Next Article in Journal
Hepatitis A Strain Linked to the European Outbreaks During Gay Events between 2016 and 2017, Identified in a Brazilian Homosexual Couple in 2017
Next Article in Special Issue
Lack of Middle East Respiratory Syndrome Coronavirus Transmission in Rabbits
Previous Article in Journal
A Novel Orthohepadnavirus Identified in a Dead Maxwell’s Duiker (Philantomba maxwellii) in Taï National Park, Côte d’Ivoire
Previous Article in Special Issue
Efficacy of an Adjuvanted Middle East Respiratory Syndrome Coronavirus Spike Protein Vaccine in Dromedary Camels and Alpacas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Host Determinants of MERS-CoV Transmission and Pathogenesis

by
W. Widagdo
,
Syriam Sooksawasdi Na Ayudhya
,
Gadissa B. Hundie
and
Bart L. Haagmans
*
Department of Viroscience, Erasmus Medical Center, 3025 Rotterdam, The Netherlands
*
Author to whom correspondence should be addressed.
Submission received: 1 March 2019 / Revised: 11 March 2019 / Accepted: 13 March 2019 / Published: 19 March 2019
(This article belongs to the Special Issue MERS-CoV)

Abstract

:
Middle East respiratory syndrome coronavirus (MERS-CoV) is a zoonotic pathogen that causes respiratory infection in humans, ranging from asymptomatic to severe pneumonia. In dromedary camels, the virus only causes a mild infection but it spreads efficiently between animals. Differences in the behavior of the virus observed between individuals, as well as between humans and dromedary camels, highlight the role of host factors in MERS-CoV pathogenesis and transmission. One of these host factors, the MERS-CoV receptor dipeptidyl peptidase-4 (DPP4), may be a critical determinant because it is variably expressed in MERS-CoV-susceptible species as well as in humans. This could partially explain inter- and intraspecies differences in the tropism, pathogenesis, and transmissibility of MERS-CoV. In this review, we explore the role of DPP4 and other host factors in MERS-CoV transmission and pathogenesis—such as sialic acids, host proteases, and interferons. Further characterization of these host determinants may potentially offer novel insights to develop intervention strategies to tackle ongoing outbreaks.

1. Introduction

Middle East respiratory syndrome coronavirus (MERS-CoV) is a novel pathogen that was isolated in late 2012 [1]. Since then, the virus has caused multiple outbreaks and infected more than 2000 individuals, [2] who then develop a respiratory infection ranging in severity from asymptomatic to fatal [3,4]. Severe-to-fatal MERS-CoV patients have a higher chance of transmitting this virus since they shed a higher amount of virus progeny in comparison to the asymptomatic-to-mild ones [5,6,7,8]. Identifying and quarantining these patients in healthcare facilities where outbreaks have occurred, together with implementing proper infection control, has been effective in reducing transmission and containing these outbreaks [9,10]. However, new MERS-CoV cases are still being reported, especially in the Arabian Peninsula [2,11]. This is partly due to the continuous zoonotic introduction of this virus to the human population in this region by dromedaries [12]. The dromedary camel is the only animal species that has been reported to transmit this virus to humans [13,14,15,16]. MERS-CoV infection in these animals merely causes mild upper respiratory tract infection [17,18], but seroepidemiological studies showed that this virus has been circulating in dromedary camels for decades, suggesting the efficient transmission of MERS-CoV in this species [19,20,21,22].
Although the clinical manifestations, as well as transmission, are remarkably different in MERS-CoV-infected humans and dromedary camels, the viruses isolated from these two species are highly similar, if not indistinguishable [12,16]. This indicates that host factors play a significant role in MERS-CoV pathogenesis and transmission. However, the identity of these host factors and how they affect the pathogenesis and transmission of MERS-CoV are generally not well understood. Dipeptidyl peptidase-4 (DPP4)—the MERS-CoV receptor, sialic acids, proteases, and interferons are all examples of potentially critical host factors that have been shown to affect MERS-CoV infection in vitro [23,24,25,26]. This review highlights the role of some MERS-CoV-interacting host factors—especially DPP4—in MERS-CoV pathogenesis and transmission.

2. MERS-CoV-Interacting Host Factors

MERS-CoV infection of a target cell is initiated by the virus attachment to the cell surface [23,27]. MERS-CoV uses the N-terminal part of its spike (S)—the so called S1 protein (Figure 1A)—to bind to two host cell surface molecules, dipeptidyl peptidase-4 (DPP4) and α2,3-sialic acids [23,24]. DPP4 is the functional receptor of MERS-CoV; its absence renders cells resistant to this virus, while its transient expression in non-susceptible cells permits viral replication [23]. DPP4 is a serine exopeptidase, which is either expressed at the cell surface or shed in a soluble form. It has the capacity to cleave-off dipeptides from polypeptides with either l-proline or l-alanine at the penultimate position. Accordingly, DPP4 is capable of cutting various substrates, such as hormones, cytokines, chemokines, and neuropeptides, allowing it to be involved in multiple physiological functions as well as pathophysiological conditions [28]. This enzymatic activity is mediated by the α/β hydrolase domain of DPP4, while MERS-CoV infection is mediated by the binding of S1 protein to the β-propeller domain of this exopeptidase (Figure 1B) [28,29,30,31]. There are 11 critical residues within the β-propeller domain that directly interact with the S1 protein [29,30,31]. These residues are quite conserved in camelids, primates, and rabbits—species shown to be susceptible to MERS-CoV [17,31,32,33]. In contrast, ferrets, rats, and mice resist MERS-CoV infection due to differences in some critical DPP4 residues [31,34,35,36]. These data illustrate that DPP4 has the capacity to determine the host range of MERS-CoV.
Other MERS-CoV-interacting host factors besides DPP4 are less extensively studied and have mostly been investigated in vitro. Glycotopes of α2,3-sialic acids coupled with 5-N-acetylated neuraminic acid are recognized by the S1 protein of MERS-CoV during attachment [24]. In the absence of these glycotopes, MERS-CoV entry is reduced but not abolished, indicating their function as an attachment factor rather than a receptor [24]. Besides α2,3-sialic acids, CEACAM5 and GRP78 have also been suggested to be attachment factors for MERS-CoV, but their roles in vivo during MERS-CoV infection are not clear at this moment [37,38]. Post attachment, MERS-CoV uses the C-terminal part of its S protein—known as S2 (Figure 1A)—to interact with host proteases, such as furin, TMPRSS2, and cathepsins [39,40,41,42]. These proteases cleave the S protein and induce conformational changes, allowing fusion between viral and host cellular membranes, resulting in the release of viral RNA into the cell cytoplasm [27]. TMPRSS2 and DPP4 are held in one complex at the cell surface by a scaffolding protein, the tetraspanin CD9, leading to a rapid and efficient entry of MERS-CoV into the susceptible cells [43]. Once fusion with host cell membranes has occurred, MERS-CoV subsequently replicates its genetic material and produces viral proteins in the cell cytoplasm to generate new virus progeny. During this stage, MERS-CoV uses its nsp3–4 polyproteins to build its replication organelles as well as its accessory proteins such as the 4a and 4b proteins to inhibit host anti-viral defense mechanisms [44,45,46,47,48,49,50,51,52,53,54]. However, the capacity of MERS-CoV accessory proteins to impede several pathways of host immune response in the lungs may be limited. MERS-CoV inoculation of macaques and genetically modified mice generally results in limited clinical manifestations; thus, adapting this virus through serial passaging or defecting the type I interferon pathway may be needed to enhance viral replication and pathogenesis in these animals [32,55,56,57,58]. These observations, together with studies showing type I interferon capacity to inhibit MERS-CoV infection in vitro [25,59], highlight the importance of the innate immune response, especially type I interferon, as an inhibiting factor for MERS-CoV.

3. Host Factors in MERS-CoV Transmission

So far MERS-CoV has been isolated from dromedary camels and humans [1,60]. Both species are not only susceptible to MERS-CoV infection, but also capable of transmitting this virus [7,12,13,14,15,16,17,18,22]. However, current data indicate that virus spread is more efficient in dromedary camels than in humans [5,7,19,20,21,61]. This difference in transmissibility could be partially due to the different tropism of MERS-CoV in these two species. In dromedaries, MERS-CoV has been shown to replicate in the nasal epithelium upon experimental in vivo infection [17], while in humans, MERS-CoV mainly replicates in the lower respiratory tract, particularly in the bronchiolar and alveolar epithelia [23,62,63,64,65]. Higher viral RNA levels in the sputum and lavage samples of MERS-CoV patients compared to nasal and throat swabs are consistent with the tropism of MERS-CoV in humans [66,67,68]. This different MERS-CoV tropism in dromedary camels and humans is in line with the localization of DPP4 in the respiratory tract tissues of these two species. In humans, DPP4 is absent in the nasal epithelium but present in the lower respiratory tract epithelium, mainly in type II pneumocytes [69,70]. In contrast, DPP4 is expressed in the nasal epithelium of dromedary camels [69]. This difference in DPP4 localization between humans and dromedary camels therefore explains MERS-CoV tropism in these two species and highlights DPP4 as an essential determinant of MERS-CoV tropism.
DPP4 localization has also been investigated in many other MERS-CoV-susceptible species. In Gambian and Egyptian fruit bats, DPP4 is expressed in the respiratory tract and intestinal epithelium, suggesting that MERS-CoV can target both tissues [71]. In line with this finding, MERS-CoV inoculation via intranasal and intraperitoneal routes in the Jamaican fruit bat led to viral RNA shedding both in the respiratory tract and the intestinal tract [72]. In contrast to frugivorous bats, DPP4 is limitedly expressed in the respiratory tract epithelium of two insectivorous bats, i.e., common pipistrelle and common serotine bats, but abundant in their intestinal epithelium [71]. Accordingly, sequences of MERS-like-CoVs were mainly obtained from rectal swabs and fecal samples of insectivorous bats [73,74,75,76,77,78,79,80]. These findings not only support insectivorous bats as the origin host of MERS-CoV [73,74,75,76,77,78,79,80], but also indicate the importance of intestinal tropism and fecal–oral transmission of MERS-like-CoV in these insectivorous bats.
Besides bats, humans, and dromedary camels, other animal species have also been proposed as potential hosts of MERS-CoV. Remarkably, DPP4 of horses, llamas, alpacas, pigs, bovines, goats, sheep, and rabbits has been demonstrated to recognize the S protein of MERS-CoV [81,82]. In most of these species, there is a preferential upper respiratory tract expression of DPP4 observed. Rabbits express DPP4 in the upper and lower respiratory tract epithelium, and thus may allow MERS-CoV to replicate in both compartments [33,83]. Horses, llamas, and pigs mainly express DPP4 in the upper respiratory tract—particularly the nasal epithelium [84]. Upon intranasal MERS-CoV inoculation, llamas, alpacas, and pigs developed upper respiratory tract infection, while horses did not seroconvert and only shed infectious virus in a limited amount [84,85,86,87,88]. The reason why horses seem to be less permissive to MERS-CoV remains to be investigated, but a chronic co-infection in the guttural pouch, a common disease among horses, might be one of the explanations. This guttural pouch infection results in excessive mucus production that might hinder MERS-CoV from attaching and entering the nasal epithelium [84,89,90]. Sheep, on the other hand, did not seem to express significant levels of DPP4 in their respiratory tract, and thus did not seroconvert nor shed infectious virus upon experimental MERS-CoV inoculation [84,88]. Comparable to sheep, goats limitedly shed infectious virus upon experimental infection and did not transmit this virus to other naïve goats upon direct contact [88]. The results of experimental MERS-CoV infection in livestock animals are in line with data from epidemiological studies. MERS-CoV seropositive llamas and alpacas are present in the field, while horses, goats, and sheep are generally found to be seronegative [22,86,87,91,92,93,94,95,96,97,98].
Given the fact that experimental in vivo infection studies and DPP4 expression analysis in different animal species revealed that dromedary camels are not the only animals in which MERS-CoV has an upper respiratory tract tropism [17,18,83,84], it is then relevant to question whether other animals can potentially spread MERS-CoV as well. New World camelids, i.e., alpacas and llamas, are able to transmit the virus to respective naïve animals upon contact [86]. Pigs and rabbits, on the other hand, hardly transmit the virus—neither by contact nor airborne routes [83,99]. Most likely, this is caused by the fact that pigs and rabbits, unlike dromedary camels, shed low levels of infectious virus upon MERS-CoV inoculation (Figure 2). This difference indicates that other host factors besides DPP4 could cause interspecies variation in MERS-CoV infection. Indeed, several glycotopes of α2,3-sialic acids that function as attachment factors of MERS-CoV are present in the nasal epithelium of dromedary camels but absent in that of rabbits and pigs (Figure 3) [24,100]. The lack of these glycotopes in pigs and rabbits might limit the susceptibility and transmission of MERS-CoV in these animals. Although the role of these glycotopes in MERS-CoV transmission still requires further investigation, it remains plausible that an efficient transmission of this virus might require the presence of both DPP4 and MERS-CoV-recognized glycotopes of α2,3-sialic acids (Figure 3).
Besides entry and attachment receptors, MERS-CoV has been demonstrated to use both cell surface and lysosomal proteases to enter its target cells [39,40,43,101]. The preference of MERS-CoV to use certain host proteases is influenced by the type of target cell and the cleavage stage of their S protein prior to infection [40]. It has also been reported that the lysosomal proteases from bat cells support coronavirus spike-mediated virus entry more efficiently than their counterparts from human cells [39]. These observations suggest that host proteases from different host species may determine the species and tissue tropism of MERS-CoV.
Because MERS-CoV has been circulating in dromedary camels for decades before emerging in the human population [19,20,21,22], it is plausible that this virus inhibits the immune response of dromedary camels more efficiently than that of other species, including pigs and rabbits. The difference in immune response among MERS-CoV-susceptible species is therefore another factor that might yield interspecies variation in permissiveness to MERS-CoV. Characterizing the difference in host proteases and immune responses among MERS-CoV-susceptible species, as performed for DPP4 and MERS-CoV-recognized α2,3-sialic acid glycotopes (Figure 3), has not yet been investigated. These data, however, may further explain interspecies variation in MERS-CoV infection and transmission.

4. Host Factors in MERS-CoV Pathogenesis

MERS-CoV causes respiratory infection in humans ranging from asymptomatic to severe pneumonia [3,4]. However, it is currently unclear what causes this intraspecies variation. Epidemiology data indicate that individuals with certain risk factors are at higher risk of developing severe MERS-CoV infection [4,102]. This implies that some host factors may dictate the outcome of MERS-CoV infection, thus rendering intraspecies variation. Two of the risk factors, i.e., smoking and chronic obstructive pulmonary disease (COPD), have been shown to upregulate DPP4 expression in the lungs [70,102,103,104], suggesting DPP4 as a possible reason for intraspecies variation observed among MERS-CoV patients. In healthy human lungs, DPP4 is almost exclusively expressed in type II pneumocytes [69,70]. Type II pneumocytes are small cuboidal cells that can regenerate alveolar epithelium upon injury, and roughly cover 2% of the alveolar surface area. Meanwhile, around 95% of the surface area of the alveolus is occupied by type I pneumocytes that are morphologically flat and responsible for gas exchange [105,106]. In the lungs of smokers and COPD patients, unlike in healthy human lungs, DPP4 is prominently expressed in both type I and II pneumocytes, indicating upregulated expression on type I pneumocytes [104]. Autopsy reports from fatal MERS-CoV patients showed that both type I and II pneumocytes expressed DPP4 and became infected by MERS-CoV, proposing a role of DPP4-expressing type I pneumocytes in MERS-CoV pathogenesis [64,107]. Damage to type I cells in the lung alveoli during viral infection may lead to diffuse alveolar damage [108]. In line with observations made in human MERS cases, common marmosets that express DPP4 in both type I and II pneumocytes have been reported to produce more infectious virus upon experimental MERS-CoV infection, compared to rhesus and cynomolgus macaques that merely expressed DPP4 in type II pneumocytes [58,109,110,111,112]. Accordingly, these common marmosets developed moderate-to-severe infection, while macaques generally developed mild transient pneumonia [32,58,109,110,111,112]. Similarly, in genetically modified mice that displayed MERS-CoV tropism for type II pneumocytes, only mild clinical manifestations were observed upon MERS-CoV infection [56,113]. Adapting MERS-CoV through serial passaging or upregulating DPP4 expression throughout the airway epithelium in mice, however, will induce severe clinical disease [55,56]. These data altogether support the role of DPP4-expressing type I pneumocytes in the pathogenesis of severe MERS-CoV infection.
The differential expression of host factors that limits the infection should also be taken into account. DPP4 in soluble form has been demonstrated to protect against MERS-CoV infection in vitro and in a mouse model [23,114]; however, its presence in the lungs and role in MERS-CoV pathogenesis remain to be investigated. The host immune response also has the capacity to inhibit MERS-CoV infection. MERS-CoV has been shown to replicate to higher levels in immunocompromised rhesus macaques [115], consistent with the observation that immunocompromised individuals have difficulties clearing MERS-CoV upon infection [68,107,116]. The survivors of MERS-CoV infection have been shown to develop virus-specific CD4+ and CD8+ T cell responses, implying the role of T cells in virus clearance [117]. However, the depletion of T cells in mice can either lead to failure in MERS-CoV clearance or improvement in clinical outcome, depending on the type of mouse model used [57,118]. Therefore, the role of adaptive immune response in MERS-CoV pathogenesis is currently unclear. On the other hand, one of the main components of the host innate immune response, type I interferon, inhibits MERS-CoV replication in susceptible cells, partly by inhibiting double membrane vesicles (DMV) formation [25,57,59,119,120]. The absence of type I interferon signaling in mice also resulted in more severe clinical manifestations and histopathological lesions upon MERS-CoV infection [57]. Advance age, which can cause delayed type I interferon response upon viral infection, is a well-known risk factor for fatal MERS-CoV infection [4,102,121,122,123]. Collectively, these data highlight the role of host innate immune response as a potent inhibitor for MERS-CoV infection.
It is indubitable that severe MERS-CoV infection is not solely driven by the pathogen. Additional underlying conditions increase MERS-CoV replication and induce severe-to-fatal clinical manifestations [4,11,103,124,125]. It is plausible that more than one underlying condition is needed to yield a fatal outcome. DPP4 upregulation in type I pneumocytes and insufficient type I interferon response might be crucial determinants for severe MERS-CoV infection (Figure 4). Further investigation of the host determinants of MERS-CoV pathogenesis may offer insights for developing novel therapeutic measures.

5. Concluding Remarks and Future Perspectives

Although MERS-CoV has been reported to undergo some genotypic changes since it emerged in the human population [12,126,127,128,129], this has not resulted in distinct phenotypic changes so far [63,126]. Therefore, host factors remain the most significant determinant in explaining inter- and intraspecies variations observed in MERS-CoV pathogenesis and transmission. DPP4 and MERS-CoV-recognized α2,3-sialic acids might partially explain these variations, since their localization has been demonstrated to be variable between MERS-CoV-susceptible species [69,71,84,100]. DPP4 expression in human lungs has also been shown to vary due to certain comorbidities [70,96,104]. Nevertheless, it is undoubtable that the inter- and intraspecies variation in MERS-CoV pathogenesis and transmission is a complex phenomenon influenced by more than one host factor. Current data suggest proteases and interferons as other critical host factors, but how they instigate inter- and intraspecies variations, as well as their role in MERS-CoV pathogenesis and transmission, still remain to be further elucidated. Characterization of the host determinants of MERS-CoV pathogenesis and transmission could potentially offer insight into this virus epidemiology and guide novel therapeutic development. It may also help to identify the most vulnerable individuals to protect against MERS-CoV infection—for example, by using vaccination.

Author Contributions

All authors contributed to the writing of the manuscript and carefully evaluated the manuscript before submission.

Funding

Our work is supported by Zoonotic Anticipation and Preparedness Initiative (Innovative Medicines Initiative grant 115760), with assistance and financial support from Innovative Medicines Initiative and the European Commission, and contributions from European Federation of Pharmaceutical Industries and Associations partners. Syriam Sooksawasdi Na Ayudhya received the Royal Thai Government Scholarship supported by the Ministry of Science and Technology of Thailand to perform her doctoral study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zaki, A.M.; van Boheemen, S.; Bestebroer, T.M.; Osterhaus, A.D.; Fouchier, R.A. Isolation of a novel coronavirus from a man with pneumonia in Saudi Arabia. N. Engl. J. Med. 2012, 367, 1814–1820. [Google Scholar] [CrossRef]
  2. World Health Organization MERS Situation Update, March 2018. Available online: http://www.emro.who.int/pandemic-epidemic-diseases/mers-cov/mers-situation-update-march-2018.html (accessed on 1 March 2019).
  3. Widagdo, W.; Okba, N.M.A.; Stalin Raj, V.; Haagmans, B.L. MERS-coronavirus: From discovery to intervention. One Health 2017, 3, 11–16. [Google Scholar] [CrossRef]
  4. The WHO Mers-CoV Research Group. State of Knowledge and Data Gaps of Middle East Respiratory Syndrome Coronavirus (MERS-CoV) in Humans. PLoS Curr. 2013, 5. [Google Scholar] [CrossRef]
  5. Moon, S.Y.; Son, J.S. Infectivity of an Asymptomatic Patient With Middle East Respiratory Syndrome Coronavirus Infection. Clin. Infect. Dis. 2017, 64, 1457–1458. [Google Scholar] [CrossRef] [PubMed]
  6. Memish, Z.A.; Al-Tawfiq, J.A.; Makhdoom, H.Q.; Al-Rabeeah, A.A.; Assiri, A.; Alhakeem, R.F.; AlRabiah, F.A.; Al Hajjar, S.; Albarrak, A.; Flemban, H.; et al. Screening for Middle East respiratory syndrome coronavirus infection in hospital patients and their healthcare worker and family contacts: A prospective descriptive study. Clin. Microbiol. Infect. 2014, 20, 469–474. [Google Scholar] [CrossRef]
  7. Drosten, C.; Meyer, B.; Muller, M.A.; Corman, V.M.; Al-Masri, M.; Hossain, R.; Madani, H.; Sieberg, A.; Bosch, B.J.; Lattwein, E.; et al. Transmission of MERS-coronavirus in household contacts. N. Engl. J. Med. 2014, 371, 828–835. [Google Scholar] [CrossRef]
  8. Kim, S.W.; Park, J.W.; Jung, H.D.; Yang, J.S.; Park, Y.S.; Lee, C.; Kim, K.M.; Lee, K.J.; Kwon, D.; Hur, Y.J.; et al. Risk factors for transmission of Middle East respiratory syndrome coronavirus infection during the 2015 outbreak in South Korea. Clin. Infect. Dis. 2017, 64, 551–557. [Google Scholar] [CrossRef] [PubMed]
  9. Normile, D. South Korea finally MERS-free. Science 2015. [Google Scholar] [CrossRef]
  10. World Health Organization Middle East Respiratory Syndrome Coronavirus (MERS-CoV), Republic of Korea—Disease Outbreak News 25 October 2015. Available online: http://www.who.int/csr/don/25-october-2015-mers-korea/en/ (accessed on 1 March 2019).
  11. World Health Organization MERS-CoV Global Summary and Assessment of Risk. July 2017. Available online: http://www.who.int/emergencies/mers-cov/risk-assessment-july-2017.pdf?ua=1 (accessed on 1 March 2019).
  12. Dudas, G.; Carvalho, L.M.; Rambaut, A.; Bedford, T. MERS-CoV spillover at the camel-human interface. Elife 2018, 7, eLife.31257. [Google Scholar]
  13. Haagmans, B.L.; Al Dhahiry, S.H.; Reusken, C.B.; Raj, V.S.; Galiano, M.; Myers, R.; Godeke, G.J.; Jonges, M.; Farag, E.; Diab, A.; et al. Middle East respiratory syndrome coronavirus in dromedary camels: An outbreak investigation. Lancet Infect. Dis. 2014, 14, 140–145. [Google Scholar] [CrossRef]
  14. Memish, Z.A.; Cotten, M.; Meyer, B.; Watson, S.J.; Alsahafi, A.J.; Al Rabeeah, A.A.; Corman, V.M.; Sieberg, A.; Makhdoom, H.Q.; Assiri, A.; et al. Human infection with MERS coronavirus after exposure to infected camels, Saudi Arabia, 2013. Emerg. Infect. Dis. 2014, 20, 1012–1015. [Google Scholar] [CrossRef] [PubMed]
  15. Paden, C.R.; Yusof, M.; Al Hammadi, Z.M.; Queen, K.; Tao, Y.; Eltahir, Y.M.; Elsayed, E.A.; Marzoug, B.A.; Bensalah, O.K.A.; Khalafalla, A.I.; et al. Zoonotic origin and transmission of Middle East respiratory syndrome coronavirus in the UAE. Zoonoses. Public Health 2018, 65, 322–333. [Google Scholar] [CrossRef]
  16. Briese, T.; Mishra, N.; Jain, K.; Zalmout, I.S.; Jabado, O.J.; Karesh, W.B.; Daszak, P.; Mohammed, O.B.; Alagaili, A.N.; Lipkin, W.I. Middle East respiratory syndrome coronavirus quasispecies that include homologues of human isolates revealed through whole-genome analysis and virus cultured from dromedary camels in Saudi Arabia. MBio 2014, 5, e01146-14. [Google Scholar]
  17. Haagmans, B.L.; van den Brand, J.M.; Raj, V.S.; Volz, A.; Wohlsein, P.; Smits, S.L.; Schipper, D.; Bestebroer, T.M.; Okba, N.; Fux, R.; et al. An orthopoxvirus-based vaccine reduces virus excretion after MERS-CoV infection in dromedary camels. Science 2016, 351, 77–81. [Google Scholar] [CrossRef]
  18. Adney, D.R.; van Doremalen, N.; Brown, V.R.; Bushmaker, T.; Scott, D.; de Wit, E.; Bowen, R.A.; Munster, V.J. Replication and shedding of MERS-CoV in upper respiratory tract of inoculated dromedary camels. Emerg. Infect. Dis. 2014, 20, 1999–2005. [Google Scholar]
  19. Meyer, B.; Muller, M.A.; Corman, V.M.; Reusken, C.B.; Ritz, D.; Godeke, G.J.; Lattwein, E.; Kallies, S.; Siemens, A.; van Beek, J.; et al. Antibodies against MERS coronavirus in dromedary camels, United Arab Emirates, 2003 and 2013. Emerg. Infect. Dis. 2014, 20, 552–559. [Google Scholar] [CrossRef] [PubMed]
  20. Muller, M.A.; Corman, V.M.; Jores, J.; Meyer, B.; Younan, M.; Liljander, A.; Bosch, B.J.; Lattwein, E.; Hilali, M.; Musa, B.E.; et al. MERS coronavirus neutralizing antibodies in camels, Eastern Africa, 1983–1997. Emerg. Infect. Dis. 2014, 20, 2093–2095. [Google Scholar]
  21. Reusken, C.B.; Messadi, L.; Feyisa, A.; Ularamu, H.; Godeke, G.J.; Danmarwa, A.; Dawo, F.; Jemli, M.; Melaku, S.; Shamaki, D.; et al. Geographic distribution of MERS coronavirus among dromedary camels, Africa. Emerg. Infect. Dis. 2014, 20, 1370–1374. [Google Scholar]
  22. Alagaili, A.N.; Briese, T.; Mishra, N.; Kapoor, V.; Sameroff, S.C.; Burbelo, P.D.; de Wit, E.; Munster, V.J.; Hensley, L.E.; Zalmout, I.S.; et al. Middle East respiratory syndrome coronavirus infection in dromedary camels in Saudi Arabia. MBio 2014, 5, e00884-14. [Google Scholar] [CrossRef]
  23. Raj, V.S.; Mou, H.; Smits, S.L.; Dekkers, D.H.; Muller, M.A.; Dijkman, R.; Muth, D.; Demmers, J.A.; Zaki, A.; Fouchier, R.A.; et al. Dipeptidyl peptidase 4 is a functional receptor for the emerging human coronavirus-EMC. Nature 2013, 495, 251–254. [Google Scholar]
  24. Li, W.; Hulswit, R.J.G.; Widjaja, I.; Raj, V.S.; McBride, R.; Peng, W.; Widagdo, W.; Tortorici, M.A.; van Dieren, B.; Lang, Y.; et al. Identification of sialic acid-binding function for the Middle East respiratory syndrome coronavirus spike glycoprotein. Proc. Natl. Acad. Sci. USA 2017, 114, E8508–E8517. [Google Scholar] [CrossRef]
  25. De Wilde, A.H.; Raj, V.S.; Oudshoorn, D.; Bestebroer, T.M.; van Nieuwkoop, S.; Limpens, R.W.; Posthuma, C.C.; van der Meer, Y.; Barcena, M.; Haagmans, B.L.; et al. MERS-coronavirus replication induces severe in vitro cytopathology and is strongly inhibited by cyclosporin A or interferon-alpha treatment. J. Gen. Virol. 2013, 94 Pt 8, 1749–1760. [Google Scholar] [CrossRef]
  26. Iwata-Yoshikawa, N.; Okamura, T.; Shimizu, Y.; Hasegawa, H.; Takeda, M.; Nagata, N. TMPRSS2 contributes to virus spread and immunopathology in the airways of murine models after coronavirus infection. J. Virol. 2019. [Google Scholar] [CrossRef]
  27. Fehr, A.R.; Perlman, S. Coronaviruses: An overview of their replication and pathogenesis. Methods. Mol. Biol. 2015, 1282, 1–23. [Google Scholar] [PubMed]
  28. Boonacker, E.; Van Noorden, C.J. The multifunctional or moonlighting protein CD26/DPPIV. Eur. J. Cell. Biol. 2003, 82, 53–73. [Google Scholar] [CrossRef] [PubMed]
  29. Wang, N.; Shi, X.; Jiang, L.; Zhang, S.; Wang, D.; Tong, P.; Guo, D.; Fu, L.; Cui, Y.; Liu, X.; et al. Structure of MERS-CoV spike receptor-binding domain complexed with human receptor DPP4. Cell Res. 2013, 23, 986–993. [Google Scholar] [CrossRef]
  30. Lu, G.; Hu, Y.; Wang, Q.; Qi, J.; Gao, F.; Li, Y.; Zhang, Y.; Zhang, W.; Yuan, Y.; Bao, J.; et al. Molecular basis of binding between novel human coronavirus MERS-CoV and its receptor CD26. Nature 2013, 500, 227–231. [Google Scholar] [CrossRef]
  31. Bosch, B.J.; Raj, V.S.; Haagmans, B.L. Spiking the MERS-coronavirus receptor. Cell Res. 2013, 23, 1069–1070. [Google Scholar] [CrossRef] [PubMed]
  32. De Wit, E.; Rasmussen, A.L.; Falzarano, D.; Bushmaker, T.; Feldmann, F.; Brining, D.L.; Fischer, E.R.; Martellaro, C.; Okumura, A.; Chang, J.; et al. Middle East respiratory syndrome coronavirus (MERS-CoV) causes transient lower respiratory tract infection in rhesus macaques. Proc. Natl. Acad. Sci. USA 2013, 110, 16598–16603. [Google Scholar] [CrossRef] [PubMed]
  33. Haagmans, B.L.; van den Brand, J.M.; Provacia, L.B.; Raj, V.S.; Stittelaar, K.J.; Getu, S.; de Waal, L.; Bestebroer, T.M.; van Amerongen, G.; Verjans, G.M.; et al. Asymptomatic Middle East respiratory syndrome coronavirus infection in rabbits. J. Virol. 2015, 89, 6131–6135. [Google Scholar] [CrossRef]
  34. Raj, V.S.; Smits, S.L.; Provacia, L.B.; van den Brand, J.M.; Wiersma, L.; Ouwendijk, W.J.; Bestebroer, T.M.; Spronken, M.I.; van Amerongen, G.; Rottier, P.J.; 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. [Google Scholar] [CrossRef] [PubMed]
  35. Coleman, C.M.; Matthews, K.L.; Goicochea, L.; Frieman, M.B. Wild-type and innate immune-deficient mice are not susceptible to the Middle East respiratory syndrome coronavirus. J. Gen. Virol. 2014, 95 Pt 2, 408–412. [Google Scholar] [CrossRef]
  36. Iwata-Yoshikawa, N.; Fukushi, S.; Fukuma, A.; Suzuki, T.; Takeda, M.; Tashiro, M.; Hasegawa, H.; Nagata, N. Non Susceptibility of Neonatal and Adult Rats against the Middle East Respiratory Syndrome Coronavirus. Jpn. J. Infect. Dis. 2016, 69, 510–516. [Google Scholar] [CrossRef] [PubMed]
  37. Chu, H.; Chan, C.M.; Zhang, X.; Wang, Y.; Yuan, S.; Zhou, J.; Au-Yeung, R.K.; Sze, K.H.; Yang, D.; Shuai, H.; et al. Middle East respiratory syndrome coronavirus and bat coronavirus HKU9 both can utilize GRP78 for attachment onto host cells. J. Biol. Chem. 2018, 293, 11709–11726. [Google Scholar] [CrossRef] [PubMed]
  38. Chan, C.M.; Chu, H.; Wang, Y.; Wong, B.H.; Zhao, X.; Zhou, J.; Yang, D.; Leung, S.P.; Chan, J.F.; Yeung, M.L.; et al. Carcinoembryonic Antigen-Related Cell Adhesion Molecule 5 Is an Important Surface Attachment Factor That Facilitates Entry of Middle East Respiratory Syndrome Coronavirus. J. Virol. 2016, 90, 9114–9127. [Google Scholar] [CrossRef]
  39. Zheng, Y.; Shang, J.; Yang, Y.; Liu, C.; Wan, Y.; Geng, Q.; Wang, M.; Baric, R.; Li, F. Lysosomal proteases are a determinant of coronavirus tropism. J. Virol. 2018, 92, e01504. [Google Scholar] [CrossRef]
  40. Park, J.E.; Li, K.; Barlan, A.; Fehr, A.R.; Perlman, S.; McCray, P.B., Jr.; Gallagher, T. Proteolytic processing of Middle East respiratory syndrome coronavirus spikes expands virus tropism. Proc. Natl. Acad. Sci. USA 2016, 113, 12262–12267. [Google Scholar] [CrossRef]
  41. 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. [Google Scholar] [CrossRef] [PubMed]
  42. Millet, J.K.; Whittaker, G.R. Host cell entry of Middle East respiratory syndrome coronavirus after two-step, furin-mediated activation of the spike protein. Proc. Natl. Acad. Sci. USA 2014, 111, 15214–15219. [Google Scholar] [CrossRef] [PubMed]
  43. Earnest, J.T.; Hantak, M.P.; Li, K.; McCray, P.B., Jr.; Perlman, S.; Gallagher, T. The tetraspanin CD9 facilitates MERS-coronavirus entry by scaffolding host cell receptors and proteases. PLoS Pathog. 2017, 13, e1006546. [Google Scholar] [CrossRef] [PubMed]
  44. Muller, C.; Hardt, M.; Schwudke, D.; Neuman, B.W.; Pleschka, S.; Ziebuhr, J. Inhibition of Cytosolic Phospholipase A2alpha Impairs an Early Step of Coronavirus Replication in Cell Culture. J. Virol. 2018, 92, e01463-17. [Google Scholar] [PubMed]
  45. Belov, G.A.; van Kuppeveld, F.J. (+)RNA viruses rewire cellular pathways to build replication organelles. Curr. Opin. Virol. 2012, 2, 740–747. [Google Scholar] [CrossRef]
  46. Belov, G.A.; Nair, V.; Hansen, B.T.; Hoyt, F.H.; Fischer, E.R.; Ehrenfeld, E. Complex dynamic development of poliovirus membranous replication complexes. J. Virol. 2012, 86, 302–312. [Google Scholar] [CrossRef]
  47. Rabouw, H.H.; Langereis, M.A.; Knaap, R.C.; Dalebout, T.J.; Canton, J.; Sola, I.; Enjuanes, L.; Bredenbeek, P.J.; Kikkert, M.; de Groot, R.J.; et al. Middle East Respiratory Coronavirus Accessory Protein 4a Inhibits PKR-Mediated Antiviral Stress Responses. PLoS Pathog. 2016, 12, e1005982. [Google Scholar] [CrossRef]
  48. Canton, J.; Fehr, A.R.; Fernandez-Delgado, R.; Gutierrez-Alvarez, F.J.; Sanchez-Aparicio, M.T.; Garcia-Sastre, A.; Perlman, S.; Enjuanes, L.; Sola, I. MERS-CoV 4b protein interferes with the NF-kappaB-dependent innate immune response during infection. PLoS Pathog. 2018, 14, e1006838. [Google Scholar] [CrossRef]
  49. Siu, K.L.; Yeung, M.L.; Kok, K.H.; Yuen, K.S.; Kew, C.; Lui, P.Y.; Chan, C.P.; Tse, H.; Woo, P.C.; Yuen, K.Y.; 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. [Google Scholar] [CrossRef]
  50. Nakagawa, K.; Narayanan, K.; Wada, M.; Popov, V.L.; Cajimat, M.; Baric, R.S.; Makino, S. The endonucleolytic RNA cleavage function of nsp1 of Middle East respiratory syndrome coronavirus promotes the production of infectious virus particles in specific human cell lines. J. Virol. 2018, 92, e01157-18. [Google Scholar] [CrossRef]
  51. Knoops, K.; Kikkert, M.; Worm, S.H.; Zevenhoven-Dobbe, J.C.; van der Meer, Y.; Koster, A.J.; Mommaas, A.M.; Snijder, E.J. SARS-coronavirus replication is supported by a reticulovesicular network of modified endoplasmic reticulum. PLoS Biol. 2008, 6, e226. [Google Scholar] [CrossRef] [PubMed]
  52. Thornbrough, J.M.; Jha, B.K.; Yount, B.; Goldstein, S.A.; Li, Y.; Elliott, R.; Sims, A.C.; Baric, R.S.; Silverman, R.H.; Weiss, S.R. Middle East Respiratory Syndrome Coronavirus NS4b Protein Inhibits Host RNase L Activation. MBio 2016, 7, e00258-16. [Google Scholar] [CrossRef] [PubMed]
  53. Menachery, V.D.; Mitchell, H.D.; Cockrell, A.S.; Gralinski, L.E.; Yount, B.L., Jr.; Graham, R.L.; McAnarney, E.T.; Douglas, M.G.; Scobey, T.; Beall, A.; et al. MERS-CoV Accessory ORFs Play Key Role for Infection and Pathogenesis. MBio 2017, 8, e00665-17. [Google Scholar] [CrossRef] [PubMed]
  54. Oudshoorn, D.; Rijs, K.; Limpens, R.; Groen, K.; Koster, A.J.; Snijder, E.J.; Kikkert, M.; Barcena, M. Expression and Cleavage of Middle East Respiratory Syndrome Coronavirus nsp3-4 Polyprotein Induce the Formation of Double-Membrane Vesicles That Mimic Those Associated with Coronaviral RNA Replication. MBio 2017, 8, e01658-17. [Google Scholar] [CrossRef]
  55. Li, K.; Wohlford-Lenane, C.L.; Channappanavar, R.; Park, J.E.; Earnest, J.T.; Bair, T.B.; Bates, A.M.; Brogden, K.A.; Flaherty, H.A.; Gallagher, T.; et al. Mouse-adapted MERS coronavirus causes lethal lung disease in human DPP4 knockin mice. Proc. Natl. Acad. Sci. USA 2017, 114, E3119–E3128. [Google Scholar] [CrossRef]
  56. Cockrell, A.S.; Yount, B.L.; Scobey, T.; Jensen, K.; Douglas, M.; Beall, A.; Tang, X.C.; Marasco, W.A.; Heise, M.T.; Baric, R.S. A mouse model for MERS coronavirus-induced acute respiratory distress syndrome. Nat. Microbiol. 2016, 2, 16226. [Google Scholar] [CrossRef]
  57. Zhao, J.; Li, K.; Wohlford-Lenane, C.; Agnihothram, S.S.; Fett, C.; Zhao, J.; Gale, M.J., Jr.; Baric, R.S.; Enjuanes, L.; Gallagher, T.; et al. Rapid generation of a mouse model for Middle East respiratory syndrome. Proc. Natl. Acad. Sci. USA 2014, 111, 4970–4975. [Google Scholar] [CrossRef]
  58. Yao, Y.; Bao, L.; Deng, W.; Xu, L.; Li, F.; Lv, Q.; Yu, P.; Chen, T.; Xu, Y.; Zhu, H.; et al. An animal model of MERS produced by infection of rhesus macaques with MERS coronavirus. J. Infect. Dis. 2014, 209, 236–242. [Google Scholar] [CrossRef] [PubMed]
  59. Falzarano, D.; de Wit, E.; Martellaro, C.; Callison, J.; Munster, V.J.; Feldmann, H. Inhibition of novel beta coronavirus replication by a combination of interferon-alpha2b and ribavirin. Sci. Rep. 2013, 3, 1686. [Google Scholar] [CrossRef]
  60. Raj, V.S.; Farag, E.A.; Reusken, C.B.; Lamers, M.M.; Pas, S.D.; Voermans, J.; Smits, S.L.; Osterhaus, A.D.; Al-Mawlawi, N.; Al-Romaihi, H.E.; et al. Isolation of MERS coronavirus from a dromedary camel, Qatar, 2014. Emerg. Infect. Dis. 2014, 20, 1339–1342. [Google Scholar] [CrossRef]
  61. Cho, S.Y.; Kang, J.M.; Ha, Y.E.; Park, G.E.; Lee, J.Y.; Ko, J.H.; Lee, J.Y.; Kim, J.M.; Kang, C.I.; Jo, I.J.; et al. MERS-CoV outbreak following a single patient exposure in an emergency room in South Korea: An epidemiological outbreak study. Lancet 2016, 388, 994–1001. [Google Scholar] [CrossRef]
  62. Chan, R.W.; Chan, M.C.; Agnihothram, S.; Chan, L.L.; Kuok, D.I.; Fong, J.H.; Guan, Y.; Poon, L.L.; Baric, R.S.; Nicholls, J.M.; 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. [Google Scholar] [CrossRef] [PubMed]
  63. Chan, R.W.; Hemida, M.G.; Kayali, G.; Chu, D.K.; Poon, L.L.; Alnaeem, A.; Ali, M.A.; Tao, K.P.; Ng, H.Y.; Chan, M.C.; 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. [Google Scholar] [CrossRef]
  64. Ng, D.L.; Al Hosani, F.; Keating, M.K.; Gerber, S.I.; Jones, T.L.; Metcalfe, M.G.; Tong, S.; Tao, Y.; Alami, N.N.; Haynes, L.M.; et al. Clinicopathologic, Immunohistochemical, and Ultrastructural Findings of a Fatal Case of Middle East Respiratory Syndrome Coronavirus Infection in United Arab Emirates, April 2014. Am. J. Pathol. 2016, 186, 652–658. [Google Scholar] [CrossRef]
  65. Hocke, A.C.; Becher, A.; Knepper, J.; Peter, A.; Holland, G.; Tonnies, M.; Bauer, T.T.; Schneider, P.; Neudecker, J.; Muth, D.; et al. Emerging human middle East respiratory syndrome coronavirus causes widespread infection and alveolar damage in human lungs. Am. J. Respir. Crit. Care Med. 2013, 188, 882–886. [Google Scholar] [CrossRef]
  66. Bermingham, A.; Chand, M.A.; Brown, C.S.; Aarons, E.; Tong, C.; Langrish, C.; Hoschler, K.; Brown, K.; Galiano, M.; Myers, R.; et al. Severe respiratory illness caused by a novel coronavirus, in a patient transferred to the United Kingdom from the Middle East, September 2012. Eurosurveillance 2012, 17, 20290. [Google Scholar]
  67. Corman, V.M.; Albarrak, A.M.; Omrani, A.S.; Albarrak, M.M.; Farah, M.E.; Almasri, M.; Muth, D.; Sieberg, A.; Meyer, B.; Assiri, A.M.; et al. Viral Shedding and Antibody Response in 37 Patients With Middle East Respiratory Syndrome Coronavirus Infection. Clin. Infect. Dis. 2016, 62, 477–483. [Google Scholar] [CrossRef]
  68. Drosten, C.; Seilmaier, M.; Corman, V.M.; Hartmann, W.; Scheible, G.; Sack, S.; Guggemos, W.; Kallies, R.; Muth, D.; Junglen, S.; et al. Clinical features and virological analysis of a case of Middle East respiratory syndrome coronavirus infection. Lancet Infect. Dis. 2013, 13, 745–751. [Google Scholar] [CrossRef]
  69. Widagdo, W.; Raj, V.S.; Schipper, D.; Kolijn, K.; van Leenders, G.J.; Bosch, B.J.; Bensaid, A.; Segales, J.; Baumgartner, W.; Osterhaus, A.D.; et al. Differential Expression of the Middle East Respiratory Syndrome Coronavirus Receptor in the Upper Respiratory Tracts of Humans and Dromedary Camels. J. Virol. 2016, 90, 4838–4842. [Google Scholar] [CrossRef]
  70. Meyerholz, D.K.; Lambertz, A.M.; McCray, P.B., Jr. Dipeptidyl Peptidase 4 Distribution in the Human Respiratory Tract: Implications for the Middle East Respiratory Syndrome. Am. J. Pathol. 2016, 186, 78–86. [Google Scholar] [CrossRef] [PubMed]
  71. Widagdo, W.; Begeman, L.; Schipper, D.; Run, P.R.V.; Cunningham, A.A.; Kley, N.; Reusken, C.B.; Haagmans, B.L.; van den Brand, J.M.A. Tissue Distribution of the MERS-Coronavirus Receptor in Bats. Sci. Rep. 2017, 7, 1193. [Google Scholar] [CrossRef]
  72. Munster, V.J.; Adney, D.R.; van Doremalen, N.; Brown, V.R.; Miazgowicz, K.L.; Milne-Price, S.; Bushmaker, T.; Rosenke, R.; Scott, D.; Hawkinson, A.; et al. Replication and shedding of MERS-CoV in Jamaican fruit bats (Artibeus jamaicensis). Sci. Rep. 2016, 6, 21878. [Google Scholar] [CrossRef] [PubMed]
  73. Annan, A.; Baldwin, H.J.; Corman, V.M.; Klose, S.M.; Owusu, M.; Nkrumah, E.E.; Badu, E.K.; Anti, P.; Agbenyega, O.; Meyer, B.; et al. Human betacoronavirus 2c EMC/2012-related viruses in bats, Ghana and Europe. Emerg. Infect. Dis. 2013, 19, 456–459. [Google Scholar] [CrossRef]
  74. Yang, L.; Wu, Z.; Ren, X.; Yang, F.; Zhang, J.; He, G.; Dong, J.; Sun, L.; Zhu, Y.; Zhang, S.; et al. MERS-related betacoronavirus in Vespertilio superans bats, China. Emerg. Infect. Dis. 2014, 20, 1260–1262. [Google Scholar] [CrossRef]
  75. Wacharapluesadee, S.; Sintunawa, C.; Kaewpom, T.; Khongnomnan, K.; Olival, K.J.; Epstein, J.H.; Rodpan, A.; Sangsri, P.; Intarut, N.; Chindamporn, A.; et al. Group C betacoronavirus in bat guano fertilizer, Thailand. Emerg. Infect. Dis. 2013, 19, 1349–1351. [Google Scholar] [CrossRef]
  76. Kim, H.K.; Yoon, S.W.; Kim, D.J.; Koo, B.S.; Noh, J.Y.; Kim, J.H.; Choi, Y.G.; Na, W.; Chang, K.T.; Song, D.; et al. Detection of Severe Acute Respiratory Syndrome-Like, Middle East Respiratory Syndrome-Like Bat Coronaviruses and Group H Rotavirus in Faeces of Korean Bats. Transbound. Emerg. Dis. 2016, 63, 365–372. [Google Scholar] [CrossRef] [PubMed]
  77. Corman, V.M.; Ithete, N.L.; Richards, L.R.; Schoeman, M.C.; Preiser, W.; Drosten, C.; Drexler, J.F. 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. [Google Scholar] [CrossRef] [PubMed]
  78. Ithete, N.L.; Stoffberg, S.; Corman, V.M.; Cottontail, V.M.; Richards, L.R.; Schoeman, M.C.; Drosten, C.; Drexler, J.F.; Preiser, W. Close relative of human Middle East respiratory syndrome coronavirus in bat, South Africa. Emerg. Infect. Dis. 2013, 19, 1697–1699. [Google Scholar] [CrossRef] [PubMed]
  79. Anthony, S.J.; Gilardi, K.; Menachery, V.D.; Goldstein, T.; Ssebide, B.; Mbabazi, R.; Navarrete-Macias, I.; Liang, E.; Wells, H.; Hicks, A.; et al. Further Evidence for Bats as the Evolutionary Source of Middle East Respiratory Syndrome Coronavirus. MBio 2017, 8, e00373-17. [Google Scholar] [CrossRef]
  80. Luo, C.M.; Wang, N.; Yang, X.L.; Liu, H.Z.; Zhang, W.; Li, B.; Hu, B.; Peng, C.; Geng, Q.B.; Zhu, G.J.; et al. Discovery of Novel Bat Coronaviruses in South China That Use the Same Receptor as Middle East Respiratory Syndrome Coronavirus. J. Virol. 2018, 92, e00116-18. [Google Scholar] [CrossRef]
  81. Barlan, A.; Zhao, J.; Sarkar, M.K.; Li, K.; McCray, P.B., Jr.; Perlman, S.; Gallagher, T. Receptor variation and susceptibility to Middle East respiratory syndrome coronavirus infection. J. Virol. 2014, 88, 4953–4961. [Google Scholar] [CrossRef] [PubMed]
  82. Van Doremalen, N.; Miazgowicz, K.L.; Milne-Price, S.; Bushmaker, T.; Robertson, S.; Scott, D.; Kinne, J.; McLellan, J.S.; Zhu, J.; Munster, V.J. Host species restriction of Middle East respiratory syndrome coronavirus through its receptor, dipeptidyl peptidase 4. J. Virol. 2014, 88, 9220–9232. [Google Scholar] [CrossRef] [PubMed]
  83. Widagdo, W.; Okba, N.M.A.; Richard, M.; de Muelder, D.; Bestebroer, T.M.; Lexmond, P.; J.M.A. v. d., B.; Haagmans, B.L.; Herfst, S. Middle East respiratory syndrome coronavirus transmission in rabbits. Unpublished work. 2019. [Google Scholar]
  84. Vergara-Alert, J.; van den Brand, J.M.; Widagdo, W.; Munoz, M.t.; Raj, S.; Schipper, D.; Solanes, D.; Cordon, I.; Bensaid, A.; Haagmans, B.L.; et al. Livestock Susceptibility to Infection with Middle East Respiratory Syndrome Coronavirus. Emerg. Infect. Dis. 2017, 23, 232–240. [Google Scholar] [CrossRef]
  85. De Wit, E.; Feldmann, F.; Horne, E.; Martellaro, C.; Haddock, E.; Bushmaker, T.; Rosenke, K.; Okumura, A.; Rosenke, R.; Saturday, G.; et al. Domestic Pig Unlikely Reservoir for MERS-CoV. Emerg. Infect. Dis. 2017, 23, 985–988. [Google Scholar] [CrossRef]
  86. Adney, D.R.; Bielefeldt-Ohmann, H.; Hartwig, A.E.; Bowen, R.A. Infection, Replication, and Transmission of Middle East Respiratory Syndrome Coronavirus in Alpacas. Emerg. Infect. Dis. 2016, 22, 1031–1037. [Google Scholar] [CrossRef]
  87. Crameri, G.; Durr, P.A.; Klein, R.; Foord, A.; Yu, M.; Riddell, S.; Haining, J.; Johnson, D.; Hemida, M.G.; Barr, J.; et al. Experimental Infection and Response to Rechallenge of Alpacas with Middle East Respiratory Syndrome Coronavirus. Emerg. Infect. Dis. 2016, 22, 1071–1074. [Google Scholar] [CrossRef]
  88. Adney, D.R.; Brown, V.R.; Porter, S.M.; Bielefeldt-Ohmann, H.; Hartwig, A.E.; Bowen, R.A. Inoculation of Goats, Sheep, and Horses with MERS-CoV Does Not Result in Productive Viral Shedding. Viruses 2016, 8, 230. [Google Scholar] [CrossRef] [PubMed]
  89. Harris, S.R.; Robinson, C.; Steward, K.F.; Webb, K.S.; Paillot, R.; Parkhill, J.; Holden, M.T.; Waller, A.S. Genome specialization and decay of the strangles pathogen, Streptococcus equi, is driven by persistent infection. Genome Res. 2015, 25, 1360–1371. [Google Scholar] [CrossRef]
  90. Waller, A.S. Strangles: A pathogenic legacy of the war horse. Vet. Rec. 2016, 178, 91–92. [Google Scholar] [CrossRef] [PubMed]
  91. Reusken, C.B.; Haagmans, B.L.; Muller, M.A.; Gutierrez, C.; Godeke, G.J.; Meyer, B.; Muth, D.; Raj, V.S.; Smits-De Vries, L.; Corman, V.M.; et al. Middle East respiratory syndrome coronavirus neutralising serum antibodies in dromedary camels: A comparative serological study. Lancet Infect. Dis. 2013, 13, 859–866. [Google Scholar] [CrossRef]
  92. Hemida, M.G.; Perera, R.A.; Wang, P.; Alhammadi, M.A.; Siu, L.Y.; Li, M.; Poon, L.L.; Saif, L.; Alnaeem, A.; Peiris, M. Middle East Respiratory Syndrome (MERS) coronavirus seroprevalence in domestic livestock in Saudi Arabia, 2010 to 2013. Eurosurveillance 2013, 18, 20659. [Google Scholar] [CrossRef]
  93. Van Doremalen, N.; Hijazeen, Z.S.; Holloway, P.; Al Omari, B.; McDowell, C.; Adney, D.; Talafha, H.A.; Guitian, J.; Steel, J.; Amarin, N.; et al. High Prevalence of Middle East Respiratory Coronavirus in Young Dromedary Camels in Jordan. Vector Borne Zoonotic Dis. 2017, 17, 155–159. [Google Scholar] [CrossRef] [PubMed]
  94. Meyer, B.; Garcia-Bocanegra, I.; Wernery, U.; Wernery, R.; Sieberg, A.; Muller, M.A.; Drexler, J.F.; Drosten, C.; Eckerle, I. Serologic assessment of possibility for MERS-CoV infection in equids. Emerg. Infect. Dis. 2015, 21, 181–182. [Google Scholar] [CrossRef]
  95. Hemida, M.G.; Chu, D.K.W.; Perera, R.; Ko, R.L.W.; So, R.T.Y.; Ng, B.C.Y.; Chan, S.M.S.; Chu, S.; Alnaeem, A.A.; Alhammadi, M.A.; et al. Coronavirus infections in horses in Saudi Arabia and Oman. Transbound. Emerg. Dis. 2017, 64, 2093–2103. [Google Scholar] [CrossRef]
  96. Ali, M.; El-Shesheny, R.; Kandeil, A.; Shehata, M.; Elsokary, B.; Gomaa, M.; Hassan, N.; El Sayed, A.; El-Taweel, A.; Sobhy, H.; et al. Cross-sectional surveillance of Middle East respiratory syndrome coronavirus (MERS-CoV) in dromedary camels and other mammals in Egypt, August 2015 to January 2016. Eurosurveillance 2017, 22, 30487. [Google Scholar] [CrossRef]
  97. David, D.; Rotenberg, D.; Khinich, E.; Erster, O.; Bardenstein, S.; van Straten, M.; Okba, N.M.A.; Raj, S.V.; Haagmans, B.L.; Miculitzki, M.; et al. Middle East respiratory syndrome coronavirus specific antibodies in naturally exposed Israeli llamas, alpacas and camels. One Health 2018, 5, 65–68. [Google Scholar] [CrossRef]
  98. Reusken, C.B.; Schilp, C.; Raj, V.S.; De Bruin, E.; Kohl, R.H.; Farag, E.A.; Haagmans, B.L.; Al-Romaihi, H.; Le Grange, F.; Bosch, B.J.; et al. MERS-CoV Infection of Alpaca in a Region Where MERS-CoV is Endemic. Emerg. Infect. Dis. 2016, 22, 1129. [Google Scholar] [CrossRef] [PubMed]
  99. Vergara-Alert, J.; Raj, V.S.; Munoz, M.; Abad, F.X.; Cordon, I.; Haagmans, B.L.; Bensaid, A.; Segales, J. Middle East respiratory syndrome coronavirus experimental transmission using a pig model. Transbound. Emerg. Dis. 2017, 64, 1342–1345. [Google Scholar] [CrossRef]
  100. Widagdo, W.; Okba, N.M.A.; Li, W.; de Jong, A.; de Swart, R.; Begeman, L.; Cunningham, A.A.; van Riel, D.; van den Brand, J.M.A.; Segales, J.; et al. Species specific binding of the MERS-coronavirus S1A protein. Unpublished work. 2019. [Google Scholar]
  101. Yang, Y.; Du, L.; Liu, C.; Wang, L.; Ma, C.; Tang, J.; Baric, R.S.; Jiang, S.; Li, F. Receptor usage and cell entry of bat coronavirus HKU4 provide insight into bat-to-human transmission of MERS coronavirus. Proc. Natl. Acad. Sci. USA 2014, 111, 12516–12521. [Google Scholar] [CrossRef] [PubMed]
  102. Nam, H.S.; Park, J.W.; Ki, M.; Yeon, M.Y.; Kim, J.; Kim, S.W. High fatality rates and associated factors in two hospital outbreaks of MERS in Daejeon, the Republic of Korea. Int. J. Infect. Dis. 2017, 58, 37–42. [Google Scholar] [CrossRef]
  103. Alraddadi, B.M.; Watson, J.T.; Almarashi, A.; Abedi, G.R.; Turkistani, A.; Sadran, M.; Housa, A.; Almazroa, M.A.; Alraihan, N.; Banjar, A.; et al. Risk Factors for Primary Middle East Respiratory Syndrome Coronavirus Illness in Humans, Saudi Arabia, 2014. Emerg. Infect. Dis. 2016, 22, 49–55. [Google Scholar] [CrossRef]
  104. Seys, L.J.M.; Widagdo, W.; Verhamme, F.M.; Kleinjan, A.; Janssens, W.; Joos, G.F.; Bracke, K.R.; Haagmans, B.L.; Brusselle, G.G. DPP4, the Middle East Respiratory Syndrome Coronavirus Receptor, is Upregulated in Lungs of Smokers and Chronic Obstructive Pulmonary Disease Patients. Clin. Infect. Dis. 2018, 66, 45–53. [Google Scholar] [CrossRef]
  105. Dahlin, K.; Mager, E.M.; Allen, L.; Tigue, Z.; Goodglick, L.; Wadehra, M.; Dobbs, L. Identification of genes differentially expressed in rat alveolar type I cells. Am. J. Respir. Cell. Mol. Biol. 2004, 31, 309–316. [Google Scholar] [CrossRef] [PubMed]
  106. Evans, M.J.; Cabral, L.J.; Stephens, R.J.; Freeman, G. Renewal of alveolar epithelium in the rat following exposure to NO2. Am. J. Pathol. 1973, 70, 175–198. [Google Scholar] [PubMed]
  107. Alsaad, K.O.; Hajeer, A.H.; Al Balwi, M.; Al Moaiqel, M.; Al Oudah, N.; Al Ajlan, A.; AlJohani, S.; Alsolamy, S.; Gmati, G.E.; Balkhy, H.; et al. Histopathology of Middle East respiratory syndrome coronovirus (MERS-CoV) infection-clinicopathological and ultrastructural study. Histopathology 2018, 72, 516–524. [Google Scholar] [CrossRef]
  108. Haagmans, B.L.; Kuiken, T.; Martina, B.E.; Fouchier, R.A.; Rimmelzwaan, G.F.; van Amerongen, G.; van Riel, D.; de Jong, T.; Itamura, S.; Chan, K.H.; et al. Pegylated interferon-alpha protects type 1 pneumocytes against SARS coronavirus infection in macaques. Nat. Med. 2004, 10, 290–293. [Google Scholar] [CrossRef]
  109. Chen, Z.; Bao, L.; Chen, C.; Zou, T.; Xue, Y.; Li, F.; Lv, Q.; Gu, S.; Gao, X.; Cui, S.; et al. Human Neutralizing Monoclonal Antibody Inhibition of Middle East Respiratory Syndrome Coronavirus Replication in the Common Marmoset. J. Infect. Dis. 2017, 215, 1807–1815. [Google Scholar] [CrossRef]
  110. Chan, J.F.; Yao, Y.; Yeung, M.L.; Deng, W.; Bao, L.; Jia, L.; Li, F.; Xiao, C.; Gao, H.; Yu, P.; et al. Treatment With Lopinavir/Ritonavir or Interferon-beta1b Improves Outcome of MERS-CoV Infection in a Nonhuman Primate Model of Common Marmoset. J. Infect. Dis. 2015, 212, 1904–1913. [Google Scholar] [CrossRef]
  111. Falzarano, D.; de Wit, E.; Feldmann, F.; Rasmussen, A.L.; Okumura, A.; Peng, X.; Thomas, M.J.; van Doremalen, N.; Haddock, E.; Nagy, L.; et al. Infection with MERS-CoV causes lethal pneumonia in the common marmoset. PLoS. Pathog. 2014, 10, e1004250. [Google Scholar] [CrossRef]
  112. Widagdo, W.; Wiersma, L.C.M.; Smits, S.L.; de Vries, R.D.; Schipper, D.; Raj, V.S.; van den Ham, H.J.; Brown, R.; Zambon, M.; Kondova, I.; et al. DPP4-expressing type I pneumocytes in a fatal human MERS-coronavirus case. Unpublished work. 2019. [Google Scholar]
  113. Li, K.; Wohlford-Lenane, C.; Perlman, S.; Zhao, J.; Jewell, A.K.; Reznikov, L.R.; Gibson-Corley, K.N.; Meyerholz, D.K.; McCray, P.B., Jr. Middle East Respiratory Syndrome Coronavirus Causes Multiple Organ Damage and Lethal Disease in Mice Transgenic for Human Dipeptidyl Peptidase 4. J. Infect. Dis. 2016, 213, 712–722. [Google Scholar] [CrossRef]
  114. Algaissi, A.; Agrawal, A.S.; Han, S.; Peng, B.H.; Luo, C.; Li, F.; Chan, T.S.; Couch, R.B.; Tseng, C.K. Elevated Human Dipeptidyl Peptidase 4 Expression Reduces the Susceptibility of hDPP4 Transgenic Mice to Middle East Respiratory Syndrome Coronavirus Infection and Disease. J. Infect. Dis. 2018, 219, 829–835. [Google Scholar] [CrossRef] [PubMed]
  115. Prescott, J.; Falzarano, D.; de Wit, E.; Hardcastle, K.; Feldmann, F.; Haddock, E.; Scott, D.; Feldmann, H.; Munster, V.J. Pathogenicity and Viral Shedding of MERS-CoV in Immunocompromised Rhesus Macaques. Front. Immunol. 2018, 9, 205. [Google Scholar] [CrossRef]
  116. Kim, S.H.; Ko, J.H.; Park, G.E.; Cho, S.Y.; Ha, Y.E.; Kang, J.M.; Kim, Y.J.; Huh, H.J.; Ki, C.S.; Jeong, B.H.; et al. Atypical presentations of MERS-CoV infection in immunocompromised hosts. J. Infect. Chemother. 2017, 23, 769–773. [Google Scholar] [CrossRef]
  117. Zhao, J.; Alshukairi, A.N.; Baharoon, S.A.; Ahmed, W.A.; Bokhari, A.A.; Nehdi, A.M.; Layqah, L.A.; Alghamdi, M.G.; Al Gethamy, M.M.; Dada, A.M.; et al. Recovery from the Middle East respiratory syndrome is associated with antibody and T-cell responses. Sci. Immunol. 2017, 2, eaan5393. [Google Scholar] [CrossRef] [PubMed]
  118. Coleman, C.M.; Sisk, J.M.; Halasz, G.; Zhong, J.; Beck, S.E.; Matthews, K.L.; Venkataraman, T.; Rajagopalan, S.; Kyratsous, C.A.; Frieman, M.B. CD8+ T Cells and Macrophages Regulate Pathogenesis in a Mouse Model of Middle East Respiratory Syndrome. J. Virol. 2017, 91, e01825-16. [Google Scholar] [CrossRef]
  119. Hart, B.J.; Dyall, J.; Postnikova, E.; Zhou, H.; Kindrachuk, J.; Johnson, R.F.; Olinger, G.G., Jr.; Frieman, M.B.; Holbrook, M.R.; Jahrling, P.B.; et al. Interferon-beta and mycophenolic acid are potent inhibitors of Middle East respiratory syndrome coronavirus in cell-based assays. J. Gen. Virol. 2014, 95 Pt 3, 571–577. [Google Scholar] [CrossRef]
  120. Oudshoorn, D.; van der Hoeven, B.; Limpens, R.W.; Beugeling, C.; Snijder, E.J.; Barcena, M.; Kikkert, M. Antiviral Innate Immune Response Interferes with the Formation of Replication-Associated Membrane Structures Induced by a Positive-Strand RNA Virus. MBio 2016, 7, e01991-16. [Google Scholar] [CrossRef]
  121. Uno, K.; Yagi, K.; Yoshimori, M.; Tanigawa, M.; Yoshikawa, T.; Fujita, S. IFN production ability and healthy ageing: Mixed model analysis of a 24 year longitudinal study in Japan. BMJ Open 2013, 3, e002113. [Google Scholar] [CrossRef]
  122. Li, G.; Ju, J.; Weyand, C.M.; Goronzy, J.J. Age-Associated Failure To Adjust Type I IFN Receptor Signaling Thresholds after T Cell Activation. J. Immunol. 2015, 195, 865–874. [Google Scholar] [CrossRef] [PubMed]
  123. Channappanavar, R.; Fehr, A.R.; Vijay, R.; Mack, M.; Zhao, J.; Meyerholz, D.K.; Perlman, S. Dysregulated Type I Interferon and Inflammatory Monocyte-Macrophage Responses Cause Lethal Pneumonia in SARS-CoV-Infected Mice. Cell Host Microbe 2016, 19, 181–193. [Google Scholar] [CrossRef]
  124. Zumla, A.; Hui, D.S.; Perlman, S. Middle East respiratory syndrome. Lancet 2015, 386, 995–1007. [Google Scholar] [CrossRef]
  125. Alfaraj, S.H.; Al-Tawfiq, J.A.; Alzahrani, N.A.; Altwaijri, T.A.; Memish, Z.A. The impact of co-infection of influenza A virus on the severity of Middle East Respiratory Syndrome Coronavirus. J. Infect. 2017, 74, 521–523. [Google Scholar] [CrossRef]
  126. Chu, D.K.W.; Hui, K.P.Y.; Perera, R.; Miguel, E.; Niemeyer, D.; Zhao, J.; Channappanavar, R.; Dudas, G.; Oladipo, J.O.; Traore, A.; et al. MERS coronaviruses from camels in Africa exhibit region-dependent genetic diversity. Proc. Natl. Acad. Sci. USA 2018, 115, 3144–3149. [Google Scholar] [CrossRef]
  127. Assiri, A.M.; Biggs, H.M.; Abedi, G.R.; Lu, X.; Bin Saeed, A.; Abdalla, O.; Mohammed, M.; Al-Abdely, H.M.; Algarni, H.S.; Alhakeem, R.F.; et al. Increase in Middle East Respiratory Syndrome-Coronavirus Cases in Saudi Arabia Linked to Hospital Outbreak With Continued Circulation of Recombinant Virus, July 1-August 31, 2015. Open Forum Infect. Dis. 2016, 3, ofw165. [Google Scholar] [CrossRef] [PubMed]
  128. Payne, D.C.; Biggs, H.M.; Al-Abdallat, M.M.; Alqasrawi, S.; Lu, X.; Abedi, G.R.; Haddadin, A.; Iblan, I.; Alsanouri, T.; Al Nsour, M.; et al. Multihospital Outbreak of a Middle East Respiratory Syndrome Coronavirus Deletion Variant, Jordan: A Molecular, Serologic, and Epidemiologic Investigation. Open Forum Infect. Dis. 2018, 5, ofy095. [Google Scholar] [CrossRef] [PubMed]
  129. Lamers, M.M.; Raj, V.S.; Shafei, M.; Ali, S.S.; Abdallh, S.M.; Gazo, M.; Nofal, S.; Lu, X.; Erdman, D.D.; Koopmans, M.P.; et al. Deletion Variants of Middle East Respiratory Syndrome Coronavirus from Humans, Jordan, 2015. Emerg. Infect. Dis. 2016, 22, 716–719. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Schematic figure depicting four structural proteins of Middle East respiratory syndrome coronavirus (MERS-CoV), i.e., S, E, M, and N proteins (A); a cartoon representation of MERS-CoV S1 protein binding to DPP4 (PDB code 4L72) (B). The S protein consists of the S1 and S2 subunits. The α/β hydrolase domain of DPP4 is indicated in red, β-propeller domain in green, while part of the MERS-CoV S1 protein is shown in blue.
Figure 1. Schematic figure depicting four structural proteins of Middle East respiratory syndrome coronavirus (MERS-CoV), i.e., S, E, M, and N proteins (A); a cartoon representation of MERS-CoV S1 protein binding to DPP4 (PDB code 4L72) (B). The S protein consists of the S1 and S2 subunits. The α/β hydrolase domain of DPP4 is indicated in red, β-propeller domain in green, while part of the MERS-CoV S1 protein is shown in blue.
Viruses 11 00280 g001
Figure 2. Schematic overview of viral RNA and infectious virus shedding of MERS-CoV-inoculated dromedary camels, pigs, and rabbits. Each data point represents the average data from previous experiments [17,33,84]. Viral RNA is measured in TCID50/mL genome equivalents, while infectious virus is expressed in TCID50/mL.
Figure 2. Schematic overview of viral RNA and infectious virus shedding of MERS-CoV-inoculated dromedary camels, pigs, and rabbits. Each data point represents the average data from previous experiments [17,33,84]. Viral RNA is measured in TCID50/mL genome equivalents, while infectious virus is expressed in TCID50/mL.
Viruses 11 00280 g002
Figure 3. Schematic representation of DPP4 expression and MERS-CoV-recognized α2,3-sialic acid glycotopes in the respiratory tract of dromedary camel, pig, rabbit, human, and sheep.
Figure 3. Schematic representation of DPP4 expression and MERS-CoV-recognized α2,3-sialic acid glycotopes in the respiratory tract of dromedary camel, pig, rabbit, human, and sheep.
Viruses 11 00280 g003
Figure 4. MERS-CoV infection in the lungs of asymptomatic-to-mild (left panel) and severe-to-fatal cases (right panel). Shown is a hypothetical model with two critical host determinants, DPP4 and interferon, differentially expressed in asymptomatic-to-mild and severe-to-fatal MERS-CoV infection.
Figure 4. MERS-CoV infection in the lungs of asymptomatic-to-mild (left panel) and severe-to-fatal cases (right panel). Shown is a hypothetical model with two critical host determinants, DPP4 and interferon, differentially expressed in asymptomatic-to-mild and severe-to-fatal MERS-CoV infection.
Viruses 11 00280 g004

Share and Cite

MDPI and ACS Style

Widagdo, W.; Sooksawasdi Na Ayudhya, S.; Hundie, G.B.; Haagmans, B.L. Host Determinants of MERS-CoV Transmission and Pathogenesis. Viruses 2019, 11, 280. https://0-doi-org.brum.beds.ac.uk/10.3390/v11030280

AMA Style

Widagdo W, Sooksawasdi Na Ayudhya S, Hundie GB, Haagmans BL. Host Determinants of MERS-CoV Transmission and Pathogenesis. Viruses. 2019; 11(3):280. https://0-doi-org.brum.beds.ac.uk/10.3390/v11030280

Chicago/Turabian Style

Widagdo, W., Syriam Sooksawasdi Na Ayudhya, Gadissa B. Hundie, and Bart L. Haagmans. 2019. "Host Determinants of MERS-CoV Transmission and Pathogenesis" Viruses 11, no. 3: 280. https://0-doi-org.brum.beds.ac.uk/10.3390/v11030280

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop