ArticlePDF AvailableLiterature Review

Comparative pathology, molecular pathogenicity, immunological features, and genetic characterization of three highly pathogenic human coronaviruses (MERS-CoV, SARS-CoV, and SARS-CoV-2)

Authors:
  • Johns Hopkins Aramco healthcare
  • Almoosa Health Group - Saudi Arabia. Wollongong University, Australia

Abstract

The last two decades have witnessed the emergence of three deadly coronaviruses (CoVs) in humans: severe acute respiratory syndrome coronavirus (SARS-CoV), Middle East respiratory syndrome coronavirus (MERS-CoV), and severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). There are still no reliable and efficient therapeutics to manage the devastating consequences of these CoVs. Of these, SARS-CoV-2, the cause of the currently ongoing coronavirus disease 2019 (COVID-19) pandemic, has posed great global health concerns. The COVID-19 pandemic has resulted in an unprecedented crisis with devastating socio-economic and health impacts worldwide. This highlights the fact that CoVs continue to evolve and have the genetic flexibility to become highly pathogenic in humans and other mammals. SARS-CoV-2 carries a high genetic homology to the previously identified CoV (SARS-CoV), and the immunological and pathogenic characteristics of SARS-CoV-2, SARS-CoV, and MERS contain key similarities and differences that can guide therapy and management. This review presents salient and updated information on comparative pathology, molecular pathogenicity, immunological features, and genetic characterization of SARS-CoV, MERS-CoV, and SARS-CoV-2; this can help in the design of more effective vaccines and therapeutics for countering these pathogenic CoVs.
7162
European Review for Medical and Pharmacological Sciences 2021; 25: 7162-7184
A.A. RABAAN1,2,3, A.A. MUTAIR4,5,6, Z.A. ALAWI7, S. ALHUMAID8,
M.A. MOHAINI9,10, J. ALDALI11, R. TIRUPATHI12,13, A.A. SULE14, T. KORITALA15,
R. ADHIKARI16, M. BILAL17, M. DHAWAN18,19, R.K. MOHAPATRA20, R. TIWARI21,
S.A. SAMI22, S. MITRA23, M.K. PANDEY24, H. HARAPAN25,26,27,
T.B. EMR AN28, K. DHAMA29
1Molecular Diagnostic Laboratory, Johns Hopkins Aramco Healthcare, Dhahran, Saudi Arabia
2Department of Public Health and Nutrition, The University of Haripur, Haripur, Pakistan
3College of Medicine, Alfaisal University, Riyadh, Saudi Arabia
4Research Center, Almoosa Specialist Hospital, Al-Ahsa, Saudi Arabia
5College of Nursing, Princess Norah Bint Abdulrahman University, Riyadh, Saudi Arabia
6School of Nursing, Wollongong University, Wollongong, NSW, Australia
7Division of Allergy and Immunology, College of Medicine, King Faisal University, Al-Ahsa, Saudi Arabia
8Administration of Pharmaceutical Care, Al-Ahsa Health Cluster, Ministry of Health, Al-Ahsa, Saudi Arabia
9Basic Sciences Department, College of Applied Medical Sciences, King Saud bin Abdulaziz
University for Health Sciences, Al-Ahsa, Saudi Arabia
10King Abdullah International Medical Research Center, Al-Ahsa, Saudi Arabia
11Pathology Organization, Imam Mohammed Ibn Saud Islamic University, Riyadh, Saudi Arabia
12Department of Medicine Keystone Health, Penn State University School of Medicine, Hershey, PA, USA
13Department of Medicine Wellspan Chambersburg and Waynesboro Hospitals, Penn State
Chambersburg, PA, USA
14Department of Informatics and Outcomes, St. Joseph Mercy Oakland Pontiac, MI, USA
15Department of Internal Medicine, Mayo Clinic Health System Mankato, Mayo Clinic College of
Medicine and Science, MN, USA
16Department of Hospital Medicine, Franciscan Health Lafayette, IN, USA
17School of Life Science and Food Engineering, Huaiyin Institute of Technology, Huaian, China
18Department of Microbiology, Punjab Agricultural University, Ludhiana, India
19The Trafford Group of Colleges, Manchester, UK
20Department of Chemistry, Government College of Engineering, Keonjhar, Odisha, India
21Department of Veterinary Microbiology and Immunology, College of Veterinary Sciences,
Uttar Pradesh Pandit DeenDayal Upadhyaya PashuChikitsa Vigyan Vishwavidyalaya Evam Go
AnusandhanSansthan (DUVASU), Mathura, India
22Department of Pharmacy, Faculty of Biological Sciences, University of Chittagong, Chittagong,
Bangladesh
23Department of Pharmacy, Faculty of Pharmacy, University of Dhaka, Dhaka, Bangladesh
24Department of Translational Medicine Center, All India Institute of Medical Sciences, Bhopal,
Madhya Pradesh, India
25Medical Research Unit, School of Medicine, Universitas Syiah Kuala, Banda Aceh, Aceh, Indonesia
26Tropical Diseases Centre, School of Medicine, Universitas Syiah Kuala, Banda Aceh, Aceh, Indonesia
27Department of Microbiology, School of Medicine, Universitas Syiah Kuala, Banda Aceh, Aceh, Indonesia
28Department of Pharmacy, BGC Trust University Bangladesh, Chittagong, Bangladesh
29Division of Pathology, ICAR-Indian Veterinary Research Institute, Izatnagar, Bareilly, Uttar
Pradesh, India
Corresponding Authors: Kuldeep Dhama, MVSc, Ph.D; e-mail: kdhama@rediffmail.com
Talha Bin Emran, Ph.D; e-mail: talhabmb@bgctub.ac.bd
Comparative pathology, molecular pathogenicity,
immunological features, and genetic
characterization of three highly pathogenic
human coronaviruses (MERS-CoV, SARS-CoV,
and SARS-CoV-2)
Comparative review of three human coronaviruses
7163
Abstract. The last two decades have wit-
nessed the emergence of three deadly coronavi-
ruses (CoVs) in humans: severe acute respiratory
syndrome coronavirus (SARS-CoV), Middle East
respiratory syndrome coronavirus (MERS-CoV),
and severe acute respiratory syndrome corona-
virus 2 (SARS-CoV-2). There are still no reliable
and efcient therapeutics to manage the devas-
tating consequences of these CoVs. Of these,
SARS-CoV-2, the cause of the currently ongoing
coronavirus disease 2019 (COVID-19) pandem-
ic, has posed great global health concerns. The
COVID-19 pandemic has resulted in an unprec-
edented crisis with devastating socio-econom-
ic and health impacts worldwide. This highlights
the fact that CoVs continue to evolve and have
the genetic exibility to become highly patho-
genic in humans and other mammals. SARS-
CoV-2 carries a high genetic homology to the
previously identied CoV (SARS-CoV), and the
immunological and pathogenic characteristics
of SARS-CoV-2, SARS-CoV, and MERS contain
key similarities and differences that can guide
therapy and management. This review presents
salient and updated information on compara-
tive pathology, molecular pathogenicity, immu-
nological features, and genetic characterization
of SARS-CoV, MERS-CoV, and SARS-CoV-2; this
can help in the design of more effective vaccines
and therapeutics for countering these pathogen-
ic CoVs.
Key Words:
Pathology, Immunology, Genetic characterization,
Coronaviruses, MERS-CoV, SARS-CoV, SARS-CoV-2,
COVID-19.
Introduction
The devastating fact that zoonotic diseases at-
tributed to coronavirus (CoV) strains can result
in pandemics came to public attention in 2003
after a severe acute respiratory syndrome coro-
navirus (SARS-CoV) outbreak. Since this re-
alization, scientists and public health ofcials
have raised concerns over health threats posed
to the human population by the three coronavi-
ruses (CoVs) SARS-CoV, Middle East respira-
tory syndrome coronavirus (MERS-CoV), and
severe acute respiratory syndrome coronavirus
2 (SA R S - C oV- 2)1-7. Among at least six strains of
human-infecting CoVs that have been identied
by studies, these three have proved to be highly
pathogenic as they trigger severe pneumonia and
systemic symptoms in humans5,8 -13. CoVs are a
complex and diverse family of enveloped, posi-
tive-sense, single-stranded RNA viruses and are
divided into four genera: alpha, beta, gamma, and
delta CoV8,14,15. Of these, beta CoVs have drawn
the most attention due to their ability to cross an-
imal-human barriers and act as signicant global
infectious agents2, 6,8 ,16. SARS-CoV, SARS-CoV-2,
and MERS-CoV have been identied as the most
important and evolving beta CoVs, and their mo-
lecular biology and immunological features re-
main to be investigated in detail1,4,5,8 ,17-19. Seasonal
variations have been observed in the pattern of
these viruses: SARS-CoV-2 outbreak occurs in
the winter, in contrast to MERS-CoV and SARS-
CoV outbreaks and triggering severe pneumo-
nia18. Moreover, these three viruses show similar
genomic composition, clinical manifestations,
and route of transmission1,4, 20. The current pan-
demic of coronavirus disease 2019 (COVID-19)
caused by SARS-CoV-2 has apparent similari-
ties with SARS9,10, 21, including disease progres-
sion, escape from the host immune system, and
subsequent acute respiratory distress syndrome
(ARDS). The International Committee on Tax-
onomy of Viruses (ICTV) designated the causal
agent of COVID-19 as “SARS-CoV-2” due to its
similarities with SARS-CoV19,22-2 4.
During the COVID-19 pandemic, the world
has experienced unprecedented challenges, with
over 4.9 million deaths and more than 243 mil-
lion conrmed cases of SARS-CoV-2 infection in
over 225 countries with a fatality rate ranging be-
tween 1.5% and 5% as of October 26, 202117,22,25.
A high case fatality rate of about 49%9,26 has been
reported in patients with an acute disease requir-
ing ventilator support and Intensive Care Unit
(ICU) admission. There have been signicant
breakthroughs in vaccine development, with sev-
eral vaccines administered globally for protection
against SARS-CoV-2. In addition, many effective
drugs and therapeutic candidates are being eval-
uated, such as antivirals, monoclonal antibodies,
cytokine inhibitors, and immunosuppressants27-32 .
SARS-CoV-2, when observed under an electron
microscope, has a structure similar to a crown (co-
rona). The mechanism of virus entry into the host is
identical to that of SARS-CoV, which binds to the
human angiotensin-converting enzyme 2 (ACE2)
receptor via its protein receptor-binding domain
(R BD)33-3 5. In contrast, MERS-CoV binds to the
DPP4 receptor to enter host cells. Genomic analysis
data has revealed that the genome sequence simi-
larity of SARS-CoV-2 to SARS-CoV and MERS-
CoV is 80% and 50%, respectively21,36 . While ex-
ploring the evolutionary potential of SARS-CoV-2,
studies have found that its genome exhibits 96%
A.A. Rabaan, A.A. Mutair, Z.A. Alawi, S. Alhumaid, M.A. Mohaini, J. Aldali, R. Tirupathi, et al
7164
similarity to that of bat-derived CoV isolated in
201337,38 . SARS-CoV-2 and SARS-CoV have 380
amino acid (AA) substitution sites. It has been hy-
pothesized that any substitution in the AA sequence
could lead to a possible novel viral protein function
with unclear pathogenesis39. The spike (S) protein
and the nucleocapsid protein are linked to higher
transmission capability and lower pathogenicity
in SARS-CoV-2. However, the mutations in the S
protein are especially crucial because the S protein
is key for the rst step of viral transmission: entry
into the cell by binding to the ACE2 receptor40-44.
SARS-CoV-2 steadily mutates during continuous
transmission among humans, and naturally occur-
ring S mutations can reduce or enhance cell entry
via the ACE2 receptor44,45. According to a recent
study46, six AA residues (D480, T487, Y442, N479,
L472, and Y4911 for SARS-CoV, and Q493, S494,
L455, F486, Y505, and N501 for SARS-CoV-2) are
essential for binding to the human ACE2 receptor.
Among these six SARS-CoV-2 AA residues, the
lack of similarity of ve residues to those of SARS-
CoV may be attributed to the deletions, insertions,
or mutations in the S1 and S2 regions, which are
responsible for evolutionary changes4 6,47. The nov-
el strain has an evolutionary path different from
those of MERS-CoV and SARS-CoV, with lin-
eage similarity to previously evaluated bat-derived
CoV. However, there are proteomic and genomic
differences between the bat and human CoVs, in-
dicating a unique immune invasion mechanism
and a distinct immunopathology associated with
host response48. The common clinical symptoms
of COVID-19 are similar to those of SARS: dry
cough (67.7%), fever (87.9%), myalgia (34.8%), fa-
tigue (69.6%), hypoxia, and progressive dyspnea
followed by damage to multiple organs. In contrast
to SARS-CoV, SARS-CoV-2 is more transmissi-
ble, but the overall mortality rate is lower than that
for SARS-CoV infection. Like MERS and SARS,
COVID-19 is likely to be more severe in elderly
people and those suffering underlying comorbid-
ities, including many chronic health conditions.
Here, we present salient and updated information
on comparative pathology, molecular pathogenic-
ity, immunological features, and genetic charac-
terization of SARS-CoV, MERS-CoV, and SARS-
CoV-2. As the current pandemic remains ongoing,
this review can contribute to the design of more
effective vaccines and therapeutics.
Early Phase of Viral Infection
In the early stages, SARS infection causes
non-specic symptoms such as myalgia, fever,
headache, and severe fatigue49. These symptoms
tend to diminish in seven days. Sequential naso-
pharyngeal aspirate samples from SARS patients
indicate a direct relationship between clinical pro-
gression and viral load50. After its peak, viral load
usually decreases rapidly, with IgG seroconver-
sion serving as an indicator of specic immunity
development. However, some patients’ clinical
conditions can worsen during this period, creat-
ing inconsistencies with viral clearance observa-
tions. Delay in viral peak can indicate absence or
hindrance of host antiviral responses necessary to
enhance viral clearance51.
A retrospective study evaluating the cause of
worsening clinical condition after viral load re-
duction highlighted the underlying association be-
tween viral clearance, immune dysregulation, and
disease development52. The host hyper-inamma-
tory response, not the cytopathic effect of the vi-
rus, may be responsible for this phenomenon36 ,51.
To some extent, rapid viral load elevation could
be the contributing factor for disease pathology.
Clinical features such as diarrhea, oxygen desat-
uration, hepatic dysfunction, and fatality indicate
that high viral load may contribute to direct organ
dysfunction49,53. Clinical specimens of various
anatomic sites of organ dysfunction have yielded
virus. For instance, stool specimen was highly re-
lated to diarrhea, with viral particles detected in
ileum and colon biopsies observed under an elec-
tron microscope54. There is extensive evidence
regarding the relationship between pathological
effects, viremia, and viral loads from these nd-
ings. Strong evidence exists of high viral loads
associated with massive inltration of the inam-
matory immune cells being signicantly linked
to worse clinical outcomes in patients54. Patients
with elevated viral load at an early stage were also
likely to have higher mortality55 ,56. Therefore, it
is essential to address the molecular pathology,
immunological characteristics, pathogenicity,
and genetic sequence of MERS-CoV, SARS-CoV,
SARS-CoV-2, and other CoVs. A few of the gen-
eral characteristics of MERS-CoV, SARS-CoV
and SARS-CoV-2 are presented in Table I.
Genetic Similarities of MERS-CoV,
SARS-CoV, and SARS-CoV-2
Among the CoV subtypes, beta CoVs cause
severe and fatal diseases in humans, while al-
pha CoVs cause mild infections. The genom-
ic sequences of MERS-CoV, SARS-CoV, and
SARS-CoV-2 are quite similar, but SARS-CoV-2
displays signicant differences in genome com-
Comparative review of three human coronaviruses
7165
position compared to its predecessors57. Genom-
ic analysis suggests that SARS-CoV-2 is closely
related to pangolin CoV (86%-92%) and bat CoV
(96%), which further suggests bats as the prima-
ry reservoir43,58 -60. Furthermore, the outbreak of
SARS-CoV-2 is thought to be linked to trading
practices in Wuhans wet market, and due to
the genetic identities between SARS-CoV-2 and
BatCoV RaTG13 (a bat-CoV), it has been hy-
pothesized that bats could be the natural source
of S A R S - CoV-2 8, 43,61. A plethora of research ev-
idence shows that pangolins may be the inter-
mediate host—there is 99% homology between
SARS-CoV-2 and the CoV strain originating from
pangolins—but bats are the natural reservoir for
the virus62, 63. Bats are generally recognized as
potential primary reservoirs for most of the RNA
viruses64. The genome of SARS-CoV-2 showed
96.2% homology to that of the bat CoV (RaTG13)
collected in the Yunnan province of China43. The
SARS-CoV-2 genome is closely related (88%) to
zoonotic bat viruses, bat-SL-CoVZXC45, and bat-
S L- C oV Z XC 2165. The most commonly identied
sequence similarity between these bat and human
viruses is in the E gene, and the least common-
ly identied similarity is in the S gene. Multiple
SARS-CoV-2 proteins have the same sequence
as the bat-SL-CoVZC45 and bat-SL-CoVZXC21,
except for the S protein and protein 1366. A team
of researchers concluded that pangolin-CoV is a
highly associated descendant of SARS-CoV-2,
suggesting that pangolins could be the natural
reservoirs for SARS-CoV-2 and bat CoV67. The se-
quence similarity (89.2%) between SARS-CoV-2
and RaTG13, in terms of the RBD, is less than
the sequence similarity (97.2%) between SARS-
CoV-2 and pangolin-CoV. Additionally, the latter
contains six complete identical RBD residues,
whereas the former contains only one identical
amino acid residue43. Notably, pangolins in Chi-
na are categorized as endangered due to their
decreasing numbers, which are close to the point
of extinction; this reduces the likelihood of pan-
golins acting as an intermediate host of SARS-
CoV-2. The selling of pangolins is against the law,
and they have not been spotted in Wuhan’s wet
markets in recent times68. Through the use of the
optimized random forest model for human se-
quences of MERS-CoV and SARS-CoV, interme-
diate hosts (Camelids and Carnivores) were con-
rmed based on evolutionary signatures. With the
same method, SARS-CoV-2 evolutionary signa-
tures identied bats as hosts, further conrming
bats as the suspected origin of the present pan-
demic69. Furthermore, a recent study70 based on
genetic similarities proposed that snakes may be
intermediate hosts, as there are similarities in co-
dons among SARS-CoV-2, bat CoV, and a snake
virus. However, this analysis was insufcient to
reach a conclusive hypothesis, as several limita-
Table I. Characteristics of MERS-CoV, SARS-CoV and SARS-CoV-2.
Features MERS-CoV SARS-CoV SARS-CoV-2
Outbreak 2012, April 2002, November 2019, December
Location of the rst case Jeddah, Saudi Arabia Guangdong, China Wuhan, China
Key hosts Bat, camel Bat, palm civets,
raccoon dogs Bat, pangolin
Active cases conrmed 2519 (from 2012 until 8096 Over 243 million
January 31, 2020) (as of October 26, 2021)
Genome length (bp) 30,119 29,751 29,903
Mortality 34.40% 10% (6.8-16.1%) 2-5%
Days took to infect the rst 1000 persons 903 130 48
Incubation period (day) 5 to 6 2 to 7 7 to 14
Basic reproduction number (R0) 1 2-4 1.4-5.5
Receptor DPP4 ACE2 ACE2
Mode of transmission Touching or consumption Believed to have Human-to-human
of camel milk or meat. spread on close contact
There is limited human-to-human from bats. transmission occurs when
transmission despite close There is evidence there is close physical
physical contact of human-to-human contact (mainly
transmission through respiratory aerosols/
droplets). The transmissions
may be possible through
fecal-oral route
and contaminated
objects/ surfaces/fomites
A.A. Rabaan, A.A. Mutair, Z.A. Alawi, S. Alhumaid, M.A. Mohaini, J. Aldali, R. Tirupathi, et al
7166
tions were present in the study71. In any case, beta
CoVs are less likely to infect reptiles by crossing
over through mammals72. These ndings made
the natural reservoir of CoV a controversial topic,
and a contingent of groups embrace the idea that
different intermediate host species are yet to be
discovered, other than bats73 -75. The disease out-
break related to SARS-CoV-2 demonstrates con-
cealed virus reservoirs in animals that may spread
into human populations occasionally76. The lower
effective number of codons and the extreme co-
don usage bias of SARS-CoV-2 in S, envelop, and
matrix protein genes suggest higher gene expres-
sion efciency than that of SARS, bat SARS, or
MERS-CoV, which is similar to Pangolin beta
CoV77. In the human host, the SARS-CoV-2 dinu-
cleotide pair, UpG and CpA dinucleotides, were
highly preferred, and CpG dinucleotide was high-
ly avoided. This strategy might imply evasion of
the human immune system78. Multiple sequence
alignments of the ACE2 receptor proteins of hu-
mans with that of dogs, cats, tigers, minks, and
other animals revealed a high homology and full
conservation of the ve AA residues, 353-KGD-
FR-357, among the species, which may throw
light on the possibility of transmission of SARS-
CoV-2 from animals to humans78.
MERS-CoV is closely related to two bat CoV
(HKU4, HKU5); it has been suggested that it
may be isolated from bats, and dromedary camels
probably act as intermediate host, as evidenced
from serological studies79,80 . In Qatar, the pres-
ence of MERS-CoV RNA was reported in swabs
obtained from dromedary camels that shared a
correlation with two human cases of MERS81. A
comprehensive evolutionary relationship analysis
depicted the origin of MERS-CoV from bats due
to the occurrence of recombination events within
S and ORF1ab genes82,83. Recombination events
were also reported in SARS-CoV as regions for
putative recombination were detected via com-
putational genomic studies84. The MERS-CoV
strains isolated from humans and camels have
been reported to share over 99% identity with
variations located in the ORF3, ORF4b, and S
genes85. SARS-CoV-2 shows 80% similarity with
SARS-CoV and 51% with the MERS-CoV86.
Most of the coding areas of SARS-CoV-2 indi-
cate a similar genomic architecture to that of the
bat-originating CoVs and SARS-CoV. The twelve
coding regions predicted are; lab, 3, E, M, 7, 8, 9,
10B, N, S, 13, and 14. The proteins encoded by
all the three CoVs are mostly similar in length87.
However, there is a signicant variation in the S
protein of SARS-CoV-2, which is longer in com-
parison to the protein encoded in the bat CoVs,
SA RS- C oV, and MERS-CoV88.
SARS-CoV-2 shares many similarities in ar-
chitecture and pathogenicity with SARS-CoV
compared to MERS-CoV. Mathematical mod-
els such as decision-tree experiments have also
shown remarkable characteristics of an AA se-
quence of SARS-CoV-2, which is different from
M E RS - C oV 12. The CoVs use a similar S protein
for binding to their respective host cells and the
same cellular protease enzyme for the activation
of the S protein89. The S protein in SARS-CoV-2
has a sequence similarity of about 77% with that
of SARS-CoV, structural proteins are more than
90% similar to SARS-CoV, and 32.79% similar to
MERS-CoV counterparts. The receptor-binding
domain (S2) of SARS-CoV-2 has a sequence sim-
ilarity of 74% with the S2 domain in SARS-CoV
and an overall similarity of about 52% with that
of S A R S - CoV90. The E protein of SARS-CoV-2 is
96.00% similar to that of SARS-CoV and 36.00%
similar to that of MERS-CoV. The M protein of
SARS-CoV-2 is 89.59% similar to that of SARS-
CoV and 39.27% similar to that of MERS-CoV.
The SARS-CoV-2 N protein is 85.41% similar to
that of SARS-CoV and 48.47% similar to that of
MERS- C oV 91.
Accessory proteins are regarded as essential
for in vitro replication of viral particles; however,
some of these proteins are associated with viral
pathogenesis92,93. The 3CLpro (nsp5) and RdRp
(nsp12) proteins of SARS-CoV-2 are prime me-
diators of replication and new virion production,
and they share high sequence identity with SARS-
CoV and M ERS-CoV94. Recent reports95,96 have
demonstrated that ORF8b and ORF3a of SARS-
CoV catalyze the induction of proinammatory
cytokines and thus play a role in regulating che-
motaxis in macrophages. ORF8b of SARS-CoV
and MERS-CoV is also involved in suppressing
the induction of interferon (IFN-I)97,9 8. Another
study demonstrated that ORF8 of SARS-CoV-2
variant binds to major histocompatibility complex
(MHC) and regulates its degradation in cell cul-
ture, indicating that immune evasion may be me-
diated by ORF8. However, SARS-CoV-2 ORF8
shows low homology to SARS-CoV ORF899.
Generally, no homologous accessory proteins
are found in CoV genera. However, some simi-
lar kinds of proteins might be present in closely
associated CoVs. For instance, SARS-CoV-2 and
SARS-CoV show over 80% similarities in OR-
F3a, 6, 7a, 7b, and 9b protein sequences.
Comparative review of three human coronaviruses
7167
Comparative Molecular Pathology of
MERS-CoV, SARS-CoV, and SARS-CoV-2
Infection
SARS-CoV is considered a zoonotic virus
that was transmitted to humans from birds pri-
or to human-to-human transmission100 . However,
in humans, various risk factors including age,
underlying metabolic disease like diabetes, and
heart disease, lead to an increase in death risk101.
SARS starts with viral infection in the respiratory
tract of people of all ages via droplet transmis-
sion of virus present in the mucus or saliva102. It
was reported that viral loads of SARS-CoV de-
creased with increased severity of the disease.
On the contrary, a similar trend is still unclear
fo r M E R S - CoV103. Clinical symptoms associated
with SARS-CoV infection include fever, chills,
diarrhea, myalgia, and fatigue104. SA R S - C oV e n -
ters into the human cell through the attachment
of viral S glycoprotein (S protein) to the ACE2
receptor. ACE2 functions as a dominant host re-
ceptor, and the presence of two co-receptors, DC-
SIGN (CD209) and L-SIGN (CD209L), are also
reported105,106. In dendritic cells, viral infection
does not occur prior to DC-SIGN binding, but this
binding may enhance SARS-CoV infection and
dissemination substantially. On the other hand,
L-SIGN is considered an alternative receptor that
may bind with its spike protein and regulate cel-
lular entry of SARS-CoV107. Changes occur in the
S glycoprotein in the endosomal environment via
the serine protease cathepsins B and L to assist
in the union process108. The S glycoprotein is not
just an essential structural protein of CoVs; it per-
forms a vital role in the association of virus with
the host cell. The S- protein is made up of two
subunits: S1 and S210 9. The S1 subunit contains the
RBD, which is responsible for binding the virus
to the host receptor, while the S2 subunit controls
membrane fusion occurring during virus-host
membrane interactions. These interactions lead
to the penetration of the viral genome into the cy-
toplasm of the host cell110. SARS-CoV-2 encodes
a longer S protein compared to SARS-CoV and
MERS-CoV, as identied by phylogenetic analy-
sis20 ,76. The RBD of SARS-CoV, MERS-CoV, and
SARS-CoV-2 binds to functional receptors pres-
ent on the cellular surface, allowing penetration of
the virus into host cells111. SARS-CoV and SARS-
CoV-2 predominately utilize angiotensin-convert-
ing enzyme 2 (ACE2) as a host receptor105, 110,111.
Additionally, viral entry by antibody-depen-
dent enhancement (ADE) has been observed112.
Through ADE, the B cell producing antibodies
may also expedite viral infection113. Surprising-
ly, ACE2 exhibits stronger afnities for SARS-
CoV-2 compared to SARS-CoV114. For instance,
the interaction between host ACE2 and SARS-
CoV-2 spike ectodomain displayed 10- to 20-fold
higher binding afnity than that for SARS-CoV
in a recent study115. Another study speculated that
SARS-CoV-2 could use other cellular receptors
and proteins to bind with host cell receptors such
as integrins116. However, there is to date insuf-
cient evidence to corroborate this assumption.
CD147-SP can be considered another entry portal
of S A R S - CoV-2117. In addition to attachment of S
proteins to functional host receptors, priming of
S proteins is necessary for invading the cellular
machinery of the host118.
Apart from lung cells, the heart, kidney, liver,
and tongue also express ACE2 receptors on their
epithelial cells119,12 0. In fact, cilia could be the entry
gate of the virus121. Surprisingly, after the S gly-
coprotein attaches to ACE2, there is a signicant
cilia loss, squamous cell metaplasia, and elevated
macrophage migration into the alveoli, causing
notable damage to alveoli in the lungs. Addition-
ally, SARS-CoV generates 7a and 3a proteins that
lead to substantial programmed death of cells in
the lungs, liver, and kidney122. Host translation
elongation factor 1 (EF-1A) and serine protease
2 strongly bind to N protein of both SARS-CoV
and MERS-CoV, and subsequently induce local
or systemic inammatory responses94. TH1 acti-
vation also causes increased inammation in the
affected organs. MERS-CoV infection is more
common in males than females123, and SARS-
CoV and SARS-CoV-2 infection follow the same
order of gender prevalence8. Clinical presentation
of infection may range from being asymptom-
atic to massive organ damage. Notably, MERS
is closely associated with metabolic syndromes
such as diabetes mellitus, obesity, and cardiovas-
cular morbidities124. The developing metabolic
syndrome in most cases alters the immunological
function, exposing the infected person to further
risk of more infections.
Many previous investigations reported that
CoV infection leads to cytopathic effects, includ-
ing cell lysis and apoptosis. Cellular fusion is
caused by the virus and usually leads to syncytia
formation. These processes are observed in the
cell due to the mobilization of vesicles that form
the replication complex and cause disruption of
Golgi complexes at the time of viral replication94.
Unlike in SARS-CoV, DPP4 CD26 is the MERS-
CoV attachment site to lung and respiratory tract
A.A. Rabaan, A.A. Mutair, Z.A. Alawi, S. Alhumaid, M.A. Mohaini, J. Aldali, R. Tirupathi, et al
7168
epithelial cells125. Notably, MERS-CoV carries a
particular RBD in its S glycogen that binds DDP4
on the host cells. DPP4 plays a signicant role
in altering glucose metabolism, activating the
T cells, modulating cytotoxicity, and regulating
apoptosis126. SARS-CoV-2 infects both the lower
and upper respiratory systems and multiple other
organs and systems, thus causing multiple patho-
logical conditions, including neurological and
gastrointestinal manifestations and kidney dam-
age127-129. ACE2 receptors are abundant in oral
mucosa, nasal secretory and ciliated cells, lower
airways, lungs, cornea, ileum, and colon. Hence,
patients suffer from collapsed lung and symptoms
of diarrhea130,131. When spike D614 is replaced by
mutant G614, S protein possesses greater stability
and a potential to grow at a temperature of 37oC,
compared to early SARS-CoV-2 isolates, which
showed a preference for 33oC132. While SARS-
CoV-2 is less pathogenic than MERS-CoV or
SARS-CoV, its human-to-human transmission is
faster1,4,133. Underlying illnesses (comorbidities)
such as heart disease, diabetes, and hypertension
have a close association with the severe patho-
genesis of SARS-CoV-2 in affected patients134.
These disorders reduce the generation of IFN and
interleukin that leads to the downregulation of
the host’s innate immunity via blockage of lym-
phocyte and macrophage functions. In healthy
people, ACE2 alters the renin-angiotensin system
through angiotensin-II breakdown into angioten-
sin-17 to prevent the development of acute lung
failure135. Acute lung injury is directly related
to a deciency in ACE2 and an increase in Ang
II136,137. Postmortem analysis of SARS-CoV-2 pa-
tients has revealed pneumocyte hyperplasia and
partial brosis leading to thickening and collapse
of alveoli93,138
The sgRNAs are presumed to be translated into
accessory and structural proteins of CoV in the
cytoplasm. A recently concluded in vitro study
indicated that the enzymatic function of the nsp14
exoribonuclease (ExoN) is crucial for replication
of SARS-CoV-2 and MERS-CoV139. By enhanc-
ing degradation and interfering with host mRNA
translation, beta CoV nsp1 inhibits the expression
of host genes and thus serves as a potent virulence
factor140. MERS-CoV nsp1 inhibits mRNA trans-
lation and induces mRNA degradation by selec-
tively targeting nuclear mRNA translation and
avoiding cytoplasmic viral mRNAs141. Current
structural analysis and related studies have un-
veiled that SARS-CoV-2 nsp1 inhibits ribosomal
mRNA entrance14 2. The delta CoVs and gamma
CoVs cannot produce nsp1 due to lack of nsp1/
nsp2 cleavage sites, though the same host shutoff
is triggered by other mechanisms that have not
been explored well.
Clinical and Immunological Features of
MERS-CoV, SARS-CoV, and SARS-CoV-2
Infection
MERS is currently a common human corona-
virus (hCoV) infection. MERS-CoV infection has
lower transmissibility than the other two CoVs
but causes severe symptoms, leading to a high-
er case fatality rate40. Like SARS-CoV patients,
patients with MERS-CoV usually show milder
symptoms initially and later develop dyspnea and
complications leading to respiratory failure, with
most of the patients (63.4%) developing lethal
pneumonia25. Organ function later deteriorates,
leading to fatality within two weeks after infec-
tion48. Prominent comorbidities associated with
mortality among MERS-CoV patients are diabe-
tes and renal failure, which result in poor health
outcomes. To better understand the pathogenesis
and immunological features of MERS-CoV, it is
essential to undertand its comparative analysis
with SARS-CoV infection1,4.
Unlike the SARS-CoV abortive mechanism
of infection, MERS-CoV multiplies in lympho-
cytes, dendritic cells, and macrophages52,5 4,55,143.
Viral genomes, nucleoprotein expression, and vi-
ral particles are detectable in virus-infected cells.
Viral multiplication in macrophages and dendritic
cells indicates that host cells are the source of vi-
ral reservoirs thwarting host immunorecognition
of the virus55. MERS-CoV has been reported to
induce greater transcriptomic changes than those
induced by SARS-CoV88. Cells carrying the virus
facilitate systemic dissemination of the infection
to lymph nodes. Naïve T cells interact with the vi-
rus and trigger adaptive immune responses. This
leads to the release of massive amounts of cyto-
kines and chemokines. The diverse activation av-
enues that trigger production of cytokines during
MERS-CoV infection cause a distinct cytokine
prole compared to SARS-CoV infection88,14 4.
The reason for productive replication is the
high number of DDP4 receptors expressed in the
dendritic and monocyte cells compared to the
expression level of ACE2 receptors. This results
in differential infection outcomes. MERS-CoV
can infect cells from different human cell lines
in ex-vivo studies36, 145. DDP4 receptors are identi-
able in endothelial and epithelial cells present in
the prostate, liver, kidney, and intestines41, 42. Dis-
Comparative review of three human coronaviruses
7169
semination of the virus throughout the body was
observed in patients with MERS-CoV infections,
explaining the high incidence of systemic events
like multi-organ failure and septic shock. Another
important immunopathological feature is the an-
tibody-dependent enhancement (ADE) conrmed
in MERS- C oV146. The underlying mechanism is
linked to the enhanced membrane fusion pro-
cess. The interactions of antibodies and RBD of S
protein tend to elevate proteolytic susceptibility,
leading to conformational changes in the target
host cells55. The binding Ab enhances virus entry
via canonical receptor-dependent pathways.
Only three cytokines (IL-6, IP-10, and IFN-γ)
display a marked increase in SARS-CoV, MERS-
CoV, a nd S A R S - CoV-2 35. SARS-CoV shows sig-
nicant IFNantagonism, while the MERS-CoV
has minor antagonist characteristics that lead to
enhanced sensitivity to IFN-I antiviral respons-
es147. Furthermore, due to differences in the vi-
ral proteins among these human CoVs, SARS-
CoV-2 is more sensitive to IFN-I-dependent
antiviral response compared to other CoVs. In
fact, the levels of IFN-I and IFN-III in patients
with SARS-CoV-2 infections are reduced, un-
like that in patients infected by SARS-CoV and
other respiratory viruses144. MERS-CoV shares
similar viral evasion strategies and IFN antago-
nism with SARS-CoV. As a result, MERS-CoV
tends to decrease upregulation of antiviral inter-
feron-stimulated gene (ISG) responses through a
novel approach, resulting in the induction of re-
pressive histone modications in the host cells.
Similar histone modications, which mediate
several biological events such as gene regulation,
were identied in patients with H5N1 u infec-
tion52 . Inhibition of transcription factor binding
is controlled by modifying the basal state of
host chromatin, where genes are packed. This
mechanism is linked to low ISG expression in
IFN-administered MERS40. As in patients with
SARS-CoV infection, levels of IFN-I in MERS-
CoV infected patients are attenuated, and their
rate of increase is slowed. The absence of IFN-I
resulted in a lack of marked lung immunopa-
thology in studies. It also improved the clinical
outcomes compared to the delayed IFN-1 group,
suggesting an atypical IFN-I effect linked to
SARS infection36 ,37, but there were adverse out-
comes linked to MERS-CoVinfection. In this
respect, early IFN administration improved the
protection of mice from severe infection, irre-
spective of down-regulation of cytokine-related
genes and ISG.
In seriously ill patients with MERS-CoV, the
inability to activate Th1 cells reduces IFN-γ
production, leading to activated natural killer
(NK) cells and CD8+ T cells. This generates an
uncontrolled immune response and attenuates
viral clearance36,4 3,148 ,149. Extensive CD8+ T-cell
responses were noted in critically ill patients
during the acute phase, indicating no benet as-
sociated with hyper-activated T-cell responses150.
The acute-phase T-cell response for SARS-CoV
corresponds to that for MERS-CoV. The absence
of T-cell response activation during the innate re-
sponse provided sufcient enhancement of host
survival and improved disease outcome. Addi-
tionally, the adaptive immune response against
MERS-CoV infection showed positive effects
relative to those observed against SARS-CoV in-
fection151.
Concentration of certain inammatory cyto-
kines and chemokines (CXCR3, SOCS5, IL-1β,
IL-8, IL-15, IL-17, CCR2, IL-1α, IP-10, TNF-α,
and IFN-γ) was reported to increase during
MERS-CoV infection144,1 52. In terms of cytokines,
MERS-CoV-associated IL-17 expression demon-
strated signicant upregulation compared to that
associated with SARS-CoV. Secretion of IL-17 by
CD4+ T cells can produce extensive pro-inam-
matory effects on host cells54 ,153. IL-17 expression
is known to aggravate respiratory syncytial virus
(RSV). In MERS-CoV, IL-17 expression tends to
induce immune-mediated pathology, resulting in
an elevated mortality rate143. Patients with MERS-
CoV infection exhibit higher and more prolonged
production of cytokines compared to those with
SARS-CoV infection20.
The SARS-CoV-2 pandemic has resulted in
devastating outcomes in the 21st century because
of high transmissibility and viral-shedding prop-
erties6,22. COVID-19 patients show clinical symp-
toms that resemble u during the onset of the dis-
ease, including myalgia, fever, and dry cough128 ,154.
Symptoms such as rhinorrhoea, pharyngalgia, al-
veolar edema, amblygeustia, shortness of breath,
dry mouth, nausea, and vomiting have also been
recorded in a small number of COVID-19 pa-
tients4,155-157. The laboratory ndings obtained
from patients infected by SARS-CoV, MERS-
CoV, and SARS-CoV-2 are markedly similar; the
most common abnormal ndings include throm-
bocytopenia and lymphocytopenia. Additionally,
signicant elevation in serum levels of alanine
aminotransferase, lactate dehydrogenase, aspar-
tate aminotransferase, and C-reactive protein
have been recorded158-16 0. In severely affected pa-
A.A. Rabaan, A.A. Mutair, Z.A. Alawi, S. Alhumaid, M.A. Mohaini, J. Aldali, R. Tirupathi, et al
7170
tients, coagulation disorders, where D-dimer lev-
el is elevated and prothrombin time is prolonged,
are commonly observed161. Meanwhile, elevation
in the creatine kinase and serum creatinine lev-
el were reported in some patients, largely those
infected with MERS-CoV158- 160. Furthermore, a
large number of COVID-19 cases with gastroin-
testinal symptoms such as extreme diarrhea have
been recorded in numerous laboratories, imply-
ing that the virus is replicating in the digestive
system and viral particles are shed via stool162,16 3.
In addition, the sheer volume of ACE2 recep-
tors in the bile duct relative to the alveolar cells
contributes to the hypothesis that infection with
SARS-CoV or SARS-CoV-2 causes serious liv-
er damage164 -166. SARS-CoV-2 infection has also
been linked to neurological symptoms in several
recent studies. Moreover, in some cases, SARS-
CoV-2-RNA has been reported in cerebrospinal
uid167,16 8. The involvement of ACE2 receptors in
the central nervous system (CNS) has been con-
nected to neurological symptoms such as stroke,
polyneuropathy, acute encephalitis, anosmia, age-
usia169, 170, and brain inammation associated with
COV ID-19168 ,171. In previous studies, patients with
COVID-19 were evaluated for symptoms such as
anosmia and dysgeusia, and a high percentage
(approximately 75%) revealed alteration in the
senses of smell and taste172 ,173.
Cerebral arteriopathies and ischemic vascular
stroke have been attributed as the direct effects
of SARS-CoV-2 on ACE2 receptors of endothelial
cells and as indirect effects of misdirected host
immune response174. Neuro-opthalmic manifes-
tations of COVID-19 syndrome are also increas-
ingly recognized175,176 . SARS-CoV-2 has recently
been shown to infect different cells of the renal
system, including tubular epithelial cells and
podocytes, through direct tropism and indirect
action by induction of cytokine storm and other
mechanisms, resulting in a variety of renal ab-
normalities, including acute kidney damage, and
higher mortality177- 180. Several studies have linked
COVID-19 to impaired kidney function during
the course of COVID-19 progression. In seriously
ill patients with COVID-19, numerous renal dis-
orders such as proteinuria, hematuria, and acute
kidney failure (AKF) have been identied181-184.
According to current observational evidence,
AKF is one of the most important causes of ill-
ness and death in SARS-CoV-2 patients, second
only to ARDS185.
Secondary bacterial and fungal infections ob-
served in patients infected with COVID-19 have
been implicated to further complicate the severity
of the disease, constituting an important factor,
especially for high-risk patients186,18 7. Even the
microbiota, such as the bacterial microbiome,
virome, and fungal microbiome, are found to af-
fect the natural course of SARS-CoV-2 infection,
along with comorbidities such as diabetes and hy-
pertension188-19 0. Additional manifestations under-
stated during the COVID-19 epidemic, including
psychological illnesses such as depression, anxi-
ety, and sleep disorders191 and skin disorders such
as urticaria, rashes, erythema, and acro-ischemic
lesions, also should be considered192. All the
above COVID-19 manifestations are either the di-
rect results of SARS-CoV-2 multiplication or of
the indirect hyperinammatory condition known
as macrophage activation syndrome or cytokine
storm. This syndrome leads to increased produc-
tion of IL6, IL7, and TNF-alpha and inammato-
ry chemokines such as CCL2, CC13, and CXC10,
as well as elevated amounts of serum ferritin,
D-dimer, and chronic reactive protein; however,
evidence for inammasome activation is not pres-
ent as IL1β production is not elevated193-195.
Chronic cough is common in long COVID
(post-COVID syndrome) after SARS-CoV-2 in-
fection and may be associated with the vagal
sensory neurons and/or neuroinammatory re-
sponse. Mechanisms of post-COVID-19 chronic
cough and optimal management are still unclear.
New anti-inammatory drugs or neuromodula-
tors (gabapentin or opioids) could be considered
for treatment; however, randomized studies are
highly recommended to analyze the safety and
efcacy of these potential treatments196. Gunst
et al197 have performed a randomized trial to un-
derstand the safety and efcacy of TMPRSS2
inhibitors in COVID-19 patients and found that
TMPRSS2 inhibitors or camostat mesylate may
be effective in the early phase of the disease and
lowers the risk of disease progression; howev-
er, this treatment is not effective for severely ill
and hospitalized patients. According to another
recent study, saliva and nasal swabs offered the
best diagnostic performance and may be used as
an alternative specimen collection method for di-
agnosing SARS-CoV-2 infection198,199.
An overview of comparative clinical mani-
festations in patients with MERS, SARS, and
COVID-19 is presented in Table II.
Pathological studies on SARS-CoV-2 demon-
strate increased inltration into an infected per-
son’s lung tissues154,206 . Viral particles/macro-
phages/inammatory cells have been identied in
Comparative review of three human coronaviruses
7171
the bronchoalveolar uid (BALF) of COVID-19
patients103,207,208. SARS-CoV-2 also targets both
types of pneumocytes (I and II), similar to SARS-
CoV20 9. Monocyte-derived macrophages are pres-
ent at high levels in the BALF, constituting 80%
of the inltrated cells observed in severely sick
patients210. Various forms of activation of mono-
cyte-originating macrophages have been noted211.
ACE2 receptor expression on the surface revealed
that entry-binding receptors for SARS-CoV-2
could be detected in alveolar macrophages, indi-
cating the possibility of this path in the entry of
SARS-CoV-2. From these ndings, there is suf-
cient evidence that monocytes are essential in
the cytokine storm and lung pathology103,212 . The
pro-inammatory classically activated phenotype
(M1) identied in wound-healing can activate the
phenotype (M2) that leads to inammatory tissue
injuries and the development of brotic lesions in
ARDS patients213, 214. These results indicate fea-
tures of SARS research ndings that are common
to those observed for SARS-CoV-2.
Studies211,214 evaluating chemokines and cy-
tokines in SARS-CoV-2 infection would aid in
clarifying the full cytokine proles of patients
with severe infection during the acute phase. This
would help elucidate the pathogenic mechanisms
that result in worse health outcomes. SARS-
CoV-2 patients demonstrate increased concen-
tration of pro-inammatory cytokines, including
IL1β, IL-2, IL-6, IL-8, IL-17, TNF-α, IP-10, MCP-
1, GM-CSF, and G-CSF215 , which may be attribut-
ed to Th-1-cell responses216. Signicant cytokine
elevation was identied in patients with severe
symptoms, and Th-1-cell and Th-2-cell-related
cytokines were detected simultaneously217. Pre-
vious investigations conrmed that the increased
level of certain pro-inammatory cytokines (e.g.,
IFN-γ, IL1β, IL-6, IL-12) in serum positively
correlated with severe lung damage and pulmo-
nary inammation in SARS-CoV patients218. T
cell-related CD molecules and lymphocyte levels
showed a negative correlation with changes in cy-
tokines in patients with SARS-CoV-2 infection.
Therefore, there is a potential correlation between
cytokine storms and adaptive immunity. In pa-
tients with mild symptoms, lymphocyte levels
are usually normal during the convalescent phase
and are undetectable later as the infection pro-
gresses219, 220. In acute stages in mild cases, lym-
phocyte elevation was not linked to elevation in
cytokines. This could be due to cellular immune
response initiation that tends to accelerate viral
clearance during the early phases, inhibiting cy-
tokine production through innate immune activa-
tion, thus alleviating disease severity221. In severe
cases, cytokine hyperactivation during the acute
phase of SARS-CoV-2 resulted in dysregulated
systemic disease inammation and deterioration,
as shown by CRP, ferritin, D-dimers, and procal-
citonin elevation217 ,22 0. Cytokine storms generate
pathogenic effects instead of protective effects
against SARS-CoV-2 infection of host cells222-225.
Excessive cytokine and chemokine activation by
macrophages has similar outcomes, as reported
in macrophage activation syndrome (MAS) and
hemophagocytic lymphohistiocytosis. In particu-
lar, many proinammatory cytokines, IL-1, IL-6,
and TNF-α, are involved in COVID-19 pathogen-
esis226. Single-cell RNA sequencing has revealed
Table II. Comparative analysis of clinical manifestations in patients with MERS, SARS, and COVID-19.
MERS-CoV SARS-CoV SARS-CoV-2
Clinical symptoms/manifestations 8,157,200 8,157,160 8,201–205
Fever 63.5-83.5% 93-97.6% 67.7-90.8%
Cough 51-70% 49-59% 45.5-62.9%
Fatigue 21-35% 31.2% 23.3-38.1%
Dyspnoea 51% 32% 21-40%
Sputum 22-43% 16-27% 21.2-37.2%
Sore throat ≤ 25% 11-25% 14.6-31.0%
Headache 11-20% 30-46% 7.9-15.2%
Gastrointestinal symptoms
(like diarrhoea) ≤ 30% ≤ 32% 3-17%
Bilateral pneumonia N.A. N.A. 58.2-81.0%
Acute respiratory dist ress syndrome (ARDS) 20–30% 20% 18–30%
Neurological manifestations such as stroke 17.4 % N.A. 17- 30%
Acute kidney injury 41–50% 7% 3-20%
N.A.* (Not Available).
A.A. Rabaan, A.A. Mutair, Z.A. Alawi, S. Alhumaid, M.A. Mohaini, J. Aldali, R. Tirupathi, et al
7172
potent interactions between immune and epithe-
lial cells, with inammatory macrophages that
express multiple cytokines in samples with criti-
cal COVID-19 conditions130. Amongst these, IL-8
and IL-6 are detected at elevated concentrations
in individuals with critical or severe COVID-19.
These high levels of both cytokines indicate lym-
phocytopenia that predicts disease progression227.
During the initial stages of SARS-CoV-2 infec-
tion, the lymphocyte count decreases in patients
with severe symptoms associated with a dramat-
ic decrease in NK cells, B cells, CD4+, CD8+ T,
and CD3+ cells228. Mild cases of SARS-CoV-2
infection demonstrate a moderate increase in
these lymphocytes220, 221. These ndings further
reveal that lymphocytes can reach comparable
levels along with only slight variation in notice-
able lymphocyte count214, 221. Changes in adaptive
immunity occur due to imbalanced activation of
Th1/Th2, and alteration of T lymphocyte func-
tion causes adverse effects on host cells, wors-
ening the course of disease. Reduced function of
CD8+ T cells serves as a predictor of the severity
of SARS-CoV-2 infection208. Moreover, the pro-
duction of endogenous IFNβ in the nasal mucosa
of critical patients can be considered a prognostic
tool of IFN therapy for managing COVID-19, as it
predicts clinical outcomes229. The eosinophil level
in COVID-19 patients was related to the reduced
anticoagulant effect and therefore may be consid-
ered for determining the prophylactic anticoagu-
lant administration strategy2 30. The high eosino-
Figure 1. Coronaviruses infect human lung epithelial cells via the ACE2 receptor. TLR3/7 and MAVS, endosomal and cyto-
plasmic sensors, respectively, are activated by viral R NA. Interferon Regulator y Factors (IR Fs) and NFkB are activated by these
receptors, which trigger the cytokine storm by inducing inammatory cytokines such as interferons (IFNs) and proinammatory
cytokines. These lead to the recruitment of many immune cells into the lungs and cause a hypersecretion of cytokines i.e., cyto-
kine storm by activated immune cells. The cytokine storm causes cytotoxic and immunological disruption in host cells, as well as
ARDS and other clinical manifestations. Moreover, to prime the adaptive immunity, dendritic cells (DCs) sample antigen moves
towards lymphoid organs. After recognising antigen on DCs or infected cells, CD8+ T cells cause apoptosis.
Comparative review of three human coronaviruses
7173
phil count is also associated with lower activity
of anti-factor Xa230. Liu and coworkers have also
suggested the importance of familial cluster (FC)
and non-familial (NF) patients during the treat-
ment of COVID-19 patients231.
An overview on pathology and pathobiology of
SARS-CoV-2 is presented in Figu re 1.
Immunocompetent infected individuals usual-
ly present mild manifestations or become asymp-
tomatic232. However, the immunocompromised,
the elderly, and individuals with underlying con-
ditions such as cardiovascular disease, diabetes,
and cancer, develop severe symptoms and clinical
disease134. A recent study linked renal, cardiovas-
cular, and respiratory diseases with greater ICU
admission and fatality rate in COVID-19 patients.
Cancer and diabetes conditions also showed a
strong correlation with severe disease outcomes
in SARS-CoV-2 infected patients. Male patients
and older people showed higher ICU admission
and mortality. The risk of COVID-19 was dra-
matically increased in older patients233. In another
report, hypertension, cardiovascular disease, and
diabetes were closely associated with severe out-
comes in COVID-19-infected individuals across
all age groups. In contrast to elderly patients,
young patients exhibited a varied prevalence
of cardiovascular comorbidities234. The De Ri-
tis ratio has been associated with poor survival
in SARS-COV-2 infected patients and is an im-
portant prognostic factor. Moreover, the De Ritis
ratio on admission was signicantly associated
with in-hospital mortality in COVID-19 patients.
However, since the sample size in the clinical
study was considerably smaller, additional inves-
tigations are needed to validate the ability of this
parameter to independently predict death in hos-
pitalized patients235. Obesity is strongly associat-
ed with immune cells, including MAIT and NK
cells236-238. Hence, it may be considered a factor
for increased risk of severe COVID-19. Popkin
et al239 have recently conducted a meta-analysis
and found 48% higher mortality in COVID-19
patients with obesity. In a recent cohort study,
Gao et al240 found a linear increase in the risk
of severity in patients with COVID-19 leading
to hospitalization and death with an increase in
body mass index. This excess risk was observed
particularly in younger people and also in black
individuals. According to another cross-sectional
study, Lega et al241 found that COVID-19 patients
with a severe psychiatric disorder (schizophre-
nia and others) died at a younger age compared
to those without any psychiatric disorder. Addi-
tionally, the vulnerability of COVID-19 patients
with psychiatric disorders may reduce the chance
of recovery. Pregnant women were affected more
by COVID-19, particularly in the second wave
of infections, which may be associated with the
emergence of increasing numbers of pathogenic
strains242. Moreover, people with physical disabil-
ities are particularly at risk and need additional
support from mental health services243. In a simi-
lar report, authors have mentioned that older peo-
ple with disabilities have been neglected during
the COVID-19 pandemic244.
Conclusions
Three highly pathogenic coronaviruses (SARS-
CoV, MERS-CoV, and SARS-CoV-2) have been
reported in humans in the last two decades. Al-
though these CoVs exhibit several similar features
in their infection process, distinctive features and
characteristics are observed in the immunopa-
thology and clinical outcomes of each. Recurrent
outbreaks of infectious and pathogenic strains of
CoVs have posed a signicant burden and danger
to humankind, such as the current COVID-19
pandemic that has resulted in an unprecedented
crisis with devastating social, health, and eco-
nomic impacts worldwide. All three CoVs share
immunological aspects that affect pathological
characteristics. The viral agents undergo rep-
lication in the host immune cells and set off an
innate immune response that leads to induction
of pro-inammatory cells and cytokines. This cy-
tokine storm can be life-threatening. Finally, the
body responds by producing protective antibodies
that clear the viral infection and also confer im-
munity against future infection with the same vi-
rus. In vit ro studies have been particularly helpful
for understanding the immunological and patho-
logical aspects of the viruses and in conducting
drug trials against these agents. However, there is
a need to advance clinical research since these vi-
ral agents can undergo further mutations and may
give rise to viral species with enhanced patho-
genicity in the future. Collaborative and intense
efforts by scientists worldwide have resulted in
advanced discoveries related to many aspects of
SARS-CoV-2 and COVID-19. In addition, eluci-
dating the immunopathology and clinical features
of CoVs will help in developing better and more
effective drugs, medicines, and vaccines to count-
er the emergence and re-emergence of pathogenic
CoVs. Prospective outcomes from clinical inves-
A.A. Rabaan, A.A. Mutair, Z.A. Alawi, S. Alhumaid, M.A. Mohaini, J. Aldali, R. Tirupathi, et al
7174
tigations of different vaccines and antiviral can-
didates provide hope to end this pandemic soon.
Continuous research efforts to better understand
the pathogenesis, molecular biology, and immu-
nological characteristics of SARS-CoV-2 and oth-
er CoVs will help to stem the tide of the ongoing
COVID-19 pandemic and formulate prevention
plans for future pandemics.
Conflicts of interest
All authors declare that there exist no commercial or
nancial relationships that could, in any way, lead to a
potential conict of interest.
Author Contributions
All the authors substantially contributed to the con-
ception, compilation of data, checking and approving
the nal version of the manuscript, and agree to be ac-
countable for its contents.
Acknowledgements
All the authors acknowledge and thank their respec-
tive Institutes and Universities.
Funding
This is a review article written by its authors and re-
quired no substantial funding to be stated.
References
1) Shaw PD, Patel N, Patil S, Samuel R, Khanna P,
Prajapati B, Sharun K, Tiwari R, Dhama K, Na-
tesan S. Comparative evaluation of the origin,
evolution, transmission, diagnosis, and vaccine
development of three highly pathogenic human
coronaviruses (Sars-cov, mers-cov and sars-
cov-2). J Exp Biol Agric Sci 2020; 8: S103-S113.
2) Coleman CM, Frieman MB. Coronaviruses: Im-
portant Emerging Human Pathogens. J Virol
2014; 88: 5209-5212.
  3)  WHO.  WHO  Director-General’s  remarks  at  the 
media  brieng  on  2019-nCov.  Who  [Internet] 
2020; 1-May-2020.
4) Rabaan AA, Al-Ahmed SH, Haque S, Sah R, Ti-
wari R, Malik YS, Dhama K, Yatoo MI, Bonilla-Al-
dana DK, Rodriguez-Morales AJ. SARS-CoV-2,
SARS-CoV, and MERS-CoV: A comparative over-
view. Infez Med 2020; 28: 174-184.
5) Rabaan AA, Al-Ahmed SH, Sah R, Alqumber MA,
Haque S, Patel SK, Pathak M, Tiwari R, Yatoo MI, 
Haq AU, Bilal M, Dhama K, Rodriguez-Morales
AJ. MERS-CoV: epidemiology, molecular dynam-
ics, therapeutics, and future challenges. Ann Clin
Microbiol Antimicrob 2021; 20: 1-14.
  6)  Dhama K, Patel SK,  Sharun K, Pathak M, Tiwari 
R, Yatoo MI, Malik  YS, Sah R, Rabaan AA,Pan-
war  PK,  Singh  KP,  Michalak  I,  Chaicumpa  W, 
Martinez-Pulgarin DF, Bonilla-Aldana DK, Rodri-
guez-Morales AJ.SARS-CoV-2 jumping the spe-
cies barrier: Zoonotic lessons from SARS, MERS
and recent advances to combat this pandemic
virus. Travel Med Infect Dis 2020; 37: 101830.
7) Berger A, Drosten C, Doerr HW, Stürmer M, Preis-
er W. Severe acute respiratory syndrome (SARS)
- Paradigm of an emerging viral infection. J Clin
Virol 2004; 29: 13-22.
  8)  Zhu Z, Lian X, Su X, Wu W, Marraro GA, Zeng Y. 
From SARS and MERS to COVID-19: A brief sum-
mary and comparison of severe acute respiratory
infections caused by three highly pathogenic hu-
man coronaviruses. Respir Res 2020; 21: 1-14.
9) Peiris JSM, Chu CM, Cheng VCC, Chan KS, Hung
IFN,  Poon  LLM,  Law  KI,  Tang  BSF,  Hon  TY W, 
Chan CS, Chan KH, Ng JSC, Zheng BJ, Ng WL,
Lai  RWM,  GuanY,  YuenKY.Clinical  progression 
and  viral  load  in a community outbreak  of  coro-
navirus-associated SARS pneumonia: A prospec-
tive study. Lancet 2003; 361: 1767-1772.
10) Prompetchara E, Ketloy C, Palaga T. Immune
responses in COVID-19 and potential vaccines:
Lessons learned from SARS and MERS epidem-
ic. Asian Pacic J Allergy Immunol 2020; 38: 1-9. 
11) Leung CHC, Gomersall CD. Middle East respiratory
syndrome. Intensive Care Med 2014; 40:1015-1017.
12)  Huh JE,  Han  S,  Yoon T. Data mining of corona-
virus: SARS-CoV-2, SARS-CoV and MERS-CoV.
BMC Res Notes 2021; 14: 1-6.
13)  Chong  ZX,  Liew  WPP,  Ong  HK,  Yong  CY,  Shit 
CS, Ho WY, Ng SYL, Yeap SK. Current diagnostic 
approaches to detect two important betacoronavi-
ruses: Middle East respiratory syndrome corona-
virus (MERS-CoV) and severe acute respiratory
syndrome coronavirus 2 (SARS-CoV-2). Pathol -
Res Pract 2021; 225: 153-565.
14) Weiss SR. Forty years with coronaviruses. J Exp
Med 2020; 217: e20200537.
15)  Suaifan GARY, Alkhawaja BA MA. RNA Coronavi-
ruses’ Outbreaks: Recent Progress on the SARS-
CoV-2 Pandemic Diagnostic Tests, Vaccination
and Therapeutics. Mini Rev Med Chem 2021;
2021: 1-17.
16)  Tiwari R, Dhama K, Sharun K, Iqbal Yatoo M, Malik
YS, Singh R, Michalak I, Sah R, Bonilla-Aldana DK, 
Rodriguez-Morales AJ. COVID-19: animals, veteri-
nary and zoonotic links. Vet Q 2020; 40: 169-182.
17) Dhama K, Khan S, Tiwari R, Sircar S, Bhat S, Ma-
lik  YS,  Singh  KP,  Chaicumpa  W,  Bonilla-Aldana 
DK, Rodriguez-Morales AJ. Coronavirus disease
2019-COVID-19. Clin Microbiol Rev 2020; 33: 1-48.
18)  Meo  SA,  Alhowikan  AM,  Khlaiwi  TAL,  Meo  IM, 
Halepoto DM, Iqbal M, Usmani AM, Hajjar W,
Ahmed N. Novel coronavirus 2019-nCoV: Preva-
lence, biological and clinical characteristics com-
parison with SARS-CoV and MERS-CoV. Eur Rev
Med Pharmacol Sci 2020; 24: 2012-2019.
Comparative review of three human coronaviruses
7175
19)  Gorbalenya AE, Baker SC, Baric RS, de Groot RJ, 
Drosten C, Gulyaeva AA, Haagmans BL, Lauber
C, Leontovich AM, Neuman BW, Penzar D,Perl-
man S, Poon  LLM,  Samborskiy DV , Sidorov IA, 
Sola I, Ziebuhr J.The species Severe acute respi-
ratory syndrome-related coronavirus: classifying
2019-nCoV and naming it SARS-CoV-2. Nat Mi-
crobiol 2020; 5: 536-544.
20) Noor R. A comparative review of pathogene-
sis and host innate immunity evasion strategies
among the severe acute respiratory syndrome
coronavirus 2 (SARS-CoV-2), severe acute respi-
ratory syndrome coronavirus (SARS-CoV) and
the Middle East respiratory syndrome coro. Arch
Microbiol 2021; 203: 1943-1951.
21) Chan JFW, Lau SKP, To KKW, Cheng VCC, Woo
PCY, Yue KY. Middle East Respiratory syndrome 
coronavirus: Another zoonotic betacoronavirus
causing  SARS-like  disease.  Clin  Microbiol  Rev 
2015; 28: 465-522.
22) WHO. WHO COVID-19 Dashboard - Up to date data
on pandemic. WHO Heal Emerg Dasboard 2021.
23)  To KKW, Tsang OTY, Leung WS, Tam AR, Wu TC, 
Lung DC, Yip CCY, Cai JP, Chan JMC, Chik TSH, 
Lau DPL, Choi, CYC, Chen LL, Chan WM, Chan 
K-H, Daniel J, Chin-Ki A, Poon RWS, Luo CT,
Cheng VCC, Chan JFW, Yuen KY. Temporal pro-
les of viral load in posterior oropharyngeal saliva 
samples and serum antibody responses during
infection by SARS-CoV-2: an observational co-
hort study. Lancet Infect Dis 2020; 20: 565-574.
24)  Lam TTY, Shum MHH, Zhu HC, Tong YG, Ni XB, 
Liao YS, Wei W, Cheung  WYM, Li WJ, Li LF, Ji-
ang BG, Wei W, Yuan TT, Zheng K, Cui XM, Li J, 
Pei GQ, Qiang X, Cheung WY, Cao W. Identifying 
SARS-CoV-2-related coronaviruses in Malayan
pangolins. Nature 2020; 583: 282-285.
25) Ioannidis JPA. The infection fatality rate of
COVID-19 inferred from seroprevalence data.
medRxiv 2020; 2020.05.13.20101253.
26) Abdallat MM, Abroug F, Al Dhahry SHS, Alha-
jri  MM,  Al- Hakeem  R,  Al  Hosani  FI,  Al  Qasrawi 
SMA, Al-Romaihi HE, Assiri A, Baillie JK,Ben
Embarek  PK,  Ben Salah  A,  Blümel B,  Briese T, 
Buchholz U, Cognat SBF, Defang GN, De La Roc-
que S, Donatelli I, Drosten C, Drury PA, Eremin
SR,Ferguson NM, Fontanet A, Formenty PBH,-
Fouchier  RAM,  Gao  CQ,  Garcia  E,  Gerber  SI, 
Guery B, Haagmans BL, Haddadin AJ, Hardiman
MC, Hensley LE, Hugonnet SAL, Hui DSC, Isla
N, Karesh  WB, Koopmans  M,  Kuehne A,  Lipkin 
WI, Ma AR, Malik M, Manuguerra JC, Memish Z, 
Mounts AW, Mumford E, Opoka L, Osterhaus A, 
John Oxenford C, Pang J, Pebody R, Peiris JSM,
Jay Plotkin B, Poumerol G, Reusken C, Rezza G, 
Roth CE, Shindo N, Shumate AM, Siwula M, Slim
A, Smallwood  C, van der Werf S, Van  Kerkhove 
MD,  Zambon  M.  State  of  knowledge  and  data 
gaps of middle east respiratory syndrome coro-
navirus (MERS-CoV) in humans. PLoS Curr 2013;
5: 1-36.
27)  Rabaan AA, Al-Ahmed SH, Sah R, Tiwari R, Ya-
too  MI,  Patel  SK,  Pathak  M,  Malik  YS,  Dhama 
K, Singh KP, Bonilla-Aldana K, Haque S, Marti-
nez-Pulgarin DF, Rodriguez-Morales AJ, Leblebi-
cioglu H. SARS-CoV-2/COVID-19 and advances
in developing potential therapeutics and vaccines
to counter this emerging pandemic. Ann Clin Mi-
crobiol Antimicrob 2020; 19: 1-37.
28)  Sharun K, Tiwari R, Iqbal Yatoo M, Patel SK, Na-
tesan  S,  Dhama  J,  Malik YS, Harapan H, Singh 
RK, Dhama K. Antibody-based immunotherapeu-
tics and use of convalescent plasma to counter
COVID-19: advances and prospects. Expert Opin
Biol Ther 2020; 20: 1033-1046.
29)  Iqbal Yatoo M, Hamid Z, Parray OR, Wani AH, Ul 
Haq A, Saxena A, Patel SK, Pathak M, Tiwari R, 
Malik YS, Sah R, Rabaan AA, Rodriguez Morales 
AJ, Dhama K. COVID-19 - Recent advancements
in identifying novel vaccine candidates and cur-
rent status of upcoming SARS-CoV-2 vaccines.
Hum Vaccines Immunother 2020; 16: 2891-2904.
30)  Iqbal  Yatoo  M,  Hamid  Z,  Rather  I,  Nazir  QUA, 
Bhat RA, Ul Haq A, Magray SN, Haq Z, Sah R,
Tiwari R, Natesan S, Bilal M, Harapan H, Dha-
ma K. Immunotherapies and immunomodulatory
approaches in clinical trials - a mini review. Hum
Vaccines Immunother 2021; 17: 1897-1909.
31) WHO Covid-19. Draft landscape of COVID-19
candidate vaccines. WHO 2020; 3.
32) Dallocchio RN, Dessì A, De Vito A, Delogu G, Ser-
ra PA, Madeddu G. Early combination treatment
with existing HIV antivirals: An effective treatment
for COVID-19? Eur Rev Med Pharmacol Sci 2021;
25: 2435-2448.
33) Wang D, Hu B, Hu C, Zhu F, Liu X, Zhang J, Wang
B, Xiang H, Cheng Z, Xiong Y, Zhao Y, Li Y, Wang 
X, Peng, Z. Clinical Characteristics of 138 Hospi-
talized Patients with 2019 Novel Coronavirus-In-
fected Pneumonia in Wuhan, China. JAMA - J Am
Med Assoc 2020; 323: 1061-1069.
34)  Guan WJ, Ni ZY, Hu Y, Liang WH, Ou CQ, He JX, 
Liu L, Shan H, Lei CL, Hui DSC,Du B, Li LJ, Zeng
G, Yuen KY , Chen RC , Tang CL , Wang T, Chen 
PY, Xiang J, Li SY, Wang JL, Liang ZJ, Peng YX, 
Wei L, Liu Y, Hu YH, Peng P, Wang  JM,  Liu  JY, 
Chen Z, Li G, Zheng ZJ, Qiu SQ, Luo J, Ye CJ, 
Zhu SY, Zhong  NS.China Medical  Treatment Ex-
pert Group for Covid-19. Clinical characteristics of
coronavirus disease 2019 in China. N Engl J Med
2020; 382: 1708-1720.
35) Wu JT, Leung K, Leung GM. Nowcasting and
Forecasting the Potential Domestic and Interna-
tional  Spread  of the 2019-nCoV Outbreak  Origi-
nating in Wuhan, China: A Modeling Study. Lan-
cet 2020; 395: 689-697.
36)  Xiao  K ,  Zhai  J,  Feng  Y,  Zhou  N,  Zhang  X,  Zou 
JJ, Li N,  Guo  Y, Li X, Shen X,  Zhang  Z, Shu F, 
Huang W, Li Y, Zhang Z, Chen RA, Wu YJ, Peng 
SM,  Huang  M,  Xie  WJ,  Cai  QH,  Ho u  FH,  Chen 
W, Xiao L, Shen Y.Isolation of SARS-CoV-2-relat-
ed coronavirus from Malayan pangolins. Nature
2020; 583: 286-289.
37)  Ruan Q, Yang K, Wang W, Jiang L, Song J. Clinical 
predictors of mortality due to COVID-19 based on
an analysis of data of 150 patients from Wuhan,
China. Intensive Care Med 2020; 46: 846-848.
38) Wu Z, McGoogan JM. Characteristics of and Im-
portant Lessons from the Coronavirus Disease
A.A. Rabaan, A.A. Mutair, Z.A. Alawi, S. Alhumaid, M.A. Mohaini, J. Aldali, R. Tirupathi, et al
7176
2019 (COVID-19) Outbreak in China: Summary of 
a Report of 72314 Cases from the Chinese Center
for Disease Control and Prevention. JAMA - J Am
Med Assoc 2020;323:1239-1242.
39)  Wu A, Peng Y, Huang B, Ding X, Wang X, Niu P, 
Meng J, Zhu Z, Zhang Z, Wang J, Sheng J, Quan 
L, Xia Z, Tan W, Cheng G, Jiang T.Genome Com-
position and Divergence of the Novel Coronavirus
(2019-nCoV) Originating in China. Cell Host Mi-
crobe 2020; 27: 325-328.
40) Hamming I, Timens W, Bulthuis MLC, Lely AT, Na-
vis GJ, van Goor H. Tissue distribution of ACE2
protein, the functional receptor for SARS corona-
virus. A rst step in understanding  SARS patho-
genesis. J Pathol 2004; 203: 631-637.
41)  Raj VS, Mou H, Smits SL, Dekkers DH, Müller MA, 
Dijkman R, Muth D, Demmers JA, Zaki A, Fouchi-
er RA, Thiel V. Drosten C, Rottier PJM, Osterhaus
ADME, Bosch BJ, Haagmans BL.Dipeptidyl pepti-
dase 4 is a functional receptor for the emerging hu-
man coronavirus-EMC. Nature 2013; 495: 251-254.
42)  Lu R,  Zhao  X,  Li J, Niu P, Yang B, Wu H, Wang 
W, Song H, Huang B, Zhu N,Bi Y, Ma X, Zhan F, 
Wang L, Hu T, Zhou H, Hu Z, Zhou W, Zhao L,
Chen J, Meng Y, Wang J, Lin Y, Yuan J, Xie Z, Ma 
J, Liu WJ, Wang D, Xu W, Holmes EC, Gao GF,
Wu G, Chen W, Shi W, Tan W.Genomic charac-
terisation and epidemiology of 2019 novel coro-
navirus: implications for virus origins and receptor
binding. Lancet 2020; 395: 565-574.
43)  Zhou P, Yang X Lou, Wang XG,  Hu B, Zhang L, 
Zhang W, Si HR, Zhu Y, Li B, Huang CL,Chen HD, 
Chen J, Luo  Y, Guo  H, Jiang RD, Liu  MQ,  Chen 
Y, Shen XR, Wang X, Zheng XS, Zhao K, Chen 
QJ, Deng F, Liu LL, Yan B, Zhan FX,  Wang YY, 
Xiao GF, Shi ZL.A pneumonia outbreak associat-
ed with a new coronavirus of probable bat origin.
Nature 2020; 579: 270-273.
44)  Ozono S, Zhang Y, Ode H, Sano  K, Tan TS, Imai 
K, Miyoshi K, Kishigami S,  Ueno  T, Iwatani  Y,Su-
zuki  T,  Tokunaga  K.SARS-CoV-2  D614G  spike 
mutation increases entry efciency with enhanced 
ACE2-binding afnity. Nat Commun 2021; 12: 1-9.
45)  Wu F, Zhao  S,  Yu B,  Chen  YM,  Wang  W, Song 
ZG,  Hu  Y,  Tao  ZW,  Tian  JH,  Pei  YY,Yuan  ML, 
Zhang YL, Dai FH, Liu Y, Wang QM, Zheng JJ, Xu 
L, Holmes EC, Zhang YZ.A  new coronavirus  as-
sociated with human respiratory disease in China.
Nature 2020;579: 265-269.
46)  Walls AC, Park Y-J, Tortorici MA, Wall A, McGuire 
AT, Veesler D. Structure, Function, and Antigenic-
ity  of  the  SARS- CoV-2  Spike  Glycoprotein.  Cell 
2020; 181: 281-292.
47) Wrapp D, Wang N, Corbett KS, Goldsmith JA,
Hsieh CL, Abiona O, Graham BS, McLellan
JS.  Cryo-EM  structure  of  the  2019-nCoV  spike 
in the prefusion conformation. bioRxiv 2020;
10.1101/2020.02.11.944462.
48) Benvenuto D, Giovanetti M, Ciccozzi A, Spoto S,
Angeletti S, Ciccozzi M. The 2019-new coronavi-
rus epidemic: Evidence for virus evolution. J Med
Virol 2020; 92: 455-459.
49)  Epidemiology Working Group for NCIP Epidemic 
Response, Chinese Center for Disease Control
and Prevention. The epidemiological characteris-
tics of an outbreak of 2019 novel coronavirus dis-
eases (COVID-19) in China. Zhonghua Liu Xing
Bing Xue Za Zhi 2020; 41: 145-151.
50)  D’Cruz RJ, Currier AW, Sampson VB. Laboratory 
Testing Methods for Novel Severe Acute Respi-
ratory Syndrome-Coronavirus-2 (SARS-CoV-2).
Front Cell Dev Biol 2020; 8: 468.
51)  Wang  WK,  Ch en  SY, Li u  IJ,  Kao  CL,  Chen  HL, 
Chiang BL, Wang JT, Sheng WH, Hsueh PR,
Yang CF, Yang PC, Chang SC. Temporal relation-
ship of viral load, ribavirin, interleukin (IL)-6, IL-8, 
and clinical progression in patients with severe
acute respiratory syndrome. Clin Infect Dis 2004;
39: 1071-1075.
52)  Chen W, Xu Z, Mu J, Yang L, Gan H, Mu F, Fan B, 
He B, Huang S, You B, Yang Y,TangX, Qiu L, Qiu 
Y, WenJ, FangJ, Wang J. Antibody response and 
viraemia during the course of severe acute respi-
ratory syndrome (SARS)-associated coronavirus
infection. J Med Microbiol 2004; 53: 435-438.
53)  Choi  HM,  Moon  SY,  Yang  HI,  Kim  KS.  Under-
standing viral infection mechanisms and patient
symptoms for the development of covid-19 thera-
peutics. Int J Mol Sci 2021; 22: 1-25.
54) Hung IFN, Cheng VCC, Wu AKL, Tang BSF, Chan
KH, Chu CM, Wong MML, Hui WT, Poon LLM, Tse
DMW,  ChanKS,  WooPCY,  LauSKP,  PeirisJSM, 
YuenKY.Viral  loads  in  clinical  specimens  and 
SARS manifestations. Emerg Infect Dis 2004; 10:
1550-1557.
55) Chu CM, Poon LLM, Cheng VCC, Chan KS, Hung
IFN, Wong MML, Chan KH, Leung WS, Tang BSF,
Chan VL,NgWL, SimTC, NgPW, LawKI, TseDMW,
PeirisJSM, YuenKY. Initial viral load  and the out-
comes of SARS. Cmaj 2004; 171: 1349-1352.
56) Fajnzylber J, Regan J, Coxen K, Corry H, Wong
C, Rosenthal A, Worrall D, Giguel F, Piechoc-
ka-Trocha  A,  Atyeo  C,  Fischinger  S,  Chan  A, 
Flaherty KT, Hall K, Dougan M, Ryan ET, Gilles-
pie E, Chishti R, Li Y, Jilg N, Hanidziar D, Baron 
RM, Baden L, Tsibris AM, Armstrong KA, Kuritz-
kes  DR,  Alter  G,  Walker  BD,  Yu  X,  Li  JZ,  Betty 
Abayneh BA, Allen P, Antille D, Balazs A, Bals J,
Barbash  M, Bartsch Y, Boucau J, Boyce S, Bra-
ley J,  Branch  K, Broderick K,  Carney  J,  Cheva-
lier J, Choudhary MC, Chowdhury N, Daley G,
Davidson S, Desjardins M, Donahue L, Drew D,
Einkauf K, Elizabeth S, Elliman A, Etemad B, Fal-
lon J, Fedirko L, Finn K, Flannery J, Forde P, Gar-
cia-Broncano P, Gettings E, Golan D, Goodman
K,  Grif n  A,  Grimmel  S,  Grinke  K,  Har tana  CA, 
Healy M, Heller H, Henault D, Holland G, Jiang
C, Jordan H, Kaplonek P, Karlson EW, Karpell M, 
Kayitesi C, Lam EC, LaValle V, Lefteri K, Lian X,
Lichterfeld M, Lingwood D, Liu H, Liu J, Lopez K,
Lu Y, Luthern S, Ly NL, MacGowan M, Magispoc 
K, Marchewka J, Martino B, McNamara R, Michell 
A, Millstrom I, Miranda N, Nambu C, Nelson S,
Noone M, Novack L, O’Callaghan C, Ommerborn 
C, Osborn M, Pacheco LC, Phan N, Pillai S, Porto
FA, Rassadkina  Y, Reissis  A,  Ruzicka F, Seiger 
K, Selleck K, Sessa L, Sharpe A, Sharr C, Shin S, 
Singh N, Slaughenhaupt S, Sheppard KS, Sun W,
Sun X, Suschana EL, Talabi O, Ticheli H, Weiss
Comparative review of three human coronaviruses
7177
ST, Wilson V, Zhu A. SARS-CoV-2 viral load is
associated with increased disease severity and
mortality. Nat Commun 2020; 11: 1-9.
57)  Ganesh B, Rajakumar T, Malathi M, Manikandan 
N, Nagaraj J, Santhakumar A, Elangovan A, Ma-
lik YS. Epidemiology and pathobiology of SARS-
CoV-2 (COVID-19) in comparison with SARS,
MERS:  An  updated  overview  of  current  knowl-
edge and future perspectives. Clin Epidemiol
Glob Heal 2021; 10: 100694.
58)  GISAID  Global  Initiative  on  Sharing  All  Inuenza 
Data. Phylogeny of SARS-like betacoronaviruses in-
cluding novel coronavirus (nCoV) 2020; 22: 30494.
59)  Malik YS, Sircar S, Bhat S, Sharun  K,  Dhama  K,
Dadar M, Tiwari R, Chaicumpa W. Emerging novel
coronavirus (2019-nCoV)—current scenario, evo-
lutionary perspective based on genome analysis
and recent developments. Vet Q 2020; 40: 68-76. 
60) Mohapatra RK, Pintilie L, Kandi V, Sarangi AK,
Das  D,  Sahu  R,  Perekhoda  L.  Th e  recent  chal-
lenges of highly contagious COVID-19, causing
respiratory infections: Symptoms, diagnosis,
transmission, possible vaccines, animal models,
and immunotherapy. Chem Biol Drug Des 2020;
96: 1187-1208.
61)  Zhou H, Chen X, Hu T, Li J, Song H, Liu Y, Wang 
P, Liu D, Yang J, Holmes E, HughesA, BiY, ShiW. 
A novel bat coronavirus reveals natural insertions
at the S1/S2 cleavage site of the Spike protein and 
a possible recombinant origin of HCoV-19. bioRx-
iv2020; 10.1101/2020.03.02.974139.
62) Zheng J. SARS-coV-2: An emerging coronavirus
that causes a global threat. Int J Biol Sci 2020; 16:
1678-1685.
63) Kaur N, Singh R, Dar Z, Bijarnia RK, Dhingra N,
Kaur T. Genetic comparison among various coro-
navirus  strains  for  the  identication  of  potential 
vaccine targets of SARS-CoV2. Infect Genet Evol
2021; 89: 104490.
64) Kumar V. Emerging Human Coronavirus Infec-
tions (SARS, MERS, and COVID-19): Where They
Are Leading Us. Int Rev Immunol 2021; 40: 5-53.
65)  Malik  YA.  Proper ties  of  coronavirus  and  SARS-
CoV-2. Malays J Pathol 2020; 42: 3-11.
66) Urra JM, Cabrera CM, Porras L, Ródenas I. Se-
lective CD8 cell reduction by SARS-CoV-2 is as-
sociated with a worse prognosis and systemic
inammation in COVID-19 patients. Clin Immunol 
2020; 217: 108486.
67)  Zhang T, Wu Q, Zhang Z. Probable Pangolin Ori-
gin of SARS-CoV-2 Associated with the COVID-19
Outbreak. Curr Biol 2020; 30: 1346-1351.
68)  Yuan S, Jiang SC, Li ZL. Analysis of Possible In-
termediate Hosts of the New Coronavirus SARS-
CoV-2. Front Vet Sci 2020; 7: 379.
69) Brierley L, Fowler A. Predicting the animal hosts of
coronaviruses from  compositional biases  of spike 
protein and whole genome sequences through ma-
chine learning. PLoS Pathog 2021; 17: e1009149.
70) Ji W, Wang W, Zhao X, Zai J, Li X. Cross-species
transmission  of  the  newly  identied  coronavirus 
2019-nCoV. J Med Virol 2020; 92: 433-440.
71)  Qian  J, Feng Y, Li J. Comments  on  “Cross-spe-
cies transmission of the newly identied coronavi-
rus 2019-nCoV.” J Med Virol 2020; 92: 1437-1439.
72)  Liu Z, Xiao X, Wei  X, Li J, Yang J, Tan H, Zhu J, 
Zhang Q, Wu J, Liu L. Composition and di vergence
of  coronavirus  s pike  proteins  and  host  ACE2  re-
ceptors predict potential intermediate hosts of
SARS-CoV-2. J Med Virol 2020; 92: 595-601.
73)  Andersen KG, Rambaut A, Lipkin WI, Holmes EC, 
Garry RF. The proximal origin of SARS-CoV-2.
Nat Med 2020; 26: 450-452.
74)  Sun P, Qie S, Liu Z, Ren J, Li K, Xi J. Clinical char-
acteristics of hospitalized patients with SARS-
CoV-2 infection: A single arm meta-analysis. J
Med Virol 2020; 92: 612-617.
75)  Zhang  YZ,  Holmes  EC.  A  Genomic Perspective 
on the Origin and Emergence of SARS-CoV-2.
Cell 2020; 181: 223-227.
76) Ramanathan K, Antognini D, Combes A, Paden M,
Zakhary B, Ogino M,  Maclaren  G, Brodie  D. Ge-
nomic characterisation and epidemiology of 2019
novel coronavirus: implications for virus origins and
receptor binding. Lancet 2020; 2020: 19-21.
77) Kandeel M, Ibrahim A, Fayez M, Al-Nazawi M.
From SARS and MERS CoVs to SARS-CoV-2:
Moving toward more biased codon usage in viral
structural and nonstructural genes. J Med Virol
2020; 92: 660-666.
78) Roy A, Guo F, Singh B, Gupta S, Paul K, Chen X,
Sharma NR, Jaishee N, Irwin DM, Shen Y. Base 
Composition and Host Adaptation of the SARS-
CoV-2: Insight From the Codon Usage Perspec-
tive. Front Microbiol 2021; 12: 548275.
79) van Boheemen S, de Graaf M, Lauber C, Beste-
broer  TM,  Raj  VS,  Zaki  AM,  Osterhaus  ADME, 
Haagmans BL, Gorbalenya AE, Snijder EJ,Fouch-
ierRAM.Genomic characterization of a newly dis-
covered coronavirus associated with acute respi-
ratory distress syndrome in humans. MBio 2012;
3: e00473-12.
80)  Meyer  B,  Müller  MA,  Corman  VM,  Reusken 
CBEM,  Ritz  D,  Godeke  GJ,  Lattwein  E,  Kallies 
S,  Siemens  A,  van  Beek  J,DrexlerJF,  MuthD, 
BoschBJ, WerneryU, KoopmansMPG, WerneryR,
Drosten C.Antibodies against MERS coronavirus
in dromedaries, United Arab Emirates, 2003 and
2013. Emerg Infect Dis 2014; 20: 552-559.
81)  Stalin Raj V, Farag EABA, Reusken CBEM, Lam-
ers MM, Pas SD, Voermans J, Smits SL, Oster-
haus ADME, Al-Mawlawi N, Al-Romaihi HE,Alha-
jriMM, El-SayedAM, MohranKA, GhobashyH, Al-
hajriF, Al-ThaniM, Al-MarriSA, El-MaghrabyMM,
KoopmansMPG, HaagmansBL.Isolation of MERS
coronavirus  from dromedary camel, Qatar, 2014. 
Emerg Infect Dis 2014; 20: 1339-1342.
82) Dudas G, Rambaut A. MERS-CoV recombination:
implications about the reservoir and potential for
adaptation. Virus Evol 2016; 2: vev023.
83)  Wang Y, Liu  D, Shi W, Lu R,  Wang  W,  Zhao  Y, 
Deng Y, Zhou  W, Ren H,  Wu  J, Wang Y, Wu G, 
GaoGF, TanaW. Origin and possible genetic re-
combination of the middle east respiratory syn-
drome coronavirus from the rst imported case in 
china: Phylogenetics and coalescence analysis.
MBio 2015; 6: e01280-15.
84)  Kenney SP, Helmy YA, Fawzy M, Elaswad A, So-
bieh A, Shehata AA. The COVID-19 Pandemic: A
Comprehensive Review of Taxonomy, Genetics,
A.A. Rabaan, A.A. Mutair, Z.A. Alawi, S. Alhumaid, M.A. Mohaini, J. Aldali, R. Tirupathi, et al
7178
Epidemiology, Diagnosis, Treatment, and Control.
J Clin Med 2020; 9: 2-29.
85)  Chu  DKW,  Hui  KPY,  Perera  RAPM,  Miguel  E, 
Niemeyer D, Zhao J, Channappanavar R, Du-
das G, Oladipo JO, Traoré A, Fassi-Fihri O, Ali A,
Demissié GF, Muth D, Chan MCW, Nicholls JM,
Meyerholz DK, Kuranga SA, Mamo G, Zhou Z, So
RTY, Hemida MG, Webby RJ, Roger F, Rambaut 
A, Poon LLM, Perlman S, Drosten C, Chevalier
V, Peiris M. MERS coronaviruses from camels in
Africa exhibit region-dependent genetic diversity.
Proc Natl Acad Sci U S A 2018; 115: 3144-3149.
86) Hu B, Guo H, Zhou P, Shi ZL. Characteristics of
SARS-CoV-2 and COVID-19. Nat Rev Microbiol
2021; 19: 141-154.
87)  Sahin AR, Erdogan A, Agaoglu PM, Dineri Y, Ca-
kirci  AY,  Senel  ME,  Okyay  RA,  Tasdogan  AM. 
2019 Novel Coronavirus (COVID-19) Outbreak: A 
Review of the Current Literature. Eurasian J Med
Oncol 2020; 4: 1-7.
88) Krishnamoorthy P, Raj AS, Roy S, Kumar NS,
Kumar H. Comparative transcriptome analysis
of SARS-CoV, MERS-CoV, and SARS-CoV-2 to
identify potential pathways for drug repurposing.
Comput Biol Med 2021; 128: 104123.
89) De Albuquerque LP, Siqueira Patriota LL de, Gon-
zatto V, Pontual EV, Guedes Paiva PM, Napoleão
TH. Coronavirus Spike (S) Protein: A Brief Review 
on Structure-Function Relationship, Host Recep-
tors, and Role in Cell Infection. Adv Res 2020; 28:
116 -124.
90) Pandey P, Rane JS, Chatterjee A, Kumar A, Khan
R,  Prakash  A,  Ray  S.  Targeting  SARS-CoV-2 
spike protein of COVID-19 with naturally occurring 
phytochemicals: an in silico study for drug devel-
opment. J Biomol Struct Dyn 2020; 2020: 1-11.
91) Kaur N, Singh R, Dar Z, Bijarnia RK, Dhingra N,
Kaur T. Genetic comparison among various coro-
navirus  strains  for  the  identication  of  potential 
vaccine targets of SARS-CoV2. Infect Genet Evol
2020; 89: 104490.
92)  Liu DX, Fung TS, Chong KKL, Shukla A, Hilgen-
feld R. Accessory proteins of SARS-CoV and oth-
er coronaviruses. Antiviral Res 2014; 109: 97-109.
93) Fung TS, Liu DX. Similarities and Dissimilarities
of COVID-19 and Other Coronavirus Diseases.
Annu Rev Microbiol 2021; 75: 1-18.
94) Naqvi AAT, Fatima K, Mohammad T, Fatima U,
Singh IK, Singh A, Atif SM, Hariprasad G, Hasan
GM, Hassan MI. Insights into SARS-CoV-2 ge-
nome, structure, evolution, pathogenesis and ther-
apies: Structural genomics approach. Biochim Bio-
phys Acta - Mol Basis Dis 2020; 1866: 165878.
95) Shi CS, Nabar NR, Huang NN, Kehrl JH.
SARS-Coronavirus Open Reading Frame-8b trig-
gers intracellular stress pathways and activates
NLRP3 inammasomes. Cell Death Discov 2019; 
5: 1-12.
96)  Yue  Y, N abar  NR,  Shi  CS,  Kamenyeva  O,  Xiao 
X, Hwang IY, Wang M, Kehrl JH.  SARS-Corona-
virus Open Reading Frame-3a drives multimodal
necrotic cell death. Cell Death Dis 2018; 9: 1-15.
97) Wong HH, Fung TS, Fang S, Huang M, Le MT,
Liu DX. Accessory proteins 8b and 8ab of severe
acute respiratory syndrome coronavirus suppress
the interferon signaling pathway by mediating
ubiquitin-dependent rapid degradation of inter-
feron regulatory factor 3. Virology 2018; 515: 165-
175.
98)  Wong LYR, Ye ZW, Lui PY, Zheng X, Yuan S, Zhu 
L, Fung SY, Yuen KS, Siu KL, Yeung ML, Cai Z, 
Woo PCY, Yuen KY, Chan CP, Jin DY. Middle East 
Respiratory Syndrome Coronavirus ORF8b Ac-
cessory Protein Suppresses Type I IFN Expres-
sion by Impeding HSP70-Dependent Activation of
IRF3 Kinase  IKKε. J  Immunol  2020;  205:  1564-
1579.
99)  V’kovski P, Kratzel A,  Steiner S, Stalder H, Thiel 
V. Coronavirus biology and replication: implica-
tions for SARS-CoV-2. Nat Rev Microbiol 2021;
19: 155-170.
100) Millán-Oñate J, Rodriguez-Morales AJ, Cama-
cho-Moreno G, Mendoza-Ramírez H,Rodrí-
guez-Sabogal IA, Álvarez-Moreno C. A new emerg-
ing zoonotic virus of concern : the 2019 novel Coro-
navirus (SARS CoV-2). Infection 2020; 24: 187-192.
101) Cappuccio FP, Siani A. Covid-19 and cardiovas-
cular  risk:  Susceptibility  to  infection  to  SARS-
CoV-2, severity and prognosis of Covid-19 and
blockade  of  the  renin-angiotensin-aldosterone 
system. An evidence-based viewpoint. Nutr Me-
tab Cardiovasc Dis 2020; 30: 1227-1235.
102) L’Huillier AG, Torriani  G,  Pigny  F, Kaiser L, Eck-
erle I. Shedding of infectious SARS-CoV-2 in
symptomatic neonates, children and adolescents.
medRxiv 2020; 10.1101/2020.04.27.20076778.
103) de Wit E, van Doremalen N, Falzarano D, VJ M.
SARS and MERS: recent insights into emerging
coronaviruses. Nat Rev Microbiol 2020; 14: 23-34.
104) Pal M, Berhanu G, Desalegn C, Kandi V. Se-
vere Acute Respiratory Syndrome Coronavirus-2
(SARS-CoV-2): An Update. Cureus 2020; 12:
e7423 .
105) Li W, Moore MJ, Vasllieva N, Sui J, Wong SK,
Berne MA, Somasundaran M, Sullivan JL, Lu-
zuriaga K, Greeneugh TC, Choe H, FarzanM.
Angiotensin-converting enzyme 2 is a functional
receptor for the SARS coronavirus. Nature 2003;
426: 450-454.
106) Gu J, Korteweg C. Pathology and pathogenesis of
severe acute respiratory syndrome. Am J Pathol
2007; 170 : 113 6 -1147.
107) Chen J, Subbarao K. The immunobiology of
SARS. Annu Rev Immunol 2007; 25: 443-472.
108) Zhou T, Tsybovsky Y, Gorman J, Rapp M, Cerut-
ti  G,  Chuang  GY,  Katsamba  PS,  Sampson  JM, 
Schön A, Bimela J, Boyington JC, Nazzari A, Olia
AS, Shi W, Sastry M, Stephens T, Stuckey J, Teng 
IT, Wang P, Wang S, Zhang B, Friesner RA, Ho
DD, Mascola JR, Shapiro L, Kwong PD. Cryo-EM
Structures of SARS-CoV-2 Spike without and with 
ACE2 Reveal a pH-Dependent Switch to Mediate
Endosomal Positioning of Receptor-Binding Do-
mains. Cell Host Microbe 2020; 28: 867-879.
109) Schoeman D, Fielding BC. Coronavirus envelope
protein: Current knowledge. Virol J 2019; 16: 1-22. 
110) Ou X, Liu Y, Lei X, Li P, Mi D, Ren L, Guo L, Guo 
R, Chen T, Hu J, Xiang Z, Mu Z, ChenX, Chen J,
Hu K, Jin Q, Wang J, Qian Z. Characterization of 
spike glycoprotein of SARS-CoV-2 on virus entry 
Comparative review of three human coronaviruses
7179
and its immune cross-reactivity with SARS-CoV.
Nat Commun 2020; 11: 1-12.
111) Hoffmann M, Kleine-Weber H, Schroeder S,
Krüger N, Herrler T, Erichsen S, Schiergens TS,
Herrler G, Wu NH, Nitsche A, MüllerMA, Dro-
stenC, PöhlmannS.SARS-CoV-2 Cell Entry De-
pends on ACE2 and TMPRSS2 and Is Blocked by 
a Clinically Proven Protease Inhibitor. Cell 2020;
181: 271-280.
112) Wan Y, Shang J, Sun S, Tai W, Chen J, Geng Q, 
He L, Chen Y, Wu J, Shi Z, ZhouY, DuL, LiF.Mo-
lecular Mechanism for Antibody-Dependent En-
hancement of Coronavirus Entry. J Virol 2019; 94:
e0 2015-19.
113) Tay MZ, Poh CM, Rénia L, MacAry PA, Ng LFP. The
trinity  of COVID-19: immunity, inammation  and in-
tervention. Nat Rev Immunol 2020; 20: 363-374.
114) Wrapp D, Wang N, Corbett KS, Goldsmith JA, Hsieh
CL, Abiona O, Graham BS, McLellan JS. Cryo-EM
structure  of  th e  2019-nCoV  spike  in  the  prefus ion 
conformation. Science 2020; 367: 1260-1263.
115) Chu  H,  Chan  JFW,  Yuen  TTT,  Shuai  H,  Yuan 
S,  Wang  Y,  Hu  B,  Yip  CCY, Tsang  JOL,  Hua ng 
X, Chai  Y, Yang D, Hou Y, Chik  KKH, Zhang X, 
Fung AYF, Tsoi HW, Cai JP, Chan WM, Ip JD, Chu 
AWH, Zhou J, Lung DC, Kok KH, To KKW, Tsang 
OTY, Chan  KH,  Yuen KY. Comparative  tropism, 
replication kinetics,  and cell  damage  proling  of 
SARS-CoV-2 and SARS-CoV with implications
for clinical manifestations, transmissibility, and
laboratory studies of COVID-19: an observational
study. The Lancet Microbe 2020; 1: e14-e23.
116) Sigrist CJ, Bridge A, Le Mercier P. A potential role
for integrins in host cell entry by SARS-CoV-2.
Antiviral Res 2020; 177: 104759.
117)  Wang K , Chen  W, Zhang  Z, Den g Y, Lian  JQ,  Du 
P,  Wei  D,  Zhang  Y,  Sun  XX,  Gong  L,Yang  X,  He 
L,  Zhang  L, Yang Z, Geng  JJ,  Chen R, Zhang  H,
Wang B, Zhu YM, Nan G, Jiang JL, Li L, Wu J, Lin P, 
Huang W, Xie L, Zheng ZH, Zhang K, Miao JL, Cui
HY, Huang M, Zhang J, Fu L, Yang XM, Zhao Z, Sun
S, Gu H, Wang Z, Wang CF, Lu Y, Liu YY, Wang QY, 
Bian H, Zhu P, Chen  ZN. CD147-spike protein is a
novel route for SARS-CoV-2 infection to host cells.
Signal Transduct Target Ther 2020; 5: 1-10.
118) Cui J, Li F, Shi ZL. Origin and evolution of patho-
genic coronaviruses. Nat Rev Microbiol 2019; 17:
181-192.
119) Xu H, Zhong L, Deng J, Peng J, Dan H, Zeng X, Li
T, Chen Q.  High expression of ACE2 receptor of 
2019-nCoV on the epithelial cells of oral mucosa.
Int J Oral Sci 2020; 12: 1-5.
120) Sun K, Gu L, Ma L, Duan Y. Atlas of ACE2 gene 
expression reveals novel insights into transmis-
sion of SARS-CoV-2. Heliyon 2021; 7: e05850.
121) Buqaileh R, Saternos H, Ley S, Aranda A, Fore-
ro K AW. Can cilia provide an entry gateway for
SARS-CoV-2 to human ciliated cells? Physiol Ge-
nomics 2021; 53: 249-258.
122) Fani M, Teimoori A, Ghafari S. Comparison of the
COVID-2019 (SARS-CoV-2) pathogenesis with
SARS-CoV and MERS-CoV infections. Future Vi-
rol 2020; 15: 317-323.
123) Khan S, El Morabet R, Khan RA, Bindajam
A, Alqadhi S, Alsubih M, Khan NA. Where we
missed? Middle East Respiratory Syndrome
(MERS-CoV) epidemiology in Saudi Arabia; 2012-
2019. Sci Total Environ 2020; 747: 141369.
124) Navand, Soltani S, Moghadami M, Hosseini P,
Nasimzadeh S, Zandi M. Diabetes and corona-
virus infections (SARS-CoV, MERS-CoV, and
SARS-CoV-2). J Acute Dis 2020; 9: 244-247.
125) Li Y, Zhang Z, Yang L, Lian X, Xie Y, Li S, Xin S, 
Cao P, Lu J. The MERS-CoV Receptor DPP4 as
a Candidate Binding Target of the SARS-CoV-2
Spike. iScience 2020; 23: 101160.
126) Deacon CF. Physiology and Pharmacology of
DPP-4 in Glucose Homeostasis and the Treat-
ment of Type 2 Diabetes. Front Endocrinol (Laus-
anne) 2019; 10: 80-99.
127) Somsen GA, van Rijn C, Kooij S, Bem RA, Bonn
D. Small droplet aerosols in poorly ventilated
spaces and SARS-CoV-2 transmission. Lancet
Respir Med 2020; 8: 658-659.
128) Dhama K,  Patel  SK,  Pathak  M,  Yatoo MI,  Tiwari 
R, Malik YS, Singh R, Sah R, Rabaan AA, Bonil-
la-Aldana DK R-MA. An update on SARS-CoV-2/
COVID-19 with particular reference to its clinical
pathology, pathogenesis, immunopathology and
mitigation strategies. Travel Med Infect Dis 2020;
37: 101755.
129) Tripathi  S,  Pandey  MK,  Malik  YS,  Bilal  M,  D ha-
ma K, Chaicumpa W, Chandra R. Novel corona-
virus (Sars-cov-2): Molecular biology, pathogen-
esis, pathobiology and advances in treatment of
covid-19 patients- an update. J Exp Biol Agric Sci
2020; 8: 683-708.
130) Chua RL, Lukassen S, Trump S, Hennig BP, Wen-
disch D, Pott F, Debnath O, Thürmann L, Kurth
F,  Völker  MT,  Kazmierski  J,  Timmermann  B, 
Twa rdziok  S, Schn eider S, Ma chle idt  F, Müller- Re-
detzky H, Maier M, Krannich A, Schmidt S, Balzer 
F,  Liebig  J,  Loske  J,  Sut torp  N,  Eils  J,  Ishaque 
N, Liebert UG, von Kalle C, Hocke A, Witzenrath 
M,  Gofnet  C,  Drosten  C,  Laudi  S,  Lehmann  I, 
Conrad C, Sander LE, Eils R.COVID-19 severi-
ty correlates with airway epithelium-immune cell
interactions identied by single-cell analysis. Nat 
Biotechnol 2020; 38: 970-979.
131) Tsai  SF,  Lu  KY,  Chuang  HM,  Liu  CA.  Survivi ng 
the Rookie Virus, Severe Acute Respiratory Syn-
drome Coronavirus 2 (SARS-CoV2): The Immu-
nopathology of a SARS-CoV2 Infection. Cell
Transplant 2021; 30: 096368972199376.
132) Laporte M, Stevaert A, Raeymaekers V, van Ber-
waer R, Martens K, Pöhlmann S, Naesens L. The
SARS-CoV-2 and other human coronavirus spike 
proteins are ne-tuned  towards temperature  and 
proteases of the human airways. PLoS Pathog
2021; 17: e1009500.
133) Halaji M, Farahani A, Ranjbar R, Heiat M, Dehkordi 
FS. Emerging coronaviruses: First SARS, second
MERS and third SARS-COV-2. epidemiological
updates of COVID-19. Infez Med 2020; 28: 6-17.
134) Arumugam VA, Thangavelu S, Fathah Z, Ravindran
P, Sanjeev AMA, Babu S, Meyyazhagan A, Yatoo MI, 
Sharun K, Tiwari R, PandeyMK, SahR, ChandraR,
DhamaK.COVID-19 and the world with co-morbidi-
ties of heart disease, hypertension and diabetes. J
Pure Appl Microbiol 2020; 14: 1623-1638.
A.A. Rabaan, A.A. Mutair, Z.A. Alawi, S. Alhumaid, M.A. Mohaini, J. Aldali, R. Tirupathi, et al
7180
135) Corvol P. The renin-angiotensin-aldosterone sys-
tem and its suppression. Ann Pediatr (Paris) 1982;
29: 657-658.
136) Oz M, Lorke DE. Multifunctional angiotensin con-
verting enzyme 2, the SARS-CoV-2 entry recep-
tor, and critical appraisal of its role in acute lung
injury. Biomed Pharmacother 2021; 136: 111193.
137) Cheng  H,  Wang  Y,  Wang  GQ.  Organ-protec tive 
effect of angiotensin-converting enzyme 2 and its
effect on the prognosis of COVID-19. J Med Virol
2020; 92: 726-730.
138) Schaller T, Hirschbühl  K,  Burkhardt K, Braun G, 
Trepel M,  rkl B, Claus R. Postmortem  Exam-
ination of Patients with COVID-19. JAMA - J Am
Med Assoc 2020; 323: 2518-2520.
139) Ogando NS, Zevenhoven-Dobbe JC, van der
Meer  Y,  Bredenbeek  PJ,  Posthuma  CC,  Snijder 
EJ. The Enzymatic Activity of the nsp14 Exoribo-
nuclease Is Critical for Replication of MERS-CoV
and SARS-CoV-2. J Virol 2020; 94: e01246-20.
140)NarayananK, Huang C, Lokugamage K, Kamitani 
W, Ikegami T, Tseng  CT, Makino  S.  Severe  acute 
respiratory syndrome coronavirus nsp1 suppresses
host gene expression, including that of type I inter-
feron, in infected cells. J Virol 2008; 82: 4471-4479.
141) Lokugamage  KG,  Narayanan  K,  Nakagawa  K, 
Terasaki  K,  Ramirez  SI,  Tseng  CTK,  Makino  S. 
Middle East Respiratory Syndrome Coronavirus
nsp1 Inhibits Host Gene Expression by Selective-
ly Targeting mRNAs Transcribed in the Nucleus
while Sparing mRNAs of Cytoplasmic Origin. J
Virol 2015; 89: 10970-10981.
142) Thoms M, Buschauer R, Ameismeier M, Koepke 
L,  Denk  T,  Hirschenberger  M,  Kratzat  H,  Hayn 
M, MacKens-Kiani T, Cheng J, Straub JH, Stür-
zel CM, Fröhlich  T, Berninghausen O, Becker T, 
Kirchhoff F, Sparrer KMJ, Beckmann R. Structural 
basis for translational shutdown and immune eva-
sion by the Nsp1 protein of SARS-CoV-2. Science
2020; 369: 1249-1256.
143) Snijder EJ, van der Meer Y, Zevenhoven-Dobbe J, 
Onderwater JJM, van der Meulen J, Koerten HK,
Mommaas AM. Ultrastructure and Origin of Mem-
brane Vesicles Associated with the Severe Acute
Respiratory Syndrome Coronavirus Replication
Complex. J Virol 2006; 80: 5927-5940.
144) Hemmat N, Asadzadeh Z, Ahangar NK, Alemo-
hammad H, Najafzadeh B, Derakhshani A, Bagh-
banzadeh A, Baghi HB, Javadrashid D, Naja S, 
Ar Gouilh M, Baradaran B.The roles of signal-
ing pathways in SARS-CoV-2 infection; lessons
learned from SARS-CoV and MERS-CoV. Arch
Virol 2021; 166: 675-696.
145) Granados A, Peci A, McGeer A, Gubbay JB. Inu-
enza and rhinovirus viral load and disease severi-
ty in upper respiratory tract infections. J Clin Virol
2017; 86: 14-19.
146) Jiang X, Rayner S, Luo MH. Does SARS-CoV-2
has a longer incubation period than SARS and
MERS? J Med Virol 2020; 92: 476-478.
147) Wen J, Cheng Y, Ling R, Dai Y, Huang B, Huang 
W,  Zhang  S,  Jiang  Y.  Antibody-dependent  en-
hancement of coronavirus. Int J Infect Dis 2020;
100: 483-489.
148) The epidemiological characteristics of an outbreak 
of 2019 novel coronavirus diseases (COVID-19) in
China. Zhonghua Liu Xing Bing Xue Za Zhi 2020;
41: 145-151.
149) Miorin L, Kehrer T, Sanchez-Aparicio MT, Zhang
K, Cohen P, Patel RS, Cupic A, Makio T, Mei M, 
Moreno E, Danziger O, White KM, Rathnasinghe
R, Uccellini M, Gao S, Aydillo T, Mena I, Yin X, 
Martin-Sancho L, Krogan NJ, Chanda SK, Schot-
saert  M,  Wozniak  RW,  Ren  Y,  Rosenberg  BR, 
Fontoura BMA, García-Sastre A.SARS-CoV-2
Orf6 hijacks Nup98 to block STAT nuclear import 
and antagonize interferon signaling. Proc Natl
Acad Sci U S A 2020; 117: 28344-28354.
150) Sinderewicz  E,  Czelejewska  W,  Jezier-
ska-Wozniak  K,  Staszkiewicz-Chodor  J,  Maksy-
mowicz W. Immune response to covid-19: Can we
benet from the sars-cov and mers-cov pandemic 
experience? Pathogens 2020; 9: 739.
151) Thieme CJ, Anft M, Paniskaki K, Blazquez-Navar-
ro  A,  Doevelaar A,  Seibert  FS,  Hoelzer B, Konik 
MJ, Berger MM, Brenner T, Tempfer C, Watzl C,
Meister TL, Pfaender S, Steinmann E, Dolff S, Ditt-
mer U, Westhoff TH, Witzke O, Stervbo U, Roch T, 
Babel  N. Robust  T  Cell  Response Toward Spike, 
Membrane, and Nucleocapsid SARS-CoV-2 Pro-
teins Is Not Associated with Recovery in Critical
COVID-19 Patients. Cell Reports Med 2020; 1:
100092.
152) Mahallawi WH, Khabour OF, Zhang Q, Makhdoum 
HM, Suliman BA. MERS-CoV infection in humans
is  associated  with  a  pro-inammatory  Th1  and 
Th17 cytokine prole. Cytokine 2018; 104: 8-13.
153) Kumar S, Nyodu R, Maurya VK, Saxena SK. Host
Immune Response and Immunobiology of Human
SARS-CoV-2 Infection. 2020; 2020: 43-53.
154) Bouvet M, Debarnot C, Imbert I, Selisko B, Snijder 
EJ, Canard B, Decroly E. In vitro reconstitution
of sars-coronavirus mRNA cap methylation. PLoS
Pathog 2010; 6: 1-13.
155) Chen L, Zhao J, Peng J, Li X, Deng X, Geng Z,
Shen Z, Guo F, Zhang Q, Jin Y,WangL, WangS.
Detection of 2019-nCoV in Saliva and Character-
ization of Oral Symptoms in COVID-19 Patients.
Cell Prolif 2020; 53: e12923.
156) Rakib A, Sami SA, Mimi NJ, Chowdhury MM, Eva 
TA, Nainu F, Paul A, Shahriar A, Tareq AM, Emon
NU, Chakraborty S, Shil S, Mily SJ, Hadda TB, Al-
malki  FA,  Emran  TB.  Immunoinformatics-guided 
design of an epitope-based vaccine against severe
acute  respiratory  syndrome  coronavirus  2  spike 
glycoprotein. Comput Biol Med 2020; 124: 103967.
157) Pormohammad A, Ghorbani S, Khatami A, Farzi
R, Baradaran B, Turner DL, Turner RJ, Mansour-
nia MA, Kyriacou DN, Bahr NC, IdrovoJ-P.Com-
parison  of  Conrmed  COVID-19 with  SARS  and 
MERS Cases - Clinical Characteristics, Laborato-
ry Findings, Radiographic Signs and Outcomes:
A Systematic Review and Meta-Analysis. Rev
Med Virol 2020; 30: e2112.
158) Gao Y, Li T, Han M, Li X, Wu D, Xu Y, Zhu Y, Liu Y, 
Wang X, Wang L. Diagnostic utility of clinical lab-
oratory data determinations for patients with the
severe COVID-19. J Med Virol 2020; 92: 791-796.
Comparative review of three human coronaviruses
7181
159) Al-Tawq JA, Hinedi K, Abbasi S, Babiker M, Sun-
ji A, Eltigani M. Hematologic, hepatic, and renal
function changes in hospitalized patients with
Middle East respiratory syndrome coronavirus.
Int J Lab Hematol 2017; 39: 272-278.
160) Wang JT, Sheng WH,  Fang CT, Chen YC, Wang 
JL, Yu CJ, Chang SC, Yang PC. Clinical Manifes-
tations, Laboratory Findings, and Treatment Out-
comes of SARS Patients. Emerg Infect Dis 2004;
10: 818-824.
161) Giannis D, Ziogas IA, Gianni P. Coagulation disor-
ders in coronavirus infected patients: COVID-19,
SARS-CoV-1, MERS-CoV and lessons from the
past. J Clin Virol 2020; 127.
162) Ma C,  Cong  Y, Zhang  H.  COVID-19 and the Di-
gestive System. Am J Gastroenterol 2020; 115:
1003-1006.
163) Zhou Z, Zhao N, Shu Y, Han S, Chen B, Shu X. Ef-
fect of Gastrointestinal Symptoms in Patients With
COVID-19. Gastroenterology 2020; 158: 2294-2297.
164) Loganathan  S,  Kuppusamy  M,  Wankhar  W, 
Gurugubelli KR, Mahadevappa VH, Lepcha L,
Choudhary AK. Angiotensin-converting enzyme 2
(ACE2): COVID-19 gate way to multiple organ fail-
ure syndromes. Respir Physiol Neurobiol 2021;
283: 103548.
165) Chai X, Hu L, Zhang Y, Han W, Lu Z, Ke A, Zhou 
J, Shi G, Fang N, Fan J, CaiJ, FanJ, LanF. Specif-
ic ACE2 expression in cholangiocytes may cause
liver damage after 2019-nCoV infection. bioRxiv
2020; 10.1101/2020.02.03.931766.
166) Lei HY, Ding YH, Nie K, Dong YM, Xu JH, Yang 
ML,  Liu  MQ,  Wei  L,  Nasser  MI,  Xu  LY,  ZhuP, 
ZhaoMY. Potential effects of SARS-CoV-2 on the 
gastrointestinal tract and liver. Biomed Pharma-
cother 2021; 133: 111064.
167) Puelles VG, Lütgehetmann M, Lindenmeyer MT,
Sperhake  JP,  Wong  MN,  Allweiss  L,  Chilla  S, 
Heinemann A, Wanner N, Liu S, Braun F, Lu S,
Pfefferle S, Schröder AS, Edler C, Gross O, Glat-
zel M, Wichmann D, Wiech T, Kluge S, Pueschel K,
Aepfelbacher M, Huber TB. Multiorgan and Renal
Tropism of SARS-CoV-2. N Engl J Med 2020; 383:
590-592.
168) Montalvan V, Lee J, Bueso T, De Toledo J, Rivas
K. Neurological manifestations of COVID-19 and
other coronavirus infections: A systematic review.
Clin Neurol Neurosurg 2020; 194: 105921.
169) Al-Washahi M, Al-Abri R. Loss of smell and taste
are newly emerging symptoms in COVID-19 pa-
tients necessitating more insights into their diag-
nostic evaluation. Oman Med J 2021; 36: e257.
170) De Vito A, Geremia N, Fiore V, Princic E, Babud-
ieri S, Madeddu G. Clinical features, laboratory
ndings  and  predictors  of  death  in  hospitalized 
patients with COVID-19 in Sardinia, Italy. Eur Rev
Med Pharmacol Sci 2020; 24: 7861-7868.
171) Neumann B, Schmidbauer ML, Dimitriadis K,
Otto S, Knier B, Niesen WD, Hosp JA, Günther
A,  Lindeman n  S,  Nagy  G,SteinbergT,  LinkerRA, 
HemmerB, BöselJ.Cerebrospinal uid ndings in 
COVID-19 patients with neurological symptoms. J
Neurol Sci 2020; 418: 117090.
172) Vaira LA, Deiana G, Lechien JR, De Vito A, Cos-
su A, Dettori M, Del Rio A, Saussez S, Madeddu
G, Babudieri S, Fois AG, Clementina C, Hopkins 
C, De Riu G, Piana AF. Correlations Between Ol-
factory Psychophysical Scores and SARS-CoV-2
Viral Load in COVID-19 Patients. Laryngoscope
2021; 42: 1560-1590.
173) Vaira LA, Hopkins C, Salzano G, Petrocelli M, Me-
lis A, Cucurullo M, Ferrari M, Gagliardini L, Pipolo
C, Deiana G, Fiore V, De Vito A, Turra N, Canu S,
Maglio A, Serra A, Bussu F, Madeddu G, Babudieri
S, Giuseppe Fois A, Pirina P, Salzano FA, De Riu P,
Biglioli F, De Riu G. Olfactory and gustatory function
impairment in COVID-19 patients: Italian objective
multicenter-study. Head Neck 2020; 42: 1560-1569. 
174) Hosseini MJ, Halaji M, Nejad JH, Ranjbar R. Cen-
tral nervous system vasculopathy associated with
severe acute respiratory syndrome coronavirus 2
(SARS-CoV-2): a novel case report from Iran. J
Neurovirol 2021; 27: 507-509.
175) Vasanthapuram  VH,  Badakere  A.  Internucle-
ar ophthalmoplegia as a presenting feature in a
COVID-19-positive patient. BMJ Case Rep 2021;
14: e241873.
176) Conrady CD, Faia LJ, Gregg KS, Rao RC. Coro-
navirus-19-Associated Retinopathy. Ocul Immu-
nol Inamm 2021; 2021: 1-2.
177) Gabarre P, Dumas G, Dupont T, Darmon M, Azou-
lay  E,  Zafrani  L.  Acu te  kidney  injury  in  criti cally 
ill patients with COVID-19. Intensive Care Med
2020; 46: 1339-1348.
178) Henry BM, Lippi G. Chronic kidney disease is as-
sociated with severe coronavirus disease 2019
(COVID-19) infection. Int Urol Nephrol 2020; 52:
119 3 -1194.
179) Patel SK, Singh R, Rana J, Tiwari R, Natesan S,
Harapan H, Arteaga-Livias K, Bonilla-Aldana DK,
Rodríguez-Morales AJ, Dhama K. The kidney and 
COVID-19 patients - Important considerations.
Travel Med Infect Dis 2020; 37: 101831.
180) Ronco C, Reis T. Kidney involvement in COVID-19
and rationale for extracorporeal therapies. Nat
Rev Nephrol 2020; 16: 308-310.
181) Gäckler A, Rohn H, Witzke O. Acute renal failure 
in COVID-19. Nephrologe 2021; 16:66-70.
182)  Hardenberg JHB, Stockmann H, Eckardt KU, Schmidt-
Ott KM. COVID-19 and acute kidney injury in the inten-
sive care unit. Nephrologe 2021; 16: 20-25.
183) Ng JH, Hirsch JS, Hazzan A, Wanchoo R, Shah
HH,  Maliecka l  DA,  Ross  DW,  Sharma  P,  Sakhi-
ya V, Fishbane S, Jhaveri KD, Abate M, Andrade
HP, Barnett  RL,  Bellucci  A, Bhaskaran  MC,  Co-
rona AG, Flores Chang BS, Finger M, Gitman
M,  Halinski  C,  Hasan  S,  Hazzan  AD,  Hong  S, 
KhaninY, KuanA, Madireddy V, Malieckal D, Muz-
ib A, Nair G, Nair VV, Ng JH, Parikh R, Sachdeva 
M, Schwarz R, Singhal PC, Uppal NN. Outcomes
Among Patients Hospitalized With COVID-19 and
Acute Kidney Injury. Am J Kidney Dis 2021; 77:
204-215.e1.
184) Martinez-Rojas MA, Vega-Vega O, Bobadilla
XNA. Is the kidney a target of SARS-CoV-2? Am J 
Physiol - Ren Physiol 2020; 318: F1454-F1462.
185) Chong WH, Saha BK. Relationship Between Se-
vere Acute Respiratory Syndrome Coronavirus 2
(SARS-CoV-2) and the Etiology of Acute Kidney
Injury (AKI). Am J Med Sci 2021; 361: 287-296.
A.A. Rabaan, A.A. Mutair, Z.A. Alawi, S. Alhumaid, M.A. Mohaini, J. Aldali, R. Tirupathi, et al
7182
186) Mitaka  H,  Kuno  T,  Takagi  H  PP.  Incidence  and 
Mortality of COVID-19-associated Pulmonary As-
pergillosis: A Systematic Review and Meta-analy-
sis. Mycoses 2021; 2021: 1-13.
187) Silva DL, Lima CM, Magalhães VCR, Baltazar LM,
Peres NTA, Caligiorne RB, Moura AS, Fereguetti
T, Martins JC, Rabelo LF, AbrahãoJS, LyonAC,
JohannS, SantosDA. Fungal and bacterial coin-
fections increase mortality of severely ill COVID-19
patients. J Hosp Infect 2021; 113: 145-154.
188) Zuo T, Liu Q, Zhang F, Yeoh YK, Wan Y, Zhan H, 
Lui GCY, Chen Z, Li AYL, Cheung CP, Chen N, Lv 
W, Ng RWY, Tso EYK, Fung KSC, Chan V, Ling L, 
Joynt G, Hui DSC, Chan FKL, Chan PKS,Ng SC.
Temporal landscape of human gut RNA and DNA
virome in SARS-CoV-2 infection and severity. Mi-
crobiome 2021; 9: 1-16.
189) Zuo T, Zhang F, Lui GCY, Yeoh YK, Li AYL, Zhan 
H, Wan Y, Chung ACK, Cheung CP, Chen N, Lai 
CKC, Chen Z, Tso EYK, Fung KSC, Chan V, Ling 
L, Joynt G, Hui DSC, Chan FKL,Chan PKS, Ng
SC. Alterations in Gut Microbiota of Patients With
COVID-19 During Time of Hospitalization. Gastro-
enterology 2020; 159: 944-955.
190) Zuo T, Zhan H, Zhang F, Liu Q, Tso EYK, Lui GCY, 
Chen N, Li A, Lu W, Chan FKL, Chan PKS, Ng SC.
Alterations in Fecal Fungal Microbiome of Patients
With COVID-19 During Time of Hospitalization until
Discharge. Gastroenterology 2020; 159: 1302-1310.
191) Elhadi  M,  Alsou  A,  Msherghi  A,  Alshareea  E, 
Ashini A, Nagib T, Abuzid N, Abodabos S, Alrifai
H, Gresea E, Yahya W, Ashour D, Abomengal S, 
Qarqab N, Albibas A, Anaiba M, Idheiraj H, Abra-
heem  H,  Fayyad  M, Alkilani  Y, Alsuwiyah  S,  El-
ghezewi A, Zaid A. Psychological Health, Sleep
Quality, Behavior, and  Internet  Use  Among  Peo-
ple During the COVID-19 Pandemic: A Cross-Sec-
tional Study. Front Psychiatry 2021; 12: 632496.
192) Cazzato G, Foti C, Colagrande A, Cimmino A,
Scarcella S, Cicco G, Sablone S, Arezzo F, Romi-
ta  P, Lettini  T,RestaL,  IngravalloG.Skin  Manifes-
tation of SARS-CoV-2: The Italian Experience. J
Clin Med 2021; 10: 1566.
193) Rabaan AA, Al-Ahmed SH, Muhammad J, Khan
A, Sule AA, Tirupathi R, Mutair AA, Alhumaid S,
Al-Omari A, Dhawan M, Tiwari R, Sharun K, Mo-
hapatra RK, Mitra S, Bilal M, Alyami SA, Emran
TB,  Moni  MA,  Dhama  K.  Role  of  Inammatory 
Cytokines  in  COVID-19  Patients:  A  Review  on 
Molecular Mechanisms, Immune Functions, Im-
munopathology and Immunomodulatory Drugs to
Counter Cytokine Storm. Vaccines 2021; 9: 436. 
194) Merad M, Martin JC. Pathological inammation in 
patients with COVID-19: a key role for monocytes 
and macrophages. Nat Rev Immunol 2020; 20:
355-362.
195) Obando-Pereda G. Can molecular mimicry ex-
plain  the  cytokine  storm  of  SARS-CoV-2?: an in 
silico approach. J Med Virol 2021; 93: 5350-5357.
196) Song W-J, Hui CKM, Hull JH, Birring SS, Mc-
Garvey L, Mazzone SB, Chung KF. Confronting
COVID-19-associated cough and the post-COVID
syndrome: role of viral neurotropism, neuroinam-
mation, and neuroimmune responses. Lancet Re-
spir Med 2021; 9: 533-544.
197) Gunst  JD,  Staerke  NB,  Pahus  MH,  Kristensen 
LH, Bodilsen J, Lohse N, Dalgaard LS, Brønnum
D, Fröbert O, Hønge B, Johansen IS, Monrad I,
Erikstrup  C, Rosendal  R, Vilstrup  E,  Mariager T, 
Bove DG, Offersen R, Shakar S, Cajander S, Jør-
gensen NP, Sritharan SS, Breining P, Jespersen
S, Mortensen KL, Jensen ML, Kolte L, Frattari
GS, Larsen CS, Storgaard M, Nielsen LP, Tolstrup
M, Sædder EA, Østergaard LJ, Ngo HTT, Jensen
MH, Højen JF, Kjolby M, Søgaard OS. Efcacy of 
the TMPRSS2 inhibitor camostat mesilate in pa-
tients hospitalized with Covid-19-a double-blind
randomized controlled trial. EClinicalMedicine
2021; 35: 100849.
198) Tan SH, Allicock O, Armstrong-Hough M, Wyllie AL. 
Saliva as a gold-standard sample for SARS-CoV-2
detection. Lancet Respir Med 2021; 9: 562-564.
199) Tsang NNY, So  HC, Ng  KY, Cowling  BJ, Leung 
GM, Ip DKM. Diagnostic performance of different
sampling approaches for SARS-CoV-2 RT-PCR
testing: a systematic review and meta-analysis.
Lancet Infect Dis 2021; 21: 1-13.
200) Kim JE, Heo JH, Kim HO, Song SH, Park SS, Park 
TH, Ahn JY, Kim MK, Choi JP. Neurological compli-
cations during treatment of middle east respiratory
syndrome. J Clin Neurol 2017; 13: 227-233.
201) Lai AL, Millet JK, Daniel S, Freed JH, Whittaker 
GR. Prevalence of comorbidities and its effects in
patients infected with SARS-CoV-2: a systematic
review and meta-analysis. Int J Infect Dis 2020;
94: 91-95.
202) Khamis AH, Jaber M, Azar A, AlQahtani F, Bisha-
wi K, Shanably A. Clinical and laboratory ndings 
of COVID-19: A systematic review and meta-anal-
ysis. J Formos Med Assoc 2020; 2020: 1-18.
203) Lee KW, Yusof Khan AHK,  Ching  SM,  Chia  PK, 
Loh WC, Abdul Rashid AM, Baharin J, Inche Mat
LN, Wan Sulaiman WA, Devaraj NK, Sivaratnam
D, Basri H, Hoo FK. Stroke and Novel Coronavirus 
Infection in Humans: A Systematic Review and
Meta-Analysis. Front Neurol 2020; 11: 579070.
204) Correia AO, Feitosa PWG, Moreira JL de S,
Nogueira SÁR, Fonseca RB, Nobre MEP. Neu-
rological manifestations of COVID-19 and other
coronaviruses: A systematic review. Neurol Psy-
chiatry Brain Res 2020; 37: 27-32.
205) Borges do Nascimento IJ, Cacic N, Abdulazeem
HM, von Groote TC,  Jayarajah  U,  Weerasekara 
I, Esfahani MA, Civile VT, Marusic A, Jeroncic
A,Carvas  Junior  N,  Pericic  TP,  Zakarija-Grkov-
ic I, Meirelles Guimarães SM, Luigi BragazziN,
Bjorklund M, So-Mahmudi A, Altujjar M, Tian M, 
ArcaniDMC, O’Mathúna DP, Marcolino MS. Novel 
Coronavirus Infection (COVID-19) in Humans: A
Scoping Review and Meta-Analysis. J Clin Med
2020; 9: 941.
206) Hu W, Yen YT, Singh  S, Kao CL, Wu-Hsieh  BA. 
SARS-CoV regulates immune function-related
gene expression in human monocytic cells. Viral
Immunol 2012; 25: 277-288.
207) Connor RF, Roper RL. Unique SARS-CoV protein
nsp1: bioinformatics, biochemistry and potential ef-
fects on virulence. Trends Microbiol 2007; 15: 51-53.
208) Cheng VCC,  Lau  SKP, Woo  PCY, Yuen KY. Se-
vere Acute Respiratory Syndrome Coronavirus as
Comparative review of three human coronaviruses
7183
an Agent of Emerging and Reemerging Infection.
Clin Microbiol Rev 2007; 20: 660-694.
209) Chu H, Chan JF, Wang Y, Yuen TT, Chai Y, Hou 
Y,  Shuai  H,  Yang  D,  Hu  B,  Huang  X,  Zhang  X, 
Cai JP, Zhou J, Yuan S, Kok KH, To KK, Chan IH, 
Zhang  AJ,  Sit  K Y,  Au  WK,  Yuen  KY.  Compara-
tive replication  and immune activation  proles of 
SARS-CoV-2 and SARS-CoV in human lungs: an
ex vivo study with implications for the pathogen-
esis of COVID-19. Clin Infect Dis 2020; 71: 1400-
1409.
210) Siddique F, Abbas RZ, Mansoor MK, Alghamdi ES,
Saeed M, Ayaz MM, Rahman M, Mahmood MS,
Iqbal A, Manzoor M, Abbas A, JavaidA, HussainI.
An Insight Into COVID-19: A 21st Century Disaster
and Its Relation to Immunocompetence and Food
Antioxidants. Front Vet Sci 2021; 7: 586637.
211) Channappanavar R, Perlman S. Pathogenic hu-
man coronavirus infections: causes and conse-
quences of cytokine storm and immunopathology. 
Semin Immunopathol 2017; 39: 529-539.
212) Law HKW, Chung YC, Hoi  YN, Sin FS, Yuk OC, 
Luk W, Nicholls JM, Peiris JSM, Lau YL. Chemo-
kine up-regulation in SARS-coronavirus-infected, 
monocyte-derived human dendritic cells. Blood
2005; 106: 2366-2374.
213) Zheng  J, Wang Y, Li K,  Meyerholz  DK,  Allamargot 
C, Perlman S. Severe Acute Respiratory Syndrome
Coronavirus 2-Induced Immune Activation and
Death of Monocyte-Derived Human Macrophages
and Dendritic Cells. J Infect Dis 2021; 223: 785-795.
214) Cheung  CY, Poon  LLM,  Ng  IHY, Luk W, Sia  SF, 
Wu MHS,  Chan  KH, Yuen KY, Gordon S, Guan 
Y,  Peiris  JSM.  Cytokine  Responses  in  Severe 
Acute Respiratory Syndrome Coronavirus-Infect-
ed Macrophages In Vitro: Possible Relevance to
Pathogenesis. J Virol 2005; 79: 7819-7826.
215) Lee P, Kim DJ. Newly emerging human corona-
viruses: Animal models and vaccine research
for SARS, MERS, and COVID-19. Immune Netw
2020; 20: 1-25.
216) Tetro JA. Is COVID-19 receiving ADE from other
coronaviruses? Microbes Infect 2020; 22: 72-73.
217) Broxmeyer HE, Sherry B, Cooper S, Lu L, Maze
R,  Beckman n  MP,  Cera mi  A,  Ralph  P.  Compar-
ative analysis  of the human macrophage inam-
matory  protein  fa mily  of  cytokines  (chemok ines) 
on proliferation of human myeloid progenitor cells.
Interacting effects involving suppression, syner-
gistic  suppression, and blocking of  suppression. 
J Immunol 1993; 150: 3448-3458.
218) Wong CK, Lam CW, Wu AK, Ip WK, Lee NL, Chan
IH, Lit LC, Hui DS, Chan MH, Chung SS, Sung JJ.
Plasma inammatory cytokines  and  chemokines 
in severe acute respiratory syndrome. Clin Exp
Immunol 2004; 136: 95-103.
219) Herold S, Steinmueller M, Von Wulffen W, Ca-
karova  L,  Pinto  R,  Plesch ka  S,  Mack  M,  Kuziel 
WA, Corazza N, Brunner T, Seeger W, Lohmey-
er J. Lung  epithelial apoptosis  in  inuenza virus 
pneumonia: The role of macrophage-expressed
TNF-related apoptosis-inducing ligand. J Exp
Med 2008; 205: 3065-3077.
220) Boonnak K,  Vogel  L,  Feldmann  F, Feldmann  H, 
Legge KL, Subbarao K. Lymphopenia Associated
with Highly Virulent H5N1 Virus Infection Due to
Plasmacytoid Dendritic Cell-Mediated Apoptosis
of T Cells. J Immunol 2014; 192: 5906-5912.
221) Cameron MJ, Ran L, Xu L, Danesh A, Berme-
jo-Martin JF, Cameron CM, Muller MP, Gold WL,
Richardson SE, Poutanen SM,Willey BM, DeVries
ME, Fang Y, Seneviratne C, Bosinger SE, Persad 
D, Wilkinson P, Greller LD, Somogyi R, Humar A, 
Keshavjee S, Louie M, Loeb MB, BruntonJ, Mc-
Geer AJ, Kelvin DJ. Interferon-Mediated Immuno-
pathological Events Are Associated with Atypical
Innate and Adaptive Immune Responses in Pa-
tients with Severe Acute Respiratory Syndrome. J
Virol 2007; 81: 8692-8706.
222) Croce L, Gangemi D, Ancona G, Liboà F, Bendotti
G, Minelli L, Chiovato L. The  cytokine storm and 
thyroid hormone changes in COVID-19. J Endo-
crinol Invest 2021; 44: 891-904.
223) Hirawat R, Sai MA, Godugu C. Targeting inam-
matory cytokine storm  to ght against COVID-19 
associated severe complications. Life Sci 2021;
267: 118923.
224) Keam S, Megawati D, Patel SK, Tiwari R, Dhama
K, Harapan H. Immunopathology and immuno-
therapeutic strategies in severe acute respiratory
syndrome coronavirus 2 infection. Rev Med Virol
2020; 30: e2123.
225) Yongzhi  X.  COVID-19-associated  cytokine  storm 
syndrome and diagnostic principles: an old and new
Issue. Emerg Microbes Infect 2021; 10: 266-276.
226) Cao X. COVID-19: immunopathology and its im-
plications for therapy. Nat Rev Immunol 2020; 20:
269-270.
227) Zhang X, Tan Y, Ling Y, Lu G, Liu F, Yi Z, Jia X, Wu 
M, Shi B, Xu S, Chen J, Wang W, Chen B, Jiang L,
Yu S, Lu J, Wang J, Xu M, Yuan Z, Zhang Q, Zhang
X, Zhao G, Chen S, Lu H. Viral and host factors
related to the clinical outcome of COVID-19. Nature
2020; 583: 437-440.
228) Ruytinx P, Proost P, Van Damme J, Struyf S. Chemo-
kine-induced macrophage polarization in inamma-
tory conditions. Front Immunol 2018; 9: 1930.
229) Menezes SM, Braz M, Llorens-Rico V, Wauters J,
Van Weyenbergh J. Endogenous IFNβ expression 
predicts outcome in critical patients with COVID-19.
The Lancet Microbe 2021; 2: e235-e236.
230) Ari S, Can V, Demir ÖF, Ari H, Ağca FV, Melek M, 
Çamci S, Dikiş ÖŞ, Huysal K, Türk T. Elevated eo-
sinophil count is related with lower anti-factor Xa
activity in COVID-19 patients. J Hematop 2020;
13: 249-258.
231) Liu S, Yuan H, Zhang B, Li W, You J, Liu J, Zhong 
Q, Zhang L, Chen L,  Li S, ZouY, ZhangS. Com-
parison of Clinical Features and CT Temporal
Changes Between Familial Clusters and Non-fa-
milial Patients With COVID-19 Pneumonia. Front
Med 2021; 8: 1-9.
232) Surendar  J,  Frohberger  SJ,  Karunakaran  I, 
Schmitt V, Stamminger W, Neumann AL, Wilhelm
C, Hoerauf A, Hübner MP. Adiponectin limits IFN-
γand  IL-17  produci ng  CD4  T  cells  in  obesity  by 
restraining cell intrinsic glycolysis. Front Immunol
2019; 10: 2555.
233) Laires PA, Dias S, Gama A, Moniz M, Pedro AR,
Soares P, Aguiar P, Nunes C. The association
A.A. Rabaan, A.A. Mutair, Z.A. Alawi, S. Alhumaid, M.A. Mohaini, J. Aldali, R. Tirupathi, et al
7184
between chronic disease and serious COVID-19
outcomes  and  its  inuence  on  risk  perception: 
Survey study and database analysis. JMIR Public
Heal Surveill 2021; 7: e22794.
234) Bae SA, Kim SR, Kim MN, Shim WJ, Park SM. Im-
pact of cardiovascular disease and risk factors on 
fatal outcomes in patients with COVID-19 accord-
ing to age: A systematic review and meta-analy-
sis. Heart 2021; 107: 373-380.
235) Zinellu A, Arru F, De Vito A, Sassu A, Valdes G,
Scano V, Zinellu E, Perra R, Madeddu G, Carru
C,PirinaP, MangoniAA, BabudieriS, FoisAG.The
De Ritis ratio as prognostic  biomarker of  in-hos-
pital mortality in COVID-19 patients. Eur J Clin
Invest 2021; 51: e13427.
236) McCarthy  C,  O’Donnell  CP, Kell y  NEW,  O’Shea 
D, Hogan AE. COVID-19 severity and obesity: are
MAIT cells a factor? Lancet Respir Med 2021; 9:
445-447.
237) O’Shea D, Hogan AE. Dysregulation of natural kill-
er cells in obesity. Cancers (Basel) 2019; 11: 573.
238) Brien AO, Kedia-Mehta N, Tobin L, Veerapen N,
Besra GS, Shea DO, Hogan AE. Targeting mi-
tochondrial dysfunction in MAIT cells limits IL-17
production in obesity. Cell Mol Immunol 2020; 17:
119 3 -1195.
239) Popkin BM, Du S, Green WD, Beck MA, Algaith 
T,  Herbst  CH,  Alsukait  RF,  Alluhidan  M,  Ala-
zemi N, Shekar M. Individuals with obesity and 
COVID-19: A global perspective on the epide-
miology and biological relationships. Obes Rev
2020; 21: e13128.
240) Gao M, Piernas C, Astbury NM, Hippisley-Cox
J, O’Rahilly S, Aveyard P, Jebb SA. Associations 
between body-mass index and COVID-19 severi-
ty in 6·9 million people in England: a prospective,
community-based, cohort study. Lancet Diabetes
Endocrinol 2021; 9: 350-359.
241) Lega I, Nisticò L, Palmieri L, Caroppo E, Noce CL,
Donfrancesco C, Vanacore N, Scattoni ML, Picar-
di A, Gigantesco A, Brusaferro S. Psychiatric
disorders among hospitalized patients deceased
with COVID-19 in Italy. EClinicalMedicine 2021;
35: 100854.
242) Kadiwar S, Smith JJ, Ledot S, Johnson M, Bianchi
P, Singh N, Montanaro C, Gatzoulis M, Shah N,
Ukor EF. Were pregnant women more affected by 
COVID-19 in the second wave of the pandemic?
Lancet 2021; 397: 1539-1540.
243) Steptoe A, Di Gessa G. Mental health and social
interactions of older people with physical disabil-
ities in England during the COVID-19 pandemic :
a longitudinal cohort study. Lancet Public Heal
2021; 6: e365-e373.
244) Kuper  H,  Shakespeare  T.  Comment  Are  old-
er people with disabilities neglected in the
COVID-19 pandemic ? Lancet Public Heal 2021;
6: e347-e348.
... Patients infected with SARS-CoV-2 share similar pathological findings with those infected with SARS-CoV and MERS-CoV [42]. There was a significant decrease in CD4 and CD8 T cell counts following cytometric analysis of peripheral blood samples [43]. ...
Article
Full-text available
Coronavirus disease 2019 (COVID-19) is a novel and highly contagious viral infection caused by Sars-cov-2 and has been associated with a hyper-inflammatory immune response. This study aimed to evaluate changes in some inflammatory cytokines as clinical biomarkers to aid in the effective management of COVID-19 progression in Port Harcourt. A case-control study design was employed in this study, where a total of one hundred and ten (110) subjects were recruited, comprising fifty-five (55) COVID-19 positive subjects and fifty-five (55) COVID-19 negative subjects (controls), all between the ages of twenty (20) and seventy (70) years old, and both male and female subjects. Five milliliters (5 ml) of whole blood was collected using standard venipuncture Original Research Article Kpaluku et al.; Int. 45 technique with sterile hypodermic syringes and needles, aseptically, and dispensed into a plain bottle for the analysis of inflammatory cytokines. For the confirmation of COVID-19 positivity, a nasopharyngeal swab was collected aseptically using the RT-PCR technique. The results of this study revealed a significantly increased level of IL-6 (p = 0.0399) among subjects with COVID-19. The results from this study indicate significant alterations in inflammatory cytokines among subjects with COVID-19. It is necessary that tertiary health care settings and isolation centres consider IL-6 as a biomarker for effective management of patients with COVID-19 disease.
... This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/). The immune response is often seriously dysfunctional in severe Coronavirus disease 2019 [1][2][3][4][5][6][7]. In particular, patients with a severe form do not resolve the course of the disease despite prolonged and high neutralizing antibody titers against SARS-CoV-2. ...
Article
Full-text available
The immune response is a key player in the course of SARS-CoV-2 infection, and is often seriously dysfunctional in severe Coronavirus Disease 2019. The hyperinflammatory status has been described to be accompanied by the appearance of autoantibodies. In a lethal COVID-19 infection, we observed the emergence of a de novo natural alloantibody which targeted the M antigen from the MNS blood group on red blood cells (RBC) without evidence of any cross-reaction with SARS-CoV-2 antigens. This IgM lambda alloantibody was unmutated and unswitched. Here, we describe for the first time the emergence of a bystander de novo natural alloantibody against RBCs in a severe COVID-19 patient, highlighting the extra-follicular humoral response reported in these cases.
... Several studies have revealed that such a highly pathogenic coronavirus resulted in severe and acute respiratory distress syndrome (ARDS), resulting in diffused inflammation in the respiratory system, difficulty in breathing, and even death. In contrast, SARS-CoV-2 has a tendency to spread more rapidly than SARS-CoV-1 and MERS; however, the fatality rate tends to be lower than 2-3 % in comparison [5,6]. ...
Article
Full-text available
Background: As the therapeutic regimens against COVID-19 remain scarce, the microRNAs (miRNAs) can be exploited to generate efficient therapeutic targets. The miRNAs have been found to play pivotal roles in the several regulatory functions influencing the prognosis of viral infection. The miRNAs have a prospective role in the up and down-regulation of the ACE2 receptors. This review examines the clinical applications, as well as the possible threats associated with the use of miRNAs to combat the deleterious consequences of SARS-CoV-2 infection. Methodology: This article was compiled to evaluate how the miRNAs are involved in the SARS-CoV-2 pathogenesis and infection, and their potential functions which could help in the development of therapeutic targets against COVID-19. The sources of the collected information include several journals, databases and scientific search engines such as Google Scholar, Pubmed, Science direct, official website of WHO, among other sites. The investigations on the online platform were conducted using the keywords miRNA biogenesis, miRNA and ACE2 interaction, therapeutic role of miRNAs against SARS-CoV-2 and miRNA therapy side effects. Results: This review has highlighted that the miRNAs can be exploited to generate potential therapeutic targets against COVID-19. Changes in the miRNA levels following viral replication are an essential component of the host's response to infection. The collection and modification of miRNA modulates may help to minimize the deleterious consequences of SARS-CoV-2 infection, such as by controlling or inhibiting the generation of cytokines and chemokines. The degradation of viral RNA by the cellular miRNAs, along with the reduced expression of ACE2 receptors, can substantially reduce the viral load. Specific miRNAs have been found to have an antiviral influence, allowing the immune system to combat the infection or forcing the virus into a latency stage. Conclusion: This review summarizes several studies revealing the involvement of miRNAs in diverse and complex processes during the infection process of SARS-CoV-2. The miRNAs can substantially reduce the viral load by degradation of viral RNA and reduced expression of ACE2 receptors, besides mitigating the deleterious consequences of the exaggerated secretion of cytokines. Extensive investigations need to be done by the scientific community to utilize the miRNA based strategies for the development of effective therapeutic targets against COVID-19.
... [9]. Although the coronavirus specifically infects the respiratory system, it also has contributed to psychological illnesses such as depression, anxiety, and sleep disorders [10,11]. Psychiatric symptoms associated with COVID-19 have also been reported, mostly among patients with preexisting psychiatric illness presenting with a relapse of their condition owing to the COVID-19 infection [6]. ...
Article
Full-text available
Background The coronavirus disease 2019 syndrome typically consists of respiratory symptoms and other general nonspecific symptoms. Psychotic manifestations of coronavirus disease 2019 attributable to severe acute respiratory syndrome coronavirus 2 infection are seldom reported. We report a case of coronavirus disease 2019 in a young West African male who had no known risk factors of psychiatric illness or past history of psychiatric disease presenting with acute psychosis. Case presentation Our patient, who was a young West African male, presented without the typical respiratory symptoms of coronavirus disease 2019 and also without a background history of psychiatric illness or any other significant stressors in his past or present social history. He had acute onset of psychotic symptoms consisting of visual and auditory hallucinations, delusions of persecution, and lack of insight. He was admitted and managed with antipsychotic medication and mood stabilizer. His laboratory workup was normal except for positive coronavirus disease 2019 polymerase chain reaction and his liver enzymes, which showed elevated gamma glutamyl transferase, a finding consistent with coronavirus disease 2019. His head computed tomography scan was also normal. The patient made a gradual recovery from his psychotic symptoms, with gain of insight 7 weeks after onset of symptoms, at which time his coronavirus disease 2019 test came back negative along with other laboratory parameters. He returned to work 12 weeks after his presentation and has been performing well. Conclusion Psychosis can be a primary presenting symptom in patients with coronavirus disease 2019, including those without respiratory symptoms.
Article
BACKGROUND: Middle East Respiratory Syndrome Coronavirus is a highly pathogenic virus that poses a significant threat to public health. OBJECTIVE: The purpose of this study is to develop and characterize novel mouse monoclonal antibodies targeting the spike protein S1 subunit of the Middle East Respiratory Syndrome Corona Virus (MERS-CoV). METHODS: In this study, three mouse monoclonal antibodies (mAbs) against MERS-CoV were generated and characterized using hybridoma technology. The mAbs were evaluated for their reactivity and neutralization activity. The mAbs were generated through hybridoma technology by the fusion of myeloma cells and spleen cells from MERS-CoV-S1 immunized mice. The resulting hybridomas were screened for antibody production using enzyme-linked immunosorbent assays (ELISA). RESULTS: ELISA results demonstrated that all three mAbs exhibited strong reactivity against the MERS-CoV S1-antigen. Similarly, dot-ELISA revealed their ability to specifically recognize viral components, indicating their potential for diagnostic applications. Under non-denaturing conditions, Western blot showed the mAbs to have robust reactivity against a specific band at 116 KDa, corresponding to a putative MERS-CoV S1-antigen. However, no reactive bands were observed under denaturing conditions, suggesting that the antibodies recognize conformational epitopes. The neutralization assay showed no in vitro reactivity against MERS-CoV. CONCLUSION: This study successfully generated three mouse monoclonal antibodies against MERS-CoV using hybridoma technology. The antibodies exhibited strong reactivity against MERS-CoV antigens using ELISA and dot ELISA assays. Taken together, these findings highlight the significance of these mAbs for potential use as valuable tools for MERS-CoV research and diagnosis (community and field-based surveillance and viral antigen detection).
Article
Full-text available
Simple Summary Many studies have evaluated the spread and transmission of SARS-CoV-2 and MERS-CoV in humans; however, the ability of the virus to infect pets, including dogs, has not been fully clarified. Accordingly, in this study, we evaluated the ability of SARS-CoV-2 and MERS-CoV to infect beagle dogs. Our results showed that dogs can be infected by both viruses. Viral shedding into nasal secretions, feces, and urine was observed, and the lung tissues from the dogs inoculated with SARS-CoV-2 or MERS-CoV showed pathological changes, as well as changes in their lactate dehydrogenase levels. Abstract The coronavirus disease 19 (COVID-19) pandemic, caused by the severe acute respiratory syndrome, coronavirus 2 (SARS-CoV-2), has resulted in unprecedented challenges to healthcare worldwide. In particular, the anthroponotic transmission of human coronaviruses has become a common concern among pet owners. Here, we experimentally inoculated beagle dogs with SARS-CoV-2 or Middle East respiratory syndrome (MERS-CoV) to compare their susceptibility to and the pathogenicity of these viruses. The dogs in this study exhibited weight loss and increased body temperatures and shed the viruses in their nasal secretions, feces, and urine. Pathologic changes were observed in the lungs of the dogs inoculated with SARS-CoV-2 or MERS-CoV. Additionally, clinical characteristics of SARS-CoV-2, such as increased lactate dehydrogenase levels, were identified in the current study.
Article
The devastating complications of coronavirus disease 2019 (COVID-19) result from an individual's dysfunctional immune response following the initial severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) infection. Multiple toxic stressors and behaviors contribute to underlying immune system dysfunction. SARS-CoV-2 exploits the dysfunctional immune system to trigger a chain of events ultimately leading to COVID-19. The current study identifies eighty immune system dysfunction-enabling toxic stressors and behaviors (hereafter called modifiable contributing factors (CFs)) that also link directly to COVID-19. Each CF is assigned to one of the five categories in the CF taxonomy shown in Section 3.3.: Lifestyle (e.g., diet, substance abuse); Iatrogenic (e.g., drugs, surgery); Biotoxins (e.g., micro-organisms, mycotoxins); Occupational/Environmental (e.g., heavy metals, pesticides); Psychosocial/Socioeconomic (e.g., chronic stress, lower education). The current study shows how each modifiable factor contributes to decreased immune system capability, increased inflammation and coagulation, and increased neural damage and neurodegeneration. It is unclear how real progress can be made in combatting COVID-19 and other similar diseases caused by viral variants without addressing and eliminating these modifiable CFs.
Article
Full-text available
Abstract: The current century has witnessed infections of pandemic proportions caused by Coron�aviruses (CoV) including severe acute respiratory syndrome-related CoV (SARS-CoV), Middle East respiratory syndrome-related CoV (MERS-CoV) and the recently identified SARS-CoV2. Significantly, the SARS-CoV2 outbreak, declared a pandemic in early 2020, has wreaked devastation and imposed intense pressure on medical establishments world-wide in a short time period by spreading at a rapid pace, resulting in high morbidity and mortality. Therefore, there is a compelling need to combat and contain the CoV infections. The current review addresses the unique features of the molecular virology of major Coronaviruses that may be tractable towards antiviral targeting and design of novel preventative and therapeutic intervention strategies. Plant-derived vaccines, in particular oral vaccines, afford safer, effectual and low-cost avenues to develop antivirals and fast response vaccines, requiring minimal infrastructure and trained personnel for vaccine administration in developing countries. This review article discusses recent developments in the generation of plant-based vaccines, therapeutic/drug molecules, monoclonal antibodies and phytochemicals to preclude and combat infections caused by SARS-CoV, MERS-CoV and SARS-CoV-2 viruses. Efficacious plant-derived antivirals could contribute significantly to combating emerging and re-emerging pathogenic CoV infections and help stem the tide of any future pandemics.
Article
Objectives: We sought to fill the research gap on the effects of statins on the risks of ischemic stroke and heart disease among individuals with human immunodeficiency virus infection, influenza, and severe acute respiratory syndrome associated-coronavirus (HIS) disorders. Methods: We enrolled a HIS cohort treated with statins (n = 4921) and a HIS cohort not treated with statins (n = 4921). The cumulative incidence of ischemic stroke and heart disease was analyzed using a time-dependent Cox proportional regression analysis. We analyzed the adjusted hazard ratio (aHR) and 95% confidence interval (CI) of ischemic stroke and heart disease for statins users relative to nonusers based on sex, age, comorbidities and medications. Results: The aHR (95% CI) was 0.38 (0.22-0.65) for ischemic stroke. The aHR (95% CI) of heart disease was 0.50 (0.46-0.55). The aHRs (95% CI) of statin users with low, medium, and high adherence (statin use covering <33%, 33%-66%, and >66%, respectively, of the study period) for the risks of ischemic stroke were 0.50 (0.27-0.92), 0.31 (0.10-1.01), and 0.16 (0.04-0.68) and for heart disease were 0.56 (0.51-0.61), 0.40 (0.33-0.48), and 0.44 (0.38-0.51), respectively, compared with statin nonusers. Conclusion: Statin use was associated with lower aHRs for ischemic stroke and heart disease in those with HIS disorders with comorbidities.
Article
Full-text available
PARP14 and PARP9 play a key role in macrophage immune regulation. SARS‐CoV‐2 is an emerging viral disease that triggers hyper‐inflammation known as cytokine storm. In this study, using in silico tools, we hypothesize about the immunological phenomena of molecular mimicry between SARS‐CoV‐2 Nsp3 and the human PARP14 and PARP9. The results showed an epitope of SARS‐CoV‐2 Nsp3 protein that contains consensus sequences for both human PARP14 and PARP9 that are antigens for MHC class 1 and 2, which can potentially induce an immune response against human PARP14 and PARP9; while its depletion causes a hyper‐inflammatory state in SARS‐CoV‐2 patients. This article is protected by copyright. All rights reserved.
Article
Full-text available
Coronavirus disease 2019 (COVID-19), caused by severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), is a severe pandemic of the current century. The vicious tentacles of the disease have been disseminated worldwide with unknown complications and repercussions. Advanced COVID-19 syndrome is characterized by the uncontrolled and elevated release of pro-inflammatory cytokines and suppressed immunity, leading to the cytokine storm. The uncontrolled and dysregulated secretion of inflammatory and pro-inflammatory cytokines is positively associated with the severity of the viral infection and mortality rate. The secretion of various pro-inflammatory cytokines such as TNF-α, IL-1, and IL-6 leads to a hyperinflammatory response by recruiting macrophages, T and B cells in the lung alveolar cells. Moreover, it has been hypothesized that immune cells such as macrophages recruit inflammatory monocytes in the alveolar cells and allow the production of large amounts of cytokines in the alveoli, leading to a hyperinflammatory response in severely ill patients with COVID-19. This cascade of events may lead to multiple organ failure, acute respiratory distress, or pneumonia. Although the disease has a higher survival rate than other chronic diseases, the incidence of complications in the geriatric population are considerably high, with more systemic complications. This review sheds light on the pivotal roles played by various inflammatory markers in COVID-19-related complications. Different molecular pathways, such as the activation of JAK and JAK/STAT signaling are crucial in the progression of cytokine storm; hence, various mechanisms, immunological pathways, and functions of cytokines and other inflammatory markers have been discussed. A thorough understanding of cytokines’ molecular pathways and their activation procedures will add more insight into understanding immunopathology and designing appropriate drugs, therapies, and control measures to counter COVID-19. Recently, anti-inflammatory drugs and several antiviral drugs have been reported as effective therapeutic drug candidates to control hypercytokinemia or cytokine storm. Hence, the present review also discussed prospective anti-inflammatory and relevant immunomodulatory drugs currently in various trial phases and their possible implications.
Article
Full-text available
The COVID-19 pandemic, driven by the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), has currently infected more than 61 million people and caused the death of more than 1.4 million around the world (November 2020 figures). As a result, many countries and health organizations have had to investigate the reasons for the pandemic and improve their health care programs to cope with the emergency while trying to develop an effective vaccine.
Article
Full-text available
Background: The trans-membrane protease serine 2 (TMPRSS2) is essential for severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) cell entry and infection. Efficacy and safety of TMPRSS2 inhibitors in patients with coronavirus disease 2019 (Covid-19) have not been evaluated in randomized trials. Methods: We conducted an investigator-initiated, double-blind, randomized, placebo-controlled multicenter trial in patients hospitalized with confirmed SARS-CoV-2 infection from April 4, to December 31, 2020. Within 48 h of admission, participants were randomly assigned in a 2:1 ratio to receive the TMPRSS2 inhibitor camostat mesilate 200 mg three times daily for 5 days or placebo. The primary outcome was time to discharge or clinical improvement measured as ≥2 points improvement on a 7-point ordinal scale. Other outcomes included 30-day mortality, safety and change in oropharyngeal viral load. ClinicalTrials.gov Identifier: NCT04321096. EudraCT Number: 2020-001,200-42. Findings: 137 patients were assigned to receive camostat mesilate and 68 to placebo. Median time to clinical improvement was 5 days (interquartile range [IQR], 3 to 7) in the camostat group and 5 days (IQR, 2 to 10) in the placebo group (P = 0·31). The hazard ratio for 30-day mortality in the camostat compared with the placebo group was 0·82 (95% confidence interval [CI], 0·24 to 2·79; P = 0·75). The frequency of adverse events was similar in the two groups. Median change in viral load from baseline to day 5 in the camostat group was -0·22 log10 copies/mL (p <0·05) and -0·82 log10 in the placebo group (P <0·05). Interpretation: Under this protocol, camostat mesilate treatment was not associated with increased adverse events during hospitalization for Covid-19 and did not affect time to clinical improvement, progression to ICU admission or mortality.
Article
Full-text available
Background There is concern about the increased risk for SARS-CoV-2 infection, COVID-19 severe outcomes and disparity of care among patients with a psychiatric disorder (PD). Based on the Italian COVID-19 death surveillance, which collects data from all the hospitals throughout the country, we aimed to describe clinical features and care pathway of patients dying with COVID-19 and a preceding diagnosis of a PD. Methods In this cross-sectional study, the characteristics of a representative sample of patients, who have died with COVID-19 in Italian hospitals between February 21st and August 3rd 2020, were drawn from medical charts, described and analysed by multinomial logistic regression according to the recorded psychiatric diagnosis: no PD, severe PD (SPD) (i.e. schizophrenia and other psychotic disorders, bipolar and related disorders), common mental disorder (CMD) (i.e. depression without psychotic features, anxiety disorders). Findings The 4020 COVID-19 deaths included in the study took place in 365 hospitals across Italy. Out of the 4020 deceased patients, 84 (2•1%) had a previous SPD, 177 (4.4%) a CMD. The mean age at death was 78.0 (95%CI 77.6–78.3) years among patients without a PD, 71.8 (95%CI 69.3–72.0) among those with an SPD, 79.5 (95%CI 78.0–81.1) in individuals with a CMD. 2253 (61.2%) patients without a PD, 62 (73.8%) with an SPD, and 136 (78.2%) with a CMD were diagnosed with three or more non-psychiatric comorbidities. When we adjusted for clinically relevant variables, including hospital of death, we found that SPD patients died at a younger age than those without a PD (adjusted OR per 1 year increment 0.96; 95% CI 0.94–0.98). Women were significantly more represented among CMD patients compared to patients without previous psychiatric history (aOR 1.56; 95% CI 1.05–2.32). Hospital admission from long-term care facilities (LTCFs) was strongly associated with having an SPD (aOR 9.02; 95% CI 4.99–16.3) or a CMD (aOR 2.09; 95% CI 1.19–3.66). Comorbidity burden, fever, admission to intensive care and time from symptoms’ onset to nasopharyngeal swab did not result significantly associated with an SPD or with a CMD in comparison to those without any PD. Interpretation Even where equal treatment is in place, the vulnerability of patients with a PD may reduce their chance of recovering from COVID-19. The promotion of personalised therapeutic projects aimed at including people with PD in the community rather than in non-psychiatric LTCFs should be prioritised.
Article
Full-text available
The high transmissibility of SARS-CoV-2 is related to abundant replication in the upper airways, which is not observed for the other highly pathogenic coronaviruses SARS-CoV and MERS-CoV. We here reveal features of the coronavirus spike (S) protein, which optimize the virus towards the human respiratory tract. First, the S proteins exhibit an intrinsic temperature preference, corresponding with the temperature of the upper or lower airways. Pseudoviruses bearing the SARS-CoV-2 spike (SARS-2-S) were more infectious when produced at 33°C instead of 37°C, a property shared with the S protein of HCoV-229E, a common cold coronavirus. In contrast, the S proteins of SARS-CoV and MERS-CoV favored 37°C, in accordance with virus preference for the lower airways. Next, SARS-2-S-driven entry was efficiently activated by not only TMPRSS2, but also the TMPRSS13 protease, thus broadening the cell tropism of SARS-CoV-2. Both proteases proved relevant in the context of authentic virus replication. TMPRSS13 appeared an effective spike activator for the virulent coronaviruses but not the low pathogenic HCoV-229E virus. Activation of SARS-2-S by these surface proteases requires processing of the S1/S2 cleavage loop, in which both the furin recognition motif and extended loop length proved critical. Conversely, entry of loop deletion mutants is significantly increased in cathepsin-rich cells. Finally, we demonstrate that the D614G mutation increases SARS-CoV-2 stability, particularly at 37°C, and, enhances its use of the cathepsin L pathway. This indicates a link between S protein stability and usage of this alternative route for virus entry. Since these spike properties may promote virus spread, they potentially explain why the spike-G614 variant has replaced the early D614 variant to become globally predominant. Collectively, our findings reveal adaptive mechanisms whereby the coronavirus spike protein is adjusted to match the temperature and protease conditions of the airways, to enhance virus transmission and pathology.
Article
Full-text available
Background: The COVID-19 pandemic has affected mental health, psychological wellbeing, and social interactions. People with physical disabilities might be particularly likely to be negatively affected, but evidence is scarce. Our aim was to evaluate the emotional and social experience of older people with physical disabilities during the early months of the COVID-19 pandemic in England. Methods: In this longitudinal cohort study, we analysed data from the English Longitudinal Study of Ageing collected in 2018-19 and June-July, 2020, from participants aged 52 years and older and living in England. Physical disability was defined as impairment in basic and instrumental activities of daily living (ADL) and impaired mobility. Depression, anxiety, loneliness, quality of life, sleep quality, and amount of real-time and written social contact were assessed online or by computer-assisted telephone interviews. Comparisons of experiences during the COVID-19 pandemic of people with and without a physical disability were adjusted statistically for pre-pandemic outcome measures, age, sex, wealth, ethnicity, presence of a spouse or partner, number of people in the household, and chronic pain. All participants with full data available for both surveys were included in the analyses. Findings: Between June 3 and July 26, 2020, 5820 participants responded, 4887 of whom had full data available for both the pre-pandemic measures and the COVID-19 survey and were included in the analysis. During the COVID-19 pandemic, significantly more people with ADL impairment had clinically significant symptoms of depression (odds ratio 1·78 [95% CI 1·44-2·19]; p<0·0001), anxiety (2·23 [1·72-2·89]; p<0·0001), and loneliness (1·52 [1·26-1·84]; p<0·0001) than people without ADL impairment. Significantly more people with ADL impairment also had impaired sleep quality (1·44 [1·20-1·72]; p<0·0001) and poor quality of life than people without ADL impairment. The results were similar when disability was defined by impaired mobility. People with ADL impairment had less frequent real-time contact (0·70 [0·55-0·89]; p=0·0037) and written social contact (0·54 [0·45-0·64]; p<0·0001) with family than people without ADL impairment. Results for social contact were similar when disability was defined by impaired mobility. Interpretation: People with physical disability might be at particular risk for emotional distress, poor quality of life, and low wellbeing during the COVID-19 pandemic, highlighting the need for additional support and targeted mental health services. Funding: Economic and Social Research Council/UK Research and Innovation, National Institute on Aging, National Institute for Health Research.
Article
Full-text available
The COVID-19 pandemic has demonstrated the serious potential for novel zoonotic coronaviruses to emerge and cause major outbreaks. The immediate animal origin of the causative virus, SARS-CoV-2, remains unknown, a notoriously challenging task for emerging disease investigations. Coevolution with hosts leads to specific evolutionary signatures within viral genomes that can inform likely animal origins. We obtained a set of 650 spike protein and 511 whole genome nucleotide sequences from 222 and 185 viruses belonging to the family Coronaviridae , respectively. We then trained random forest models independently on genome composition biases of spike protein and whole genome sequences, including dinucleotide and codon usage biases in order to predict animal host (of nine possible categories, including human). In hold-one-out cross-validation, predictive accuracy on unseen coronaviruses consistently reached ~73%, indicating evolutionary signal in spike proteins to be just as informative as whole genome sequences. However, different composition biases were informative in each case. Applying optimised random forest models to classify human sequences of MERS-CoV and SARS-CoV revealed evolutionary signatures consistent with their recognised intermediate hosts (camelids, carnivores), while human sequences of SARS-CoV-2 were predicted as having bat hosts (suborder Yinpterochiroptera), supporting bats as the suspected origins of the current pandemic. In addition to phylogeny, variation in genome composition can act as an informative approach to predict emerging virus traits as soon as sequences are available. More widely, this work demonstrates the potential in combining genetic resources with machine learning algorithms to address long-standing challenges in emerging infectious diseases.
Article
Middle East respiratory syndrome coronavirus (MERS-CoV) and severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) are two common betacoronaviruses, which are still causing transmission among the human population worldwide. The major difference between the two coronaviruses is that MERS-CoV is now causing sporadic transmission worldwide, whereas SARS-CoV-2 is causing a pandemic outbreak globally. Currently, different guidelines and reports have highlighted several diagnostic methods and approaches which could be used to screen and confirm MERS-CoV and SARS-CoV-2 infections. These methods include clinical evaluation, laboratory diagnosis (nucleic acid-based test, protein-based test, or viral culture), and radiological diagnosis. With the presence of these different diagnostic approaches, it could cause a dilemma to the clinicians and diagnostic laboratories in selecting the best diagnostic strategies to confirm MERS-CoV and SARS-CoV-2 infections. Therefore, this review aims to provide an up-to-date comparison of the advantages and limitations of different diagnostic approaches in detecting MERS-CoV and SARS-CoV-2 infections. This review could provide insights for clinicians and scientists in detecting MERS-CoV and SARS-CoV-2 infections to help combat the transmission of these coronaviruses.
Article
COVID‐19‐associated pulmonary aspergillosis (CAPA) has been reported worldwide. However, basic epidemiological characteristics have not been well established. In this systematic review and meta‐analysis, we aimed to determine the incidence and mortality of CAPA in critically ill patients with COVID‐19 to improve guidance on surveillance and prognostication. Observational studies reporting COVID‐19‐associated pulmonary aspergillosis were searched with PubMed and Embase databases, followed by an additional manual search in April 2021. We performed a one‐group meta‐analysis on the incidence and mortality of CAPA using a random‐effect model. We identified 28 observational studies with a total of 3,148 patients to be included in the meta‐analysis. Among the 28 studies, 23 were conducted in Europe, two in Mexico, and one each in China, Pakistan, and the United States. Routine screening for secondary fungal infection was employed in 13 studies. The modified AspICU algorithm was utilized in 15 studies and was the most commonly used case definition and diagnostic algorithm for pulmonary aspergillosis. The incidence and mortality of CAPA in the ICU were estimated to be 10.2% (95% CI, 8.0 ‐ 12.5; I2 = 82.0%) and 54.9% (95% CI, 45.6 ‐ 64.2; I2 = 62.7%), respectively. In conclusion, our estimates may be utilized as a basis for surveillance of CAPA and prognostication in the ICU. Large, prospective cohort studies based on the new case definitions of CAPA are warranted to validate our estimates.