected to diminish the accumulation or translation of genomic and all upstream subgenomic RNAs. Adapted with permission from Ref. 9, American Society for Microbiology.

PowerPoint slide

The surface structural spike glycoprotein (S) is of particular interest for antiviral development because of its critical role in the virus–cell receptor interaction. S is composed of the amino-terminal receptor-binding S1 and carboxy-terminal membrane fusion S2 subunits. Cleavage at the protease site at the S1–S2 junction is required to activate membrane fusion, virus entry and syncytium formation9. Binding of the S1 subunit receptor-binding domain (RBD) to the host receptor triggers conformational changes in the S2 subunit (the stalk region of S) to bring the viral and cell membranes into close proximity and enable fusion92. Monoclonal antibodies (mAbs) against the S1 subunit RBD and fusion inhibitors targeting the S2 subunit have potent anti-CoV activity in vitro and/or in vivo92,93,94,95,96,97,98,99,100. The key functional host receptors utilized by human pathogenic CoVs include angiotensin-converting enzyme 2 (ACE2; used by SARS-CoV and human CoV (HCoV)-NL63), dipeptidyl peptidase 4 (DPP4; used by MERS-CoV), aminopeptidase N (used by HCoV-229E), and O-acetylated sialic acid (used by HCoV-OC43 and HCoV-HKU1)101,102,103,104,105,106. The host receptor is important in determining the pathogenicity, tissue tropism and host range of the virus. mAbs or agents that target the host receptor are potential anti-CoV agents so long as they do not induce immunopathological effects in animal models.

CoVs enter the host cell using the endosomal pathway and/or the cell surface non-endosomal pathway9 (Fig. 2). Low pH and the pH-dependent endosomal cysteine protease cathepsins help to overcome the energetically unfavourable membrane fusion reaction and facilitate endosomal cell entry of CoVs107,108. Other host proteases, such as transmembrane protease serine 2 (TMPRSS2) and TMPRSS11D (also known as airway trypsin-like protease), cleave S into the S1 and S2 subunits to activate S for cell surface non-endosomal virus entry at the plasma membrane109. Inhibitors of these proteases can abrogate this proteolytic cleavage and partially block cell entry109. MERS-CoV S is also activated by furin, a serine endoprotease that has been implicated in the processing of fusion proteins and cell entry of other RNA viruses, including HIV, avian influenza A/H5N1 virus, Ebola virus, Marburg virus and flaviviruses110. Furin is also involved in MERS-CoV S1/S2 cleavage during egress from the infected cell110. Monotherapy and/or combinatorial treatment with inhibitors of host proteases involved in the various cell entry pathways have potent anti-CoV activity in vitro and should be further evaluated in animal studies109,111.

CoVs disassemble inside the host cell and release the nucleocapsid and viral RNA into the cytoplasm, after which ORF1a/b is translated into pp1a and pp1ab, and the genomic RNA is replicated91. The numerous NSPs produced by the cleavage of pp1a and pp1ab form the replication–transcription complex. Attachment of the hydrophobic domains of the CoV replication–transcription complex to the limiting membrane derived from the endoplasmic reticulum produces typical CoV replication structures including double-membrane vesicles (DMVs) and convoluted membranes112,113. Novel agents, such as K22, that target membrane-bound CoV RNA synthesis inhibit DMV formation of a broad range of human and animal CoVs112. The full-length positive strand of genomic RNA is transcribed to form a full-length negative-strand template for the synthesis of new genomic RNAs and overlapping subgenomic negative-strand templates9,91. Subgenomic mRNAs are then synthesized and translated to produce the structural and accessory proteins9,91. The helical nucleocapsid, formed by the assembly of nucleocapsid protein (N) and genomic RNA, then interacts with S, envelope protein (E), and membrane protein (M) to form the assembled virion9. The virion is released into the extracellular compartment by exocytosis and the viral replication cycle is repeated9. Small interfering RNAs (siRNAs) targeting these structural genes could be useful in the treatment of CoV infections, and further optimization of the in vivo delivery of siRNAs may enable their clinical use.

Approaches to anti-CoV drug screening. The only two human-pathogenic CoVs known before the SARS epidemic were HCoV-229E and HCoV-OC43, which usually cause self-limiting upper respiratory tract infections2. Therefore, researchers and research facilities, especially those involved in antiviral development, were underprepared when SARS-CoV suddenly emerged in 2003. Subsequently, three general approaches were used to discover potential anti-CoV treatment options for human-pathogenic CoVs — especially SARS-CoV and the emerging MERS-CoV — that are associated with more severe disease than the other HCoVs are9,114,115.

The first approach to drug discovery is to test existing broad-spectrum antiviral drugs that have been used to treat other viral infections by using standard assays that measure the effects of these drugs on the cytopathicity, virus yield and plaque formation of live and/or pseudotyped CoVs. Examples of drugs identified using this approach include interferon alfa, interferon beta, interferon gamma, ribavirin and inhibitors of cyclophilin8,74,116,117,118,119,120,121,122. These drugs have the obvious advantage of being readily available with known pharmacokinetic and pharmacodynamic properties, side effects and dosing regimens. However, they do not have specific anti-CoV effects and may be associated with severe adverse effects.

The second anti-CoV drug discovery approach involves the screening of chemical libraries comprising large numbers of existing compounds or databases that contain information on transcriptional signatures in different cell lines122,123,124,125,126,127. This approach provides rapid, high-throughput screening of many readily available compounds that can then be further evaluated by antiviral assays. Various classes of drugs have been identified in these drug repurposing programmes, including many that have important physiological and/or immunological effects such as those that affect neurotransmitter regulation, the oestrogen receptor, kinase signalling, lipid or sterol metabolism, protein processing and DNA synthesis or repair122,123,124,125,126,127. The major disadvantage of this approach is that although many of the identified drugs exhibit anti-CoV activities in vitro, most are not clinically useful because they are either associated with immunosuppressive effects or they have anti-CoV half-maximal effective concentration (EC50) values that markedly exceed the peak serum concentration (Cmax) levels that are achievable at therapeutic dosages. A notable exception, which was found to be effective in a non-human primate model and in non-randomized clinical trials, is the anti-HIV protease inhibitor lopinavir–ritonavir76,77,128 (Table 1).

The third approach for anti-CoV drug discovery involves the de novo development of novel, specific agents based on the genomic and biophysical understanding of the individual CoVs. Examples include siRNA molecules or inhibitors that target specific viral enzymes involved in the viral replication cycle, mAbs that target the host receptor, inhibitors of host cellular proteases, inhibitors of virus endocytosis by the host cell, human or humanized mAbs that target the S1 subunit RBD and antiviral peptides that target the S2 subunit (Fig. 2). Although most of these drugs have potent in vitro and/or in vivo anti-CoV activity, their pharmacokinetic and pharmacodynamic properties and side-effect profiles have yet to be evaluated in animal and human trials. Furthermore, the development of these candidate drugs into clinically useful therapeutic options with reliable delivery modes for patients usually takes years.

Overall, these three drug discovery approaches are often used together during emerging CoV outbreaks to identify candidate drug compounds that can be broadly classified into virus-based and host-based treatment options.

Virus-based anti-CoV treatment options

Viral nucleosides, nucleotides and nucleic acids. Nucleosides and nucleotides are the building blocks of viral nucleic acids (Fig. 2). Drugs that target nucleosides or nucleotides and/or viral nucleic acids generally have broad-spectrum activity against a wide range of CoVs and other viruses (Table 2). Mycophenolate mofetil is an anti-rejection drug that inhibits inosine monophosphate dehydrogenase and the synthesis of guanine monophosphate122. The active compound, mycophenolic acid, exhibits antiviral activity in vitro against various viruses, including hepatitis B virus (HBV), hepatitis C virus (HCV) and arboviruses122. Mycophenolic acid was identified as a potential anti-MERS-CoV drug using high-throughput screening and has potent anti-MERS-CoV activity in vitro122. However, a subsequent study in a non-human primate model showed that MERS-CoV-infected common marmosets treated with mycophenolate mofetil had a worse outcome with more severe disease and higher viral loads in necropsied lung and extrapulmonary tissues than untreated animals did128. Renal transplant recipients who were on maintenance mycophenolate mofetil therapy also developed severe or fatal MERS129,130. Thus, the usual dosage of mycophenolate mofetil monotherapy is unlikely to be useful for prophylaxis or treatment of CoV infections.

Table 2. Representative virus-based treatment strategies for CoV infections.

Targeted viral components Examples Mechanism of action Status Comments Refs
Viral nucleic acids
Nucleosides and/or nucleotides Mycophenolic acid Inhibitor of IMPDH and guanine monophosphate synthesis Marketed

• Broad spectrum: MERS-CoV, HBV, HCV, arboviruses (JEV, WNV, YFV, dengue virus and CHIKV)

• Worsened outcome in MERS-CoV-infected common marmosets

• Unlikely to be useful as monotherapy, but combinatorial therapy with interferon beta-1b is synergistic in vitro

122,128,192
mRNA Ribozyme An antisense RNA with catalytic activity that specifically recognizes the base sequence GUC in the loop region on the mRNA of CoVs Preclinical

• Narrow spectrum

• Optimal delivery method in humans is uncertain

131
Host cell membrane-bound viral replication complex K22 Inhibitor of membrane-bound RNA synthesis and double membrane vesicle formation Preclinical

• Broad spectrum: SARS-CoV, MERS-CoV, HCoV-229E and animal CoVs

• No animal or human data available

112
Long viral dsRNA DRACO A chimeric protein with a viral dsRNA-binding domain and a pro-apoptotic domain that selectively induces apoptosis in cells containing viral dsRNA Preclinical

• Broad spectrum: adenoviruses, arenaviruses, bunyaviruses, dengue virus, IAV, picornaviruses, rhinoviruses and reoviruses

• Anti-CoV activity has yet to be demonstrated

132
Viral enzymes
PLpro GRL0617, compound 4 Inhibitors of PLpro activity Preclinical

• Narrow spectrum

• No animal or human data available

137,138,139,140
3CLpro Lopinavir, N3, CE-5 and GRL-001 Inhibitors of 3CLpro activity Preclinical

• Broad spectrum: SARS-CoV, MERS-CoV, HCoV-229E, HCoV-NL63 and animal CoVs

• Marketed: lopinavir–ritonavir

• Improved outcome of MERS-CoV-infected common marmosets

• Improved outcome of SARS patients in non-randomized trials

76,77,123,128,143,144,145,146,275,276
RdRp Ribavirin Guanosine analogue that inhibits viral RNA synthesis and mRNA capping Marketed

• Broad spectrum: many viral infections, especially SARS, MERS, RSV, HCV and viral haemorrhagic fevers

• Active against SARS-CoV and MERS-CoV at high doses in vitro

• Benefits in SARS and MERS patients are uncertain

• Side effects are common and may be severe with high-dose reigmens

10,21,86,87,88,89,117,277,278,279,280
BCX4430 Adenosine analogue that acts as a non-obligate RNA chain terminator to inhibit viral RNA polymerase function Preclinical

• Broad spectrum: SARS-CoV, MERS-CoV, IAV, filoviruses, togaviruses, bunyaviruses, arenaviruses, paramyxoviruses, picornaviruses and flaviviruses

• No animal or human data are available for CoVs

149
Fleximer nucleoside analogues of acyclovir Doubly flexible nucleoside analogues based on the acyclic sugar scaffold of acyclovir and the flex-base moiety in fleximers that inhibit RdRp Preclinical

• Active against MERS-CoV and HCoV-NL63

• Further modification of existing nucleoside analogues with different fleximers is possible

• No animal or human data available

150
siRNA* Short chains of dsRNA that interfere with the expression of RdRp Preclinical

• Narrow spectrum

• Optimal delivery method in humans is uncertain

151,152
Helicase Bananins and 5-hydroxychromone derivatives Inhibits helicase unwinding and ATPase activities Preclinical

• Possibly broad spectrum: helicase is relatively conserved among CoVs

• High risk of toxicity

153,154
SSYA10-001 and ADKs Inhibits helicase unwinding without affecting ATPase activity Preclinical

• Broad spectrum: SARS-CoV, MERS-CoV and animal CoVs

• Likely to be less toxic than bananins and 5-hydroxychromone derivatives

155,156,281
Viral spike glycoprotein
RBD of the S1 subunit of S MERS-4, MERS-27, m336, m337, m338, REGN3051 and REGN3048 mAbs mAbs against the RBD of the S1 subunit that block virus–host cell binding Preclinical

• Narrow spectrum

• May reduce the need for convalescent-phase plasma therapy

• Protective effects demonstrated in animal models

94,95,96,97,100,160,161,162
S2 subunit of S HR2P and P1 peptides Antiviral peptides that inhibit fusion of S with host cell receptor Preclinical

• Narrow spectrum

• Enfuvirtide, an anti-HIV antiviral peptide fusion inhibitor, has been successfully marketed

92,93,99,160,161,162
Oligosaccharides on S Griffithsin A carbohydrate-binding agent that specifically binds to oligosaccharides on S, thereby blocking virus–host cell binding Preclinical

• Broad spectrum: SARS-CoV, MERS-CoV, HCoV-229E, HCoV-OC43, HIV, HCV and Ebola virus

• Well tolerated in rodents

173,174
S expression siRNA* Short chains of dsRNA that interfere with the expression of SARS-CoV S Preclinical

• Narrow spectrum

• SARS-CoV-infected rhesus macaques had better clinical, virological, and histological parameters

• Optimal delivery method in humans is uncertain

169,170,171,172
Viral envelope, membrane, nucleocapsid and accessory proteins
E siRNA* Short chains of dsRNA that interfere with the expression of SARS-CoV E Preclinical

• Narrow spectrum

• Optimal delivery method in humans is uncertain

179
Hexamethylene amiloride Viroporin inhibitor that inhibits the ion channel activity of CoV E Preclinical

• Inhibited ion channel activities of SARS-CoV, HCoV-229E and some animal CoVs

• Analogue of the potassium-sparing diuretic drug amiloride

181,182
M siRNA* Short chains of dsRNA that interfere with the expression of SARS-CoV M Preclinical

• Narrow spectrum

• Optimal delivery method in humans is uncertain

179
N PJ34, intrabodies and siRNA* Reduces the RNA-binding affinity of N and viral replication Preclinical

• Narrow spectrum

• Optimal delivery method in humans is uncertain

179,183,282
Accessory proteins siRNA* Short chains of dsRNA that interfere with the expression of proteins from SARS-CoV ORF3a, ORF7a and ORF7b Preclinical

• Narrow spectrum

• Optimal delivery method in humans is uncertain

180
Lipid membrane LJ001 and JL103 Membrane-binding photosensitizers that induce singlet oxygen modifications of specific phospholipids Preclinical

• Broad spectrum: enveloped viruses (IAV, filoviruses, poxviruses, arenaviruses, bunyaviruses, paramyxoviruses, flaviviruses and HIV-1)

• Anti-CoV activity has yet to be demonstrated

184,185,186,187

3CLpro, 3C-like protease; ADK, aryl diketoacid; CHIKV, Chikungunya virus; CoV, coronavirus; DRACO, double-stranded RNA activated caspase oligomerizer; dsRNA, double-stranded RNA; E, envelope protein; HBV, hepatitis B virus; HCoV, human coronavirus; HCV, hepatitis C virus; IAV, influenza A virus; IMPDH, inosine-monophosphate dehydrogenase; JEV, Japanese encephalitis virus; M, membrane protein; mAb, monoclonal antibody; MERS, Middle East respiratory syndrome; N, nucleocapsid protein; ORF, open reading frame; PLpro, papain-like protease; RBD, receptor-binding domain; RdRp, RNA-dependent RNA polymerase; RSV, respiratory syncytial virus; S, spike glycoprotein; SARS, severe acute respiratory syndrome; siRNA, small interfering RNA; WNV, West Nile virus; YFV, yellow fever virus.

*Only siRNAs that have been evaluated in published reports are included. siRNAs directed against other parts of the CoV genome would also be expected to diminish the accumulation or translation of genomic and upstream subgenomic RNAs.

Intrabodies are antibodies that work within the cell to bind to intracellular proteins.

Ribozymes (also known as catalytic RNA or RNA enzymes) are RNA molecules that catalyse specific biochemical reactions. A chimeric DNA–RNA hammerhead ribozyme that specifically recognizes the base sequence GUC, which is present in the loop region of SARS-CoV mRNA, substantially reduces the expression of recombinant SARS-CoV RNA in vitro131. However, ribozymes are rapidly degraded in vivo and delivery methods would have to be optimized in humans before ribozymes could become clinically useful.

Agents targeting the specific host cell membrane-bound CoV replication complex have also been investigated. One such compound, K22, inhibits membrane-bound CoV RNA synthesis and is active against a broad range of CoVs in vitro112. In cell culture, K22 exerts potent anti-CoV activity during an early step of the viral replication cycle and impairs formation of DMVs112. HCoV-229E escape mutants that are resistant to K22 have substitutions in the potential membrane-spanning domains in nsp6, a membrane-spanning integral component of the CoV replication complex that is involved in DMV formation, including nsp6H121L and nsp6M159V (Ref. 112). The emergence of K22 resistance should be monitored in subsequent in vivo studies.

Recently, a new class of broad-spectrum antivirals that targets long viral double-stranded RNA (dsRNA) has been reported. For example, dsRNA-activated caspase oligomerizer (DRACO) is a chimeric protein with a viral dsRNA-binding domain and a pro-apoptotic domain that selectively induces apoptosis in cells that contain viral dsRNA but spares uninfected host cells132. DRACO is active against many RNA viruses in vitro and/or in vivo132. If an effective mode of DRACO delivery can be achieved, a broad-spectrum anti-CoV drug that targets highly conserved CoV RNA sequences might become a reality.

Viral enzymes. All of the major enzymes and proteins of CoVs that are involved in viral replication are potentially druggable targets (Table 2). The SARS-CoV and MERS-CoV PLpro enzymes exhibit proteolytic, deubiquitylating and deISGylating activities133,134,135. Crystallography has facilitated the characterization of these PLpro enzymes and the identification of PLpro inhibitors136. Numerous SARS-CoV PLpro inhibitors belonging to different classes have been identified, including small-molecule inhibitors, thiopurine compounds, natural products, zinc ion and zinc conjugate inhibitors and naphthalene inhibitors137. However, some of these drugs only inhibit the enzymatic activities of PLpro without inhibiting viral replication, or vice versa137,138,139. None has been validated in animal or human studies137,138. Furthermore, most PLpro inhibitors have narrow-spectrum activity because of the structural differences among the PLpro enzymes of different CoVs140,141. For example, most SARS-CoV PLpro inhibitors are inactive against MERS-CoV because of the structurally different, flexible blocking loop 2 (BL2) domains in the PLpro enzymes of SARS-CoV and MERS-CoV140.

3CLpro is the other major CoV protease that cleaves the large replicase polyproteins during viral replication. SARS-CoV 3CLpro can be targeted by numerous classes of protease inhibitors, including zinc or mercury conjugates, C2-symmetric diols, peptidomimetic-α,β-unsaturated esters, anilides, benzotriazole, N-phenyl-2-acetamide, biphenyl sulfone, glutamic acid and glutamine peptides with a trifluoromethylketone group, pyrimidinone and pyrazole analogues142. Some of these 3CLpro inhibitors demonstrate broad-spectrum in vitro activities against CoVs with highly similar key residues for substrate recognition at their 3CLpro enzymes143,144. Among these 3CLpro inhibitors, the most readily available one is lopinavir, a protease inhibitor used to treat HIV infections that is usually marketed as a ritonavir-boosted form (lopinavir–ritonavir). Lopinavir and/or lopinavir–ritonavir have anti-CoV activity in vitro, as well as in MERS-CoV-infected non-human primates and in non-randomized trials of SARS patients76,77,123,128,145. It is postulated that the 3CLpro-inhibiting activity of lopinavir–ritonavir contributes at least partially to its anti-CoV effects146. It remains to be seen if resistance emerges, as it has in patients with HIV infection, when lopinavir–ritonavir is routinely used to treat CoV infections.

RdRp is an essential part of the CoV replication–transcription complex and is involved in the production of genomic and subgenomic RNAs. Ribavirin is a guanosine analogue with broad-spectrum antiviral activity and has been used in the treatment of severe respiratory syncytial virus infection, HCV infection and viral haemorrhagic fevers. Its exact mechanism of action is undetermined, but inhibition of mRNA capping and induction of mutations in RNA-dependent viral replication are considered to be important for RNA viruses, including CoVs147. High-dose ribavirin has been used to treat SARS patients, but the benefits were unclear8,10,21,72,74,75,117. It exhibits moderate anti-MERS-CoV activity at high doses in vitro and in MERS-CoV-infected rhesus macaques, but there was no obvious survival benefit in small cohorts of MERS patients86,87,88,89,121,148. Moreover, the severe side effects associated with the use of high-dose ribavirin limit its clinical application in patients with severe CoV infections8,74. Recently, a novel adenosine analogue, BCX4430 (Immucillin-A), was developed149. It acts as a non-obligate RNA chain terminator to inhibit viral RNA polymerases of a wide range of RNA viruses, including CoVs such as SARS-CoV and MERS-CoV as well as filoviruses such as Ebola and Marburg viruses149. Its development for human use has been fast-tracked to increase the number of treatment options for the recent Ebola virus epidemic in West Africa. Existing nucleoside analogues, such as acyclovir, could be modified by incorporating fleximers, which have increased binding affinity and can overcome resistance caused by point mutations in biologically important binding sites150. These acyclic fleximer nucleoside analogues inhibit MERS-CoV and HCoV-NL63 in vitro at micromolar concentrations150. Notably, resistance to nucleoside analogues due to mutations in RdRp has been reported for other RNA viruses, and should be monitored when these agents are used to treat CoV infections. In addition to nucleoside analogues, siRNA molecules targeting SARS-CoV RdRp have been used to inhibit SARS-CoV in vitro151,152.

Helicase catalyses the unwinding of duplex oligonucleotides into single strands in an ATP-dependent reaction during the CoV replication cycle. Helicase inhibitors are attractive anti-CoV treatment options because the helicases of different CoVs are highly homologous. Based on their mechanisms of action, CoV helicase inhibitors can be broadly categorized into two groups. The first group includes bananins and 5-hydroxychromone derivatives, which inhibit the unwinding and ATPase activity of SARS-CoV helicase, resulting in inhibition of viral replication in vitro153,154. However, the toxicity resulting from the inhibition of cellular ATPases or kinases by these compounds has limited their development for human use. The second group of CoV helicase inhibitors includes compounds that selectively inhibit the unwinding activity but not the ATPase activity of CoV helicase. An example is SSYA10-001, a triazole that inhibits a broad range of CoVs, including SARS-CoV, MERS-CoV and mouse hepatitis virus155,156. The toxicity of SSYA10-001 should be evaluated in animal models.

Viral spike glycoprotein. The membrane-anchored glycoprotein, S, is a major immunogenic antigen and is essential for the interaction between the virus and the host cell receptor (Fig. 2). Adoptive transfer of sera containing anti-MERS-CoV-S antibodies blocked virus attachment and accelerated viral clearance from the lungs of MERS-CoV infected BALB/c mice that were recently transduced by adenoviral vectors expressing human DPP4 (Ref. 157) (Table 2). Small cohorts of SARS patients who received convalescent-phase plasma containing neutralizing antibodies that probably targeted CoV S had significantly higher discharge rates by 3 weeks after symptom onset and a lower mortality rate83,84. However, the use of convalescent-phase plasma therapy during emerging CoV outbreaks is limited by the good will of convalescent patients with high serum neutralizing antibody titres. Disease worsening associated with immune enhancement that results from treatment with products containing low antibody titres has been reported in cell line and animal studies158,159. To overcome these problems, mAbs targeting different regions of SARS-CoV S have been generated by immunization of human immunoglobulin transgenic mice, cloning of small chain variable regions from naive and convalescent patients as well as from immortalization of convalescent S-specific B cells160. Most of these mAbs target specific epitopes on the S1 subunit RBD to inhibit virus–cell receptor binding, whereas others bind to the S2 subunit to interrupt virus–cell fusion160. Regardless of their binding sites and mechanisms, these mAbs exhibit neutralizing activities and reduced viral titres in vitro and/or in small animal models. Similarly, several mAbs targeting different epitopes on the S1 subunit RBD of MERS-CoV S have been developed94,95,96,97,100. These monoclonal antibodies bind to the RBD with 10-fold to >450-fold higher affinity than does human DPP4, resulting in broader and higher neutralizing activity in vitro. Importantly, combination therapy with two or more synergistically acting humanized or human mAbs targeting non-cross-resistant epitopes or different regions of S may help to reduce the frequency with which viruses mutate to escape antibody-mediated neutralization94. Treatment with these mAbs showed protective effects in MERS-CoV-infected human DPP4-transgenic mice and mice transduced by adenoviral vectors expressing human DPP4 (Refs 100,161,162). Their safety profiles and treatment effects in patients should be further evaluated.

Antiviral peptides targeting different regions of S are another promising therapeutic strategy. The S2 subunits or stalk regions of both SARS-CoV and MERS-CoV S are class I viral fusion proteins that each contain an N-terminal fusion peptide, heptad repeat 1 (HR1) and HR2 domains, a transmembrane domain and a cytoplasmic domain92. Antiviral peptides analogous to the N terminus, pre-transmembrane domain or the loop region separating the HR1 and HR2 domains of SARS-CoV inhibited virus plaque formation by 40–80% at micromolar concentrations163,164. Similarly, antiviral peptides spanning the HR2 domain of MERS-CoV inhibit S-mediated cell–cell fusion and viral entry into cells in vitro92,93. A peptide called HP2P-M2 that is derived from the HR2 domain, if administered intranasally before or after viral challenge, protected C57BL/6 mice and mice deficient for V(D)J recombination-activating protein 1 (RAG1) that were recently transduced by adenoviral vectors expressing human DPP4 from MERS-CoV infection with 10-fold to >1,000-fold reduction in viral titres in the lung; this protection was enhanced by combining this peptide with interferon beta99. Combining antiviral peptides targeting different regions of the S2 subunit may be synergistic in vitro and overcome the theoretical risk of drug resistance165. Importantly, an analogous fusion inhibitor, enfuvirtide, which binds to glycoprotein 41 of HIV to block membrane fusion and HIV cell entry, has been successfully marketed for treatment of HIV-1 infection166. Primary resistance to enfuvirtide is rare and can be overcome by modifying the drug such that it contains secondary compensatory mutations167,168. This example of successful drug development includes measures to counteract drug resistance and therefore favours antiviral peptides over anti-CoV S siRNAs for further evaluation in vivo; siRNAs have remained in preclinical development despite their reported antiviral activities in vitro and in SARS-CoV-infected rhesus macaques owing to the lack of reliable drug delivery methods in humans169,170,171,172.

Another class of anti-CoV agents that target S to inhibit CoV entry is the carbohydrate-binding agents. Griffithsin is an antiviral protein originally isolated from the red alga Griffithsia spp.173. It binds specifically to oligosaccharides on viral surface glycoproteins such as S and HIV glycoprotein 120. It inhibits a broad range of CoVs, including SARS-CoV, HCoV-229E, HCoV-OC43 and HCoV-NL63 in vitro and in SARS-CoV-infected mice173,174. The optimal delivery modes and safety profiles of these agents in humans should be further evaluated.

Viral envelope, membrane, nucleocapsid and accessory proteins. E, M and N and some of the accessory proteins are not only essential for virion assembly but may also have additional functions that suppress the host immune response to facilitate viral replication. For example, the accessory proteins 4a and 4b, and possibly also M and accessory protein 5 of MERS-CoV, exhibit interferon antagonist activities, and SARS-CoV N acts as a viral suppressor of RNA silencing and suppresses RNA interference triggered by either short hairpin RNAs or siRNAs175,176,177,178 (Table 2). siRNAs targeting E, M, N, ORF3a, ORF7a or ORF7b of SARS-CoV inhibited viral replication in vitro179,180. However, similar to anti-CoV S siRNAs, none of these siRNAs is ready for human use until better delivery methods become available.

Alternatively, an increasing number of agents that target specific binding sites or functions of these proteins are being generated through crystallography and functional assays. Examples include the viroporin inhibitor hexamethylene amiloride, which reduces the ion channel activity of E in SARS-CoV and HCoV-229E, and PJ34, which binds to a distinct ribonucleotide-binding pocket at the N-terminal domain of N in HCoV-OC43 (Refs 181,182,183). However, these agents are likely to be narrow-spectrum as the binding sites and functions of these proteins are unique to individual CoVs.

Novel lipophilic thiazolidine derivatives, such as LJ001 and JL103, are membrane-binding photosensitizers that produce singlet oxygen molecules to induce changes in the properties of lipid membranes and prevent fusion between viral and target cell membranes. They exhibit broad-spectrum activities against numerous enveloped viruses and may be active against CoVs184,185,186,187.

Host-based anti-CoV treatment options

Broad-spectrum host innate immune response. The host innate interferon response is crucial for the control of viral replication after infection188. Although CoVs are able to suppress the interferon response for immune evasion, they remain susceptible to interferon treatment in vitro189,190. The interferon response can be augmented by the administration of recombinant interferons or interferon inducers (Table 3). Recombinant interferon alfa and interferon beta inhibit the replication of both SARS-CoV and MERS-CoV in vitro and in animal models8,99,116,121,122,128,148,191,192. Various combinations of interferon alfa or interferon beta with other antivirals such as ribavirin and/or lopinavir–ritonavir have been used to treat patients with SARS or MERS. Overall, combination treatments consisting of interferons and ribavirin did not consistently improve outcomes8,9,74,86,87,89. The apparent discrepancy between in vitro findings and in vivo outcomes may be related to the high EC50/Cmax ratios of these drugs and the delay between symptom onset and drug administration8,121,122. This delay is especially relevant for MERS patients, as they have a much shorter median time interval between symptom onset and death than do SARS patients9,58. The use of recombinant interferon beta-1b, which has the lowest EC50/Cmax ratio against MERS-CoV among tested preparations of recombinant interferons, should be evaluated in combination with other effective antivirals in clinical trials at early stages of the infection122,128.

Table 3. Representative host-based treatment strategies for CoV infections.

Targeted host factors Examples Mechanism of action Status Comments Refs
Broad-spectrum host innate immune response
Interferon response Recombinant interferons (interferon alfa, interferon beta, interferon gamma) Exogenous interferons Marketed

• Broad spectrum against many CoVs and other viruses

• Recombinant interferon beta was more potent than interferon alfa for SARS-CoV and MERS-CoV in vitro

• Interferon alfa reduced viral titres in lungs of SARS-CoV-infected mice and cynomolgus macaques

• Intranasal interferon beta administered before or after MERS-CoV challenge reduced viral titres in the lungs of Ad5-hDPP4 C57BL/6 and Rag1−/− mice by 10–100 fold

• Subcutaneous interferon beta-1b improved outcomes of MERS-CoV-infected common marmosets

• Benefits for SARS patients are uncertain

• Benefits of interferon alfa-2a, interferon alfa-2b and interferon beta-1a for MERS patients are uncertain

8,9,74,86,87,89,99,116,121,122,128,148,191,192,215
Poly(I:C) Induces interferon production Phase II clinical trials

• Reduced MERS-CoV load in Ad5-hDPP4 BALB/c mice

• Used in Phase II clinical trials of patients with malignant gliomas

157,193,194
Nitazoxanide A thiazolide that induces the host innate immune response by potentiation of interferon alfa and interferon beta production by fibroblasts and activation of PKR Marketed

• Broad spectrum: canine CoV, IAV, IBV, RSV, PIF, Sendai virus, rhinovirus, norovirus, rotavirus, Dengue virus, JEV, YFV, HBV, HCV and HIV

• Used in patients with parasitic infections and in Phase II and III clinical trials of HCV infection and influenza

• Activity against human-pathogenic CoVs has yet to be determined

195
Other host signalling pathways involved in viral replication
Cyclophilins Cyclosporine, alisporivir Cyclophilin inhibitor that could modulate the interaction of cyclophilins with SARS-CoV nsp1 and the calcineurin–NFAT pathway Marketed

• Broad spectrum: CoVs (SARS-CoV, MERS-CoV, HCoV-NL63, HCoV-229E, and animal CoVs), HIV, HCV, HPV, vaccinia virus and VSV

• Alisporivir does not have the immunosuppressive effects of cyclosporine and may therefore be a more suitable antiviral candidate

118,119,120,200
Kinase signalling pathways Trametinib, selumetinib, everolimus, rapamycin, dasatinib and imatinib Kinase signalling inhibitors that block the ABL1, ERK–MAPK and/or PI3K–AKT–mTOR pathways, which may block early viral entry and/or post-entry events Marketed

• Active against SARS-CoV and MERS-CoV

• May be associated with immunopathology

124,125
Host receptors utilized by CoVs for viral entry
ACE2 P4 and P5 peptides and NAAE ACE2-derived peptides or small molecules targeting ACE2 that block SARS-CoV S-mediated cell fusion Marketed

• Narrow spectrum: SARS-CoV

• May affect important biological functions such as blood pressure regulation

202,203
DPP4 Anti-DPP4 mAb clones 2F9 and YS110 Anti-DPP4 mAbs that block MERS-CoV S-mediated cell fusion Phase I clinical trial

• Narrow spectrum: MERS-CoV

• May affect important biological functions such as glucose metabolism and immunological responses

• mAb clone YS110 was used in a Phase I clinical trial of patients with advanced malignancies

102,201,227
Host proteases utilized by CoVs for viral entry
Endosomal protease (for example, cathepsins) E64D, K11777 and the small molecule 5705213 Cathepsin inhibitors that block endosomal protease-mediated cleavage and the endosomal entry pathway Preclinical

• Broad spectrum: CoVs (SARS-CoV and MERS-CoV), filoviruses (Ebola virus) and paramyxoviruses (Hendra and Nipah viruses)

• Combination with TMPRSS2 inhibitors necessary for complete inhibition of MERS-CoV in vitro

111,124,127,283
Surface protease (for example, TMPRSS2) Camostat mesylate TMPRSS2 inhibitor that blocks the TMPRSS2-mediated cell surface entry pathway Marketed

• Broad spectrum: CoVs (SARS-CoV, MERS-CoV and HCoV-229E), IAV and PIF

• Combination with cathepsin inhibitors is necessary for complete inhibition of MERS-CoV in vitro

• Used to treat patients with chronic pancreatitis

109,111,207,208,284,285,286
Other host proteases (for example, furin) dec-RVKR-CMK Furin inhibitor that blocks furin-mediated cleavage of S Preclinical Active against MERS-CoV and may be active against other CoVs that utilize furin for S cleavage 110
Endocytosis
Clathrin-mediated endocytosis Chlorpromazine An antipsychotic that also affects the assembly of clathrin-coated pits at the plasma membrane Marketed

• Broad spectrum: SARS-CoV, MERS-CoV, HCV and alphaviruses

• Clinical benefit uncertain owing to a high EC50/Cmax ratio at the usual therapeutic dosages

123
Clathrin-mediated endocytosis Ouabain and bufalin ATP1A1-binding cardiotonic steroids that inhibit clathrin-mediated endocytosis Marketed

• Active against MERS-CoV at nanomolar concentrations in vitro

• May have risk of toxicity

209
Endosomal acidification Chloroquine An antimalarial that sequesters protons in lysosomes to increase the intracellular pH Marketed

• Broad spectrum: CoVs (SARS-CoV, MERS-CoV, HCoV-229E and HCoV-OC43), HIV, flaviviruses and Ebola, Hendra and Nipah viruses in vitro

• Not active against SARS-CoV-infected mice

123,210,211,212,213,214,215

ACE2, angiotensin-converting enzyme 2; Ad5-hDPP4, adenovirus type 5 expressing human dipeptidyl peptidase 4; ATP1A1, ATPase subunit α1; Cmax, peak serum concentration; CoV, coronavirus; dec-RVKR-CMK, decanoyl-Arg-Val-Lys-Arg-chloromethylketone; DPP4; dipeptidyl peptidase 4; EC50, half-maximal effective concentration; ERK, extracellular signal-regulated kinase; HBV, hepatitis B virus; HCoV, human coronavirus; HCV, hepatitis C virus; HPV, human papillomavirus; IAV, influenza A virus; IBV, influenza B virus; JEV, Japanese encephalitis virus; mAb, monoclonal antibody; MAPK, mitogen-activated protein kinase; MERS, Middle East respiratory syndrome; mTOR, mammalian target of rapamycin; NAAE, N-(2-aminoethyl)-1-aziridine-ethanamine; NFAT, nuclear factor of activated T cells; nsp1, non-structural protein 1; PI3K, phosphoinositide 3-kinase; PIF, parainfluenza virus; PKR, protein kinase R; poly(I:C), polyinosinic:polycytidylic acid; RSV, respiratory syncytial virus; S, spike glycoprotein; SARS, severe acute respiratory syndrome; TMPRSS2, transmembrane protease serine 2; VSV, vesicular stomatitis virus; YFV, yellow fever virus.

Polyinosinic:polycytidylic acid (poly(I:C)) is a synthetic analogue of dsRNA that strongly induces type I interferons. It substantially reduced the MERS-CoV load in BALB/c mice that were transduced by adenoviral vectors expressing human DPP4 shortly before poly(I:C) administration, although its effects in standard cell culture protection assays are not published157. Intramuscular injection of poly(I:C) stabilized with poly-L-lysine and carboxymethylcellulose seemed to be well tolerated by patients with malignant gliomas in Phase II clinical trials193,194. Nitazoxanide is another potent type I interferon inducer that has been used in humans for parasitic infections195. It is a synthetic nitrothiazolyl–salicylamide derivative that exhibits broad-spectrum antiviral activities against both RNA and DNA viruses including canine CoV, influenza viruses, HBV, HCV, HIV, rotavirus, norovirus and flaviviruses195. It has been evaluated in Phase II and Phase III clinical trials for the treatment of HCV infection and influenza and has a good safety profile195,196,197. Other innate immunomodulators that have anti-SARS-CoV effects in animal models include the antimicrobial peptide rhesus θ-defensin 1 and protein cage nanoparticles that elicit a host immune response in inducible bronchus-associated lymphoid tissue198,199. The combined use of interferon inducers and innate immunomodulators with effective antiviral agents may be synergistic and should be evaluated in animal models.

Other host signalling pathways involved in viral replication. In addition to direct potentiation of the interferon response, other cell signalling pathways have been identified as potential anti-CoV treatment targets (Table 3). Cyclophilins interact with SARS-CoV nsp1 to modulate the calcineurin pathway, which is important in the T cell-mediated adaptive immune response120. The calcineurin inhibitor cyclosporine inhibits a broad range of CoVs in vitro118,119,120. However, its clinical application is limited by its immunosuppressive effects and high EC50/Cmax ratio at standard therapeutic dosages. The antiviral activities of newer, non-immunosuppressive calcineurin inhibitors, which are active against HCoV-NL63, should be evaluated for SARS-CoV and MERS-CoV200. Similarly, agents that modulate other cellular signalling pathways, such as kinase signalling pathway inhibitors, also exhibit anti-CoV activities and are commercially available124,125. However, their toxicities may limit their use in patients with severe CoV infections.

Host factors utilized by CoVs for viral replication. CoVs utilize specific host factors for virus entry and replication (Fig. 2). The host receptor can be targeted by specific monoclonal or polyclonal antibodies, peptides or functional inhibitors (Table 3). For example, anti-DPP4 mAbs inhibit MERS-CoV cell entry in vitro201. YS110 is a recombinant humanized IgG1 anti-DPP4 mAb that seems to be well tolerated in patients with advanced solid tumours201. For the treatment of SARS-CoV, small-molecule entry inhibitors such as N-(2-aminoethyl)-1-aziridine-ethanamine (NAAE) inhibit the catalytic activity of ACE2 and SARS-CoV S-mediated cell–cell fusion in vitro202. Synthetic peptides analogous to critical segments of ACE2 also have anti-SARS-CoV activity at micromolar concentrations in vitro203. However, none of these receptor-directed compounds has yet been tested in patients with CoV infections. Their anti-CoV activity is likely to be narrow-spectrum, as different CoVs utilize different host cell receptors. Furthermore, the risks of immunopathology must be assessed, especially given the multiple essential biological and immunological functions of these receptors.

The entry of CoVs into host cells via the endosomal and/or cell surface pathways is facilitated by host proteases that cleave and activate S. Cathepsins are cysteine proteases that are involved in the endosomal pathway and can be inhibited by cathepsin inhibitors such as K11777 and its related vinylsulfone analogues111. These compounds seem to be safe and effective against various parasitic infections in animal models, and have broad-spectrum activities against enveloped RNA viruses such as CoVs (SARS-CoV, MERS-CoV, HCoV-229E and HCoV-NL63), filoviruses (Ebola and Marburg viruses) and paramyxoviruses111,204,205,206. TMPRSS2 is a serine protease that mediates the cell surface entry pathway; camostat mesylate is a synthetic low-molecular-weight serine protease inhibitor that has been used to treat chronic pancreatitis in humans with minimal side effects207,208. This molecule inhibits SARS-CoV and MERS-CoV in vitro and improves survival of SARS-CoV-infected mice109,111. Furin, another ubiquitously expressed host protease, is also important in MERS-CoV S-mediated entry. Blocking furin with decanoyl-Arg-Val-Lys-Arg-chloromethylketone inhibits MERS-CoV entry and cell–cell fusion in vitro110.

Another group of candidate anti-CoV drugs target the endocytosis of CoV during cell entry. Chlorpromazine is an antipsychotic drug used in the treatment of schizophrenia that also affects the assembly of clathrin-coated pits at the plasma membrane. It is active against HCV, alphaviruses and numerous CoVs, including SARS-CoV and MERS-CoV, in vitro123. Cardiotonic steroids that bind sodium/potassium-transporting ATPase subunit α1, such as ouabain and bufalin, also inhibit clathrin-mediated endocytosis of MERS-CoV at nanomolar concentrations209. However, the use of these clathrin-mediated endocytosis inhibitors in patients with CoV infections is limited by either very high EC50/Cmax ratios or toxicity. Alternatively, endocytosis can also be suppressed by a high pH. Chloroquine is an anti-malarial drug that sequesters protons into lysosomes to increase the intracellular pH. It has broad-spectrum antiviral activities against numerous CoVs (SARS-CoV, MERS-CoV, HCoV-229E and HCoV-OC43) and other RNA viruses in vitro123,210,211,212,213,214. However, it did not substantially reduce viral replication in SARS-CoV-infected mice, possibly because the cell surface pathway was not simultaneously blocked215. The anti-CoV effects, pharmacokinetic and pharmacodynamic profiles and toxicity of the combinations of different protease and endocytosis inhibitors that target these different cell entry pathways should be further evaluated in vivo.

Development of MERS-CoV vaccines

Rapid diagnostics and effective vaccines are often complementary to antiviral treatment in the control of epidemics caused by emerging viruses (Box 1). Although there has not been any new human SARS case for over 10 years, sporadic cases and clusters of MERS continue to occur in the Middle East owing to the persistence of zoonotic sources in endemic areas, and these cases spread to other regions. Effective MERS-CoV vaccines are essential for interrupting the chain of transmission from animal reservoirs and infected humans to susceptible hosts. Live-attenuated vaccines, which have been previously evaluated in animal models for SARS, might be associated with disseminated infection in immunocompromised patients. Inactivated vaccines might be associated with immunopathology during animal challenge. These are unfavourable approaches for MERS vaccine development because a substantial proportion of patients with severe MERS have comorbidities or systemic immunocompromising conditions. Other vaccination strategies for MERS that use DNA plasmids, viral vectors, nanoparticles, virus-like particles and recombinant protein subunits are in development and some have been evaluated in animal models157,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233 (Table 4). The availability of a safe and effective MERS-CoV vaccine for dromedary camels and non-immune individuals at high risk of camel contact in endemic regions such as the Middle East and the greater Horn of Africa would be an important measure for controlling the ongoing epidemic.

Table 4. MERS-CoV candidate vaccines in development.

Vaccine type Examples Vaccine design strategy Comments Refs
Live attenuated virus rMERS-CoV-ΔE Deletion of the gene encoding MERS-CoV E rendered the mutant virus replication-competent and propagation-defective

• Attenuated SARS-CoV-ΔE mutant virus induced protection in mice and hamsters

• No animal data are available for a rMERS-CoV-ΔE-based vaccine yet

• Risk of disseminated infection in immunocompromised patients

218,287,288,289,290,291,292,293,294,295
DNA plasmid MERS-CoV S DNA DNA plasmids that encode full-length MERS-CoV S

• BALB/cJ mice vaccinated with MERS-CoV S-encoding DNA developed neutralizing anti-MERS-CoV antibodies

• The neutralizing antibody titre was boosted 10-fold after vaccination with S1 protein

• Rhesus macaques vaccinated sequentially with MERS-CoV S-encoding DNA and S1 protein had reduced CT scan abnormalities

219
Viral vectors MVA-MERS-S, Ad5-MERS-S, Ad5-MERS-S1, Ad5-S and Ad41-S Viral vectors (MVA or Ad) that express full-length MERS-CoV S or the S1 subunit of MERS-CoV S

• Both MVA and Ad vector-based vaccines induced neutralising anti-MERS-CoV antibodies in BALB/c mice

• A MVA-MERS-S vaccine conferred mucosal immunity and induced serum neutralizing anti-MERS-CoV antibodies in dromedary camels

• Mucosal (intragastric) administration of Ad5-S or Ad41-S vaccines induced the production of antigen-specific IgG and neutralizing antibodies, but not antigen-specific T cell responses, in BALB/c mice

• Systemic (intramuscular) administration of Ad5-S or Ad41-S vaccines induced antigen-specific neutralizing IgG antibodies, as well as T cell responses in splenic and pulmonary lymphocytes

• Increased immunopathology with severe hepatitis in SARS-CoV-infected ferrets that were previously vaccinated with an MVA-based vaccine expressing full-length SARS-CoV S

220,221,222,223,224,298
Nanoparticles MERS-CoV S-containing nanoparticles Purified MERS-CoV S-containing nanoparticles produced in insect (Sf9) cells that were infected with specific recombinant baculovirus containing the gene encoding MERS-CoV S

• BALB/c mice vaccinated with MERS-CoV or SARS-CoV S-containing nanoparticles developed neutralizing antibodies specific to the viral S

• Adjuvant matrix M1 or alum is required to elicit an optimal neutralizing antibody response

225
Virus-like particles VRP-S VEE virus-like replicon particles containing MERS-CoV S Vaccination of BALB/c mice transduced with Ad5-hDPP4 with VRP-S reduced viral titres in lungs to nearly undetectable levels by day 1 after inoculation with MERS-CoV 157
Recombinant protein subunits S(RBD)-Fc, S1(358–588)-Fc, S(377–588)-Fc and rRBD Full-length MERS-CoV S or the RBD subunit of MERS-CoV S

• Vaccinated BALB/c mice and/or rabbits developed neutralizing antibodies

• Protective effects may be enhanced by combination with different adjuvants

• Non-neutralizing epitopes in full-length S-based vaccines may induce antibody-mediated disease enhancement

226,227,228,229,230,231,232,233

Ad, adenovirus; CoV, coronavirus; CT, computerized tomography; E, envelope protein; hDPP4, human dipeptidyl peptidase 4; IgG, immunoglobulin G; MERS, Middle East respiratory syndrome; MVA, modified vaccinia virus Ankara; RBD, receptor-binding domain; rRBD, recombinant RBD; S, spike glycoprotein; SARS,severe acute respiratory syndrome; S(RBD)-Fc, RBD of S fused to the antibody crystallizable fragment; S1(358–588)-Fc, amino acid residues 358–588 of the S1 subunit of S fused to the antibody crystallizable fragment; VEE, Venezuelan equine encephalitis; VRP, virus replicon particle.

Box 1: The complementary roles of novel rapid diagnostics and antiviral agents.

As demonstrated in the severe acute respiratory syndrome (SARS) and Middle East respiratory syndrome (MERS) epidemics, rapid and accurate laboratory diagnosis is essential for the clinical management and epidemiological control of coronavirus (CoV) infections. Real-time reverse transcription (RT)-PCR assays, which can quantify viral loads, have facilitated studies on viral shedding patterns and optimization of treatment and infection control strategies. The peak viral load in SARS was found to occur at day 10 after symptom onset and helped to predict the timing of clinical deterioration and the need for intensive supportive care18,249. Point-of-care nucleic amplification tests such as RT-loop-mediated isothermal amplification and RT-isothermal recombinase polymerase amplification are suitable for field evaluation, especially in resource-limited areas250,251. Similarly, assays that detect abundantly expressed CoV antigens, such as the nucleocapsid protein, can be used for fast and high-throughput laboratory diagnosis without requiring biosafety level 3 containment252,253. The rapid availability of complete genome sequences of most human and animal CoVs has minimized the time required for the design of new RT-PCR assays, source identification and molecular surveillance for emerging CoVs254. This was well illustrated in the MERS epidemic, in which highly sensitive and specific RT-PCR assays targeting unique gene regions such as the region upstream of the envelope (E) gene (upE region) were quickly developed after the complete genome sequence of MERS-CoV strains isolated from humans became available255,256. Comparative genomic studies quickly identified bats and camels carrying CoVs that were highly similar to MERS-CoV strains isolated from humans, and these two animals were determined to be the likely CoV reservoirs40,42,43,44,46,91. Continuous surveillance and analysis of MERS-CoV genomes obtained from patients and animals in different areas in the Middle East are important for detecting mutations that may increase the ability of the virus to be efficiently transmitted from person to person71. Data analyses from the sequencing of small RNAs and the use of locked nucleic acid probes have allowed the development of new assays that target short but abundantly expressed gene regions from CoV genomes, such as the leader sequences257. The increasing number of complete CoV genomes and diagnostic gene targets has enabled the development of multiplex assays that simultaneously detect multiple CoVs or multiple gene targets of a particular CoV257. The increasing use of these multiplex assays in clinical laboratories worldwide will enhance our understanding of the changing epidemiology of CoV infections and enable the stratification of at-risk patients and contact groups for early treatment and prophylaxis.

In addition to improving acute clinical diagnosis, diagnostic advances have facilitated other aspects of the control of CoV epidemics and anti-CoV drug development. The isolation of infectious virus particles from clinical specimens in cell culture has a limited role in the acute diagnosis of CoV infection, as most human-pathogenic CoVs are either difficult or dangerous to culture258. Nevertheless, recent advances have enhanced their use in CoV pathogenesis studies, which are important for identifying new treatment targets. The previously unculturable HCoV-HKU1 can now be isolated from primary differentiated human tracheal bronchial epithelial cells and human alveolar type II cells that are cultured at an air–liquid interface259,260,261,262. Ex vivo organ culture enables the identification of important viral and host factors that are involved in the severe pulmonary and extrapulmonary manifestations of SARS and MERS263,264,265,266,267. The number of available human and animal cell lines from various organs is increasing, and these provide insights into tissue and species tropism258,268,269. Similarly, detection of specific anti-CoV antibodies in paired acute and convalescent sera samples are mainly useful for seroepidemiological studies and contact tracing, but not for acute diagnosis41,49,51,270. Novel assays such as the spike glycoprotein (S) pseudoparticle neutralization assay, which do not require biosafety level 3 containment, enable high-throughput antibody detection in large-scale seroepidemiological studies and outbreak investigations271.

Outlook and challenges

Animal models for testing anti-CoV drugs. Suitable animal models are especially important for testing anti-CoV drugs because most of these drugs have not been used in humans. In contrast to the limited number of animal models established for the mild infections caused by HCoV-229E, HCoV-OC43, HCoV-NL63 and HCoV-HKU1, various small animal and non-human primate models have been evaluated for studies of the pathogenesis and treatment of SARS and MERS234,235,236,237. The identification of ACE2 and DPP4 as the functional receptors for SARS-CoV and MERS-CoV, respectively, was essential to the development of animal models that are representative of severe human disease101,102. A number of different non-human primates were found to be permissive to SARS-CoV, but none of them consistently reproduced characteristics of the severe human disease, and mortality was not observed237. These models were predominantly useful to fulfil Koch's postulates238. Small animals — including young and aged BALB/c and C57BL/6 mice, knockout mice with deficiencies in T, B and/or NK cells, golden Syrian hamsters and ferrets — could be productively infected with SARS-CoV, but few of them developed clinically apparent disease237. The best available small animal models for SARS utilize transgenic mice that express human ACE2 and/or mouse-adapted SARS-CoV strains that are capable of causing lethal disease in mice239,240,241. The limited availability of these ACE2-transgenic mice and mouse-adapted virus strains remains a major obstacle to testing anti-SARS-CoV drugs.

Similar to SARS, non-human primate models were also used to fulfil Koch's postulates and investigate the pathogenesis of MERS. Rhesus macaques developed only a mild, self-limiting disease and were not optimal for the evaluation of treatments for MERS148,242,243. By contrast, MERS-CoV-infected common marmosets developed a disseminated and fatal infection that closely resembled severe human disease128,244. However, the availability of common marmosets is limited and experiments on these small primates are technically demanding. Therefore, other small animal models for MERS were evaluated. Unlike with SARS-CoV, most small animals — including BALB/c mice, golden Syrian hamsters, ferrets and rabbits — were not susceptible to MERS-CoV infection245,246,247. Intranasal inoculation of adenoviral vectors expressing human DPP4 followed by MERS-CoV inoculation was a novel method that rapidly rendered mice susceptible to MERS-CoV infection, but the disease was relatively mild and confined to the respiratory tract157. Transgenic mice expressing human DPP4 develop severe pulmonary and disseminated infection and are currently the best available small animal model for MERS248. Potential anti-MERS-CoV treatment options identified in in vitro antiviral assays should be further evaluated in these transgenic mice.

Generic challenges in the clinical development of novel anti-CoV drugs. There are a number of virological and patient-associated factors that pose major challenges in the clinical development of novel anti-CoV drugs. First, CoVs are one of the most diverse and rapidly mutating groups of viruses, and novel CoVs emerge repeatedly at unpredictable times. Therefore, most anti-CoV drugs that specifically target the replication apparatus of an existing CoV may not be effective against another novel CoV. This is especially applicable to viral enzyme inhibitors, mAbs and antiviral peptides that target S, as well as agents that target the host cell receptor. Second, there are a limited number of animal models available for infections caused by HCoV-229E, HCoV-OC43, HCoV-NL63 and HCoV-HKU1. Even for SARS and MERS, experiments using suitable animal models such as mice with transgenes encoding ACE2 or DPP4, and non-human primates, are only available in a few designated research biosafety level 3 laboratories, and these experiments are technically demanding. Last and most important, the mild clinical severity of infections caused by HCoV-229E, HCoV-OC43, HCoV-NL63 and HCoV-HKU1, together with the absence of new SARS cases, have made recruitment of patients into clinical trials difficult and reduced the incentives for pharmaceutical companies to develop specific antiviral drugs for these CoV infections. The continuing threat of MERS-CoV to global health security 3 years after its first discovery presents a golden opportunity to tackle current obstacles in the development of new anti-CoV drugs. It is prudent that a well-organized, multidisciplinary, international collaborative network consisting of clinicians, virologists and drug developers, coupled to political commitment, is formed to carry out clinical trials using anti-CoV drugs that have already been shown to be safe and effective in vitro and/or in animal models.

Prioritization of virus-based and host-based treatment options for clinical development. Despite the report of a large number of virus-based and host-based treatment options with potent in vitro activities for SARS and MERS, only a few are likely to fulfil their potential in the clinical setting in the foreseeable future. Most drugs have one or more major limitations that prevent them from proceeding beyond the in vitro stage. First, many drugs have high EC50/Cmax ratios at clinically relevant dosages. Examples of such drugs include cyclosporine, chlorpromazine and interferon alfa. Second, some have severe side effects or cause immunosuppression. For example, the use of high-dose ribavirin may be associated with haemolytic anaemia, neutropenia, teratogenicity and cardiorespiratory distress. MERS-CoV-infected common marmosets treated with mycophenolate mofetil developed a fatal infection with even higher viral loads in their lungs and extrapulmonary tissues than untreated controls did128. Agents targeting host signalling pathways or receptors may induce immunopathology. Furthermore, the lack of a reliable drug delivery method in vivo is particularly problematic for siRNAs and other agents that have not been previously used in humans.

Looking ahead, the most feasible options that should be further evaluated in clinical trials for the ongoing MERS epidemic include monotherapy or combinational therapies that include lopinavir–ritonavir, interferon beta-1b and/or mAbs and antiviral peptides targeting MERS-CoV S. These agents have protective effects against MERS in non-human primate or mouse models. Moreover, they are either marketed drugs (in the case of lopinavir–ritonavir and interferon beta-1b) or they have been successfully developed for other infections (such as palivizumab, which is used for respiratory syncytial virus infection, and enfuvirtide, which is used for HIV infection). In the long term, the development of novel, broad-spectrum, pan-CoV antiviral drugs that are active against a wide range of CoVs may become the ultimate treatment strategy for circulating and emerging CoV infections.

Supplementary information

Supplementary information S1 (figure) (1.6MB, pdf)

Global Distribution of SARS and MERS (PDF 1587 kb)

Supplementary information S2 (table) (139.6KB, pdf)

Comparisons of the clinical and laboratory features between SARS and MERS (PDF 139 kb)

Acknowledgements

The research of J. F.-W. C. and K.-Y. Y. is partly supported by funding from Respiratory Viral Research Foundation Limited; the Consultancy Service for Enhancing Laboratory Surveillance of Emerging Infectious Disease of the Department of Health; the Health and Medical Research Fund (15140762) of the Food and Health Bureau; the National Natural Science Foundation of China/Research Grants Council Joint Research Scheme (N_HKU728/14); and the Theme-based Research Scheme (T11 707/15 R) of the Research Grants Council, Hong Kong Special Administrative Region of the People's Republic of China. A.Z. acknowledges support from the National Institutes of Health Research (NIHR), Biomedical Research Centre at UCL Hospital, London, UK. The funding sources had no role in study design, data collection, analysis, or interpretation or writing of the report.

Glossary

Zoonotic virus

A virus that can transmit an infectious disease from animals (usually vertebrates) to humans.

Spike glycoprotein (S)

A major immunogenic antigen of coronaviruses that is essential for interactions between a virus and host cell receptor, and is an important therapeutic target.

Syncytium

A multinucleated cell resulting from the fusion of the host membranes of neighbouring cells infected by various viruses, including CoVs.

Convalescent-phase plasma

Plasma that contains neutralizing antibodies against a microorganism and is obtained from patients recovering from the infection.

Viroporin

A small integral membrane protein that is localized primarily within the endoplasmic reticulum and plasma membranes of host cells, and has the characteristic ability to form ion channels or pores.

Protein cage nanoparticles

Nanoscale delivery platforms, made of biomaterials and/or proteins, that are used for various biomedical applications including the delivery of therapeutic cargo molecules.

Koch's postulates

Criteria used to establish a causative relationship between a microorganism and a disease.

Biographies

Alimuddin Zumla is professor of infectious diseases and international health in the division of Infection and Immunity at University College London (UCL), UK, a consultant infectious diseases physician at UCL Hospitals NHS Foundation Trust, UK, and a founding director of the University of Zambia–UCL Medical School (UNZA–UCLMS) Research and Training Programme. He has published over 400 PubMed articles and 21 textbooks. His research has defined the epidemiology and clinical features and improved diagnostics, drug treatment and management protocols of lethal respiratory infectious diseases in Sub-Saharan Africa, the Middle East, the United Kingdom and Europe.

Jasper F. W. Chan is a clinical assistant professor at the State Key Laboratory of Emerging Infectious Diseases, the Carol Yu Centre for Infection, the Research Centre of Infection and Immunology and the Department of Microbiology at the University of Hong Kong (HKU) and is a specialist in clinical microbiology and infection at Queen Mary Hospital in Hong Kong, China, and HKU-Shenzhen Hospital in Shenzhen, China. He received his M.B.B.S. from HKU in 2005 and completed postgraduate specialist training to become a fellow of the Royal College of Physicians (Edinburgh, UK), the Royal College of Pathologists (UK), the Hong Kong College of Pathologists and the Hong Kong Academy of Medicine. He has published over 120 peer-reviewed articles on emerging infectious diseases, particularly those caused by novel coronaviruses.

Esam I. Azhar is Head of the Special Infectious Agents Unit, and Associate Professor of Medical Virology at the King Fahd Medical Research Centre, Kind Abdul Aziz University, Jeddah, Saudi Arabia. He obtained his M.Sc. and Ph.D. from the University of Sheffield, UK, and is a Fellow of the Royal College of Physicians (Edinburgh, UK). He has published over 70 publications and has made seminal contributions to research on Alkhurma viral haemorrhagic fever, dengue haemorrhagic fever virus, Rift Valley fever virus and Middle East respiratory syndrome coronavirus.

David S. C. Hui is the Stanley Ho Professor of Respiratory Medicine and the Director of the Stanley Ho Centre for Emerging Infectious Diseases at the Chinese University of Hong Kong (CUHK). He graduated from the University of New South Wales, Australia, in 1985, and then trained in Respiratory and Sleep Medicine in Sydney, Australia. He has published well over 240 peer-reviewed journal articles and 23 book chapters since joining the CUHK in 1998. His research interests include the clinical management of severe acute respiratory infections, the safety of respiratory therapy and hospital infection control in the post-SARS (severe acute respiratory syndrome) era.

Kwok-Yung Yuen is the Henry Fok Professor and Chair of Infectious Diseases at The University of Hong Kong (HKU). He also serves as the director of the clinical diagnostic microbiology services at Queen Mary Hospital in Hong Kong, China, and HKU-Shenzhen Hospital in Shenzhen, China, as the co-director of the State Key Laboratory of Emerging Infectious Diseases of China in Hong Kong and as the founding co-director of the HKU-Pasteur Research Centre in Hong Kong. He is an elected Academician of the Chinese Academy of Engineering (Basic Medicine and Health), a founding member of the Academy of Sciences of Hong Kong, and a Fellow of the Royal College of Physicians (London and Edinburgh, UK; as well as Ireland), Royal College of Surgeons (Glasgow, UK) and Royal College of Pathologists (UK). His has published over 800 peer-reviewed articles in the areas of emerging infections and microbial discovery. He and his team have discovered and characterized over 60 novel human and animal viruses, including human coronavirus HKU1 and the bat and human severe acute respiratory syndrome (SARS) coronaviruses.

PowerPoint slides

Competing interests

J. F.-W. C. has received travel grants from Pfizer Corporation Hong Kong and Astellas Pharma Hong Kong Corporation Limited. The funding sources had no role in study design, data collection, analysis or interpretation or writing of the report. The corresponding author had full access to all the data in the study and had final responsibility for the decision to submit for publication.

Footnotes

Alimuddin Zumla, Jasper F. W. Chan, Esam I. Azhar, David S. C. Hui and Kwok-Yung Yuen: All authors contributed equally to this work.

References

  • 1.Woo PC, Lau SK, Huang Y, Yuen KY. Coronavirus diversity, phylogeny and interspecies jumping. Exp. Biol. Med. (Maywood) 2009;234:1117–1127. doi: 10.3181/0903-MR-94. [DOI] [PubMed] [Google Scholar]
  • 2.Chan JF, et al. Is the discovery of the novel human betacoronavirus 2c EMC/2012 (HCoV-EMC) the beginning of another SARS-like pandemic? J. Infect. 2012;65:477–489. doi: 10.1016/j.jinf.2012.10.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Chan JF, Lau SK, Woo PC. The emerging novel Middle East respiratory syndrome coronavirus: the 'knowns' and 'unknowns'. J. Formos. Med. Assoc. 2013;112:372–381. doi: 10.1016/j.jfma.2013.05.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Chan JF, To KK, Tse H, Jin DY, Yuen KY. Interspecies transmission and emergence of novel viruses: lessons from bats and birds. Trends Microbiol. 2013;21:544–555. doi: 10.1016/j.tim.2013.05.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Chan JF, To KK, Chen H, Yuen KY. Cross-species transmission and emergence of novel viruses from birds. Curr. Opin. Virol. 2015;10:63–69. doi: 10.1016/j.coviro.2015.01.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Woo PC, et al. Discovery of seven novel mammalian and avian coronaviruses in the genus deltacoronavirus supports bat coronaviruses as the gene source of alphacoronavirus and betacoronavirus and avian coronaviruses as the gene source of gammacoronavirus and deltacoronavirus. J. Virol. 2012;86:3995–4008. doi: 10.1128/JVI.06540-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Lau SK, et al. Discovery of a novel coronavirus, China Rattus coronavirus HKU24, from Norway rats supports the murine origin of Betacoronavirus 1 and has implications for the ancestor of Betacoronavirus lineage A. J. Virol. 2015;89:3076–3092. doi: 10.1128/JVI.02420-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Cheng VC, Lau SK, Woo PC, Yuen KY. Severe acute respiratory syndrome coronavirus as an agent of emerging and reemerging infection. Clin. Microbiol. Rev. 2007;20:660–694. doi: 10.1128/CMR.00023-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Chan JF, et al. Middle East respiratory syndrome coronavirus: another zoonotic betacoronavirus causing SARS-like disease. Clin. Microbiol. Rev. 2015;28:465–522. doi: 10.1128/CMR.00102-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Lee N, et al. A major outbreak of severe acute respiratory syndrome in Hong Kong. N. Engl. J. Med. 2003;348:1986–1994. doi: 10.1056/NEJMoa030685. [DOI] [PubMed] [Google Scholar]
  • 11.Tsang KW, et al. A cluster of cases of severe acute respiratory syndrome in Hong Kong. N. Engl. J. Med. 2003;348:1977–1985. doi: 10.1056/NEJMoa030666. [DOI] [PubMed] [Google Scholar]
  • 12.Zhao Z, et al. Description and clinical treatment of an early outbreak of severe acute respiratory syndrome (SARS) in Guangzhou, PR China. J. Med. Microbiol. 2003;52:715–720. doi: 10.1099/jmm.0.05320-0. [DOI] [PubMed] [Google Scholar]
  • 13.Xu RH, et al. Epidemiologic clues to SARS origin in China. Emerg. Infect. Dis. 2004;10:1030–1037. doi: 10.3201/eid1006.030852. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Peiris JS, et al. Coronavirus as a possible cause of severe acute respiratory syndrome. Lancet. 2003;361:1319–1325. doi: 10.1016/S0140-6736(03)13077-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Kuiken T, et al. Newly discovered coronavirus as the primary cause of severe acute respiratory syndrome. Lancet. 2003;362:263–270. doi: 10.1016/S0140-6736(03)13967-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Drosten C, et al. Identification of a novel coronavirus in patients with severe acute respiratory syndrome. N. Engl. J. Med. 2003;348:1967–1976. doi: 10.1056/NEJMoa030747. [DOI] [PubMed] [Google Scholar]
  • 17.Ksiazek TG, et al. A novel coronavirus associated with severe acute respiratory syndrome. N. Engl. J. Med. 2003;348:1953–1966. doi: 10.1056/NEJMoa030781. [DOI] [PubMed] [Google Scholar]
  • 18.Peiris JS, et al. Clinical progression and viral load in a community outbreak of coronavirus-associated SARS pneumonia: a prospective study. Lancet. 2003;361:1767–1772. doi: 10.1016/S0140-6736(03)13412-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Woo PC, et al. Relative rates of non-pneumonic SARS coronavirus infection and SARS coronavirus pneumonia. Lancet. 2004;363:841–845. doi: 10.1016/S0140-6736(04)15729-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Hsu LY, et al. Severe acute respiratory syndrome (SARS) in Singapore: clinical features of index patient and initial contacts. Emerg. Infect. Dis. 2003;9:713–717. doi: 10.3201/eid0906.030264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Booth CM, et al. Clinical features and short-term outcomes of 144 patients with SARS in the greater Toronto area. JAMA. 2003;289:2801–2809. doi: 10.1001/jama.289.21.JOC30885. [DOI] [PubMed] [Google Scholar]
  • 22.Wong RS, Hui DS. Index patient and SARS outbreak in Hong Kong. Emerg. Infect. Dis. 2004;10:339–341. doi: 10.3201/eid1002.030645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.World Health Organization. Summary of probable SARS cases with onset of illness from 1 November 2002 to 31 July 2003. WHO emergencies preparedness, response[online].
  • 24.Du L, et al. Analysis on the characteristics of blood serum Ab-IgG detective result of severe acute respiratory syndrome patients in Guangzhou, China. Zhonghua Liu Xing Bing Xue Za Zhi. 2004;25:925–928. [PubMed] [Google Scholar]
  • 25.Guan Y, et al. Isolation and characterization of viruses related to the SARS coronavirus from animals in southern China. Science. 2003;302:276–278. doi: 10.1126/science.1087139. [DOI] [PubMed] [Google Scholar]
  • 26.Wang M, et al. SARS-CoV infection in a restaurant from palm civet. Emerg. Infect. Dis. 2005;11:1860–1865. doi: 10.3201/eid1112.041293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Song HD, et al. Cross-host evolution of severe acute respiratory syndrome coronavirus in palm civet and human. Proc. Natl Acad. Sci. USA. 2005;102:2430–2435. doi: 10.1073/pnas.0409608102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Tu C, et al. Antibodies to SARS coronavirus in civets. Emerg. Infect. Dis. 2004;10:2244–2248. doi: 10.3201/eid1012.040520. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Lau SK, et al. Severe acute respiratory syndrome coronavirus-like virus in Chinese horseshoe bats. Proc. Natl Acad. Sci. USA. 2005;102:14040–14045. doi: 10.1073/pnas.0506735102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Li W, et al. Bats are natural reservoirs of SARS-like coronaviruses. Science. 2005;310:676–679. doi: 10.1126/science.1118391. [DOI] [PubMed] [Google Scholar]
  • 31.Ge XY, et al. Isolation and characterization of a bat SARS-like coronavirus that uses the ACE2 receptor. Nature. 2013;503:535–538. doi: 10.1038/nature12711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.He B, et al. Identification of diverse alphacoronaviruses and genomic characterization of a novel severe acute respiratory syndrome-like coronavirus from bats in China. J. Virol. 2014;88:7070–7082. doi: 10.1128/JVI.00631-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Woo PC, Lau SK, Yuen KY. Infectious diseases emerging from Chinese wet-markets: zoonotic origins of severe respiratory viral infections. Curr. Opin. Infect. Dis. 2006;19:401–407. doi: 10.1097/01.qco.0000244043.08264.fc. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Shi Z, Hu Z. A review of studies on animal reservoirs of the SARS coronavirus. Virus Res. 2008;133:74–87. doi: 10.1016/j.virusres.2007.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Zaki AM, van Boheemen S, Bestebroer TM, Osterhaus AD, Fouchier RA. Isolation of a novel coronavirus from a man with pneumonia in Saudi Arabia. N. Engl. J. Med. 2012;367:1814–1820. doi: 10.1056/NEJMoa1211721. [DOI] [PubMed] [Google Scholar]
  • 36.de Groot RJ, et al. Middle East respiratory syndrome coronavirus (MERS-CoV): announcement of the Coronavirus Study Group. J. Virol. 2013;87:7790–7792. doi: 10.1128/JVI.01244-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Hijawi B, et al. Novel coronavirus infections in Jordan, April 2012: epidemiological findings from a retrospective investigation. East. Mediterr. Health J. 2013;19:S12–S18. doi: 10.26719/2013.19.supp1.S12. [DOI] [PubMed] [Google Scholar]
  • 38.Hui DS, Perlman S, Zumla A. Spread of MERS to South Korea and China. Lancet Respir. Med. 2015;3:509–510. doi: 10.1016/S2213-2600(15)00238-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.World Health Organization. Middle East Respiratory Syndrome coronavirus (MERS-CoV) — Jordan[online], (2015).
  • 40.Woo PC, et al. Comparative analysis of twelve genomes of three novel group 2c and group 2d coronaviruses reveals unique group and subgroup features. J. Virol. 2007;81:1574–1585. doi: 10.1128/JVI.02182-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Reusken CB, et al. Middle East respiratory syndrome coronavirus neutralising serum antibodies in dromedary camels: a comparative serological study. Lancet Infect. Dis. 2013;13:859–866. doi: 10.1016/S1473-3099(13)70164-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Haagmans BL, et al. Middle East respiratory syndrome coronavirus in dromedary camels: an outbreak investigation. Lancet Infect. Dis. 2014;14:140–145. doi: 10.1016/S1473-3099(13)70690-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Azhar EI, et al. Evidence for camel-to-human transmission of MERS coronavirus. N. Engl. J. Med. 2014;370:2499–2505. doi: 10.1056/NEJMoa1401505. [DOI] [PubMed] [Google Scholar]
  • 44.Lau SK, et al. Genetic characterization of Betacoronavirus lineage C viruses in bats reveals marked sequence divergence in the spike protein of Pipistrellus bat coronavirus HKU5 in Japanese pipistrelle: implications for the origin of the novel Middle East respiratory syndrome coronavirus. J. Virol. 2013;87:8638–8650. doi: 10.1128/JVI.01055-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Corman VM, et al. Rooting the phylogenetic tree of middle East respiratory syndrome coronavirus by characterization of a conspecific virus from an African bat. J. Virol. 2014;88:11297–11303. doi: 10.1128/JVI.01498-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Wang Q, et al. Bat origins of MERS-CoV supported by bat coronavirus HKU4 usage of human receptor CD26. Cell Host Microbe. 2014;16:328–337. doi: 10.1016/j.chom.2014.08.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Wernery U, et al. Acute middle East respiratory syndrome coronavirus infection in livestock Dromedaries, Dubai, 2014. Emerg. Infect. Dis. 2015;21:1019–1022. doi: 10.3201/eid2106.150038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Penttinen PM, et al. Taking stock of the first 133 MERS coronavirus cases globally — is the epidemic changing? Euro Surveill. 2013;18:20596. doi: 10.2807/1560-7917.ES2013.18.39.20596. [DOI] [PubMed] [Google Scholar]
  • 49.Muller MA, et al. Presence of Middle East respiratory syndrome coronavirus antibodies in Saudi Arabia: a nationwide, cross-sectional, serological study. Lancet Infect. Dis. 2015;15:559–564. doi: 10.1016/S1473-3099(15)70090-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Assiri A, et al. Hospital outbreak of Middle East respiratory syndrome coronavirus. N. Engl. J. Med. 2013;369:407–416. doi: 10.1056/NEJMoa1306742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Al-Abdallat MM, et al. Hospital-associated outbreak of Middle East respiratory syndrome coronavirus: a serologic, epidemiologic, and clinical description. Clin. Infect. Dis. 2014;59:1225–1233. doi: 10.1093/cid/ciu359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Memish ZA, Zumla AI, Al-Hakeem RF, Al-Rabeeah AA, Stephens GM. Family cluster of Middle East respiratory syndrome coronavirus infections. N. Engl. J. Med. 2013;368:2487–2494. doi: 10.1056/NEJMoa1303729. [DOI] [PubMed] [Google Scholar]
  • 53.Health Protection Agency. Evidence of person-to-person transmission within a family cluster of novel coronavirus infections, United Kingdom, February 2013. Euro Surveill.18, 20427 (2013). [DOI] [PubMed]
  • 54.Oboho IK, et al. 2014 MERS-CoV outbreak in Jeddah — a link to health care facilities. N. Engl. J. Med. 2015;372:846–854. doi: 10.1056/NEJMoa1408636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Hui DS, Peiris M. Middle East respiratory syndrome. Am. J. Respir. Crit. Care Med. 2015;192:278–279. doi: 10.1164/rccm.201506-1221ED. [DOI] [PubMed] [Google Scholar]
  • 56.Zumla A, Hui DS, Perlman S. Middle East respiratory syndrome. Lancet. 2015;386:995–1007. doi: 10.1016/S0140-6736(15)60454-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Hui DS, Zumla A. Advancing priority research on the Middle East respiratory syndrome coronavirus. J. Infect. Dis. 2014;209:173–176. doi: 10.1093/infdis/jit591. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Hui DS, Memish ZA, Zumla A. Severe acute respiratory syndrome vs. the Middle East respiratory syndrome. Curr. Opin. Pulm. Med. 2014;20:233–241. doi: 10.1097/MCP.0000000000000046. [DOI] [PubMed] [Google Scholar]
  • 59.Arabi YM, et al. Clinical course and outcomes of critically ill patients with Middle East respiratory syndrome coronavirus infection. Ann. Intern. Med. 2014;160:389–397. doi: 10.7326/M13-2486. [DOI] [PubMed] [Google Scholar]
  • 60.Assiri A, et al. Epidemiological, demographic, and clinical characteristics of 47 cases of Middle East respiratory syndrome coronavirus disease from Saudi Arabia: a descriptive study. Lancet Infect. Dis. 2013;13:752–761. doi: 10.1016/S1473-3099(13)70204-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Eckerle I, Muller MA, Kallies S, Gotthardt DN, Drosten C. In-vitro renal epithelial cell infection reveals a viral kidney tropism as a potential mechanism for acute renal failure during Middle East Respiratory Syndrome (MERS) Coronavirus infection. Virol. J. 2013;10:359. doi: 10.1186/1743-422X-10-359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Saad M, et al. Clinical aspects and outcomes of 70 patients with Middle East respiratory syndrome coronavirus infection: a single-center experience in Saudi Arabia. Int. J. Infect. Dis. 2014;29:301–306. doi: 10.1016/j.ijid.2014.09.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Al-Tawfiq JA, et al. Middle East respiratory syndrome coronavirus: a case-control study of hospitalized patients. Clin. Infect. Dis. 2014;59:160–165. doi: 10.1093/cid/ciu226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Drosten C, et al. Transmission of MERS-coronavirus in household contacts. N. Engl. J. Med. 2014;371:828–835. doi: 10.1056/NEJMoa1405858. [DOI] [PubMed] [Google Scholar]
  • 65.Breban R, Riou J, Fontanet A. Interhuman transmissibility of Middle East respiratory syndrome coronavirus: estimation of pandemic risk. Lancet. 2013;382:694–699. doi: 10.1016/S0140-6736(13)61492-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Cauchemez S, et al. Middle East respiratory syndrome coronavirus: quantification of the extent of the epidemic, surveillance biases, and transmissibility. Lancet Infect. Dis. 2014;14:50–56. doi: 10.1016/S1473-3099(13)70304-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Poletto C, et al. Assessment of the Middle East respiratory syndrome coronavirus (MERS-CoV) epidemic in the Middle East and risk of international spread using a novel maximum likelihood analysis approach. Euro Surveill. 2014;19:20824. doi: 10.2807/1560-7917.es2014.19.23.20824. [DOI] [PubMed] [Google Scholar]
  • 68.Anderson RM, et al. Epidemiology, transmission dynamics and control of SARS: the 2002–2003 epidemic. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2004;359:1091–1105. doi: 10.1098/rstb.2004.1490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Wallinga J, Teunis P. Different epidemic curves for severe acute respiratory syndrome reveal similar impacts of control measures. Am. J. Epidemiol. 2004;160:509–516. doi: 10.1093/aje/kwh255. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Cotten M, et al. Transmission and evolution of the Middle East respiratory syndrome coronavirus in Saudi Arabia: a descriptive genomic study. Lancet. 2013;382:1993–2002. doi: 10.1016/S0140-6736(13)61887-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Cotten M, et al. Spread, circulation, and evolution of the Middle East respiratory syndrome coronavirus. mBio. 2014;5:e01062-13. doi: 10.1128/mBio.01062-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Hui DS, Sung JJ. Severe acute respiratory syndrome. Chest. 2003;124:12–15. doi: 10.1378/chest.124.1.12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Cheng VC, Tang BS, Wu AK, Chu CM, Yuen KY. Medical treatment of viral pneumonia including SARS in immunocompetent adult. J. Infect. 2004;49:262–273. doi: 10.1016/j.jinf.2004.07.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Cheng VC, Chan JF, To KK, Yuen KY. Clinical management and infection control of SARS: lessons learned. Antiviral Res. 2013;100:407–419. doi: 10.1016/j.antiviral.2013.08.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Wong SS, Yuen KY. The management of coronavirus infections with particular reference to SARS. J. Antimicrob. Chemother. 2008;62:437–441. doi: 10.1093/jac/dkn243. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Chan KS, et al. Treatment of severe acute respiratory syndrome with lopinavir/ritonavir: a multicentre retrospective matched cohort study. Hong Kong Med. J. 2003;9:399–406. [PubMed] [Google Scholar]
  • 77.Chu CM, et al. Role of lopinavir/ritonavir in the treatment of SARS: initial virological and clinical findings. Thorax. 2004;59:252–256. doi: 10.1136/thorax.2003.012658. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Loutfy MR, et al. Interferon alfacon-1 plus corticosteroids in severe acute respiratory syndrome: a preliminary study. JAMA. 2003;290:3222–3228. doi: 10.1001/jama.290.24.3222. [DOI] [PubMed] [Google Scholar]
  • 79.Wang H, et al. Fatal aspergillosis in a patient with SARS who was treated with corticosteroids. N. Engl. J. Med. 2003;349:507–508. doi: 10.1056/NEJM200307313490519. [DOI] [PubMed] [Google Scholar]
  • 80.Griffith JF, et al. Osteonecrosis of hip and knee in patients with severe acute respiratory syndrome treated with steroids. Radiology. 2005;235:168–175. doi: 10.1148/radiol.2351040100. [DOI] [PubMed] [Google Scholar]
  • 81.Tsang OT, et al. Coronavirus-positive nasopharyngeal aspirate as predictor for severe acute respiratory syndrome mortality. Emerg. Infect. Dis. 2003;9:1381–1387. doi: 10.3201/eid0911.030400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Lee N, et al. Effects of early corticosteroid treatment on plasma SARS-associated Coronavirus RNA concentrations in adult patients. J. Clin. Virol. 2004;31:304–309. doi: 10.1016/j.jcv.2004.07.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Cheng Y, et al. Use of convalescent plasma therapy in SARS patients in Hong Kong. Eur. J. Clin. Microbiol. Infect. Dis. 2005;24:44–46. doi: 10.1007/s10096-004-1271-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Soo YO, et al. Retrospective comparison of convalescent plasma with continuing high-dose methylprednisolone treatment in SARS patients. Clin. Microbiol. Infect. 2004;10:676–678. doi: 10.1111/j.1469-0691.2004.00956.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Mair-Jenkins J, et al. The effectiveness of convalescent plasma and hyperimmune immunoglobulin for the treatment of severe acute respiratory infections of viral etiology: a systematic review and exploratory meta-analysis. J. Infect. Dis. 2015;211:80–90. doi: 10.1093/infdis/jiu396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Omrani AS, et al. Ribavirin and interferonα-2a for severe Middle East respiratory syndrome coronavirus infection: a retrospective cohort study. Lancet Infect. Dis. 2014;14:1090–1095. doi: 10.1016/S1473-3099(14)70920-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Al-Tawfiq JA, Momattin H, Dib J, Memish ZA. Ribavirin and interferon therapy in patients infected with the Middle East respiratory syndrome coronavirus: an observational study. Int. J. Infect. Dis. 2014;20:42–46. doi: 10.1016/j.ijid.2013.12.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Khalid M, et al. Ribavirin and interferon-α-2b as primary and preventive treatment for Middle East respiratory syndrome coronavirus (MERS-CoV): a preliminary report of two cases. Antivir. Ther. 2015;20:87–91. doi: 10.3851/IMP2792. [DOI] [PubMed] [Google Scholar]
  • 89.Shalhoub S, et al. IFN-α2a or IFN-β1a in combination with ribavirin to treat Middle East respiratory syndrome coronavirus pneumonia: a retrospective study. J. Antimicrob. Chemother. 2015;70:2129–2132. doi: 10.1093/jac/dkv085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Spanakis N, et al. Virological and serological analysis of a recent Middle East respiratory syndrome coronavirus infection case on a triple combination antiviral regimen. Int. J. Antimicrob. Agents. 2014;44:528–532. doi: 10.1016/j.ijantimicag.2014.07.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.van Boheemen S, et al. Genomic characterization of a newly discovered coronavirus associated with acute respiratory distress syndrome in humans. mBio. 2012;3:e00473-12. doi: 10.1128/mBio.00473-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Lu L, et al. Structure-based discovery of Middle East respiratory syndrome coronavirus fusion inhibitor. Nat. Commun. 2014;5:3067. doi: 10.1038/ncomms4067. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Gao J, et al. Structure of the fusion core and inhibition of fusion by a heptad repeat peptide derived from the S protein of Middle East respiratory syndrome coronavirus. J. Virol. 2013;87:13134–13140. doi: 10.1128/JVI.02433-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Jiang L, et al. Potent neutralization of MERS-CoV by human neutralizing monoclonal antibodies to the viral spike glycoprotein. Sci. Transl. Med. 2014;6:234ra59. doi: 10.1126/scitranslmed.3008140. [DOI] [PubMed] [Google Scholar]
  • 95.Ying T, et al. Exceptionally potent neutralization of Middle East respiratory syndrome coronavirus by human monoclonal antibodies. J. Virol. 2014;88:7796–7805. doi: 10.1128/JVI.00912-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Tang XC, et al. Identification of human neutralizing antibodies against MERS-CoV and their role in virus adaptive evolution. Proc. Natl Acad. Sci. USA. 2014;111:E2018–E2026. doi: 10.1073/pnas.1402074111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Du L, et al. A conformation-dependent neutralizing monoclonal antibody specifically targeting receptor-binding domain in middle East respiratory syndrome coronavirus spike protein. J. Virol. 2014;88:7045–7053. doi: 10.1128/JVI.00433-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Yuan K, et al. Suppression of SARS-CoV entry by peptides corresponding to heptad regions on spike glycoprotein. Biochem. Biophys. Res. Commun. 2004;319:746–752. doi: 10.1016/j.bbrc.2004.05.046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Channappanavar R, et al. Protective effect of intranasal regimens containing peptidic Middle East Respiratory Syndrome coronavirus fusion inhibitor against MERS-CoV infection. J. Infect. Dis. 2003;212:1894–1903. doi: 10.1093/infdis/jiv325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Pascal KE, et al. Pre- and postexposure efficacy of fully human antibodies against Spike protein in a novel humanized mouse model of MERS-CoV infection. Proc. Natl Acad. Sci. USA. 2015;112:8738–8743. doi: 10.1073/pnas.1510830112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Li W, et al. Angiotensin-converting enzyme 2 is a functional receptor for the SARS coronavirus. Nature. 2003;426:450–454. doi: 10.1038/nature02145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Raj VS, et al. Dipeptidyl peptidase 4 is a functional receptor for the emerging human coronavirus-EMC. Nature. 2013;495:251–254. doi: 10.1038/nature12005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Huang X, et al. Human coronavirus HKU1 spike protein uses O-acetylated sialic acid as an attachment receptor determinant and employs hemagglutinin-esterase protein as a receptor-destroying enzyme. J. Virol. 2015;89:7202–7213. doi: 10.1128/JVI.00854-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Yeager CL, et al. Human aminopeptidase N is a receptor for human coronavirus 229E. Nature. 1992;357:420–422. doi: 10.1038/357420a0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Vlasak R, Luytjes W, Spaan W, Palese P. Human and bovine coronaviruses recognize sialic acid-containing receptors similar to those of influenza C viruses. Proc. Natl Acad. Sci. USA. 1988;85:4526–4529. doi: 10.1073/pnas.85.12.4526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Hofmann H, et al. Human coronavirus NL63 employs the severe acute respiratory syndrome coronavirus receptor for cellular entry. Proc. Natl Acad. Sci. USA. 2005;102:7988–7993. doi: 10.1073/pnas.0409465102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Gierer S, et al. The spike protein of the emerging betacoronavirus EMC uses a novel coronavirus receptor for entry, can be activated by TMPRSS2, and is targeted by neutralizing antibodies. J. Virol. 2013;87:5502–5511. doi: 10.1128/JVI.00128-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Qian Z, Dominguez SR, Holmes KV. Role of the spike glycoprotein of human Middle East respiratory syndrome coronavirus (MERS-CoV) in virus entry and syncytia formation. PLoS. ONE. 2013;8:e76469. doi: 10.1371/journal.pone.0076469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Shirato K, Kawase M, Matsuyama S. Middle East respiratory syndrome coronavirus infection mediated by the transmembrane serine protease TMPRSS2. J. Virol. 2013;87:12552–12561. doi: 10.1128/JVI.01890-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Millet JK, Whittaker GR. Host cell entry of Middle East respiratory syndrome coronavirus after two-step, furin-mediated activation of the spike protein. Proc. Natl Acad. Sci. USA. 2014;111:15214–15219. doi: 10.1073/pnas.1407087111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Zhou Y, et al. Protease inhibitors targeting coronavirus and filovirus entry. Antiviral Res. 2015;116:76–84. doi: 10.1016/j.antiviral.2015.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Lundin A, et al. Targeting membrane-bound viral RNA synthesis reveals potent inhibition of diverse coronaviruses including the middle East respiratory syndrome virus. PLoS Pathog. 2014;10:e1004166. doi: 10.1371/journal.ppat.1004166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Knoops K, et al. SARS-coronavirus replication is supported by a reticulovesicular network of modified endoplasmic reticulum. PLoS. Biol. 2008;6:e226. doi: 10.1371/journal.pbio.0060226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Barnard DL, Kumaki Y. Recent developments in anti-severe acute respiratory syndrome coronavirus chemotherapy. Future Virol. 2011;6:615–631. doi: 10.2217/fvl.11.33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Kilianski A, Baker SC. Cell-based antiviral screening against coronaviruses: developing virus-specific and broad-spectrum inhibitors. Antiviral Res. 2014;101:105–112. doi: 10.1016/j.antiviral.2013.11.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Cinatl J, et al. Treatment of SARS with human interferons. Lancet. 2003;362:293–294. doi: 10.1016/S0140-6736(03)13973-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.So LK, et al. Development of a standard treatment protocol for severe acute respiratory syndrome. Lancet. 2003;361:1615–1617. doi: 10.1016/S0140-6736(03)13265-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Pfefferle S, et al. The SARS–coronavirus–host interactome: identification of cyclophilins as target for pan-coronavirus inhibitors. PLoS Pathog. 2011;7:e1002331. doi: 10.1371/journal.ppat.1002331. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.de Wilde AH, et al. MERS-coronavirus replication induces severe in vitro cytopathology and is strongly inhibited by cyclosporin A or interferon-α treatment. J. Gen. Virol. 2013;94:1749–1760. doi: 10.1099/vir.0.052910-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Tanaka Y, Sato Y, Sasaki T. Suppression of coronavirus replication by cyclophilin inhibitors. Viruses. 2013;5:1250–1260. doi: 10.3390/v5051250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Falzarano D, et al. Inhibition of novel β coronavirus replication by a combination of interferon-α2b and ribavirin. Sci. Rep. 2013;3:1686. doi: 10.1038/srep01686. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Chan JF, et al. Broad-spectrum antivirals for the emerging Middle East respiratory syndrome coronavirus. J. Infect. 2013;67:606–616. doi: 10.1016/j.jinf.2013.09.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.de Wilde AH, et al. Screening of an FDA-approved compound library identifies four small-molecule inhibitors of Middle East respiratory syndrome coronavirus replication in cell culture. Antimicrob. Agents Chemother. 2014;58:4875–4884. doi: 10.1128/AAC.03011-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Dyall J, et al. Repurposing of clinically developed drugs for treatment of Middle East respiratory syndrome coronavirus infection. Antimicrob. Agents Chemother. 2014;58:4885–4893. doi: 10.1128/AAC.03036-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Kindrachuk J, et al. Antiviral potential of ERK/MAPK and PI3K/AKT/mTOR signaling modulation for Middle East respiratory syndrome coronavirus infection as identified by temporal kinome analysis. Antimicrob. Agents Chemother. 2015;59:1088–1099. doi: 10.1128/AAC.03659-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Liu, Q. et al. Testing of Middle East respiratory syndrome coronavirus replication inhibitors for their ability to block viral entry. Antimicrob. Agents Chemother.10.1128/AAC.03977-14 (2014). [DOI] [PMC free article] [PubMed]
  • 127.Elshabrawy HA, et al. Identification of a broad-spectrum antiviral small molecule against severe acute respiratory syndrome coronavirus and Ebola, Hendra, and Nipah viruses by using a novel high-throughput screening assay. J. Virol. 2014;88:4353–4365. doi: 10.1128/JVI.03050-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Chan JF, et al. Treatment with lopinavir/ritonavir or interferon-β1b improves outcome of MERS-CoV infection in a non-human primate model of common marmoset. J. Infect. Dis. 2015;212:1904–1913. doi: 10.1093/infdis/jiv392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Faure E, et al. Distinct immune response in two MERS-CoV-infected patients: can we go from bench to bedside? PLoS ONE. 2014;9:e88716. doi: 10.1371/journal.pone.0088716. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.AlGhamdi M, Mushtaq F, Awn N, Shalhoub S. MERS CoV infection in two renal transplant recipients: case report. Am. J. Transplant. 2015;15:1101–1104. doi: 10.1111/ajt.13085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Fukushima A, et al. Development of a chimeric DNA–RNA hammerhead ribozyme targeting SARS virus. Intervirology. 2009;52:92–99. doi: 10.1159/000215946. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Rider TH, et al. Broad-spectrum antiviral therapeutics. PLoS ONE. 2011;6:e22572. doi: 10.1371/journal.pone.0022572. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Mielech, A. M., Kilianski, A., Baez-Santos, Y. M., Mesecar, A. D. & Baker, S. C. MERS-CoV papain-like protease has deISGylating and deubiquitinating activities. Virology 450–451, 64–70 (2014). [DOI] [PMC free article] [PubMed]
  • 134.Barretto N, et al. The papain-like protease of severe acute respiratory syndrome coronavirus has deubiquitinating activity. J. Virol. 2005;79:15189–15198. doi: 10.1128/JVI.79.24.15189-15198.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Mielech AM, Chen Y, Mesecar AD, Baker SC. Nidovirus papain-like proteases: multifunctional enzymes with protease, deubiquitinating and deISGylating activities. Virus Res. 2014;194:184–190. doi: 10.1016/j.virusres.2014.01.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Hilgenfeld R. From SARS to MERS: crystallographic studies on coronaviral proteases enable antiviral drug design. FEBS J. 2014;281:4085–4096. doi: 10.1111/febs.12936. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Baez-Santos YM, St John SE, Mesecar AD. The SARS-coronavirus papain-like protease: structure, function and inhibition by designed antiviral compounds. Antiviral Res. 2015;115:21–38. doi: 10.1016/j.antiviral.2014.12.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Baez-Santos YM, et al. X-ray structural and biological evaluation of a series of potent and highly selective inhibitors of human coronavirus papain-like proteases. J. Med. Chem. 2014;57:2393–2412. doi: 10.1021/jm401712t. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Ratia K, et al. A noncovalent class of papain-like protease/deubiquitinase inhibitors blocks SARS virus replication. Proc. Natl Acad. Sci. USA. 2008;105:16119–16124. doi: 10.1073/pnas.0805240105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Lee H, et al. Inhibitor recognition specificity of MERS-CoV papain-like protease may differ from that of SARS-CoV. ACS Chem. Biol. 2015;10:1456–1465. doi: 10.1021/cb500917m. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Chaudhuri R, et al. Comparison of SARS and NL63 papain-like protease binding sites and binding site dynamics: inhibitor design implications. J. Mol. Biol. 2011;414:272–288. doi: 10.1016/j.jmb.2011.09.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Adedeji AO, Sarafianos SG. Antiviral drugs specific for coronaviruses in preclinical development. Curr. Opin. Virol. 2014;8:45–53. doi: 10.1016/j.coviro.2014.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Yang H, et al. Design of wide-spectrum inhibitors targeting coronavirus main proteases. PLoS Biol. 2005;3:e324. doi: 10.1371/journal.pbio.0030324. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Ren Z, et al. The newly emerged SARS-like coronavirus HCoV-EMC also has an 'Achilles' heel': current effective inhibitor targeting a 3C-like protease. Protein Cell. 2013;4:248–250. doi: 10.1007/s13238-013-2841-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Chen F, et al. In vitro susceptibility of 10 clinical isolates of SARS coronavirus to selected antiviral compounds. J. Clin. Virol. 2004;31:69–75. doi: 10.1016/j.jcv.2004.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Nukoolkarn V, Lee VS, Malaisree M, Aruksakulwong O, Hannongbua S. Molecular dynamic simulations analysis of ritonavir and lopinavir as SARS-CoV 3CLpro inhibitors. J. Theor. Biol. 2008;254:861–867. doi: 10.1016/j.jtbi.2008.07.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Olschlager S, Neyts J, Gunther S. Depletion of GTP pool is not the predominant mechanism by which ribavirin exerts its antiviral effect on Lassa virus. Antiviral Res. 2011;91:89–93. doi: 10.1016/j.antiviral.2011.05.006. [DOI] [PubMed] [Google Scholar]
  • 148.Falzarano D, et al. Treatment with interferon-α2b and ribavirin improves outcome in MERS-CoV-infected rhesus macaques. Nat. Med. 2013;19:1313–1317. doi: 10.1038/nm.3362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Warren TK, et al. Protection against filovirus diseases by a novel broad-spectrum nucleoside analogue BCX4430. Nature. 2014;508:402–405. doi: 10.1038/nature13027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Peters HL, et al. Design, synthesis and evaluation of a series of acyclic fleximer nucleoside analogues with anti-coronavirus activity. Bioorg. Med. Chem. Lett. 2015;25:2923–2926. doi: 10.1016/j.bmcl.2015.05.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Lu A, et al. Attenuation of SARS coronavirus by a short hairpin RNA expression plasmid targeting RNA-dependent RNA polymerase. Virology. 2004;324:84–89. doi: 10.1016/j.virol.2004.03.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Wang Z, et al. Inhibition of severe acute respiratory syndrome virus replication by small interfering RNAs in mammalian cells. J. Virol. 2004;78:7523–7527. doi: 10.1128/JVI.78.14.7523-7527.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Tanner JA, et al. The adamantane-derived bananins are potent inhibitors of the helicase activities and replication of SARS coronavirus. Chem. Biol. 2005;12:303–311. doi: 10.1016/j.chembiol.2005.01.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Kim MK, et al. 2,6-bis-arylmethyloxy-5-hydroxychromones with antiviral activity against both hepatitis C virus (HCV) and SARS-associated coronavirus (SCV) Eur. J. Med. Chem. 2011;46:5698–5704. doi: 10.1016/j.ejmech.2011.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Adedeji AO, et al. Severe acute respiratory syndrome coronavirus replication inhibitor that interferes with the nucleic acid unwinding of the viral helicase. Antimicrob. Agents Chemother. 2012;56:4718–4728. doi: 10.1128/AAC.00957-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Adedeji AO, et al. Evaluation of SSYA10-001 as a replication inhibitor of severe acute respiratory syndrome, mouse hepatitis, and Middle East respiratory syndrome coronaviruses. Antimicrob. Agents Chemother. 2014;58:4894–4898. doi: 10.1128/AAC.02994-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Zhao J, et al. Rapid generation of a mouse model for Middle East respiratory syndrome. Proc. Natl Acad. Sci. USA. 2014;111:4970–4975. doi: 10.1073/pnas.1323279111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Yang ZY, et al. Evasion of antibody neutralization in emerging severe acute respiratory syndrome coronaviruses. Proc. Natl Acad. Sci. USA. 2005;102:797–801. doi: 10.1073/pnas.0409065102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Weingartl H, et al. Immunization with modified vaccinia virus Ankara-based recombinant vaccine against severe acute respiratory syndrome is associated with enhanced hepatitis in ferrets. J. Virol. 2004;78:12672–12676. doi: 10.1128/JVI.78.22.12672-12676.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Coughlin MM, Prabhakar BS. Neutralizing human monoclonal antibodies to severe acute respiratory syndrome coronavirus: target, mechanism of action, and therapeutic potential. Rev. Med. Virol. 2012;22:2–17. doi: 10.1002/rmv.706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Li Y, et al. A humanized neutralizing antibody against MERS-CoV targeting the receptor-binding domain of the spike protein. Cell. Res. 2015;25:1237–1249. doi: 10.1038/cr.2015.113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Corti D, et al. Prophylactic and postexposure efficacy of a potent human monoclonal antibody against MERS coronavirus. Proc. Natl Acad. Sci. USA. 2015;112:10473–10478. doi: 10.1073/pnas.1510199112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Sainz B, et al. Inhibition of severe acute respiratory syndrome-associated coronavirus (SARS-CoV) infectivity by peptides analogous to the viral spike protein. Virus Res. 2006;120:146–155. doi: 10.1016/j.virusres.2006.03.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Zheng BJ, et al. Synthetic peptides outside the spike protein heptad repeat regions as potent inhibitors of SARS-associated coronavirus. Antiviral Ther. 2005;10:393–403. [PubMed] [Google Scholar]
  • 165.Liu IJ, et al. Identification of a minimal peptide derived from heptad repeat (HR) 2 of spike protein of SARS-CoV and combination of HR1-derived peptides as fusion inhibitors. Antiviral Res. 2009;81:82–87. doi: 10.1016/j.antiviral.2008.10.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Kilby JM, et al. Potent suppression of HIV-1 replication in humans by T-20, a peptide inhibitor of gp41-mediated virus entry. Nat. Med. 1998;4:1302–1307. doi: 10.1038/3293. [DOI] [PubMed] [Google Scholar]
  • 167.Greenberg ML, Cammack N. Resistance to enfuvirtide, the first HIV fusion inhibitor. J. Antimicrob. Chemother. 2004;54:333–340. doi: 10.1093/jac/dkh330. [DOI] [PubMed] [Google Scholar]
  • 168.Izumi K, et al. Mechanism of resistance to S138A substituted enfuvirtide and its application to peptide design. Int. J. Biochem. Cell Biol. 2013;45:908–915. doi: 10.1016/j.biocel.2013.01.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Zhang Y, et al. Silencing SARS-CoV Spike protein expression in cultured cells by RNA interference. FEBS Lett. 2004;560:141–146. doi: 10.1016/S0014-5793(04)00087-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Wu CJ, Huang HW, Liu CY, Hong CF, Chan YL. Inhibition of SARS-CoV replication by siRNA. Antiviral Res. 2005;65:45–48. doi: 10.1016/j.antiviral.2004.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Tang Q, Li B, Woodle M, Lu PY. Application of siRNA against SARS in the rhesus macaque model. Methods Mol. Biol. 2008;442:139–158. doi: 10.1007/978-1-59745-191-8_11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Li BJ, et al. Using siRNA in prophylactic and therapeutic regimens against SARS coronavirus in Rhesus macaque. Nat. Med. 2005;11:944–951. doi: 10.1038/nm1280. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.O'Keefe BR, et al. Broad-spectrum in vitro activity and in vivo efficacy of the antiviral protein griffithsin against emerging viruses of the family Coronaviridae. J. Virol. 2010;84:2511–2521. doi: 10.1128/JVI.02322-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Barton C, et al. Activity of and effect of subcutaneous treatment with the broad-spectrum antiviral lectin griffithsin in two laboratory rodent models. Antimicrob. Agents Chemother. 2014;58:120–127. doi: 10.1128/AAC.01407-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Yang Y, et al. The structural and accessory proteins M, ORF 4a, ORF 4b, and ORF 5 of Middle East respiratory syndrome coronavirus (MERS-CoV) are potent interferon antagonists. Protein Cell. 2013;4:951–961. doi: 10.1007/s13238-013-3096-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Cui L, et al. The nucleocapsid protein of coronaviruses acts as a viral suppressor of RNA silencing in mammalian cells. J. Virol. 2015;89:9029–9043. doi: 10.1128/JVI.01331-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Niemeyer D, et al. Middle East respiratory syndrome coronavirus accessory protein 4a is a type I interferon antagonist. J. Virol. 2013;87:12489–12495. doi: 10.1128/JVI.01845-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Siu KL, et al. Middle east respiratory syndrome coronavirus 4a protein is a double-stranded RNA-binding protein that suppresses PACT-induced activation of RIG-I and MDA5 in the innate antiviral response. J. Virol. 2014;88:4866–4876. doi: 10.1128/JVI.03649-13. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 179.He ML, et al. Development of interfering RNA agents to inhibit SARS-associated coronavirus infection and replication. Hong Kong Med. J. 2009;3:28–31. [PubMed] [Google Scholar]
  • 180.Akerstrom S, Mirazimi A, Tan YJ. Inhibition of SARS-CoV replication cycle by small interference RNAs silencing specific SARS proteins, 7a/7b, 3a/3b and S. Antiviral Res. 2007;73:219–227. doi: 10.1016/j.antiviral.2006.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Pervushin K, et al. Structure and inhibition of the SARS coronavirus envelope protein ion channel. PLoS Pathog. 2009;5:e1000511. doi: 10.1371/journal.ppat.1000511. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Wilson L, Gage P, Ewart G. Hexamethylene amiloride blocks E protein ion channels and inhibits coronavirus replication. Virology. 2006;353:294–306. doi: 10.1016/j.virol.2006.05.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Lin SY, et al. Structural basis for the identification of the N-terminal domain of coronavirus nucleocapsid protein as an antiviral target. J. Med. Chem. 2014;57:2247–2257. doi: 10.1021/jm500089r. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Wolf MC, et al. A broad-spectrum antiviral targeting entry of enveloped viruses. Proc. Natl Acad. Sci. USA. 2010;107:3157–3162. doi: 10.1073/pnas.0909587107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Vigant F, et al. A mechanistic paradigm for broad-spectrum antivirals that target virus-cell fusion. PLoS Pathog. 2013;9:e1003297. doi: 10.1371/journal.ppat.1003297. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Hollmann A, Castanho MA, Lee B, Santos NC. Singlet oxygen effects on lipid membranes: implications for the mechanism of action of broad-spectrum viral fusion inhibitors. Biochem. J. 2014;459:161–170. doi: 10.1042/BJ20131058. [DOI] [PubMed] [Google Scholar]
  • 187.Hollmann A, et al. Effects of singlet oxygen generated by a broad-spectrum viral fusion inhibitor on membrane nanoarchitecture. Nanomedicine. 2015;11:1163–1167. doi: 10.1016/j.nano.2015.02.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Menachery VD, et al. Pathogenic influenza viruses and coronaviruses utilize similar and contrasting approaches to control interferon-stimulated gene responses. mBio. 2014;5:e01174–e01114. doi: 10.1128/mBio.01174-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189.Lau SK, et al. Delayed induction of proinflammatory cytokines and suppression of innate antiviral response by the novel Middle East respiratory syndrome coronavirus: implications for pathogenesis and treatment. J. Gen. Virol. 2013;94:2679–2690. doi: 10.1099/vir.0.055533-0. [DOI] [PubMed] [Google Scholar]
  • 190.Josset L, et al. Cell host response to infection with novel human coronavirus EMC predicts potential antivirals and important differences with SARS coronavirus. mBio. 2013;4:e00165–e00113. doi: 10.1128/mBio.00165-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Haagmans BL, et al. Pegylated interferon-α protects type 1 pneumocytes against SARS coronavirus infection in macaques. Nat. Med. 2004;10:290–293. doi: 10.1038/nm1001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Hart BJ, et al. Interferon-β and mycophenolic acid are potent inhibitors of Middle East respiratory syndrome coronavirus in cell-based assays. J. Gen. Virol. 2014;95:571–577. doi: 10.1099/vir.0.061911-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Rosenfeld MR, et al. A multi-institution Phase II study of poly-ICLC and radiotherapy with concurrent and adjuvant temozolomide in adults with newly diagnosed glioblastoma. Neuro. Oncol. 2010;12:1071–1077. doi: 10.1093/neuonc/noq071. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 194.Okada H, et al. Induction of CD8+ T-cell responses against novel glioma-associated antigen peptides and clinical activity by vaccinations with α-type 1 polarized dendritic cells and polyinosinic-polycytidylic acid stabilized by lysine and carboxymethylcellulose in patients with recurrent malignant glioma. J. Clin. Oncol. 2011;29:330–336. doi: 10.1200/JCO.2010.30.7744. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Rossignol JF. Nitazoxanide: a first-in-class broad-spectrum antiviral agent. Antiviral Res. 2014;110:94–103. doi: 10.1016/j.antiviral.2014.07.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Haffizulla J, et al. Effect of nitazoxanide in adults and adolescents with acute uncomplicated influenza: a double-blind, randomised, placebo-controlled, Phase 2b/3 trial. Lancet Infect. Dis. 2014;14:609–618. doi: 10.1016/S1473-3099(14)70717-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Rossignol JF, Kabil SM, El-Gohary Y, Elfert A, Keeffe EB. Clinical trial: randomized, double-blind, placebo-controlled study of nitazoxanide monotherapy for the treatment of patients with chronic hepatitis C genotype 4. Aliment. Pharmacol. Ther. 2008;28:574–580. doi: 10.1111/j.1365-2036.2008.03781.x. [DOI] [PubMed] [Google Scholar]
  • 198.Wohlford-Lenane CL, et al. Rhesus theta-defensin prevents death in a mouse model of severe acute respiratory syndrome coronavirus pulmonary disease. J. Virol. 2009;83:11385–11390. doi: 10.1128/JVI.01363-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Wiley JA, et al. Inducible bronchus-associated lymphoid tissue elicited by a protein cage nanoparticle enhances protection in mice against diverse respiratory viruses. PLoS. ONE. 2009;4:e7142. doi: 10.1371/journal.pone.0007142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Carbajo-Lozoya J, et al. Human coronavirus NL63 replication is cyclophilin A-dependent and inhibited by non-immunosuppressive cyclosporine A-derivatives including Alisporivir. Virus Res. 2014;184:44–53. doi: 10.1016/j.virusres.2014.02.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Ohnuma K, et al. Inhibition of Middle East respiratory syndrome coronavirus infection by anti-CD26 monoclonal antibody. J. Virol. 2013;87:13892–13899. doi: 10.1128/JVI.02448-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Huentelman MJ, et al. Structure-based discovery of a novel angiotensin-converting enzyme 2 inhibitor. Hypertension. 2004;44:903–906. doi: 10.1161/01.HYP.0000146120.29648.36. [DOI] [PubMed] [Google Scholar]
  • 203.Han DP, Penn-Nicholson A, Cho MW. Identification of critical determinants on ACE2 for SARS-CoV entry and development of a potent entry inhibitor. Virology. 2006;350:15–25. doi: 10.1016/j.virol.2006.01.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Chen YT, et al. In vitro and in vivo studies of the trypanocidal properties of WRR-483 against Trypanosoma cruzi. PLoS Negl. Trop. Dis. 2010;4:e825. doi: 10.1371/journal.pntd.0000825. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Ndao M, et al. A cysteine protease inhibitor rescues mice from a lethal Cryptosporidium parvum infection. Antimicrob. Agents Chemother. 2013;57:6063–6073. doi: 10.1128/AAC.00734-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 206.Vermeire JJ, Lantz LD, Caffrey CR. Cure of hookworm infection with a cysteine protease inhibitor. PLoS Negl. Trop. Dis. 2012;6:e1680. doi: 10.1371/journal.pntd.0001680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 207.Sai JK, et al. Efficacy of camostat mesilate against dyspepsia associated with non-alcoholic mild pancreatic disease. J. Gastroenterol. 2010;45:335–341. doi: 10.1007/s00535-009-0148-1. [DOI] [PubMed] [Google Scholar]
  • 208.Talukdar R, Tandon RK. Pancreatic stellate cells: new target in the treatment of chronic pancreatitis. J. Gastroenterol. Hepatol. 2008;23:34–41. doi: 10.1111/j.1440-1746.2007.05056.x. [DOI] [PubMed] [Google Scholar]
  • 209.Burkard C, et al. ATP1A1-mediated Src signaling inhibits coronavirus entry into host cells. J. Virol. 2015;89:4434–4448. doi: 10.1128/JVI.03274-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Savarino A, Boelaert JR, Cassone A, Majori G, Cauda R. Effects of chloroquine on viral infections: an old drug against today's diseases? Lancet Infect. Dis. 2003;3:722–727. doi: 10.1016/S1473-3099(03)00806-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Keyaerts E, Vijgen L, Maes P, Neyts J, Van Ranst M. In vitro inhibition of severe acute respiratory syndrome coronavirus by chloroquine. Biochem. Biophys. Res. Commun. 2004;323:264–268. doi: 10.1016/j.bbrc.2004.08.085. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Vincent MJ, et al. Chloroquine is a potent inhibitor of SARS coronavirus infection and spread. Virol. J. 2005;2:69. doi: 10.1186/1743-422X-2-69. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Kono M, et al. Inhibition of human coronavirus 229E infection in human epithelial lung cells (L132) by chloroquine: involvement of p38 MAPK and ERK. Antiviral Res. 2008;77:150–152. doi: 10.1016/j.antiviral.2007.10.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Madrid PB, et al. A systematic screen of FDA-approved drugs for inhibitors of biological threat agents. PloS ONE. 2013;8:e60579. doi: 10.1371/journal.pone.0060579. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Barnard DL, et al. Evaluation of immunomodulators, interferons and known in vitro SARS-coV inhibitors for inhibition of SARS-coV replication in BALB/c mice. Antiviral Chem. Chemother. 2006;17:275–284. doi: 10.1177/095632020601700505. [DOI] [PubMed] [Google Scholar]
  • 216.Zhang N, Jiang S, Du L. Current advancements and potential strategies in the development of MERS-CoV vaccines. Expert Rev. Vaccines. 2014;13:761–774. doi: 10.1586/14760584.2014.912134. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Papaneri AB, et al. Middle East respiratory syndrome: obstacles and prospects for vaccine development. Expert Rev. Vaccines. 2015;14:949–962. doi: 10.1586/14760584.2015.1036033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Almazan F, et al. Engineering a replication-competent, propagation-defective Middle East respiratory syndrome coronavirus as a vaccine candidate. mBio. 2013;4:e00650–00613. doi: 10.1128/mBio.00650-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Wang L, et al. Evaluation of candidate vaccine approaches for MERS-CoV. Nat. Commun. 2015;6:7712. doi: 10.1038/ncomms8712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Song F, et al. Middle East respiratory syndrome coronavirus spike protein delivered by modified vaccinia virus Ankara efficiently induces virus-neutralizing antibodies. J. Virol. 2013;87:11950–11954. doi: 10.1128/JVI.01672-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 221.Volz A, et al. Protective efficacy of recombinant modified vaccinia virus Ankara delivering Middle East respiratory syndrome coronavirus spike glycoprotein. J. Virol. 2015;89:8651–8656. doi: 10.1128/JVI.00614-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Kim E, et al. Immunogenicity of an adenoviral-based Middle East respiratory syndrome coronavirus vaccine in BALB/c mice. Vaccine. 2014;32:5975–5982. doi: 10.1016/j.vaccine.2014.08.058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Guo X, et al. Systemic and mucosal immunity in mice elicited by a single immunization with human adenovirus type 5 or 41 vector-based vaccines carrying the spike protein of Middle East respiratory syndrome coronavirus. Immunology. 2015;145:476–484. doi: 10.1111/imm.12462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Czub M, Weingartl H, Czub S, He R, Cao J. Evaluation of modified vaccinia virus Ankara based recombinant SARS vaccine in ferrets. Vaccine. 2005;23:2273–2279. doi: 10.1016/j.vaccine.2005.01.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 225.Coleman CM, et al. Purified coronavirus spike protein nanoparticles induce coronavirus neutralizing antibodies in mice. Vaccine. 2014;32:3169–3174. doi: 10.1016/j.vaccine.2014.04.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 226.Zhao G, et al. A safe and convenient pseudovirus-based inhibition assay to detect neutralizing antibodies and screen for viral entry inhibitors against the novel human coronavirus MERS-CoV. Virol. J. 2013;10:266. doi: 10.1186/1743-422X-10-266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Du L, et al. Identification of a receptor-binding domain in the S protein of the novel human coronavirus Middle East respiratory syndrome coronavirus as an essential target for vaccine development. J. Virol. 2013;87:9939–9942. doi: 10.1128/JVI.01048-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Mou H, et al. The receptor binding domain of the new Middle East respiratory syndrome coronavirus maps to a 231-residue region in the spike protein that efficiently elicits neutralizing antibodies. J. Virol. 2013;87:9379–9383. doi: 10.1128/JVI.01277-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 229.Du L, et al. A truncated receptor-binding domain of MERS-CoV spike protein potently inhibits MERS-CoV infection and induces strong neutralizing antibody responses: implication for developing therapeutics and vaccines. PloS ONE. 2013;8:e81587. doi: 10.1371/journal.pone.0081587. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Ma C, et al. Intranasal vaccination with recombinant receptor-binding domain of MERS-CoV spike protein induces much stronger local mucosal immune responses than subcutaneous immunization: Implication for designing novel mucosal MERS vaccines. Vaccine. 2014;32:2100–2108. doi: 10.1016/j.vaccine.2014.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 231.Ma C, et al. Searching for an ideal vaccine candidate among different MERS coronavirus receptor-binding fragments-The importance of immunofocusing in subunit vaccine design. Vaccine. 2014;32:6170–6176. doi: 10.1016/j.vaccine.2014.08.086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Lan J, et al. Tailoring subunit vaccine immunity with adjuvant combinations and delivery routes using the Middle East respiratory coronavirus (MERS-CoV) receptor-binding domain as an antigen. PloS ONE. 2014;9:e112602. doi: 10.1371/journal.pone.0112602. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Zhang, N. et al. Identification of an ideal adjuvant for receptor-binding domain-based subunit vaccines against Middle East respiratory syndrome coronavirus. Cell. Mol. Immunol.10.1038/cmi.2015.03 (2015). [DOI] [PMC free article] [PubMed]
  • 234.Lassnig C, et al. Development of a transgenic mouse model susceptible to human coronavirus 229E. Proc. Natl Acad. Sci. USA. 2005;102:8275–8280. doi: 10.1073/pnas.0408589102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Dijkman R, et al. Seroconversion to HCoV-NL63 in rhesus macaques. Viruses. 2009;1:647–656. doi: 10.3390/v1030647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Jacomy H, Fragoso G, Almazan G, Mushynski WE, Talbot PJ. Human coronavirus OC43 infection induces chronic encephalitis leading to disabilities in BALB/C mice. Virology. 2006;349:335–346. doi: 10.1016/j.virol.2006.01.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 237.Sutton TC, Subbarao K. Development of animal models against emerging coronaviruses: from SARS to MERS coronavirus. Virology. 2015;479–480:247–258. doi: 10.1016/j.virol.2015.02.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Fouchier RA, et al. Aetiology: Koch's postulates fulfilled for SARS virus. Nature. 2003;423:240. doi: 10.1038/423240a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 239.McCray PB, et al. Lethal infection of K18-hACE2 mice infected with severe acute respiratory syndrome coronavirus. J. Virol. 2007;81:813–821. doi: 10.1128/JVI.02012-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 240.Tseng CT, et al. Severe acute respiratory syndrome coronavirus infection of mice transgenic for the human Angiotensin-converting enzyme 2 virus receptor. J. Virol. 2007;81:1162–1173. doi: 10.1128/JVI.01702-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Roberts A, et al. A mouse-adapted SARS-coronavirus causes disease and mortality in BALB/c mice. PLoS Pathog. 2007;3:e5. doi: 10.1371/journal.ppat.0030005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Munster VJ, de Wit E, Feldmann H. Pneumonia from human coronavirus in a macaque model. N. Engl. J. Med. 2013;368:1560–1562. doi: 10.1056/NEJMc1215691. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Yao Y, et al. An animal model of MERS produced by infection of rhesus macaques with MERS coronavirus. J. Infect. Dis. 2014;209:236–242. doi: 10.1093/infdis/jit590. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Falzarano D, et al. Infection with MERS-CoV causes lethal pneumonia in the common marmoset. PLoS Pathog. 2014;10:e1004250. doi: 10.1371/journal.ppat.1004250. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Raj VS, et al. Adenosine deaminase acts as a natural antagonist for dipeptidyl peptidase 4-mediated entry of the Middle East respiratory syndrome coronavirus. J. Virol. 2014;88:1834–1838. doi: 10.1128/JVI.02935-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 246.de Wit E, et al. The Middle East respiratory syndrome coronavirus (MERS-CoV) does not replicate in Syrian hamsters. PLoS One. 2013;8:e69127. doi: 10.1371/journal.pone.0069127. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 247.Coleman CM, Matthews KL, Goicochea L, Frieman MB. Wild-type and innate immune-deficient mice are not susceptible to the Middle East respiratory syndrome coronavirus. J. General Virol. 2014;95:408–412. doi: 10.1099/vir.0.060640-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 248.Agrawal AS, et al. Generation of a transgenic mouse model of Middle East respiratory syndrome coronavirus infection and disease. J. Virol. 2015;89:3659–3670. doi: 10.1128/JVI.03427-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 249.Hung IF, et al. Viral loads in clinical specimens and SARS manifestations. Emerg. Infect. Dis. 2004;10:1550–1557. doi: 10.3201/eid1009.040058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 250.Abd El Wahed A, Patel P, Heidenreich D, Hufert FT, Weidmann M. Reverse transcription recombinase polymerase amplification assay for the detection of middle East respiratory syndrome coronavirus. PLoS Curr. 2013;5:62df1c7c75ffc96cd59034531e2e8364. doi: 10.1371/currents.outbreaks.62df1c7c75ffc96cd59034531e2e8364. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 251.Shirato K, et al. Detection of Middle East respiratory syndrome coronavirus using reverse transcription loop-mediated isothermal amplification (RT-LAMP) Virol. J. 2014;11:139. doi: 10.1186/1743-422X-11-139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Song D, et al. Development and validation of a rapid immunochromatographic assay for detection of Middle East respiratory syndrome coronavirus antigen in dromedary camels. J. Clin. Microbiol. 2015;53:1178–1182. doi: 10.1128/JCM.03096-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 253.Chen Y, et al. A sensitive and specific antigen detection assay for Middle East respiratory syndrome coronavirus. Emerg. Microbes Infect. 2015;4:e26. doi: 10.1038/emi.2015.26. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 254.Sridhar S, et al. A systematic approach to novel virus discovery in emerging infectious disease outbreaks. J. Mol. Diagn. 2015;17:230–241. doi: 10.1016/j.jmoldx.2014.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 255.Corman VM, et al. Assays for laboratory confirmation of novel human coronavirus (hCoV-EMC) infections. Euro Surveill. 2012;17:20334. doi: 10.2807/ese.17.49.20334-en. [DOI] [PubMed] [Google Scholar]
  • 256.Corman VM, et al. Detection of a novel human coronavirus by real-time reverse-transcription polymerase chain reaction. Euro Surveill. 2012;17:20285. doi: 10.2807/ese.17.39.20285-en. [DOI] [PubMed] [Google Scholar]
  • 257.Chan JF, et al. Development and evaluation of novel real-time RT-PCR assays with locked nucleic acid probes targeting the leader sequences of human pathogenic coronaviruses. J. Clin. Microbiol. 2015;53:2722–2726. doi: 10.1128/JCM.01224-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Chan JF, et al. Differential cell line susceptibility to the emerging novel human betacoronavirus 2c EMC/2012: implications for disease pathogenesis and clinical manifestation. J. Infect. Dis. 2013;207:1743–1752. doi: 10.1093/infdis/jit123. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 259.Pyrc K, et al. Culturing the unculturable: human coronavirus HKU1 infects, replicates, and produces progeny virions in human ciliated airway epithelial cell cultures. J. Virol. 2010;84:11255–11263. doi: 10.1128/JVI.00947-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 260.Dijkman R, et al. Isolation and characterization of current human coronavirus strains in primary human epithelial cell cultures reveal differences in target cell tropism. J. Virol. 2013;87:6081–6090. doi: 10.1128/JVI.03368-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Dominguez SR, Travanty EA, Qian Z, Mason RJ. Human coronavirus HKU1 infection of primary human type II alveolar epithelial cells: cytopathic effects and innate immune response. PLoS ONE. 2013;8:e70129. doi: 10.1371/journal.pone.0070129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.Dominguez SR, et al. Isolation, propagation, genome analysis and epidemiology of HKU1 betacoronaviruses. J. General Virol. 2014;95:836–848. doi: 10.1099/vir.0.059832-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Zhou J, et al. Active replication of Middle East respiratory syndrome coronavirus and aberrant induction of inflammatory cytokines and chemokines in human macrophages: implications for pathogenesis. J. Infect. Dis. 2014;209:1331–1342. doi: 10.1093/infdis/jit504. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 264.Chu, H. et al. Middle East respiratory syndrome coronavirus efficiently infects human primary T lymphocytes and activates the extrinsic and intrinsic apoptosis pathways. J. Infect. Dis.10.1093/infdis/jiv380 (2015). [DOI] [PMC free article] [PubMed]
  • 265.Chu H, et al. Productive replication of Middle East respiratory syndrome coronavirus in monocyte-derived dendritic cells modulates innate immune response. Virology. 2014;454–455:197–205. doi: 10.1016/j.virol.2014.02.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.Chan RW, et al. Tropism of and innate immune responses to the novel human betacoronavirus lineage C virus in human ex vivo respiratory organ cultures. J. Virol. 2013;87:6604–6614. doi: 10.1128/JVI.00009-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Chan RW, et al. Tropism and replication of Middle East respiratory syndrome coronavirus from dromedary camels in the human respiratory tract: an in-vitro and ex-vivo study. Lancet Respir.Med. 2014;2:813–822. doi: 10.1016/S2213-2600(14)70158-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Muller MA, et al. Human coronavirus EMC does not require the SARS-coronavirus receptor and maintains broad replicative capability in mammalian cell lines. mBio. 2012;3:e00515-12. doi: 10.1128/mBio.00515-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 269.Eckerle I, et al. Replicative capacity of MERS coronavirus in livestock cell lines. Emerg. Infect. Dis. 2014;20:276–279. doi: 10.3201/eid2002.131182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 270.Chan KH, et al. Cross-reactive antibodies in convalescent SARS patients' sera against the emerging novel human coronavirus EMC (2012) by both immunofluorescent and neutralizing antibody tests. J. Infect. 2013;67:130–140. doi: 10.1016/j.jinf.2013.03.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 271.Perera RA, et al. Seroepidemiology for MERS coronavirus using microneutralisation and pseudoparticle virus neutralisation assays reveal a high prevalence of antibody in dromedary camels in Egypt, June 2013. Euro Surveill. 2013;18:20574. doi: 10.2807/1560-7917.ES2013.18.36.20574. [DOI] [PubMed] [Google Scholar]
  • 272.Hung IF, et al. Hyperimmune IV immunoglobulin treatment: a multicenter double-blind randomized controlled trial for patients with severe 2009 influenza A(H1N1) infection. Chest. 2013;144:464–473. doi: 10.1378/chest.12-2907. [DOI] [PubMed] [Google Scholar]
  • 273.Hung IF, et al. Convalescent plasma treatment reduced mortality in patients with severe pandemic influenza A (H1N1) 2009 virus infection. Clin. Infect. Dis. 2011;52:447–456. doi: 10.1093/cid/ciq106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 274.WHO MERS-CoV Research Group. State of knowledge and data gaps of Middle East respiratory syndrome coronavirus (MERS-CoV) in humans. PLoS Curr.5, ecurrents.outbreaks.0bf719e352e7478f8ad85fa30127ddb8 (2013). [DOI] [PMC free article] [PubMed]
  • 275.Kilianski A, Mielech AM, Deng X, Baker SC. Assessing activity and inhibition of Middle East respiratory syndrome coronavirus papain-like and 3C-like proteases using luciferase-based biosensors. J. Virol. 2013;87:11955–11962. doi: 10.1128/JVI.02105-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.Agnihothram S, et al. A mouse model for Betacoronavirus subgroup 2c using a bat coronavirus strain HKU5 variant. mBio. 2014;5:e00047–00014. doi: 10.1128/mBio.00047-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Reichard O, Yun ZB, Sonnerborg A, Weiland O. Hepatitis C viral RNA titers in serum prior to, during, and after oral treatment with ribavirin for chronic hepatitis C. J. Med. Virol. 1993;41:99–102. doi: 10.1002/jmv.1890410203. [DOI] [PubMed] [Google Scholar]
  • 278.Hall CB, Walsh EE, Hruska JF, Betts RF, Hall WJ. Ribavirin treatment of experimental respiratory syncytial viral infection. A controlled double-blind study in young adults. JAMA. 1983;249:2666–2670. doi: 10.1001/jama.1983.03330430042027. [DOI] [PubMed] [Google Scholar]
  • 279.Ascioglu S, Leblebicioglu H, Vahaboglu H, Chan KA. Ribavirin for patients with Crimean-Congo haemorrhagic fever: a systematic review and meta-analysis. J. Antimicrob. Chemother. 2011;66:1215–1222. doi: 10.1093/jac/dkr136. [DOI] [PubMed] [Google Scholar]
  • 280.Bausch DG, Hadi CM, Khan SH, Lertora JJ. Review of the literature and proposed guidelines for the use of oral ribavirin as postexposure prophylaxis for Lassa fever. Clin. Infect. Dis. 2010;51:1435–1441. doi: 10.1086/657315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 281.Lee C, et al. Aryl diketoacids (ADK) selectively inhibit duplex DNA-unwinding activity of SARS coronavirus NTPase/helicase. Bioorg. Med. Chem. Lett. 2009;19:1636–1638. doi: 10.1016/j.bmcl.2009.02.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 282.Liao HI, et al. mRNA display design of fibronectin-based intrabodies that detect and inhibit severe acute respiratory syndrome coronavirus nucleocapsid protein. J. Biol. Chem. 2009;284:17512–17520. doi: 10.1074/jbc.M901547200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 283.Simmons G, et al. Inhibitors of cathepsin L prevent severe acute respiratory syndrome coronavirus entry. Proc. Natl Acad. Sci. USA. 2005;102:11876–11881. doi: 10.1073/pnas.0505577102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 284.Hatesuer B, et al. Tmprss2 is essential for influenza H1N1 virus pathogenesis in mice. PLoS Pathog. 2013;9:e1003774. doi: 10.1371/journal.ppat.1003774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 285.Abe M, et al. TMPRSS2 is an activating protease for respiratory parainfluenza viruses. J. Virol. 2013;87:11930–11935. doi: 10.1128/JVI.01490-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 286.Bertram S, et al. TMPRSS2 activates the human coronavirus 229E for cathepsin-independent host cell entry and is expressed in viral target cells in the respiratory epithelium. J. Virol. 2013;87:6150–6160. doi: 10.1128/JVI.03372-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 287.Lokugamage KG, et al. Chimeric coronavirus-like particles carrying severe acute respiratory syndrome coronavirus (SCoV) S protein protect mice against challenge with SCoV. Vaccine. 2008;26:797–808. doi: 10.1016/j.vaccine.2007.11.092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 288.Lu X, et al. Immune responses against severe acute respiratory syndrome coronavirus induced by virus-like particles in mice. Immunology. 2007;122:496–502. doi: 10.1111/j.1365-2567.2007.02676.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 289.See RH, et al. Comparative evaluation of two severe acute respiratory syndrome (SARS) vaccine candidates in mice challenged with SARS coronavirus. J. General Virol. 2006;87:641–650. doi: 10.1099/vir.0.81579-0. [DOI] [PubMed] [Google Scholar]
  • 290.Spruth M, et al. A double-inactivated whole virus candidate SARS coronavirus vaccine stimulates neutralising and protective antibody responses. Vaccine. 2006;24:652–661. doi: 10.1016/j.vaccine.2005.08.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 291.Qin E, et al. Immunogenicity and protective efficacy in monkeys of purified inactivated Vero-cell SARS vaccine. Vaccine. 2006;24:1028–1034. doi: 10.1016/j.vaccine.2005.06.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 292.Zhou J, et al. Immunogenicity, safety, and protective efficacy of an inactivated SARS-associated coronavirus vaccine in rhesus monkeys. Vaccine. 2005;23:3202–3209. doi: 10.1016/j.vaccine.2004.11.075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 293.Lamirande EW, et al. A live attenuated severe acute respiratory syndrome coronavirus is immunogenic and efficacious in golden Syrian hamsters. J. Virol. 2008;82:7721–7724. doi: 10.1128/JVI.00304-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 294.Netland J, et al. Immunization with an attenuated severe acute respiratory syndrome coronavirus deleted in E protein protects against lethal respiratory disease. Virology. 2010;399:120–128. doi: 10.1016/j.virol.2010.01.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 295.Fett C, DeDiego ML, Regla-Nava JA, Enjuanes L, Perlman S. Complete protection against severe acute respiratory syndrome coronavirus-mediated lethal respiratory disease in aged mice by immunization with a mouse-adapted virus lacking E protein. J. Virol. 2013;87:6551–6559. doi: 10.1128/JVI.00087-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 296.Stadler K, et al. SARS — beginning to understand a new virus. Nat. Rev. Microbiol. 2003;1:209–218. doi: 10.1038/nrmicro775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 297.Peiris JS, Guan Y, Yuen KY. Severe acute respiratory syndrome. Nat. Med. 2004;10:88–97. doi: 10.1038/nm1143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 298.Haagmans BL, et al. Co-circulation of three camel coronavirus species and recombination of MERS-CoVs in Saudi Arabia. Science. 2016;351:77–81. doi: 10.1126/science.aad1283. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary information S1 (figure) (1.6MB, pdf)

Global Distribution of SARS and MERS (PDF 1587 kb)

Supplementary information S2 (table) (139.6KB, pdf)

Comparisons of the clinical and laboratory features between SARS and MERS (PDF 139 kb)


Articles from Nature Reviews. Drug Discovery are provided here courtesy of Nature Publishing Group

RESOURCES