ArticlePDF AvailableLiterature Review

Abstract and Figures

In late December 2019, a vtiral pneumonia with an unknown agent was reported in Wuhan, China. A novel coronavirus was identified as the causative agent. Because of the human-to-human transmission and rapid spread; coronavirus disease 2019 (COVID-19) has rapidly increased to an epidemic scale and poses a severe threat to human health; it has been declared a public health emergency of international concern (PHEIC) by the World Health Organization (WHO). This review aims to summarize the recent research progress of COVID-19 molecular features and immunopathogenesis to provide a reference for further research in prevention and treatment of SARS coronavirus2 (SARS-CoV-2) infection based on the knowledge from researches on SARS-CoV and Middle East respiratory syndrome-related coronavirus (MERS-CoV).
Content may be subject to copyright.
COVID-19: Molecular and
Cellular Response
Shamila D. Alipoor
1
, Esmaeil Mortaz
2,3
*, Hamidreza Jamaati
4
, Payam Tabarsi
2
,
Hasan Bayram
5
, Mohammad Varahram
6
and Ian M. Adcock
7,8
1
Molecular Medicine Department, Institute of Medical Biotechnology, National Institute of Genetic Engineering and
Biotechnology (NIGEB), Tehran, Iran,
2
Clinical Tuberculosis and Epidemiology Research Center, National Research Institute
of Tuberculosis and Lung Diseases, Shahid Beheshti University of Medical Sciences, Tehran, Iran,
3
Department of
Immunology, School of Medicine, Shahid Beheshti University of Medical Sciences, Tehran, Iran,
4
Chronic Respiratory
Diseases Research Center, National Research Institute of Tuberculosis and Lung Diseases (NRITLD), Shahid Beheshti
University of Medical Sciences, Tehran, Iran,
5
Department of Pulmonary Medicine, Koc University School of Medicine, Koc
University Research Center for Translational Medicine (KUTTAM), Istanbul, Turkey,
6
Mycobacteriology Research Center,
National Research Institute of Tuberculosis and Lung Diseases (NRITLD), Shahid Beheshti University of Medical Sciences,
Tehran, Iran,
7
National Heart and Lung Institute, Imperial College London and the NIHR Imperial Biomedical Research
Centre, London, United Kingdom,
8
Priority Research Centre for Asthma and Respiratory Disease, Hunter Medical Research
Institute, University of Newcastle, Newcastle, NSW, Australia
In late December 2019, a vtiral pneumonia with an unknown agent was reported in
Wuhan, China. A novel coronavirus was identied as the causative agent. Because of the
human-to-human transmission and rapid spread; coronavirus disease 2019 (COVID-19)
has rapidly increased to an epidemic scale and poses a severe threat to human health; it
has been declared a public health emergency of international concern (PHEIC) by the
World Health Organization (WHO). This review aims to summarize the recent research
progress of COVID-19 molecular features and immunopathogenesis to provide a
reference for further research in prevention and treatment of SARS coronavirus2
(SARS-CoV-2) infection based on the knowledge from researches on SARS-CoV and
Middle East respiratory syndrome-related coronavirus (MERS-CoV).
Keywords: COVID-19, cytokines, IL-6, IL-8, ACE2
INTRODUCTION
Coronavirus disease 2019 (COVID-19) is caused by a member of the Coronaviridae family. It was
initially reported in Wuhan, China and has expanded rapidly in the world and is now at pandemic
level (Lippi and Plebani, 2020). The novel b-CoV was named as SARS-CoV-2by the International
Virus Classication Commission (Wu F. et al., 2020).
Coronaviruses are enveloped positive-stranded RNA viruses, which replicate in the host cell
cytoplasm. They possess a 5capped RNA and also contain the longest RNA among all RNA viruses
with a length of ~30kbp containing 14 open reading frames (ORFs) (Marra et al., 2003;Perlman and
Netland, 2009). Coronaviruses are classied into four genera (a,b,g, and d) based on phylogeny (Su
et al., 2016).
The rst human coronavirus was described in 1965 from patients with the common cold for
which there is still no vaccine. Counting SARS coronavirus2 (SARS-Cov2), there are currently seven
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 5630851
Edited by:
Rachel L. Roper,
The Brody School of Medicine at East
Carolina University, United States
Reviewed by:
Teneema Kuriakose,
St. Jude Childrens Research Hospital,
United States
Monica Miranda-Saksena,
Westmead Institute for Medical
Research, Australia
*Correspondence:
Esmaeil Mortaz
emortaz@gmail.com
Specialty section:
This article was submitted to
Virus and Host,
a section of the journal
Frontiers in Cellular
and Infection Microbiology
Received: 17 May 2020
Accepted: 08 January 2021
Published: 11 February 2021
Citation:
Alipoor SD, Mortaz E, Jamaati H,
Tabarsi P, Bayram H, Varahram M and
Adcock IM (2021) COVID-19:
Molecular and Cellular Response.
Front. Cell. Infect. Microbiol. 11:563085.
doi: 10.3389/fcimb.2021.563085
REVIEW
published: 11 February 2021
doi: 10.3389/fcimb.2021.563085
human coronaviruses strains known to infect humans which
belong to the aand bcoronavirus genera (Table 1)(Su et al.,
2016;Rithanya and Brundha, 2020).
Coronaviruses gained notoriety with the outbreak of Severe
Acute Respiratory Syndrome (SARS) in 2002-2003. This led to
the isolation and identication of HCoV-NL63 and HCoV-
HKU1. Further, the emergence of the middle east respiratory
syndrome coronavirus (MERS-CoV) in 2012 revealed that these
pathogens frequently cross the species and can pose a serious risk
to human health.
Bats are the main reservoirs of a large variety of viruses
especially human a- and b-coronaviruses (Hu et al., 2015). Some
SARS-like coronavirus (SL-CoVs) isolated from the bats have
high genomic sequence similarity and receptor usage compared
to SARS-CoV. This suggests that the spread of a bat coronavirus
to man presents a major global health risk (Ge et al., 2013;
Menachery et al., 2015;Yang X. L. et al., 2015). In the case of the
novel b-CoV strain, genome sequencing revealed (Lu et al., 2020)
the new virus has 81% identity with the sequence of the bat-
derived SARS.
The majority of COVID 19 cases (about 80%) are
asymptomatic or show mild symptoms but a low percentage
experience severe respiratory failure (Surveillances, 2020).
Interestingly, the viral load in asymptomatic patients was
similar to that in symptomatic patients. In a study in china it
was reported that in people with normal CT scans and no clinical
symptoms, who were in close contact with conrmed virus-
positive patients; nasal and throat swab tests were positive on
days 7, 10, and 11 after contact (Zou et al., 2020). Furthermore,
SARS-CoV-2 RNA was detectable in stool, saliva and urine
samples as well as in gastrointestinal tissue in patients with
COVID-19 in China. Thus, the digestive tract should also be
considered as a route of infection (Xiao et al., 2020). These
ndings emphasize the potential of viral transmission of
asymptomatic patients and indicate the urgent need for
strategies revolving around case detection and isolation
(Surveillances, 2020).
Since Dec 2019, when the rst case of disease was reported,
there have been over 80 M conrmed cases and 1.7 M deaths
have been reported across 235 countries (https://www.who.int/
emergencies/diseases/novel-coronavirus-2019). Given its
properties and rapid spread, there is an emergent need to
expand our knowledge of the molecular features and immune
pathogenesis of COVID-19. This review summarizes recent
ndings on the potential mechanisms and clinical features of
COVID-19 and relies on our knowledge of SARS-CoV and
MERS-CoV. This work aims to provide a reference for further
research in the prevention and the treatment of SARS-CoV-
2 infection
VIROLOGY AND GENOME
The SARS-CoV-2 genome consist of 29,903nt (nucleotides) and
has been assigned GenBank accession number MN908947. RNA
from the virus is closely related to two bat derived SARS-like
coronaviruses SL-CoVZC45 and SL-CoVZXC21, with a
nucleotide identity of 88.1%, but is more distant from SARS-
CoV (about 79%) and MERS-CoV (about 50%) (Lu et al., 2020;
Wu F. et al., 2020).
The genome of SARS-Cov-2, similar to other CoVs, contains
ten open reading frames (ORFs). The rst ORF covers two-
thirds of the viral RNA, which encodes polypeptides of the viral
replicase-transcriptase complex (Fehr and Perlman, 2015). The
remaining ORFs translate four main structural proteins: spike
(S), envelope (E), nucleocapsid (N), and membrane (M)
proteins. The genome is packaged into a helical nucleocapsid
surrounded by a host-derived lipid bilayer (Fehr and
Perlman, 2015).
STRUCTURAL DETERMINANTS IN THE
PATHOGENICITY OF SARS-COV2
A key to tackling the new pandemic is a clear understanding of
the factors that determine the virus pathogenicity including the
mechanism underlying receptor recognition, cell entry, and the
processing of viral proteins inside cells.
Virus Receptor
Receptor recognition is crucial in viral infectivity, pathogenesis
and cell tropism and can be considered as the primary
determinant of virus pathogenicity.
Corona viruses use a variety of receptors to enter the cells.
Dipeptidyl protease 4 (DPP4) or CD26 is used by MERS-CoV
TABLE 1 | Human Corona virus overview.
Name Discovery/location Group Receptor Symptom
HCOV-229E 1966 aAminopeptidase N
(hAPN, CD13)
Common cold
HCoV-OC43 1967 ß 9-O-acetylated sialic
acid
Common cold
SARS-CoV 2003/china ß ACE2/CD209L Severe respiratory illness
HCoV-NL63 2004/Netherlands aACE2 Lower respiratory tract infection, croup and bronchiolitis in children and immunocompromised
patients
HCoV-HKU1 2005/HongKong ß 9-O-acetylated sialic acid Chronic pulmonary disease in Hong Kong
MERS-CoV 2012/Middle East ß DPP4 Severe respiratory illness
Sars-Cov2 2019/China ß ACE2 Severe respiratory illness
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 5630852
(Wang et al., 2013) and angiotensin converting enzyme 2 (ACE2)
by SARS-CoV (Jeffers et al., 2004) and NL63-Cov (Hofmann
et al., 2005). CD209L is also used by SARS-CoV as an alternative
receptor (Jeffers et al., 2004)(Table 1). Based on viral genome
analysis of the new virus and high similarity to SARS-CoV, it is
likely that SARS-CoV-2 uses ACE2 as its receptor for cellular
invasion (Xu H. et al., 2020;Wan Y. et al., 2020). Further
evidence for ACE2 being the SARS-CoV-2 receptor is provided
by the fact that Baby Hamster Kidney broblasts (BHk) cells
transfected with human ACE2 become susceptible to SARS-
CoV-2 infection (Walls et al., 2020).
The ACE2 enzyme is involved in the reninangiotensin
aldosterone system (RAAS) activation. The RAAS includes a
cascade of vasoactive peptides that regulate the blood volume
and systemic vascular resistance in a prolonged manner
(Dronavalli et al., 2008). ACE2 is an exo-peptidase that
catalyzes the conversion of angiotensin (Ang I) I to the
nonapeptide Ang-(1-9) or the conversion of angiotensin II to
the heptapeptide Ang-(1-7) (Dronavalli et al., 2008).
Previous studies demonstrated a positive correlation between
the expression pattern of ACE2 and the infectivity of SARS-CoV.
In addition, some ACE2 variants interact with the SARS-CoV-2
or NL635 S-proteins with a lower-afnity (Mathewson et al.,
2008). Therefore, the pattern of ACE2 expression in different
tissues can determine tropism, susceptibility, symptoms, and
outcome of SARS-CoV-2 infection (Chen J. et al., 2020). ACE2
is expressed on the mucosa of oral cavity and is highly enriched
in epithelial cells of the tongue which highlights the role of the
oral cavity for infection with SARS-CoV-2 (Xu H. et al., 2020).
ACE2 is also highly expressed in vascular endothelial cells of the
heart and the kidney, and it affects cardiac function (Luft, 2014;
Badae et al., 2019).
COVID-19 has a higher mortality rate in males compared to
females (Xie et al., 2020) despite both genders having a similar
infection rate (The Sex, Gender, and COVID-19 Project, 2020).
There are differences in the reported male-biased mortality
between countries ranging from 59%75% of total mortality
(Pradhan and Olsson, 2020;Jin et al., 2020). Biological (immune
responses) and behavioral factors (e.g., smoking, mask wearing
and other lifestyle habits) may be responsible for placing men at
a greater risk of infection by SARS-CoV-2 or the consequences of
COVID-19 infection (Galasso et al., 2020). However, in a few
countries such as India, Nepal, Vietnam, and Slovenia the
COVID-19 case fatality rate is higher in women than men
(The Sex, Gender, and COVID-19 Project, 2020). These
differential ndings may be due to the incomplete data in the
case identication by sex or demographic factors such as co-
morbidities that increase the health risk for women in these
regions (Dehingia and Raj, 2020).
A recent study reported that Asian males have higher
expression of ACE2 (Zhao et al., 2020). It was also previously
reported that German cases showed mild clinical symptoms for
SARS without severe illness (Cao et al., 2020). However, the
pattern of ACE2 expression and function in different populations
remains to be investigated (Cao et al., 2020). For example, the
rate of COVID 19 related death is signicantly higher among the
Afro-Caribbean and South-East Asian people. Data to date
indicates that the conrmed cases of COVID-19 in black
counties is more than 3-fold-, and the rate of death is 6-fold-,
higher than that in white counties in the USA (Yancy, 2020;
Thebault R and Williams, 2020). This stark difference in the rate
of disease and outcome in the black population may be explained
in part, but not fully, by concomitant comorbidities (Giudicessi
et al., 2020). Comorbidities such as hypertension, metabolic
syndrome, diabetes, and cardiovascular disease (CVD) are the
main risk factors for worse outcome and mortality in patients
with COVID-19 (Bansal, 2020) and these are prevalent in Afro-
Caribbean and south-East Asian populations (Owczarek
et al., 2018).
Spike Glycoprotein (S Protein)
The viral spike glycoprotein (S protein) binds to the host cellular
receptor and is therefore, considered as the main determinant of
cell tropism and pathogenesis (Kuba et al., 2010). In corona
viruses; the spike protein is a large transmembrane protein and is
highly glycosylated. Spike proteins assemble as homo-trimers on
the virion surface to form a crown-like appearance or corona
(Belouzard et al., 2012).
The S protein is composed of an extracellular and a
transmembrane domain (TM), as well as a short cytoplasmic
tail region (CP) (Figure 1A). The extracellular domain of the S
protein is composed of two subunits (S1 and S2) which are
responsible for host-cell receptor recognition and membrane
fusion, respectively. S1 binds to ACE2 via the receptor-binding
domain (RBD) (Bosch et al., 2003).
A three-dimensional (3-D) atomic-scale map of the RBD of
SARS-CoV-2 in complex with human ACE2 has been obtained
(https://www.rcsb.org/structure/6VYB) (Figure 1B)(Shang
et al., 2020a). This analysis shows that the RBD of the SARS
CoV-2 S protein is more compact and binds human ACE-2 four
times more strongly than the SARS-CoV S protein (Shang et al.,
2020a). The RBD is the most variable part of the coronavirus
genome. Six RBD amino acids are critical for ACE2 binding (Tai
et al., 2020). Five of these six residues differ between SARS-CoV-
2 and SARS-CoV, which accounts for the higher binding afnity
of the SARS-CoV-2 S protein for human ACE2 (Andersen
et al., 2020).
Even though the SARS-CoV-2 RBD has a higher binding
afnity than SARS-CoV RBD for human ACE2, the binding
afnity of the entire SARS-CoV-2 S protein for human ACE2 is
lower than, or comparable to, that of the SARS-CoV S protein
(Shang et al., 2020b). This paradox can be explain based on the
dynamic state of the RBD. Cryo-electron microscopy studies
captured two states of the RBD: either buried (lying state) or
exposed (standing state), illustrating an inherently exible RBD
readily recognized by its receptor (Yuan et al., 2017)
(Figure 1C).
The RBD in the SARS-CoV S protein is mostly in the
standing-up state, however, it is mostly in the lying-down state
in SARS-CoV-2 and is not exposed and unable to bind ACE2.
Thus, although the SARS-CoV-2 RBD is more potent, it is less
exposed than the SARS-CoV RBD and the overall entry
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 5630853
efciencies of SARS-CoV-2 and SARS-CoV are comparable
(Shang et al., 2020b). The hidden RBD may also be responsible
for inefcient immune response and prolonged recovery time as
well as for the long incubation period. Many SARS-CoV-2
patients develop low levels of neutralizing antibodies and suffer
prolonged illness.
Proteolytic Processing
Host protease activation is the other signicant determinant of
coronavirus infection and pathogenesis. Coronavirus entry is
tightly regulated by the expression and activation of host proteases.
The presence of a proteolytic cleavage site within the S protein
mediates membrane fusion and viral infectivity. Sequence
variation around this cleavage site impacts severely on cellular
tropism and pathogenesis of CoVs (Kido et al., 2012;Coutard
et al., 2020). In the case of H5N1 hemagglutinin HA, insertion of
a multi-basic motif created a typical furin-like cleavage site in the
S protein which was responsible for the hyper-virulence of the
virus during the Hong Kong 1997 outbreak (Claas et al., 1998).
MERS-CoV also uses furin to enter the cells (Millet and
Whittaker, 2014).
In the case of the inuenza viruses, the highly pathogenic
forms contain a furin-like cleavage site that can be cleaved by
different host proteases allowing the virus to infect a broad range
of cells (Kido et al., 2012). The low-pathogenicity forms are only
cleaved by trypsin-like proteases and so virus infectivity is
limited to the respiratory and/or intestinal organs due to tissue
distribution of the activating protease(s) (Sun et al., 2010;
Coutard et al., 2020).
Interestingly, a furin-like cleavage site has been identied in
the SARS-CoV-2 S protein that is absent in the other SARS-like
CoVs and may be responsible for its high pathogenicity (Coutard
et al., 2020).
Sequence analysis of the S protein in SARS-CoV-2 reveals the
presence of a four amino acid residue insertion; RRAR;
producing a furin-sensitive motif at the boundary between the
S1 and S2 subunits (Figure 1). Interestingly this furin cleavage
site is conserved among all the isolated and sequenced SARS-
CoV-2 virions (Shang et al., 2020a) and its abrogation reduced the
efciency of viral entry (Shang et al., 2020b;Letko et al., 2020).
Processing of the S1/S2 site occurs during viral packaging in
infected cells, presumably by furin in the Golgi compartment
(Shang et al., 2020b;Letko et al., 2020).
The S protein is further cleaved by host proteases at the S2
site located immediately upstream of the fusion peptide. This
cleavage is proposed to activate the protein for membrane fusion
A
BC
FIGURE 1 |(A) Schematic of genome encoding spike protein in SARS-CoV2. SP, signal peptide; NTD, N-terminal domain; RBD, receptor-binding domain; RBM,
receptor-binding motif; FP, fusion peptide; HR1 and HR2, heptad repeat regions 1 and 2; TM, transmembrane; CP, cytoplasmic domain. (B) Schematic drawing of
the structure of coronavirus spike. S1, receptor-binding subunit (homotrimers); S2, membrane fusion subunit; TM, transmembrane anchor; IC, intracellular tail.
(C) SARS-CoV-2 spike ectodomain structure (open state) (details provided at https://www.rcsb.org/structure/6VYB).
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 5630854
via extensive irreversible conformational changes (Letko
et al., 2020).
Transmembrane serine protease2 (TMPRSS2) and lysosomal
cathepsins both have cumulative effects with furin on activating
SARS-CoV-2 entry. However, the furin pre-activation in
producer cells allows SARS-CoV-2 to be less dependent on
target cell proteases and increases its ability to enter cells with
relatively low expressions of TMPRSS2 or lysosomal cathepsins
(Shang et al., 2020b;Hoffmann et al., 2020). On the other hand,
prior cleavage at the S1/S2 site increases cleavage at the S2site
(Kuba et al., 2010).
TMPRSS2 activity is important for SARS-CoV-2 cell entry
(Braun and Sauter, 2019). The expression of TMPRSS2 and
TMPRSS4 in ACE2+ mature enterocytes in the human small
intestine facilitates virus entry into these cells. This property
makes the intestine a potential site for SARS-CoV-2 infection,
which may progress to a systemic disease (Zang et al., 2020).
It is suggested that camostat mesylate, a TMPRSS2 inhibitor
approved for clinical use, may block viral entry and could be
considered as a potential treatment option (Hoffmann et al.,
2020)(Figure 2). However, in the case of SARS-CoV, the virus
can use endosomal cysteine proteases including cathepsins B
and L for S protein priming in TMPRSS2 negative cells
(Simmons et al., 2005).
MECHANISM OF VIRAL CELL ENTRY
A key to preventing the spread of SARS CoV-2 and the resultant
pandemic is to gain a clear understanding of its mechanism of
cell entry. The virus surface S protein mediates this process by
binding to ACE2. Cleavage of the S glycoprotein between the S1
and S2 domains starts during viral packaging. This process is
completed by the type II transmembrane serine protease
TMPRSS2 (Kreye et al., 2020) which results in activation of the
S2 subdomain.
The S2 subdomain then mediates the fusion of the viral
genome with the host cell membrane to create a pore allowing
the viral RNA and RNA-associated proteins to gain access to the
cytoplasm. Another possibility is that ACE2/SARS-CoV-2
complex undergoes endocytosis. The rapid endocytosis of
SARS-CoV-2 occurs through clathrin-mediated endocytosis
(Millet and Whittaker, 2014). However, it is not completely
clear how the viral genome of SARS-CoV-2 gains access to the
cytoplasm (Figure 3).
However; after entering the cell, the viral RNA is released into
the cytoplasm and translates viral proteins followed by viral
genome replication (Kuba et al., 2010). The newly formed
envelope glycoproteins are inserted into the membrane of the
endoplasmic reticulum or Golgi, forming nucleocapsids by
assembling the genomic RNA and nucleocapsid protein. Viral
particles then incorporate into the endoplasmic reticulum-Golgi
intermediate compartment (ERGIC) and the vesicles containing
the virus particles fuse with the plasma membrane to release the
virus (Chockalingam et al., 2016)(Figure 3).
IMMUNOLOGICAL OUTCOME AND
PATHOLOGICAL EFFECT OF COVID-19
During viral infections, the innate immune system acts as a rst
line defense to prevent viral invasion or replication. This innate
immune response utilizes pattern recognition receptors (PRRs)
to detect specic viral components such as viral RNA. Following
viral entry into the cell, single stranded RNA viruses are
recognized by PRRs such as Toll-like receptor TLR7 and TLR8,
RIG-I (retinoic acid-inducible gene I)-like receptors (RLRs), and
the NOD-like receptor, CARD-containing-2 (NLRC2), that are
expressed by airway epithelial cells and innate immune cells
including alveolar and tissue macrophages. RLRs are also able to
detect double-stranded (ds) RNA structures (Streicher and
Jouvenet, 2019). Sensing of ssRNA by PRRs results in the
FIGURE 2 | Blocking SARS coronavirus2 (SARSCoV2) cell entry. SARS-CoV-2 binds to ACE2 and then uses the transmembrane Serine Protease 2 (TMPRSS2)
for S protein priming and entry to cell. A TMPRSS2 inhibitor can blocked viral entry by binding to and inhibition of enzyme activity.
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 5630855
production of the Type-I and -III antiviral interferons (IFNs)
and chemokines. The activation of this IFN-mediated antiviral
response is the rst major defense mechanism against viral
infections (Streicher and Jouvenet, 2019).
Although a rapid and well-coordinated immune response is
necessary for a potent defense against viral infection, an excessive
inammatory response may lead to tissue damage at the systemic
level. The massive production of cytokines and chemokines
detected during COVID-19 infection, the so-called cytokine
storm, is mainly responsible for the broad and uncontrolled
tissue damage observed. The cytokine storm resembles the
cytokine release syndrome (CRS) and results in plasma leakage,
vascular permeability and disseminated vascular coagulation.
These excessive proinammatory host responses are major
factors in the pathological outcomes such as acute lung injury
(ALI) and acute respiratory distress syndrome (ARDS) seen in
severe SARS-CoV-2 infected patients (Bhaskar et al., 2020).
In addition to the dominant manifestation of respiratory
symptoms, some patients have severe cardiovascular damage
and individuals with underlying cardiovascular disease (CVD)
have an increased risk of death (Zheng Y.Y. et al., 2020). The
mechanism underling the acute myocardial injury might be
related to the high expression of ACE2 in the CVS (de Abajo
et al., 2020).
On the other hand, the combination of cytokine storm
together with respiratory dysfunction and hypoxaemia may be
a mechanism by which COVID-19 results in damage to
myocardial cells (Zheng Y.Y. et al., 2020).
Patients also show lymphopenia and pneumonia with
characteristic pulmonary ground-glass opacity changes on
chest CT (Wu et al., 2020a;Zhang et al., 2020). Other forms of
severity including myocarditis, arrhythmia, cardiogenic shock as
well as acute kidney injury have been reported in 10%30% of
ICU patients with conrmed COVID-19 (Murthy et al., 2020). In
particular, neurological manifestations (Mao et al., 2020a) and
pneumonia were reported in most hospitalized COVID-19
patients (Huang et al., 2020).
The Pathogenicity of COVID-19
Induced Pneumonia
Among the critically ill patients admitted to ICU, severe disease
with respiratory failure due to mucus plugs; severe pneumonia
and ARDS is reported in 60%70% of patients whilst sepsis and
septic shock is seen in 30% of patients (Halacli et al., 2020).
SARS-CoV-2 viruses preferentially infect type II epithelial
cells within the lungs. In addition, some in vitro studies, ex-vivo,
and in silico studies suggest that mucociliary and goblet cells are
also primary target cells for infection (Hui et al., 2020;Sungnak
et al., 2020;Mason, 2020). Upon infection, type II epithelial cells
undergo programmed cell death as a part of the virus replicative
cycle (Anderson et al., 2010). Since these cells are the main player
of surfactant production, the reduced surfactant in the alveoli
FIGURE 3 | The mechanism of SARS-CoV-2 cell entry and replication. The coronavirus spike (S) protein binds to angiotensin converting enzyme 2 (ACE2)
receptors. Cleavage of the S glycoprotein between the S1 and S2 domains, which begins during viral packaging by furin in the Golgi compartments, is completed by
the protease trans-membrane serine Protease 2 (TMPRSS2) and lysosomal Cathepsin which enables cell membrane-viral fusion and viral RNA release. This process
may either create a pore allowing the viral RNA and RNA-associated proteins to gain access to the cytoplasm (1) or, alternatively, the ACE2/SARS-CoV-2 complex
may be internalized by endocytosis and be uncoated in the acidic lysosomal environment to enable release of the single stranded viral RNA (ssRNA) into the cytosol
(2). The viral genome is then replicated and translated into viral proteins by the host cell machinery. The newly formed envelope glycoproteins are processed within
the Golgi. Further assembling of the genomic RNA and nucleocapsid protein results in the formation of viral particles which are released via vesicular exocytosis.
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 5630856
leads the alveoli to collapse which subsequently facilitate
pneumonia and ARDS in the severe patients (Ayala-Nunez
et al., 2019). In addition, viral infection of the airway epithelial
cells can cause high levels of virus-induced pyroptosis
with associated vascular leakage. Pyroptosis is a form of
programmed cell death which can be triggered by cytopathic
viruses (Chen et al., 2019).
In SARS patients, despite a decrease in nasopharyngeal viral
titers 1015 days after the onset of symptoms, the pathological
outcomes and alveolar damage continue to worsen (Peiris et al.,
2003). This suggests that much of the pathological damage seen
is due to the host immune response to infection. It is likely that
similar events are occurring with SARS-CoV-2. However, the
viral nasopharyngeal load may be very different from that in the
lungs as autopsy samples showed a high concentration of virus in
different organs including the lung and intestine (Gu et al., 2005;
Farcas et al., 2005).
It was also demonstrated that a cytokine storm and its
subsequently inammatory events trigger endothelial activation and
induce endotheliitis as well as progressive microvascular pulmonary
thrombosis in lung which lead to the disseminated intravascular
coagulation (DIC) and impaired lung microcirculation (Luks and
Swenson, 2020).
Overall, the main histological ndings in the lungs represents
patchy necrosis; hyaline membrane formation and hyperplasia of
type II pneumocystis that represent diffuse alveolar damage and
injury to the gas-exchange surfaces (Anderson et al., 2010). The
pathological lung damages in the novel viral disease may be due
to either directly viral destruction of alveolar and bronchial
epithelial cells or a cytokine storm (Xu Z. et al., 2020).
However, respiratory distress in COVID-19 patients may be
due to the viral access to CNS and induced damage in the
respiratory centers of the brain, making it more complicated to
manage these patients (Baig et al., 2020).
The Neurological Pathogenicity of
SARS-CoV-2
36.4% of patients with COVID-19 develop neurological
symptoms (Mao et al., 2020a;Mao et al., 2020b)andthe
neuro-invasive properties of SARS CoV-2 is now accepted.
The retrospective studies as well as case reports from different
region in the world; indicate that Covid-19 affects CNS in several
ways and leads to a broad spectrum of neurological symptoms
from a simple headache to more serious encephalitis (Sheraton
et al., 2020).
Autopsy reports have revealed edema and partial neuronal
degeneration in brain tissue in severe patients (Channappanavar
and Perlman, 2017) and the presence of SARS-CoV-2 RNA in
the cerebrospinal uid (CSF) was conrmed by genome
sequencing (Babaha and Rezaei, 2020). However, the
pathophysiological characteristics of SARS-CoV-2-associated
encephalitis are not fully understood.
The new virus may induces nerve damage through the several
mechanisms including direct infection or immune injury (Wu
et al., 2020b) which may result in edema and alterations in
consciousness (Ye et al., 2020).
Coronaviruses are able to reach to the CNS via the synapse
connected routes and retrograde transport (Perlman et al., 1990;
Ye et al., 2020). In the case of murine models of SARS or MERS-
CoVs infection, intranasal infection enables the virus to access
the brain via the olfactory nerves and rapidly infect specic brain
areas such as the brain stem and thalamus (Netland et al., 2008)
and cause to critical neuronal histopathological changes in these
areas. Since SARS-CoV-2 is mainly spread through the
respiratory system, retrograde transport through the olfactory
nerve may be the main route for transfer of virus to the central
nervous system (CNS) (Mori, 2015;Desforges et al., 2019). The
occurrence of loss of smell or hyposmia during the early phase of
COVID-19 infections should be taken into consideration for the
involvement of the CNS (Baig et al., 2020).
Another pathway proposed for the entry of SARS-CoV-2 into
the nervous system is via the blood circulation and disruption of
the blood-brain barrier (BBB). Noroviruses use different
mechanisms for disruption of BBB including direct infection of
BBB endothelial cells or enhancing BBB permeability by
alteration the expression of matrix metalloproteinases (MMPs)
and tight junction proteins (Al-Obaidi et al., 2018). Interestingly,
the integrity of the epithelialendothelial barrier is critically
compromised in critical COVID-19 cases (Li H. et al., 2020).
A Trojan horse mechanism has also been proposed to be used
by SARS-CoV-2 to reach the CNS (Park, 2020). Here, ACE2-
expressing CD68+CD169+ macrophages act as the carrier cell
and may contribute to viral spread in COVID-19 patients and
induce an enhanced inammatory response during SARS-CoV-2
infection (Park, 2020).
The highly specialized functions and limited regenerative
capacity of neurons means that chronic and latent CNS SARS-
CoV-2 infection may have long-term detrimental consequences
(Andries and Pensaert, 1980). SARS-CoV-2 could trigger CNS
degeneration and may be a potential trigger of future
neurodegenerative diseases such as Parkinsonsdiseaseor
multiple sclerosis (Serrano-Castro et al., 2020). Considering the
large number of patients involved globally, the risk of future
neurological disorders is worrying and the clinicians should pay
more attention to the neurologic symptoms in COVID-19
patients especially in the early phase of infection.
CELLULAR AND MOLECULAR
MECHANISMS OF DISEASE
Based on data from the hospitalized patients, the majority of
COVID 19 cases (about 80%) are asymptomatic or show mild
symptoms while the rest experience severe respiratory failure
(Surveillances, 2020). This may be explained by the fact that the
onset and development of COVID-19 depends upon virus
infectivity and the strength of the individuals immune
response. The viral factors include virus type, titer, viability
and mutations, whilst the host immune factors including age,
gender and genetics (such as HLA genes) which together
determine the duration and severity of the disease (Rithanya
and Brundha, 2020).
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 5630857
SARS-CoV-2 tends to have a long incubation period: 515
days on average. This long incubation period is due to its ability
to escape from host immune detection at the early stages of
infection (Prompetchara et al., 2020). At the initiation of
infection, inhaled SARS-CoV-2 binds to ACE2 on nasal
epithelial cells and starts replicating. Despite the low viral
burden and local propagation of virus, the virus can be
detected by nasal swabs by RT-PCR (Mason, 2020). There is a
low immune response at this step. The virus then replicates and
migrates through the conducting airways and a more robust
innate immune response is triggered. At this time, the disease
COVID-19 clinically manifests itself and early markers of the
innate immune response such as CXCL10 are detectable
(Mason, 2020).
CXCL10 is known as a disease marker in SARS (Xu Z. et al.,
2020) and its expression is signicantly increased in alveolar type
II cells in response to both SARS-CoV (Qian et al., 2013) and to
inuenza (Wang J. et al., 2011). Higher levels of interleukin IL-6
and IL-10, and lower levels of CD4+T and CD8+T are also
observed in patients with COVID-19 and these correlate with the
severity of disease (Wan S. et al., 2020).
In the remainder of the review, we discuss the potential
mechanisms by which SARS-CoV-2 modulates the host
immune system predominantly based on the strategies used by
SARS-CoV and MERS.
Innate Immune Responses
The innate immune response is the rst line of defense against
viral infection and has a determinant role in protective or
destructive responses upon infection (Kindler et al., 2016).
Upon viral infection, type I interferon (IFN) responses and its
downstream cascade are initiated in order to control viral
replication and induce an effective adaptive immune response
(Prompetchara et al., 2020).
In the case of SARS-COV and MERS; the virus modulate anti-
viral IFN responses by using different strategies that interfere
with IFN production and its downstream signaling pathways
(Kindler et al., 2016). This dampening procedure is closely
associated with disease severity. In two MERS-CoV patients
with different severities, the type I IFN response was
signicantly lower in the patient with a poor outcome (death)
than in the patient who recovered (Shokri et al., 2019). IFN
production requires the phosphorylation and activation of IFN-
regulatory factor3 (IRF3) which is inhibited following infection
with SARS-CoV (Kindler et al., 2016)(Figure 4).
Nuclear translocation of IRF3 and its subsequent stimulation
of IFN bexpression is inhibited by viral ORF4a, ORF4b, and
ORF5 proteins (Yang et al., 2013;Yang Y. et al., 2015). In
addition, the MERS-CoV accessory protein 4a blocks IFN
induction through direct interaction with double-stranded
RNA (Niemeyer et al., 2013). Based on the genomic similarity
between SARS-CoV-2, SARS-CoV and MERS-CoV it is likely
that SARS-CoV-2 utilizes similar strategies to modulate the type
I IFN response (Dandekar and Perlman, 2005).
However, sequence analysis showed several changes to
SARS-CoV-2 that indicate that SARS-CoV-2 is more sensitive
to type I IFN (Lokugamage et al., 2020). ORF3b from SARS-CoV
encodes a 154 amino acid (aa) protein that suppress the type I
IFN responses by inhibition of IRF3 phosphorylation. In SARS-
CoV-2 a premature stop codon in ORF3b results in a truncated
20aa protein which lacks an equivalent function to that seen with
SARS-CoV (Lokugamage et al., 2020). SARS-CoV-2 has similar
viral replication kinetics to SARS-CoV in Vero cells (Lokugamage
et al., 2020) but is much more sensitive to type I IFN pre-treatment
and a signicant reduction in viral replication is seen following
type I IFN treatment (Lokugamage et al., 2020b). Together these
observations suggest that type I IFN may be a potential treatment
for COVID-19 as seen in animal models of SARS-CoV and
MERS-CoV infection (Channappanavar et al., 2019). In
contrast, a delayed induction of type I IFN compromises the
early viral control and drives the inux of hyper-inammatory
neutrophils and monocytes-macrophages which lead to a mass
production of pro-inammatory cytokines and may evoke a
cytokine storm (Channappanavar and Perlman, 2017).
A cytokine storm leads to an uncontrolled systemic
inammatory response that may trigger a severe attack on the
body by its own immune system. As with SARS-CoV and MERS-
CoV infection, this may lead to ARDS, multiple organ failure and
death in severe cases of SARS-CoV-2 infection (Channappanavar
and Perlman, 2017;Tabarsi, 2020). In severe COVID-19
patients, serum levels of proinammatory cytokines including
IL-2, IL6, IL-7, IL-10, G-CSF, IP-10, MCP-1, MIP-1aand TNFa
are highly elevated (Xu Z. et al., 2020). Furthermore, increased
total blood neutrophils and decreased total blood lymphocytes in
patients within ICU compared with patients not in ICU care
correlated with disease severity and death (Prompetchara
et al., 2020).
The previous studies and clinical evidence also strongly
indicate the importance of the NLR family pyrin domain
containing 3 (NLPR3) inammasome in the pathologic effects
of severe COVID-19 (Shah, 2020). In patients with reduced
immune tness whose innate immune system is unable to clear
the virus, NLRP3 activation may occur (Freeman and Swartz,
2020). A sustained and unregulated NLRP3-dependent
inammatory response leads to the severe clinical symptoms
including necrosis, fever, release of damage-associated molecular
patterns (DAMP) and severe inammation (van den Berg and Te
Velde, 2020). Direct activation of NLRP3 induces pyroptosis and
cell death in human cells. Additionally, hyper-activation of
NLRP3 may result in coagulopathy, neutrophil inltration,
Th17 and macrophage activation and a cytokine storm in
severe COVID-19 patients (Merad and Martin, 2020). Genetic
variations in host inammasome pathways may also inuence
disease outcome and may be responsible for the heterogeneous
response of patients with COVID-19 and the array of clinical
severity (Freeman and Swartz, 2020).
Adaptive Immune Responses
Corona viruses including SARS and MERS have ability to infect
immune cells which plays a key role in the disease pathogenesis
(Dandekar and Perlman, 2005). In addition, MERS-CoV induces
rapid apoptosis of macrophages by the limiting of the early
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 5630858
induction of IFN (Shokri et al., 2019). In the case of SARSCoV,
infection of lymphocytes has been proposed to play the major
role in viral-induced pathogenicity. SARS-CoV-infected
lymphocytes, similar to feline infectious peritonitis virus
(FIPV)-infected macrophages in domestic cats, might transport
the virus to distant organs resulting in systemic infection (de
Groot-Mijnes et al., 2005).
Upon virus entry, immunogenic peptides are presented to T
cells in association with human leukocyte antigen (HLA) on the
surface of antigen presenting cells (APC). Polymorphisms in the
HLA system may inuence the susceptibility and outcome of
SARS-CoV infection (Yuan et al., 2014). Some HLA alleles
including HLA-B*4601, HLA-B*0703, HLA-DRB1*1202 HLA-
DRB4*01010101, and HLA-Cw*0801 are associated with
susceptibility to SARS-CoV infection (Wang S.-F. et al., 2011).
Whereas, the HLA-DR0301, HLA-Cw1502, and HLA-A*0201
alleles are protective against severe SARS infection (Hajeer et al.,
2016). Furthermore, polymorphisms within the MBL (mannose-
binding lectin) gene are also associated with susceptibility to
SARS-CoV infection (Tu et al., 2015).
In the case of MERS-CoV infection, HLA-DRB1*11:01 and
HLA-DQB1*02:0 alleles increased the risk of infection (Wang
S-F. et al., 2011). These alleles should be assessed in COVID-19
patients (Li X. et al., 2020). In addition, antigen presentation via
the class I and II major histocompatibility complex (MHC) was
down-regulated in MERS-CoV infection. This would markedly
diminish T cell activation in response to the virus (Shokri et al.,
2019). During viral infection, antigen presentation triggers
both humeral and cellular immunity mediated by virus-specic
B and T cells.
B Cell Responses
B cell proling by RNA-seq showed a specic pattern of new B
cell-receptor changes (IGHV323 and IGHV37) in COVID-19
patients. It is also indicated that the number of naïve B cells is
decreased while the number of peripheral blood plasma cells was
remarkably increased (Wen et al., 2020).
With SARS infection, the prole of neutralizing antibodies are
predominantly IgM and IgG. Specic B and T cells epitopes are
commonly mapped against the structural S and N proteins
(Berry et al., 2010;Liu et al., 2017). SARS-CoV infection
induces seroconversion 4 days after the onset of disease and
SARS-specic IgM antibodies are found in most patients up to
week 12 post-infection while the IgG antibody can persist for
much longer. Indeed, long lasting specic IgG was detected up to
2 years post infection (Li et al., 2003).
In COVID-19, SARS-CoV-2 specic antibodies may be
produced to neutralize the virus. In one patient, serology
reports showed the peak of specic IgM at day 9 after the
onset of the disease and switching to IgG by week 2 (Zhou P.
FIGURE 4 | The innate immune modulation mechanisms by SARS-CoV and Middle East respiratory syndrome-related coronavirus (MERS-CoV). Double stranded
(ds)RNA, a by-product of RNA virus replication in the cytoplasm, is sensed by the pattern recognition receptors such as retinoic-acid inducible gene I (RIG-I) and
melanoma differentiation-associated protein 5 (MDA5) which subsequently leads to activation of the kinase TANK Binding Kinase 1 (TBK1). These kinases then
phosphorylate interferon (IFN) regulatory factor 3 (IRF3) and trafc to the nucleus to active the transcription of IFNs. Viral proteins actively modulate this pathway.
Open reading frame (ORF)3b, nucleocapsid (N), and non-structural protein (NSP)1 affect the signal transduction pathway that activates IRF3. In addition, the papain-
like protease (PLP) blocks the phosphorylation of IRF3 and its activation. Viral protein 4a suppress the activation of RIG-I/MDA5 signaling and blocks the induction of
IFNs through interaction with dsRNA.
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 5630859
et al., 2020). Interestingly, sera from ve patients with conrmed
COVID-19 show some cross-reactivity with SARS-CoV.
Furthermore, all sera from patients were able to neutralize
SARS-CoV-2 in an in vitro plaque assay, suggesting a possible
successful mounting of the humoral responses (Zhou P
et al., 2020).
In a recent study of 285 patients with mild to severe disease,
96.8% of tested patients achieved seroconversion of IgG or IgM
within 20 days after the onset of symptoms with the titer
plateauing within 6 days after seroconversion. Moreover, 100%
of patients had positive virus-specic IgG approximately 1719
days after symptom onset. In addition, 94.1% patients were
positive for virus-specic IgM approximately 2022 days after
symptom onset (Liu et al., 2020;Long et al., 2020). Antibody
titers in the severe group were higher than those with milder
disease. Interestingly, a number of individuals with negative
nucleic acid results and no symptoms had positive IgG and/or
IgM tests (Long et al., 2020). This highlights the importance of
serological testing as a complement to the RT-PCR test in
surveying for asymptomatic patients in close contact with
other (Liu et al., 2020;Long et al., 2020).
T Cell Responses
During viral infection, T helper (Th) cells play an important role
in the adaptive immunity. The cytokine microenvironment
generated by antigen presenting cells directs T cell responses.
In SARS-CoV infection a strong Th1 cell response and higher
levels of neutralizing antibodies were observed in the mild-
moderate group, while in the fatal group a higher level of
Th2 cytokines (IL-4, IL-5, IL-10) was detected (Li et al., 2008).
Current evidence indicates that a Th1 response is crucial for the
successful control of SARS-CoV and MERS-CoV and this may
also be essential for SARS-CoV-2 (Li et al., 2008).
The balance between naïve and memory T cells is crucial to
controlling infection. Naïve T cells are responsible for the defense
against new, previously unrecognized infection by the
production of cytokines. In contrast, memory T cells promote
antigen-specic immune responses. Disruption of the balance in
favor of naïve T cells could strongly promote hyper
inammation. On the other hand, the reduction in memory T
cells could contribute to COVID-19 relapse, which is reported in
a number of recovered cases of COVID-19 (Zhou L. et al., 2020;
Yuan et al., 2020;Chen D. et al., 2020).
In severe patients with COVID 19, CD8+ T cell responses
were more frequent and robust than CD4+ T cell responses and
an early rise in CD8+ T cell numbers was correlated with disease
severity (Prompetchara et al., 2020). CD8+ T cells contained a
large number of cytotoxic granules which may have induced
severe immune injury in the patients (Xu Z. et al., 2020).
Additionally, a recent report of a 50 year old male with covid-
19whodiedshowedasignicantly reduced number of
peripheral blood CD4+ and CD8+ T cells. These cells had a
hyper-activated appearance and were double positive for HLA-
DR and CD38 (Xu Z. et al., 2020).
COVID-19 also induces T-cell exhaustion (Diao et al., 2020).
T cell exhaustion is a state of T cell dysfunction that arises during
many chronic infections as well as during persistent viral
infections (Wherry, 2011). Peripheral blood T cells from
COVID-19 patients expressed high levels of the exhaustion
markers PD-1 and Tim-3 (Diao et al., 2020).
Peripheral blood total counts of T cells, CD4+, and CD8+ T
cells were signicantly lower in ICU patients than non-ICU cases
with COVID-19 and counts were negatively correlated with
patient survival (Diao et al., 2020). It seems that the cytokine
storm may promote apoptosis or necrosis of T cells and thereby
reduce their numbers (Xi-zhi and Thomas, 2017). Furthermore,
the levels of TNF-a, IL-6, and IL-10 were signicantly increased
in infected patients and raisedevenmoreinpatientswho
required ICU treatment. Interestingly, the concentrations of
these cytokines were negatively correlated with total T cell
counts (Diao et al., 2020).
Besides suppression of T cell proliferation, IL-10 can induce T
cell exhaustion by increased levels of PD-1 and Tim-3 (Brooks
et al., 2006). Inhibition of IL-10 also reduced the degree of T cell
exhaustion in animal models of chronic infection (Ejrnaes et al.,
2006). TNF-ais a pro-inammatory cytokine that promotes T
cell apoptosis via the TNF receptor TNFR-1 signaling pathway.
The expression of TNFR-1 is increased in aged T cells (Aggarwal
et al., 1999). IL-6 contributes to host defense by stimulating acute
phase responses or immune reactions in response to infections
and tissue injury. Deregulated and continual synthesis of IL-6
plays a pathological role in chronic inammation and infection
(Jones and Jenkins, 2018). The level of IL-6 in severe patients
continue to increase over time and is higher in non-survivors
(Zhou F et al., 2020).
It is suggested that the different mortality rates in COVID-19
patients is due to the difference in the response to infection
(Bienvenu et al., 2020). Analysis of viral RNA in COVID-19
patients indicated that the males show delayed viral clearance
(Xu K. et al., 2020;Zheng S. et al., 2020). Furthermore, it is
reported that male patients had higher plasma levels of innate
immune cytokines including IL-8 and IL-18 along with activated
non-classical monocytes. In contrast, female patients had a more
robust T cell activation during SARS-CoV-2 infection. A poor T
cell response may be responsible for the worse outcome in male
patients, while in female patients, higher levels of innate immune
cytokines were associated with worse disease (Takahashi
et al., 2020).
COVID-19 AND POST
INFECTION IMMUNITY
Another question regarding this new disease is that whether the
infection induces persistent immune memory that could protect
the recovered individual against reinfection. The durability of
protective antibodies induced by SARS-CoV-2 or the antibody
titers that will protect against reinfection is still unclear. It is
reported that antibodies disappear rapidly after recovery
particularly in patients with mild disease. For example, IgG has
a half-life of approximately 21 days in SARS-CoV-2 patients.
However; even in the absence of specic serum antibodies, the
presence of memory B and T cells may be maintained (Cox and
Brokstad, 2020).
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 56308510
In a recent study, COVID-19 induced memory lymphocytes
with an antiviral protective immune function. This longitudinal
study assessed the immune response in patients who recovered
from mild symptomatic COVID-19. These recovered individuals
developed SARS-CoV-2-specic IgG antibody and neutralizing
plasma in addition to virus-specic memory B and T cells that
had ability to expand over three months following the onset of
symptoms (Rodda et al., 2020). Furthermore, following antigen
re-exposure the memory T cells secreted IFN-gand were able to
clonally expand whilst memory B cells expressed antibodies
capable of neutralizing the virus (Rodda et al., 2020).
Studies of patients infected with SARS-CoV in 2003 suggested
that the infection induced durable T cell responses lasting for up
to 6 years but no prolonged memory B cells. Importantly, these T
cells could cross-react with the SARS-CoV-2 virus after 17 years,
but it is not clear that whether they can provide protection
against COVID-19 (Le Bert et al., 2020). On the other hand,
recent reports demonstrated that T cell reactivity against SARS-
CoV-2 exists in many unexposed people. It is hypothesized that
this might be due to immunity to common cold coronaviruses
that could inuence COVID-19 disease severity (Sette and
Crotty, 2020).
POTENTIAL PHARMACOLOGIC
STRATEGIES IN THE TREATMENT
AND PROTECTION
At present, there are no specic antiviral drugs or a vaccine
available to treat COVID-19. Currently, broad-spectrum
antiviral drugs like nucleoside analogues and HIV-protease
inhibitors are being used to attenuate viral infection (Lu,
2020). Other treatment regimens include a combination of
oral oseltamivir, lopinavir and ritonavir and intravenous
administration of ganciclovir for 314 days (Chen N. et al.,
2020). There is an ongoing clinical trial for examination of the
safety and efcacy of lopinavir-ritonavir and IFN-2b in patients
with COVID-19 (Lu, 2020;Habibzadeh and Stoneman, 2020).
Although the antivirals remdesivir and chloroquine controlled
SARS-CoV-2 infection in vitro, these have shown variable results
in the clinic (Wang et al., 2020). EIDD-2801 is a new drug used
to control and treat seasonal and pandemic inuenza virus
infections. It has high therapeutic efcacy in humans and may
be considered as a potential drug for the treatment of COVID-19
infection (Toots et al., 2019).
Therapies based on the ACE2 receptor have raised concerns
about the use of Renin-Angiotensin-Aldosterone System (RAAS)
inhibitors that may alter ACE2 function and expression.
Angiotensin II receptor blockers (ARBs) and other RAAS
inhibitors increased ACE2 expression in the lung in animal
models (Gurwitz, 2020). Data from cohort and cross sectional
studies in patients with heart failure or cardiovascular disease
showed that the effect of these inhibitors on ACE2 is not uniform
even in response to a given drug class (Shearer et al., 2013). In
addition, recent evidence suggests that ACE inhibitors do not
affect COVID-19 infection or severity (Bryan Williams, 2020).
Angiotensin receptor 1 (AT1R) blockers, such as losartan,
have also been suggested as a therapeutic approach for
reducing the severity and mortality of SARS-CoV-2 infections
(Gurwitz, 2020).
Interestingly, human recombinant soluble ACE2 (hrsACE2)
blocked SARS-CoV-2 infections in engineered human tissues,
which suggests promise as a treatment capable of stopping early
infection of the novel coronavirus. In this regard, APN01
developed by APEIRON is currently in clinical trials (Zoufaly
et al., 2020). SARS-CoV-2 uses the serine protease TMPRSS2 for
S protein priming. It is suggested that previously approved
TMPRSS2 inhibitors such as nafamostat mesylate and
camostat mesylate (Hoffmann et al., 2020)willblockviral
entry and may be a possible treatment option (Tay et al., 2020).
The central role of cytokine dysregulation in the pathogenicity
of seriously ill COVID-19 patients has indicated the potential of
cytokine-targeted therapy in managing disease progression
(Rahmati M. 2020). Drugs such as IL-6 inhibitors (tocilizumab,
sarilumab, and siltuximab) or IL1-binhibitors are possible
interventions in this area.
Tocilizumab (Actemra) is a humanized anti-IL-6 receptor
antibody that was successful in the treatment of rheumatoid
arthritis (RA) and juvenile idiopathic arthritis (Burmester et al.,
2016;Yokota et al., 2016). Whether tocilizumab restores T cell
counts in COVID-19 patients by suppressing IL-6 signaling
needs to be investigated (Diao et al., 2020;Atal and Fatima,
2020). Interleukin-1 targeting with a high-dose of anakinra in
patients with COVID-19 was safe and showed clinical
improvement in 72% of patients in a retrospective cohort
study (Cavalli et al., 2020). Anakinra is a recombinant
interleukin-1 (IL-1) receptor antagonist that has anti-
inammatory and immunomodulatory effect used to treat of
inammatory arthritides (Furst, 2004).
The NLRP3 inammasome with its downstream pathways is
also an attractive target for therapy of COVID-19 (Pontali et al.,
2020;Jamilloux et al., 2020;Soy et al., 2020). The rst clinical trial
study of using tranilast (an NLRP inammasome inhibitor)
to treat COVID-19 is ongoing and registered in the Chinese
clinical trial registry (http://www.chictr.org.cn/showprojen.aspx?
proj=49738).
SARS-CoV-2 is more sensitive to type I IFN in vitro and in
vivo than SARS suggesting that type I IFNs may be a potential
treatment for protection against COVID19. Interestingly,
Bacillus CalmetteGuerin (BCG) vaccination may reduce the
mortality associated with COVID-19. BCG vaccine stimulates
the production of IFNs by the innate immune system in a TLR2-
dependent manner (Ades, 2014;Kativhu and Libraty, 2016) and
may provide protection against COVID19. IFN-bmay prove
efcacious as a vaccine adjuvant to boost immune cell function
and INF production (Ades, 2014).
Remdesivir an antiviral compound that was rstly introduced
for Ebola virus (Wang et al., 2020) and was successful in
shortening the time to recovery in adults hospitalized with
COVID-19 (Lu, 2020;Beigel et al., 2020;Harapan et al., 2020).
There are ongoing clinical trials in a number of countries for
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 56308511
remdesivir as a potential treatment for COVID-19. Remdesivir,
made by Gilead, is currently going through the FDA approval
process and is authorized in the United States for use under an
Emergency Use Authorization (EUA). Currently, the drug is
given intravenously through daily infusions in the hospital.
However, an inhaled formulation given through a nebulizer,
may potentially allow for easier administration outside the
hospital at earlier stages of the disease.
These approaches would be valuable to investigate for
COVID-19 and open a window for nding new ways to
protect against and treat this deadly epidemic. There are
currently 109 ongoing clinical trials for COVID-19 registered
on clinicaltrials.gov. These include numerous pharmacological
and biological approaches and the use of natural products.
CONCLUSION
The COVID-19 pandemic is rapidly spreading across the world
and has caused more deaths compared with SARS or MERS.
Despite the high rate of mortality and morbidity, no medication
or vaccine has been consistently shown to be effective. SARS-
CoV-2 belongs to the Coronaviridae family which was also
responsible for previous widespread outbreaks of SARSCoV in
2002 and MERSCoV in 2008. The rapid spread, induction of
severe infection, cross-species transmission and unpredicted
behavior of coronaviruses result in them being a continuous
threat to human health. This is important due to the existence
of many animal reservoirs for CoVs and the lack of an
approved treatment.
There is an imperative need to design and develop effective
therapeutic and preventive strategies. This will require using the
accumulated knowledge of how the host immune response system
responds to SARS and MERS but importantly will require analysis
of immune responses within the lungs of COVID-19 patients.
Techniques such as single cell sequencing and cellular indexing of
transcriptomes and epitopes by sequencing (CITE-seq) will
tremendously increase our understanding of disease pathology
and enable knowledge-based decisions for the adoption of new
therapeutic immune modalities.
AUTHOR CONTRIBUTIONS
SA wrote rst draft. EM, HJ, PT, IMA, MV, and HB have revised
equally the rst draft. All authors contributed to the article and
approved the submitted version.
FUNDING
IMA is nancially supported by the British Heart Foundation
(PG/14/27/30679), the Dunhill Medical Trust (R368/0714), the
Welcome Trust (093080/Z/10/Z), the EPSRC (EP/T003189/1),
the Community Jameel Imperial College COVID-19 Excellence
Fund (G26290) and by the UK MRC (MR/T010371/1).
REFERENCES
Ades, P. A. (2014). A controversial step forward: A commentary on the 2013 ACC/
AHA guideline on the treatment of blood cholesterol to reduce atherosclerotic
cardiovascular risk in adults. Cor. Artery Dis. 25 (4), 360363. doi: 10.1097/
MCA.0000000000000086
Aggarwal, S., Gollapudi, S., and Gupta, S. (1999). Increased TNF-a-induced
apoptosis in lymphocytes from aged humans: changes in TNF-areceptor
expression and activation of caspases. J. Immunol. 162 (4), 21542161.
Al-Obaidi, M., Bahadoran, A., Wang, S., Manikam, R., Raju, C. S., and Sekaran, S.
(2018). Disruption of the blood brain barrier is vital property of neurotropic
viral infection of the central nervous system. Acta Virol. 62 (1), 1627. doi:
10.4149/av_2018_102
Andersen, K. G., Rambaut, A., Lipkin, W. I., Holmes, E. C., and Garry, R. F. (2020).
The proximal origin of SARS-CoV-2. Nat. Med. 26 (4), 450452. doi: 10.1038/
s41591-020-0820-9
Anderson, D. J., Politch, J. A., Nadolski, A. M., Blaskewicz, C. D., Pudney, J., and
Mayer, K. H. (2010). Targeting Trojan Horse leukocytes for HIV prevention.
AIDS 24 (2), 163187. doi: 10.1097/QAD.0b013e32833424c8
Andries, K., and Pensaert, M. B. (1980). Immunouorescence studies on the
pathogenesis of hemagglutinating encephalomyelitis virus infection in pigs
after oronasal inoculation. Am. J. Vet. Res. 41 (9), 13721378.
Atal, S., and Fatima, Z. (2020). IL-6 Inhibitors in the Treatment of Serious
COVID-19: A Promising Therapy? Pharm. Med. 32, 221223. doi: 10.1007/
s40290-020-00342-z
Ayala-Nunez, N. V., Follain, G., Delalande, F., Hirschler, A., Partiot, E., Hale, G. L.,
et al. (2019). Zika virus enhances monocyte adhesion and transmigration
favoring viral dissemination to neural cells. Nat. Commun. 10 (1), 4430.
doi: 10.1038/s41467-019-12408-x
Babaha, F., and Rezaei, N. (2020). Primary immunodeciency diseases in COVID-
19 pandemic: A predisposing or protective factor? Am. J. Med. Sci. 360 (6),
740741. doi: 10.1016/j.amjms.2020.07.027
Badae, N. M., El Naggar, A. S., and El Sayed, S. M. (2019). Is the cardioprotective
effect of the ACE2 activator diminazene aceturate more potent than the ACE
inhibitor enalapril on acute myocardial infarction in rats? Can. J. Physiol.
Pharmacol. 97 (7), 638646. doi: 10.1139/cjpp-2019-0078
Baig, A. M., Khaleeq, A., Ali, U., and Syeda, H. (2020). Evidence of the COVID-19
Virus Targeting the CNS: Tissue Distribution, Host-Virus Interaction, and
Proposed Neurotropic Mechanisms. ACS Chem. Neurosci. 11 (7), 995998.
doi: 10.1021/acschemneuro.0c00122
Bansal, M. (2020). Cardiovascular disease and COVID-19. Diabetes & Metabolic
Syndrome. Clin. Res. Rev. 14 (3), 247250. doi: 10.1016/j.dsx.2020.03.013
Beigel, J. H., Tomashek, K. M., Dodd, L. E., Mehta, A. K., Zingman, B. S., Kalil,
A. C., et al. (2020). Remdesivir for the Treatment of Covid-19 - Preliminary
Report. N. Engl. J. Med. 383 (19), 18131826. doi: 10.1056/NEJMoa2007764
Belouzard, S., Millet, J. K., Licitra, B. N., and Whittaker, G. R. (2012). Mechanisms
of coronavirus cell entry mediated by the viral spike protein. Viruses 4 (6),
10111033. doi: 10.3390/v4061011
Berry, J. D., Hay, K., Rini, J. M., Yu, M., Wang, L., Plummer, F. A., et al. (Eds.)
(2020). Neutralizing epitopes of the SARS-CoV S-protein cluster independent
of repertoire, antigen structure or mAb technology,in MAbs (Taylor &
Francis), 5366. doi: 10.4161/mabs.2.1.10788
Bhaskar, S., Sinha, A., Banach, M., Mittoo, S., Weissert, R., Kass, J. S., et al. (2020).
Cytokine Storm in COVID-19Immunopathological Mechanisms, Clinical
Considerations, and Therapeutic Approaches: The REPROGRAM Consortium
Position Paper. Front. Immunol. 11, 1648. doi: 10.3389/mmu.2020.01648
Bienvenu,L.A.,Noonan,J.,Wang,X.,andPeter,K.(2020).Highermortalityof
COVID-19 in males: Sex differences in immune response and cardiovascular
comorbidities. Cardiovasc. Res. 116 (14), 21972206. doi: 10.1093/cvr/
cvaa284
Bosch, B. J., van der Zee, R., de Haan, C. A., and Rottier, P. J. (2003). The
coronavirus spike protein is a class I virus fusion protein: structural and
functional characterization of the fusion core complex. J. Virol. 77 (16), 8801
8811. doi: 10.1128/JVI.77.16.8801-8811.2003
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 56308512
Braun, E., and Sauter, D. (2019). Furin-mediated protein processing in infectious
diseases and cancer. Clin. Trans. Immunol. 8 (8), e1073. doi: 10.1002/cti2.1073
Brooks, D. G., Trilo, M. J., Edelmann, K. H., Teyton, L., McGavern, D. B., and
Oldstone, M. B. (2006). Interleukin-10 determines viral clearance or
persistence in vivo. Nat. Med. 12 (11), 13011309. doi: 10.1038/nm1492
Bryan Williams, Y. Z. (2020). Hypertension, renin-angiotensin-aldosterone
system inhibition, and COVID-19. Lancet 36 (20), 3113131134.
doi: 10.1016/S0140-6736(20)31030-8
Burmester, G. R., Rigby, W. F., van Vollenhoven, R. F., Kay, J., Rubbert-Roth, A.,
Kelman, A., et al. (2016). Tocilizumab in early progressive rheumatoid
arthritis: FUNCTION, a randomised controlled trial. Ann. Rheum. Dis. 75
(6), 10811091. doi: 10.1136/annrheumdis-2015-207628
Cao, Y., Li, L., Feng, Z., Wan, S., Huang, P., Sun, X., et al. (2020). Comparative
genetic analysis of the novel coronavirus (2019-nCoV/SARS-CoV-2) receptor
ACE2 in different populations. Cell Discov. 6 (1), 14. doi: 10.1038/s41421-020-
0147-1
Cavalli, G., De Luca, G., Campochiaro, C., Della-Torre, E., Ripa, M., Canetti, D.,
et al. (2020). Interleukin-1 blockade with high-dose anakinra in patients with
COVID-19, acute respiratory distress syndrome, and hyperinammation: a
retrospective cohort study. Lancet Rheumatol. 2 (6), e325e331. doi: 10.1016/
S2665-9913(20)30127-2
Channappanavar, R., and Perlman, S. (2017). Pathogenic human coronavirus
infections: causes and consequences of cytokine storm and immunopathology.
In: Seminars in immunopathology (Springer) 529539.
Channappanavar, R., Fehr, A. R., Zheng, J., Wohlford-Lenane, C., Abrahante, J. E.,
Mack, M., et al. (2019). IFN-I response timing relative to virus replication
determines MERS coronavirus infection outcomes. J. Clin. Invest. 129 (9). doi:
10.1172/JCI126363
Chen, D., Xu, W., Lei, Z., Huang, Z., Liu, J., Gao, Z., et al. (2020). Recurrence of
positive SARS-CoV-2 RNA in COVID-19: A case report. Int. J. Infect. Dis. 93,
297299. doi: 10.1016/j.ijid.2020.03.003
Chen, I. Y., Moriyama, M., Chang, M. F., and Ichinohe, T. (2019). Severe Acute
Respiratory Syndrome Coronavirus Viroporin 3a Activates the NLRP3
Inammasome. Front. Microbiol. 10, 50. doi: 10.3389/fmicb.2019.00050
Chen, J., Jiang, Q., Xia, X., Liu, K., Yu, Z., Tao, W., et al. (2020). Individual
variation of the SARS-CoV-2 receptor ACE2 gene expression and regulation.
Aging Cell 19 (7), 3168. doi: 10.1111/acel.13168
Chen, N., Zhou, M., Dong, X., Qu, J., Gong, F., Han, Y., et al. (2020).
Epidemiological and clinical characteristics of 99 cases of 2019 novel
coronavirus pneumonia in Wuhan, China: a descriptive study. Lancet 395
(10223), 507513. doi: 10.1016/S0140-6736(20)30211-7
Chockalingam, A. K., Hamed, S., Goodwin, D. G., Rosenzweig, B. A., Pang, E.,
Boyne, M. T. II, et al. (2016). The Effect of Oseltamivir on the Disease
Progression of Lethal Inuenza A Virus Infection: Plasma Cytokine and
miRNA Responses in a Mouse Model. Dis. Markers 2016, 9296457.
doi: 10.1155/2016/9296457
Claas, E. C., Osterhaus, A. D., Van Beek, R., De Jong, J. C., Rimmelzwaan, G. F.,
Senne, D. A., et al. (1998). Human inuenza A H5N1 virus related to a highly
pathogenic avian inuenza virus. Lancet 351 (9101), 472477.
Coutard, B., Valle, C., de Lamballerie, X., Canard, B., Seidah, N., and Decroly, E.
(2020). The spike glycoprotein of the new coronavirus 2019-nCoV contains a
furin-like cleavage site absent in CoV of the same clade. Antiviral Res. 176,
104742. doi: 10.1016/j.antiviral.2020.104742
Cox, R. J., and Brokstad, K. A. (2020). Not just antibodies: B cells and T cells
mediate immunity to COVID-19. Nat. Rev. Immunol. 20 (10), 581582. doi:
10.1038/s41577-020-00436-4
Dandekar, A. A., and Perlman, S. (2005). Immunopathogenesis of coronavirus
infections: implications for SARS. Nat. Rev. Immunol. 5 (12), 917927. doi:
10.1038/nri1732
de Abajo, F. J., Rodrguez-Martn, S., Lerma, V., Meja-Abril, G., Aguilar, M. N.,
Garca-Luque, A., et al. (2020). Use of renin angiotensin aldosterone system
inhibitors and risk of COVID-19 requiring admission to hospital: a case-
population study. Lancet 39 (10238), 17051714. doi: 10.1016/S0140-6736(20)
31030-8
de Groot-Mijnes, J. D., van Dun, J. M., van der Most, R. G., and de Groot, R. J.
(2005). Natural history of a recurrent feline coronavirus infection and the role
of cellular immunity in survival and disease. J. Virol. 79 (2), 10361044.
doi: 10.1128/JVI.79.2.1036-1044.2005
Dehingia, N., and Raj, A. (2020). Sex differences in COVID-19 case fatality: do we
know enough? Lancet Glob. Health 9 (1), e14e15. doi: 10.1016/S2214-109X
(20)30464-2
Desforges, M., Le Coupanec, A., Dubeau, P., Bourgouin, A., Lajoie, L., Dube, M.,
et al. (2019). Human Coronaviruses and Other Respiratory Viruses:
Underestimated Opportunistic Pathogens of the Central Nervous System?
Viruses 12 (1). doi: 10.3390/v12010014
Diao, B., Wang, C., Tan, Y., Chen, X., Liu, Y., Ning, L., et al. (2020).
Reduction and Functional Exhaustion of T Cells in Patients with
Coronavirus Disease 2019 (COVID-19). Front Immunol. 11,827.doi:
10.3389/mmu.2020.00827
Dronavalli, S., Duka, I., and Bakris, G. L. (2008). The pathogenesis of diabetic
nephropathy. Nat. Clin. Pract. Endocrinol. Metab. 4 (8), 444452. doi: 10.1038/
ncpendmet0894
Ejrnaes, M., Filippi, C. M., Martinic, M. M., Ling, E. M., Togher, L. M., Crotty, S.,
et al. (2006). Resolution of a chronic viral infection after interleukin-10
receptor blockade. J. Exp. Med. 203 (11), 24612472. doi: 10.1084/
jem.20061462
Farcas, G. A., Poutanen, S. M., Mazzulli, T., Willey, B. M., Butany, J., Asa, S. L.,
et al. (2005). Fatal severe acute respiratory syndrome is associated with
multiorgan involvement by coronavirus. J. Infect. Dis. 191 (2), 193197. doi:
10.1086/426870
Fehr, A. R., and Perlman, S. (2015). Coronaviruses: an overview of their replication
and pathogenesis. Coronaviruses (Springer), 123.
Freeman, T., and Swartz, T. (2020). Targeting the NLRP3 Inammasome in Severe
COVID-19. Front. Immunol. 11(518). doi: 10.3389/mmu.2020.01518
Furst, D. E. (2004). Anakinra: review of recombinant human interleukin-I receptor
antagonist in the treatment of rheumatoid arthritis. Clin. Ther. 26 (12), 1960
1975. doi: 10.1016/j.clinthera.2004.12.019
Galasso, V., Pons, V., Profeta, P., Becher, M., Brouard, S., and Foucault, M. (2020).
Gender differences in COVID-19 attitudes and behavior: Panel evidence from
eight countries. Proc. Natl. Acad. Sci. 117 (44), 2728527291. doi: 10.1073/
pnas.2012520117
Ge, X. Y., Li, J. L., Yang, X. L., Chmura, A. A., Zhu, G., Epstein, J. H., et al.
(2013). Isolation and characterization of a bat SARS-like coronavirus
that uses the ACE2 receptor. Nature 503 (7477), 535538. doi: 10.1038/
nature12711
Giudicessi, J. R., Roden, D. M., Wilde, A. A. M., and Ackerman, M. J. (2020).
Genetic Susceptibility for COVID-19-Associated Sudden Cardiac Death in
African Americans. Heart Rhythm 17 (9), 14871492. doi: 10.1016/
j.hrthm.2020.04.045
Gu, J., Gong, E., Zhang, B., Zheng, J., Gao, Z., Zhong, Y., et al. (2005). Multiple
organ infection and the pathogenesis of SARS. J. Exp. Med. 202 (3), 415424.
doi: 10.1084/jem.20050828
Gurwitz, D. (2020). Angiotensin receptor blockers as tentative SARSCoV2
therapeutics. Drug Dev. Res. 81, 537540. doi: 10.1002/ddr.21656
Habibzadeh, P., and Stoneman, E. K. (2020). The Novel Coronavirus: A Birds Eye
View. Int. J. Occup. Environ. Med. 11 (2), 6571. doi: 10.15171/
ijoem.2020.1921
Hajeer, A. H., Balkhy, H., Johani, S., Yousef, M. Z., and Arabi, Y. (2016).
Association of human leukocyte antigen class II alleles with severe Middle
East respiratory syndrome-coronavirus infection. Ann. Thorac. Med. 11 (3),
211. doi: 10.4103/1817-1737.185756
Halacli, B., Kaya, A., and I
SKI
T, A. T. (2020). Critically ill COVID-19 patient.
Turk. J. Med. Sci. 50 (SI-1), 585591. doi: 10.3906/sag-2004-122
Harapan, H., Itoh, N., Yuka, A., Winardi, W., Keam, S., Te, H., et al. (2020).
Coronavirus disease 2019 (COVID-19): A literature review. J. Infect. Public
Health 13 (20), 667673. doi: 10.1016/j.jiph.2020.03.019
Hoffmann, M., Kleine-Weber, H., Schroeder, S., Kruger, N., Herrler, T., Erichsen,
S., et al. (2020). SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and
Is Blocked by a Clinically Proven Protease Inhibitor. Cell 181 (2), 271280.e8.
doi: 10.1016/j.cell.2020.02.052
Hofmann, H., Pyrc, K., van der Hoek, L., Geier, M., Berkhout, B., and Pöhlmann,
S. (2005). Human coronavirus NL63 employs the severe acute respiratory
syndrome coronavirus receptor for cellular entry. Proc. Natl. Acad. Sci. 102
(22), 79887993. doi: 10.1073/pnas.0409465102
Hu, B., Ge, X., Wang, L. F., and Shi, Z. (2015). Bat origin of human coronaviruses.
Virol. J. 12, 221. doi: 10.1186/s12985-015-0422-1
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 56308513
Huang, C., Wang, Y., Li, X., Ren, L., Zhao, J., Hu, Y., et al. (2020). Clinical features
of patients infected with 2019 novel coronavirus in Wuhan, China. Lancet 395
(10223), 497506. doi: 10.1016/S0140-6736(20)30183-5
Hui, K. P. Y., Cheung, M. C., Perera, R., Ng, K. C., Bui, C. H. T., Ho, J. C. W., et al.
(2020). Tropism, replication competence, and innate immune responses of the
coronavirus SARS-CoV-2 in human respiratory tract and conjunctiva: an
analysis in ex-vivo and in-vitro cultures. Lancet Respir. Med. 12 (221).
doi: 10.1016/S2213-2600(20)30193-4
Jamilloux, Y., Henry, T., Belot, A., Viel, S., Fauter, M., El Jammal, T., et al. (2020).
Should we stimulate or suppress immune responses in COVID-19? Cytokine
and anti-cytokine interventions. Autoimmun. Rev. 19 (7), 102567. doi: 10.1016/
j.autrev.2020.102567
Jeffers, S. A., Tusell, S. M., Gillim-Ross, L., Hemmila, E. M., Achenbach, J. E.,
Babcock, G. J., et al. (2004). CD209L (L-SIGN) is a receptor for severe acute
respiratory syndrome coronavirus. Proc. Natl. Acad. Sci. 101 (44), 15748
15753. doi: 10.1073/pnas.0403812101
Jin, J.-M., Bai, P., He, W., Wu, F., Liu, X.-F., Han, D.-M., et al. (2020). Gender
differences in patients with COVID-19: Focus on severity and mortality. Front.
Public Health 8, 152. doi: 10.3389/fpubh.2020.00152
Jones, S. A., and Jenkins, B. J. (2018). Recent insights into targeting the IL-6
cytokine family in inammatory diseases and cancer. Nat. Rev. Immunol. 18
(12), 773789. doi: 10.1038/s41577-018-0066-7
Kativhu, C. L., and Libraty, D. H. (2016). A model to explain how the Bacille
Calmette Guerin (BCG) vaccine drives interleukin-12 production in neonates.
PLoS One 11 (8), e0162148. doi: 10.1371/journal.pone.0162148
Kido, H., Okumura, Y., Takahashi, E., Pan, H. Y., Wang, S., Yao, D., et al. (2012).
Role of host cellular proteases in the pathogenesis of inuenza and inuenza-
induced multiple organ failure. Biochim. Biophys. Acta 1824 (1), 186194.
doi: 10.1016/j.bbapap.2011.07.001
Kindler, E., Thiel, V., and Weber, F. (2016). Interaction of SARS and MERS
coronaviruses with the antiviral interferon response. Adv. Virus Res. 96, 219
243. doi: 10.1016/bs.aivir.2016.08.006
Kreye, J., Reincke, S. M., Kornau, H. C., Sanchez-Sendin, E., Corman, V. M., Liu,
H., et al. (2020). A SARS-CoV-2 neutralizing antibody protects from lung
pathology in a COVID-19 hamster model. bioRxiv. doi: 10.1101/
2020.08.15.252320
Kuba, K., Imai, Y., Ohto-Nakanishi, T., and Penninger, J. M. (2010). Trilogy of
ACE2: A peptidase in the reninangiotensin system, a SARS receptor, and a
partner for amino acid transporters. Pharmacol. Ther. 128 (1), 119128. doi:
10.1016/j.pharmthera.2010.06.003
Le Bert, N., Tan, A. T., Kunasegaran, K., Tham, C. Y., Hafezi, M., Chia, A., et al.
(2020). SARS-CoV-2-specic T cell immunity in cases of COVID-19 and
SARS, and uninfected controls. Nature 584 (7821), 457462. doi: 10.1038/
s41586-020-2550-z
Letko, M., Marzi, A., and Munster, V. (2020). Functional assessment of cell entry
and receptor usage for SARS-CoV-2 and other lineage B betacoronaviruses.
Nat. Microbiol. 5 (4), 562569. doi: 10.1038/s41564-020-0688-y
Li, G., Chen, X., and Xu, A. (2003). Prole of specic antibodies to the SARS-
associated coronavirus. N. Engl. J. Med. 349 (5), 508509. doi: 10.1056/
NEJM200307313490520
Li, C. K.-F., Wu, H., Yan, H., Ma, S., Wang, L., Zhang, M., et al. (2008). T cell
responses to whole SARS coronavirus in humans. J. Immunol. 181 (8), 5490
5500. doi: 10.4049/jimmunol.181.8.5490
Li, H., Liu, L., Zhang, D., Xu, J., Dai, H., Tang, N., et al. (2020). SARS-CoV-2 and
viral sepsis: observations and hypotheses. Lancet 395 (10235), 15171520.
doi: 10.1016/S0140-6736(20)30920-X
Li, X., Geng, M., Peng, Y., Meng, L., and Lu, S. (2020). Molecular immune
pathogenesis and diagnosis of COVID-19. J. Pharm. Anal. 10 (2), 102106.
doi: 10.1016/j.jpha.2020.03.001
Lippi, G., and Plebani, M. (2020). Laboratory abnormalities in patients with
COVID-2019 infection. Clin. Chem. Lab. Med. (CCLM) 58 (7), 11311134.
doi: 10.1515/cclm-2020-0198
Liu, W. J., Zhao, M., Liu, K., Xu, K., Wong, G., Tan, W., et al. (2017). T-cell
immunity of SARS-CoV: Implications for vaccine development against MERS-
CoV. Antiviral Res. 137, 8292. doi: 10.1016/j.antiviral.2016.11.006
Liu, A., Li, Y., Peng, J., Huang, Y., and Xu, D. (2020). Antibody responses against
SARS-CoV-2 in COVID-19 patients. J. Med. Virol. 93, 144148. doi: 10.1002/
jmv.26241
Lokugamage, K. G., Hage, A., de Vries, M., Valero-Jimenez, A. M., Schindewolf, C.,
Dittmann, M., et al (2020). Type I interferon susceptibility distinguishes SARS-
CoV-2 from SARS-CoV. J. Virol.94(23).
Long, Q. X., Liu, B. Z., Deng, H. J., Wu, G. C., Deng, K., Chen, Y. K., et al. (2020).
Antibody responses to SARS-CoV-2 in patients with COVID-19. Nat. Med. 26
(6), 845848. doi: 10.1038/s41591-020-0897-1
Lu, R., Zhao, X., Li, J., Niu, P., Yang, B., Wu, H., et al. (2020). Genomic
characterisation and epidemiology of 2019 novel coronavirus: implications
for virus origins and receptor binding. Lancet 395 (10224), 565574. doi:
10.1016/S0140-6736(20)30251-8
Lu, H. (2020). Drug treatment options for the 2019-new coronavirus (2019-
nCoV). Biosci. Trends 14 (1), 6971. doi: 10.5582/bst.2020.01020
Luft, F. C. (2014). ACE in the hole. J. Mol. Med. (Berl) 92 (8), 793795.
doi: 10.1007/s00109-014-1172-z
Luks, A. M., and Swenson, E. R. (2020). COVID-19 Lung Injury and High Altitude
Pulmonary Edema: A False Equation with Dangerous Implications. Ann. Am.
Thor. Soc. doi: 10.1513/AnnalsATS.202004-327FR
Mao,L.,Jin,H.,Wang,M.,Hu,Y.,Chen,S.,He,Q.,etal.(2020a).
Neurologic Manifestations of Hospitalized Patients With Coronavirus
Disease 2019 in Wuhan, China. JAMA Neurol. doi: 10.1001/jamaneurol.
2020.1127
Mao, L., Wang, M., Chen, S., He, Q., Chang, J., Hong, C., et al. (2020b).
Neurological manifestations of hospitalized patients with COVID-19 in
Wuhan, China: a retrospective case series study. MedRxiv. doi: 10.1101/
2020.02.22.20026500SUPPORTING
Marra, M. A., Jones, S. J., Astell, C. R., Holt, R. A., Brooks-Wilson, A.,
Buttereld, Y. S., et al. (2003). The genome sequence of the SARS-
associated coronavirus. Science 300 (5624), 13991404. doi: 10.1126/
science.1085953
Mason, R. J. (2020). Pathogenesis of COVID-19 from a cell biology perspective.
Eur. Respir. J. 55 (4). doi: 10.1183/13993003.00607-2020
Mathewson, A. C., Bishop, A., Yao, Y., Kemp, F., Ren, J., Chen, H., et al. (2008).
Interaction of severe acute respiratory syndrome-coronavirus and NL63
coronavirus spike proteins with angiotensin converting enzyme-2. J. Gen.
Virol. 89 (Pt 11), 2741. doi: 10.1099/vir.0.2008/003962-0
Menachery, V. D., Yount, B. L.Jr., Debbink, K., Agnihothram, S., Gralinski, L. E.,
Plante, J. A., et al. (2015). A SARS-like cluster of circulating bat coronaviruses
showspotentialforhumanemergence.Nat. Med. 21 (12), 15081513.
doi: 10.1038/nm.3985
Merad, M., and Martin, J. C. (2020). Pathological inammation in patients with
COVID-19: a key role for monocytes and macrophages. Nat. Rev. Immunol. 20
(6), 355362. doi: 10.1038/s41577-020-0331-4
Millet, J. K., and Whittaker, G. R. (2014). Host cell entry of Middle East respiratory
syndrome coronavirus after two-step, furin-mediated activation of the spike
protein. Proc. Natl. Acad. Sci. 111 (42), 1521415219. doi: 10.1073/
pnas.1407087111
Mori, I. (2015). Transolfactory neuroinvasion by viruses threatens the human
brain. Acta Virol. 59 (4), 338349. doi: 10.4149/av_2015_04_338
Murthy, S., Gomersall, C. D., and Fowler, R. A. (2020). Care for critically ill
patients with COVID-19. JAMA 323 (15), 14991500. doi: 10.1001/
jama.2020.3633
Netland, J., Meyerholz, D. K., Moore, S., Cassell, M., and Perlman, S. (2008).
Severe acute respiratory syndrome coronavirus infection causes neuronal death
in the absence of encephalitis in mice transgenic for human ACE2. J. Virol. 82
(15), 72647275. doi: 10.1128/JVI.00737-08
Niemeyer, D., Zillinger, T., Muth, D., Zielecki, F., Horvath, G., Suliman, T., et al.
(2013). Middle East respiratory syndrome coronavirus accessory protein 4a is a
type I interferon antagonist. J. Virol. 87 (22), 1248912495. doi: 10.1128/
JVI.01845-13
Owczarek, K., Szczepanski, A., Milewska, A., Baster, Z., Rajfur, Z., Sarna, M., et al.
(2018). Early events during human coronavirus OC43 entry to the cell. Sci. Rep.
8 (1), 112. doi: 10.1038/s41598-018-25640-0
Park, M. D. (2020). Macrophages: a Trojan horse in COVID-19? Nat. Rev.
Immunol. 20 (6), 351. doi: 10.1038/s41577-020-0317-2
Peiris, J. S. M., Chu, C.-M., Cheng, V. C.-C., Chan, K., Hung, I., Poon, L. L., et al.
(2003). Clinical progression and viral load in a community outbreak of
coronavirus-associated SARS pneumonia: a prospective study. Lancet 361
(9371), 17671772.
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 56308514
Perlman, S., and Netland, J. (2009). Coronaviruses post-SARS: update on
replication and pathogenesis. Nat. Rev. Microbiol. 7 (6), 439450.
doi: 10.1038/nrmicro2147
Perlman, S., Evans, G., and Afifi, A. (1990). Effect of olfactory bulb ablation on
spread of a neurotropic coronavirus into the mouse brain. J. Exp. Med. 172 (4),
11271132. doi: 10.1084/jem.172.4.1127
Pipeline Review (2020). APEIRONs respiratory drug product to start pilot
clinical trial to treat coronavirus disease COVID-19 in China,in Pipeline
Review (VIENNA, Austria I). Available at: https://pipelinereviewcom/
indexphp/2020022673884/Proteins-and-Peptides/APEIRONs-respiratory-
drug-product-to-start-pilot-clinical-trial-to-treat-coronavirus-disease-
COVID-19-in-Chinahtml.
Pontali, E., Volpi, S., Antonucci, G., Castellaneta, M., Buzzi, D., Tricerri, F., et al.
(2020). Safety and efcacy of early high-dose IV anakinra in severe COVID-19
lung disease. J. Allergy Clin. Immunol. doi: 10.1016/j.jaci.2020.05.002
Pradhan, A., and Olsson, P.-E. (2020). Sex differences in severity and mortality
from COVID-19: are males more vulnerable? Biol. Sex Dif. 11 (1), 111. doi:
10.1186/s13293-020-00330-7
Prompetchara, E., Ketloy, C., and Palaga, T. (2020). Immune responses in
COVID-19 and potential vaccines: Lessons learned from SARS and MERS
epidemic. Asian Pac. J. Allergy Immunol. 38 (1), 19. doi: 10.12932/AP-
200220-0772
Qian, Z., Travanty, E. A., Oko, L., Edeen, K., Berglund, A., Wang, J., et al. (2013).
Innate immune response of human alveolar type II cells infected with severe
acute respiratory syndrome-coronavirus. Am. J. Respir. Cell Mol. Biol. 48 (6),
742748. doi: 10.1165/rcmb.2012-0339OC
Rahmati M, M. M. (2020). Cytokine-Targeted Therapy in Severely ill COVID-19
Patients: Options and Cautions. EJMO 4 (2), 179180. doi: 10.14744/
ejmo.2020.72142
Rithanya, M., and Brundha, M. (2020). Molecular Immune Pathogenesis and
Diagnosis of COVID-19-A Review. Int.J.Curr.Res.Rev.12 (21), 69.
doi: 10.31782/IJCRR.2020.SP37
Rodda, L. B., Netland, J., Shehata, L., Pruner, K. B., Morawski, P. M., Thouvenel,
C., et al. (2020). Functional SARS-CoV-2-specic immune memory persists
after mild COVID-19. medRxiv. doi: 10.1101/2020.08.11.20171843
Serrano-Castro, P., Estivill-Torr ؛s, G., Cabezudo-Garca, P., Reyes-Bueno, J.,
Petersen, N. C., Aguilar-Castillo, M., et al. (2020). Impact of SARS-CoV-2
infection on neurodegenerative and neuropsychiatric diseases: a delayed
pandemic? Neurolia (English Ed.) 35 (4). doi: 10.1016/j.nrleng.2020.04.002
Sette, A., and Crotty, S. (2020). Pre-existing immunity to SARS-CoV-2: the
knowns and unknowns. Nat. Rev. Immunol. 20 (8), 457458. doi: 10.1038/
s41577-020-0389-z
Shah, A. (2020). Novel Coronavirus-Induced NLRP3 Inammasome Activation: A
Potential Drug Target in the Treatment of COVID-19. Front. Immunol. 11,
1021. doi: 10.3389/mmu.2020.01021
Shang, J., Ye, G., Shi, K., Wan, Y., Luo, C., Aihara, H., et al. (2020a). Structural
basis of receptor recognition by SARS-CoV-2. Nature 581 (7807), 221224.
doi: 10.1038/s41586-020-2179-y
Shang, J., Wan, Y., Luo, C., Ye, G., Geng, Q., Auerbach, A., et al. (2020b). Cell entry
mechanisms of SARS-CoV-2. Proc. Natl. Acad. Sci. U. S. A. 117 (21), 11727
11734. doi: 10.1073/pnas.2003138117
Shearer, F., Lang, C., and Struthers, A. (2013). Reninangiotensinaldosterone
system inhibitors in heart failure. Clin. Pharmacol. Ther. 94 (4), 459467. doi:
10.1038/clpt.2013.135
Sheraton, M., Deo, N., Kashyap, R., and Surani, S. (2020). A Review of
Neurological Complications of COVID-19. Cureus 12 (5), e8192. doi:
10.7759/cureus.8192
Shokri, S., Mahmoudvand, S., Taherkhani, R., and Farshadpour, F. (2019).
Modulation of the immune response by Middle East respiratory syndrome
coronavirus. J. Cell. Physiol. 234 (3), 21432151. doi: 10.1002/jcp.27155
Simmons, G., Gosalia, D. N., Rennekamp, A. J., Reeves, J. D., Diamond, S. L., and
Bates, P. (2005). Inhibitors of cathepsin L prevent severe acute respiratory
syndrome coronavirus entry. Proc. Natl. Acad. Sci. U. S. A. 102 (33), 11876
11881. doi: 10.1073/pnas.0505577102
Soy, M., Keser, G., Atagunduz, P., Tabak, F., Atagunduz, I., and Kayhan, S. (2020).
Cytokine storm in COVID-19: pathogenesis and overview of anti-
inammatory agents used in treatment. Clin. Rheumatol. 39 (7), 20852094.
doi: 10.1007/s10067-020-05190-5
Streicher, F., and Jouvenet, N. (2019). Stimulation of Innate Immunity by Host
and Viral RNAs. Trends Immunol. 40 (12), 11341148. doi: 10.1016/
j.it.2019.10.009
Su, S., Wong, G., Shi, W., Liu, J., Lai, A. C., Zhou, J., et al. (2016). Epidemiology,
genetic recombination, and pathogenesis of coronaviruses. Trends Microbiol.
24 (6), 490502. doi: 10.1016/j.tim.2016.03.003
Sun, X., Tse, L. V., Ferguson, A. D., and Whittaker, G. R. (2010). Modications to
the hemagglutinin cleavage site control the virulence of a neurotropic H1N1
inuenza virus. J. Virol. 84 (17), 86838690. doi: 10.1128/JVI.00797-10
Sungnak, W., Huang, N., Becavin, C., Berg, M., Queen, R., Litvinukova, M., et al.
(2020). SARS-CoV-2 entry factors are highly expressed in nasal epithelial cells
together with innate immune genes. Nat. Med. 26 (5), 681687. doi: 10.1038/
s41591-020-0868-6
Surveillances, V. (2020). The Epidemiological Characteristics of an Outbreak of
2019 Novel Coronavirus Diseases (COVID-19)China, 2020. China CDC
Wkly. 2 (8), 113122. doi: 10.46234/ccdcw2020.032
Tabarsi, P., Mortaz, E., Varahram, M., Folkerts, G., and Adcock, I. M. (2020). The
Immune Response and Immunopathology of COVID-19. Front. Immunol. 11,
2037. doi: 10.3389/mmu.2020.02037
Tai, W., He, L., Zhang, X., Pu, J., Voronin, D., Jiang, S., et al. (2020). Characterization
of the receptor-bindingdomain (RBD) of 2019 novel coronavirus: implication for
development of RBD protein as a viral attachment inhibitor and vaccine. Cell.
Mol. Immunol. 17 (6), 613620. doi: 10.1038/s41423-020-0400-4
Takahashi, T., Ellingson, M. K., Wong, P., Israelow, B., Lucas, C., Klein, J., et al.
(2020). Sex differences in immune responses that underlie COVID-19 disease
outcomes. Nature 588, 588320. doi: 10.1038/s41586-020-2700-3
Tay, M. Z., Poh, C. M., Renia, L., MacAry, P. A., and Ng, L. F. P. (2020). The trinity
of COVID-19: immunity, inammation and intervention. Nat. Rev. Immunol.
20 (6), 363374. doi: 10.1038/s41577-020-0311-8
The Sex, Gender, and COVID-19 Project (2020). The COVID-10 sex-disaggregated
data tracker. Available at: https://globalhealth5050.org/covid19/ (Accessed
Nov 16, 2020).
Thebault R, B. T. A., and Williams, V. (2020). The coronavirus is infecting and
killing black Americans at an alarmingly high rate (Washington Post). (April 7,
2020).
Toots, M., Yoon, J.-J., Cox, R. M., Hart, M., Sticher, Z. M., Makhsous, N., et al.
(2019). Characterization of orally efcacious inuenza drug with high
resistance barrier in ferrets and human airway epithelia. Sci. Trans. Med. 11
(515), eaax5866. doi: 10.1126/scitranslmed.aax5866
Tu, X., Chong, W. P., Zhai, Y., Zhang, H., Zhang, F., Wang, S., et al. (2015).
Functional polymorphisms of the CCL2 and MBL genes cumulatively increase
susceptibility to severe acute respiratory syndrome coronavirus infection.
J. Infect. 71 (1), 101109. doi: 10.1016/j.jinf.2015.03.006
van den Berg, D. F., and Te Velde, A. A. (2020). Severe COVID-19: NLRP3
Inammasome Dysregulated. Front. Immunol. 11(1580). doi: 10.3389/mmu.
2020.01580
Walls, A. C., Park, Y.-J., Tortorici, M. A., Wall, A., McGuire, A. T., and Veesler, D.
(2020). Structure, function, and antigenicity of the SARS-CoV-2 spike
glycoprotein. Cell 183 (6), 281292. doi: 10.1016/j.cell.2020.02.058
Wan, S., Yi, Q., Fan, S., Lv, J., Zhang, X., Guo, L., et al. (2020). Characteristics of
lymphocyte subsets and cytokines in peripheral blood of 123 hospitalized
patients with 2019 novel coronavirus pneumonia (NCP). MedRxiv 221 (11).
doi: 10.1093/infdis/jiaa150
Wan, Y., Shang, J., Graham, R., Baric, R. S., and Li, F. (2020). Receptor Recognition
by the Novel Coronavirus from Wuhan: an Analysis Based on Decade-Long
Structural Studies of SARS Coronavirus. J. Virol. 94 (7), e0012700120.
doi: 10.1128/JVI.00127-20
Wang, N., Shi, X., Jiang, L., Zhang, S., Wang, D., Tong, P., et al. (2013). Structure of
MERS-CoV spike receptor-binding domain complexed with human receptor
DPP4. Cell Res. 23 (8), 986. doi: 10.1038/cr.2013.92
Wang, M., Cao, R., Zhang, L., Yang, X., Liu, J., Xu, M., et al. (2020). Remdesivir and
chloroquine effectively inhibit the recently emerged novel coronavirus (2019-
nCoV) in vitro. Cell Res. 30 (3), 269271. doi: 10.1038/s41422-020-0282-0
Wang,J.,Nikrad,M.P.,Phang,T.,Gao,B.,Alford,T.,Ito,Y.,etal.(2011).Innate
immune response to inuenza A virus in differentiated human alveolar type II cells.
Am.J.Respir.CellMol.Biol.45 (3), 582591. doi: 10.1165/rcmb.2010-0108OC
Wang, S.-F., Chen, K.-H., Chen, M., Li, W.-Y., Chen, Y.-J.,Tsao, C.-H., et al. (2011).
Human-leukocyte antigen class I Cw 1502 and class II DR 0301 genotypes are
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 56308515
associated with resistance to severe acute respiratory syndrome (SARS) infection.
Viral Immunol. 24 (5), 421426. doi: 10.1089/vim.2011.0024
Wen, W., Su, W., Tang, H., Le, W., Zhang, X., Zheng, Y., et al. (2020). Immune cell
proling of COVID-19 patients in the recovery stage by single-cell sequencing.
Cell Discov. 6 (1), 118. doi: 10.1101/2020.03.23.20039362
Wherry, E. J. (2011). T cell exhaustion. Nat. Immunol. 12 (6), 492499. doi:
10.1038/ni.2035
Wu, Y., Guo, W., Liu, H., Qi, B., Liang, K., Xu, H., et al. (2020a). Clinical outcomes
of 402 patients with COVID-19 from a single center in Wuhan, China. J. Med.
Virol 92, 27512757. doi: 10.1002/jmv.26168
Wu, Y., Xu, X., Chen, Z., Duan, J., Hashimoto, K., Yang, L., et al. (2020b). Nervous
system involvement after infection with COVID-19 and other coronaviruses.
Brain Behav. Immun. 87, 1822. doi: 10.1016/j.bbi.2020.03.031
Wu, F., Zhao, S., Yu, B., Chen, Y.-M., Wang, W., Song, Z.-G., et al. (2020). A new
coronavirus associated with human respiratory disease in China. Nature 579
(7798), 265269. doi: 10.1038/s41586-020-2008-3
Xiao, F., Tang, M., Zheng, X., Liu, Y., Li, X., and Shan, H. (2020). Evidence for
Gastrointestinal Infection of SARS-CoV-2. Gastroenterology. doi: 10.1053/
j.gastro.2020.02.055
Xie, P., Ma, W., Tang, H., and Liu, D. (2020). Severe COVID-19: A Review of
Recent Progress With a Look Toward the Future. Front. Public Health 8, 189.
doi: 10.3389/fpubh.2020.00189
Xi-zhi, J. G., and Thomas, P. G. (Eds.) (2017). New fronts emerge in the inuenza
cytokine storm,in Seminars in immunopathology (Springer).
Xu, H., Zhong, L., Deng, J., Peng, J., Dan, H., Zeng, X., et al. (2020). High
expression of ACE2 receptor of 2019-nCoV on the epithelial cells of oral
mucosa. Int. J. Oral. Sci. 12 (1), 15. doi: 10.1038/s41368-020-0074-x
Xu, K., Chen, Y., Yuan, J., Yi, P., Ding, C., Wu, W., et al. (2020). Factors associated
with prolonged viral RNA shedding in patients with COVID-19. Clin. Infect.
Dis 71 (15), 799806. doi: 10.1093/cid/ciaa351
Xu, Z., Shi, L., Wang, Y., Zhang, J., Huang, L., Zhang, C., et al. (2020). Pathological
ndings of COVID-19 associated with acute respiratory distress syndrome.
Lancet Respir. Med. 8 (4), 420422. doi: 10.1016/S2213-2600(20)30076-X
Yancy, C. W. (2020). COVID-19 and African Americans. JAMA 323(19), 1891
1892. doi: 10.1001/jama.2020.6548
Yang, Y., Zhang, L., Geng, H., Deng, Y., Huang, B., Guo, Y., et al. (2013). The
structural and accessory proteins M, ORF 4a, ORF 4b, and ORF 5 of Middle
East respiratory syndrome coronavirus (MERS-CoV) are potent interferon
antagonists. Protein Cell 4 (12), 951961. doi: 10.1007/s13238-013-3096-8
Yang, X. L., Hu, B., Wang, B., Wang, M. N., Zhang, Q., Zhang, W., et al. (2015).
Isolation and Characterization of a Novel Bat Coronavirus Closely Related to
the Direct Progenitor of Severe Acute Respiratory Syndrome Coronavirus.
J. Virol. 90 (6), 32533256. doi: 10.1128/JVI.02582-15
Yang, Y., Ye, F., Zhu, N., Wang, W., Deng, Y., Zhao, Z., et al. (2015). Middle East
respiratory syndrome coronavirus ORF4b protein inhibits type I interferon
production through both cytoplasmic and nuclear targets. Sci. Rep. 5, 17554.
doi: 10.1038/srep17554
Ye, M., Ren, Y., and Lv, T. (2020). Encephalitis as a clinical manifestation of
COVID-19. Brain Behav. Immun. 88, 945946. doi: 10.1016/j.bbi.2020.04.017
Yokota, S., Itoh, Y., Morio, T., Origasa, H., Sumitomo, N., Tomobe, M., et al.
(2016). Tocilizumab in systemic juvenile idiopathic arthritis in a real-world
clinical setting: results from 1 year of postmarketing surveillance follow-up of
417 patients in Japan. Ann. Rheum. Dis. 75 (9), 16541660. doi: 10.1136/
annrheumdis-2015-207818
Yuan, F. F., Velickovic, Z., Ashton, L. J., Dyer, W. B., Geczy, A. F., Dunckley, H.,
et al. (2014). Inuence of HLA gene polymorphisms on susceptibility and
outcome post infection with the SARS-CoV virus. Virol. Sin. 29 (2), 128130.
doi: 10.1007/s12250-014-3398-x
Yuan, Y., Cao, D., Zhang, Y., Ma, J., Qi, J., Wang, Q., et al. (2017). Cryo-EM
structures of MERS-CoV and SARS-CoV spike glycoproteins reveal the
dynamic receptor binding domains. Nat. Commun. 8, 15092. doi: 10.1038/
ncomms15092
Yuan, B., Liu, H. Q., Yang, Z. R., Chen, Y. X., Liu, Z. Y., Zhang, K., et al. (2020).
Recurrence of positive SARS-CoV-2 viral RNA in recovered COVID-19
patients during medical isolation observation. Sci. Rep. 10 (1), 11887.
doi: 10.1038/s41598-020-68782-w
Zang, R., Castro, M. F. G., McCune, B. T., Zeng, Q., Rothlauf, P. W., Sonnek, N.
M., et al. (2020). TMPRSS2 and TMPRSS4 promote SARS-CoV-2 infection of
human small intestinal enterocytes. Sci. Immunol. 5 (47). doi: 10.1126/
sciimmunol.abc3582
Zhang, G., Hu, C., Luo, L., Fang, F., Chen, Y., Li, J., et al. (2020). Clinical features
and outcomes of 221 patients with COVID-19 in Wuhan, China. MedRxiv 127
(104367). doi: 10.1101/2020.03.02.20030452
Zhao, Y., Zhao, Z., Wang, Y., Zhou, Y., Ma, Y., and Zuo, W. (2020). Single-cell
RNA expression proling of ACE2, the putative receptor of Wuhan 2019-
nCov. bioRxiv. doi: 10.1101/2020.01.26.919985
Zheng, S., Fan, J., Yu, F., Feng, B., Lou, B., Zou, Q., et al. (2020). Viral load
dynamics and disease severity in patients infected with SARS-CoV-2 in
Zhejiang province, China, January-March 2020: retrospective cohort study.
bmj 369 (21). doi: 10.1136/bmj.m1443
Zheng, Y. Y., Ma, Y. T., Zhang, J. Y., and Xie, X. (2020). COVID-19 and the
cardiovascular system. Nat. Rev. Cardiol. 17 (5), 259260. doi: 10.1038/s41569-
020-0360-5
Zhou, F., Yu, T., Du, R., Fan, G., Liu, Y., Liu, Z., et al. (2020). Clinical course and
risk factors for mortality of adult inpatients with COVID-19 in Wuhan, China:
a retrospective cohort study. Lancet 395 (10229), 10541062. doi: 10.1016/
S0140-6736(20)30566-3
Zhou, L., Liu, K., and Liu, H. G. (2020). [Cause analysis and treatment strategies of
recurrencewith novel coronavirus pneumonia (COVID-19) patients after
discharge from hospital]. Zhonghua Jie He He Hu Xi Za Zhi 43 (4), 281284.
doi: 10.3760/cma.j.cn112147-20200229-00219
Zhou, P., Yang, X.-L., Wang, X.-G., Hu, B., Zhang, L., Zhang, W., et al. (2020). A
pneumonia outbreak associated with a new coronavirus of probable bat origin.
Nature 579 (7798), 270273. doi: 10.1038/s41586-020-2012-7
Zou, L., Ruan, F., Huang, M., Liang, L., Huang, H., Hong, Z., et al. (2020). SARS-
CoV-2 viral load in upper respiratory specimens of infected patients. N. Engl. J.
Med. 382 (12), 11771179. doi: 10.1056/NEJMc2001737
Zoufaly, A., Poglitsch, M., Aberle, J. H., Hoepler, W., Seitz, T., Traugott, M., et al.
(2020). Human recombinant soluble ACE2 in severe COVID-19. Lancet
Respirat. Med. 8 (11), 11541158. doi: 10.1016/S2213-2600(20)30418-5
Conict of Interest: The authors declare that the research was conducted in the
absence of any commercial or nancial relationships that could be construed as a
potential conict of interest.
Copyright © 2021 Alipoor, Mortaz, Jamaati, Tabarsi, Bayram, Varahram and
Adcock. This is an open-access article distributed under the terms of the Creative
Commons Attribution License (CC BY). The use, distribution or reproduction in other
forums is permitted, provided the original author(s) and the copyright owner(s) are
credited and that the original publication in this journal is cited, in accordance with
accepted academic practice. No use, distribution or reproduction is permitted which
does not comply with these terms.
Alipoor et al. COVID-19 and Response
Frontiers in Cellular and Infection Microbiology | www.frontiersin.org February 2021 | Volume 11 | Article 56308516
... SARS-CoV-2 binds to ACE2 through its S protein -a glycosylated type I transmembrane fusion protein -and is then provided entry into the cell through the cleavage created by S proteins and the use of type II transmembrane serine protease TMPRSS2. Once inside the cell, the virus' RNA creates viral proteins and particles, which then fuse with the plasma membrane to release the virus [5]. Additionally, mutations in the Spike protein of SARS-CoV-2, including deletions and substitutions, have implications for viral transmission, treatment effectiveness and diagnostic procedures, indicating a fitness advantage that enhances virus transmission and severity [17]. ...
Article
Full-text available
Aim: Describe the utility of an inverse reinforcement learning pathway to develop a novel model to predict and manage the spread of COVID-19. Materials & methods: Convolutional neural network (CNN) with multilayer perceptron (MLP) modeling functions utilized inverse reinforcement learning to predict COVID-19 outcomes based on a comprehensive array of factors. Results: Our model demonstrates a sensitivity of 0.67 in the receiver operating characteristic curve and can correctly identify approximately 67% of the positive cases. Conclusion: We demonstrate the ability to augment clinical decision-making with a novel artificial intelligence (AI) solution that accurately predicted the susceptibility of transplant patients to COVID-19. This enables physicians to administer treatment and take appropriate preventative measures based on patients' risk factors.
... The differences in SARS-CoV-2 response between patients may be due to virus infectivity and immune system inter-personal variability. Type I IFNs and their signaling pathways are activated during infection, triggering a adaptive immune response, in which B and T cells play the main role, producing antibodies and controlling viral dissemination [12]. ...
... The high expression of ACE2 on the small bowel enterocyte brush border and co-expression of TMPRSS2 supports potential enteric infection by SARS-CoV-2 [98,99]. Following the viral entry into the host cell, single-stranded RNA is recognized by the innate immune system via PRRs such as TLR7 and TLR8, RIG-I (retinoic acid-inducible gene I)-like receptors (RLRs), and the NOD-like receptor, CARD-containing-2 (NLRC2) [100]. In addition to the secretion of type-I and -III antiviral interferons (IFNs) and chemokines during early response [55], pro-inflammatory cytokines and chemokines are also produced by epithelial cells during SARS-CoV-2 infection, leading to the characteristic "cytokine storm" and ARDS of severe COVID-19 [101]. ...
Article
Full-text available
The gut microbiome plays a critical role in maintaining overall health and immune function. However, dysbiosis, an imbalance in microbiome composition, can have profound effects on various aspects of human health, including susceptibility to viral infections. Despite numerous studies investigating the influence of viral infections on gut microbiome, the impact of gut dysbiosis on viral infection and pathogenesis remains relatively understudied. The clinical variability observed in SARS-CoV-2 and seasonal influenza infections, and the presence of natural HIV suppressors, suggests that host-intrinsic factors, including the gut microbiome, may contribute to viral pathogenesis. The gut microbiome has been shown to influence the host immune system by regulating intestinal homeostasis through interactions with immune cells. This review aims to enhance our understanding of how viral infections perturb the gut microbiome and mucosal immune cells, affecting host susceptibility and response to viral infections. Specifically, we focus on exploring the interactions between gamma delta (γδ) T cells and gut microbes in the context of inflammatory viral pathogenesis and examine studies highlighting the role of the gut microbiome in viral disease outcomes. Furthermore, we discuss emerging evidence and potential future directions for microbiome modulation therapy in the context of viral pathogenesis.
... La-related Protein 1 (1e9)/angiotensin 2 receptor (AT2R) and Ang-(1e7)/Mas1 receptor downstream signaling but higher Ang-II/AT1R signaling pathway [31]. Overall, dysregulation of the RAAS leads to inflammation, vasoconstriction, fibrosis, oxidation, and capillary permeability which can all be a factor in triggering cancer progression and development [26e30, 32,33]. The formation of cancer stem cells may also result from the actions of Ang-II, resulting in its link to carcinogenesis, metastasis and relapse [34]. ...
Article
Severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) has shown diverse life-threatening effects, most of which are considered short-term. In addition to its short-term effects, which has claimed many millions of lives since 2019, the long-term complications of this virus are still under investigation. Similar to many oncogenic viruses, it has been hypothesized that SARS-CoV-2 employs various strategies to cause cancer in different organs. These include leveraging the renin angiotensin system, altering tumor suppressing pathways by means of its nonstructural proteins, and triggering inflammatory cascades by enhancing cytokine production in the form of a "cytokine storm" paving the way for the emergence of cancer stem cells in target organs. Since infection with SARS-CoV-2 occurs in several organs either directly or indirectly, it is expected that cancer stem cells may develop in multiple organs. Thus, we have reviewed the impact of coronavirus disease 2019 (COVID-19) on the vulnerability and susceptibility of specific organs to cancer development. It is important to note that the cancer-related effects of SARS-CoV-2 proposed in this article are based on the ability of the virus and its proteins to cause cancer but that the long-term consequences of this infection will only be illustrated in the long run.
... Wabah Covid-19 memiliki dampak yang dapat mengubah tatanan kehidupan di keluarga maupun masyarakat seperti kesehatan, kehidupan sosial, ekonomi, sumber daya manusia, sumber daya alam, dan lain-lain (Gunadi et al., 2020). Pandemi Covid-19 mulai terjadi pada akhir tahun 2019 yang menyebabkan beberapa negara harus di lockdown untuk mencegah penyebaran virus (Alipoor et al., 2021). Upaya pencegahan menjadi sangat penting seperti diagnosis, pengobatan dan vaksinasi. ...
Article
Pandemi Covid-19 menuntut masyarakat di dunia untuk hidup sehat, hingga saat ini belum ada obat atau vaksin yang efektif untuk menghentikan laju penyebaran Covid-19. Indonesia memiliki kearifan lokal untuk memanfaatkan tanaman sebagai alternatif pengobatan. Penggunaan tanaman sebagai pengobatan tradisional menjadi salah satu solusi hidup sehat di masa pandemi Covid-19. Tanaman memiliki banyak khasiat yang bisa dimanfaatkan untuk meningkatkan sistem imun tubuh dan mencegah penyebaran Covid-19. Namun informasi tersebut masih banyak yang belum terdokumentasi dan menyebar kepada seluruh masyarakat dalam berbagai macam kalangan usia. Penelitian studi etnomedisin telah dilakukan untuk membantu mendokumentasikan dan memberikan informasi terkait manfaat tanaman yang ada di sekitar kita khususnya untuk pencegahan Covid-19. Metode yang digunakan adalah snowball sampling, hasil yang didapatkan adalah terdapat >100 jenis tanaman yang dimanfaatkan oleh masyarakat untuk pengobatan tradisional. Tanaman seperti jahe, kunyit, dan tempuyung memiliki potensi untuk mencegah penyebaran coronavirus. Hasil penelitian yang diperoleh disosialisasikan kepada masyarakat luas secara nasional melalui sarana zoom meeting.
... 51 Barge (2021) performed homology modelling of the transmembrane serine protease 2 (TMPRSS2), responsible for the viral entry. The hit compounds can block the TMPRSS2 activity and resist the entry of the SARS-CoV-2 virus into targeted human cells by reducing the virus's infectivity and transmissibility (Alipoor et al. 2021) (Fig. 5). ...
... The viral S protein binds to angiotensin-converting enzyme-2 (ACE-2) receptors on the surface of target cells and thus attacks the cell. 3,6 T-cell lymphopenia is usually seen in the severe form of the disease, which may be due to the infiltration of leukocytes into the airways or excessive T-cell activity resulting in the expression of proapoptotic molecules. [7][8][9][10] In animal models of COVID-19, lymphopenia causes overactive T cells. ...
Article
Full-text available
COVID-19, caused by SARS-CoV-2, requires new approaches to control the disease. Programmed cell death protein (PD-1) and cytotoxic T-lymphocyte–associated protein 4 (CTLA-4) play important roles in T-cell exhaustion in severe COVID-19. This study evaluated the frequency of whole blood lymphocytes expressing PD-1 and CTLA-4 in COVID-19 patients upon admission to the intensive care unit (ICU) (i.e., severe) or infection ward (i.e., moderate) and after 7 days of antiviral therapy. COVID-19 patients were treated with either favipiravir or Kaletra (FK group, 11 severe and 11 moderate) or dexamethasone plus remdesivir (DR group, 7 severe and 10 moderate) for 7 days in a pilot study. Eight healthy control subjects were also enrolled. The frequency of PD-1+ and CTLA-4+ lymphocytes in whole blood was evaluated by flow cytometry. Patients on DR therapy had shorter hospital stays than those on FK therapy. The frequency of PD-1+ lymphocytes in the FK group at baseline differed between COVID-19 patients and healthy controls, while the frequency of both PD-1+ and CTLA-4+ cells increased significantly 7 days of FK therapy. The response was similar in both moderate and severe patients. In contrast, the frequency of PD-1+ and CTLA-4+ lymphocytes varied significantly between patients and healthy controls before DR treatment. DR therapy enhanced PD-1+ but not the CTLA-4+ frequency of these cells after 7 days. We show that the frequency of PD-1 and CTAL-4-bearing lymphocytes during hospitalization was increased in Iranian ICU COVID-19 patients who received FK treatment, but that the frequency of CTLA-4+ cells was higher at baseline and did not increase in patients who received DR. The effectiveness of DR treatment may reflect differences in T-cell activation or exhaustion status, particularly in CTLA-4-expressing cells.
Article
Full-text available
The COVID-19 pandemic caused by SARS-CoV-2 has resulted in millions of confirmed cases and deaths worldwide. Understanding the biological mechanisms of SARS-CoV-2 infection is crucial for the development of effective therapies. This study conducts differential expression (DE) analysis, pathway analysis, and differential network (DN) analysis on RNA-seq data of four lung cell lines, NHBE, A549, A549.ACE2, and Calu3, to identify their common and unique biological features in response to SARS-CoV-2 infection. DE analysis shows that cell line A549.ACE2 has the highest number of DE genes, while cell line NHBE has the lowest. Among the DE genes identified for the four cell lines, 12 genes are overlapped, associated with various health conditions. The most significant signaling pathways varied among the four cell lines. Only one pathway, “cytokine-cytokine receptor interaction”, is found to be significant among all four cell lines and is related to inflammation and immune response. The DN analysis reveals considerable variation in the differential connectivity of the most significant pathway shared among the four lung cell lines. These findings help to elucidate the mechanisms of SARS-CoV-2 infection and potential therapeutic targets.
Article
Full-text available
Introduction: The new coronavirus has caused a pandemic that has infected hundreds of millions of people around the world since its outbreak. But the cardiovascular damage caused by the new coronavirus is unknown. We have analyzed the current global scenario and the general pattern of growth. After summarizing the known relationship between cardiovascular diseases and new coronary pneumonia, relevant articles are analyzed through bibliometrics and visualization. Methods: Following our pre-designed search strategy, we selected publications on COVID-19 and cardiovascular disease in the Web of Science database. In our relevant bibliometric visualization analysis, a total of 7,028 related articles in the WOS core database up to 20th October 2022 were summarized, and the most prolific authors, the most prolific countries, and the journals and institutions that published the most articles were summarized and quantitatively analyzed. Results: SARS-CoV-2 is more infectious than SARS-CoV-1 and has significant involvement in the cardiovascular system in addition to pulmonary manifestations, with a difference of 10.16% (20.26%/10.10%) in the incidence of cardiovascular diseases. The number of cases increases in winter and decreases slightly in summer with temperature changes, but the increase in cases tends to break out of seasonality across the region as mutant strains emerge. The co-occurrence analysis found that with the progress of the epidemic, the research keywords gradually shifted from ACE2 and inflammation to the treatment of myocarditis and complications, indicating that the research on the new crown epidemic has entered the stage of prevention and treatment of complications. Conclusion: When combined with the current global pandemic trend, how to improve prognosis and reduce human body damage could become a research focus. At the same time, timely detection, prevention, and discovery of new mutant strains have also become key tasks in the fight against the epidemic, and full preparations have been made to prevent the spread of the next wave of mutant strains, and still need to continue to pay attention to the differential performance of the variant “omicron.”
Article
Full-text available
The emergence of SARS-CoV-2 led to pandemic spread of coronavirus disease 2019 (COVID-19), manifesting with respiratory symptoms and multi-organ dysfunction. Detailed characterization of virus-neutralizing antibodies and target epitopes is needed to understand COVID-19 pathophysiology and guide immunization strategies. Among 598 human monoclonal antibodies (mAbs) from ten COVID-19 patients, we identified 40 strongly neutralizing mAbs. The most potent mAb CV07-209 neutralized authentic SARS-CoV-2 with IC50 of 3.1 ng/ml. Crystal structures of two mAbs in complex with the SARS-CoV-2 receptor-binding domain at 2.55 and 2.70 Å revealed a direct block of ACE2 attachment. Interestingly, some of the near-germline SARS-CoV-2 neutralizing mAbs reacted with mammalian self-antigens. Prophylactic and therapeutic application of CV07-209 protected hamsters from SARS-CoV-2 infection, weight loss and lung pathology. Our results show that non-self-reactive virus-neutralizing mAbs elicited during SARS-CoV-2 infection are a promising therapeutic strategy.
Article
Full-text available
SARS-CoV-2, a novel coronavirus (CoV) that causes COVID-19, has recently emerged causing an ongoing outbreak of viral pneumonia around the world. While distinct from SARS-CoV, both group 2B CoVs share similar genome organization, origins to bat CoVs, and an arsenal of immune antagonists. In this report, we evaluate type I interferon (IFN-I) sensitivity of SARS-CoV-2 relative to the original SARS-CoV. Our results indicate that while SARS-CoV-2 maintains similar viral replication to SARS-CoV, the novel CoV is much more sensitive to IFN-I. In Vero E6 and in Calu3 cells, SARS-CoV-2 is substantially attenuated in the context of IFN-I pretreatment, whereas SARS-CoV is not. In line with these findings, SARS-CoV-2 fails to counteract phosphorylation of STAT1 and expression of ISG proteins, while SARS-CoV is able to suppress both. Comparing SARS-CoV-2 and influenza A virus in human airway epithelial cultures, we observe the absence of IFN-I stimulation by SARS-CoV-2 alone but detect the failure to counteract STAT1 phosphorylation upon IFN-I pretreatment, resulting in near ablation of SARS-CoV-2 infection. Next, we evaluated IFN-I treatment postinfection and found that SARS-CoV-2 was sensitive even after establishing infection. Finally, we examined homology between SARS-CoV and SARS-CoV-2 in viral proteins shown to be interferon antagonists. The absence of an equivalent open reading frame 3b (ORF3b) and genetic differences versus ORF6 suggest that the two key IFN-I antagonists may not maintain equivalent function in SARS-CoV-2. Together, the results identify key differences in susceptibility to IFN-I responses between SARS-CoV and SARS-CoV-2 that may help inform disease progression, treatment options, and animal model development. IMPORTANCE With the ongoing outbreak of COVID-19, differences between SARS-CoV-2 and the original SARS-CoV could be leveraged to inform disease progression and eventual treatment options. In addition, these findings could have key implications for animal model development as well as further research into how SARS-CoV-2 modulates the type I IFN response early during infection.
Article
The severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) virus is causing a global pandemic, and cases continue to rise. Most infected individuals experience mildly symptomatic coronavirus disease 2019 (COVID-19), but it is unknown whether this can induce persistent immune memory that could contribute to immunity. We performed a longitudinal assessment of individuals recovered from mild COVID-19 to determine whether they develop and sustain multifaceted SARS-CoV-2-specific immunological memory. Recovered individuals developed SARS-CoV-2-specific immunoglobulin (IgG) antibodies, neutralizing plasma, and memory B and memory T cells that persisted for at least 3 months. Our data further reveal that SARS-CoV-2-specific IgG memory B cells increased over time. Additionally, SARS-CoV-2-specific memory lymphocytes exhibited characteristics associated with potent antiviral function: memory T cells secreted cytokines and expanded upon antigen re-encounter, whereas memory B cells expressed receptors capable of neutralizing virus when expressed as monoclonal antibodies. Therefore, mild COVID-19 elicits memory lymphocytes that persist and display functional hallmarks of antiviral immunity.
Article
The high mortality rate of Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) infection is a critical concern of the coronavirus disease 2019 (COVID-19) pandemic. Strikingly, men account for the majority of COVID-19 deaths, with current figures ranging from 59% to 75% of total mortality. However, despite clear implications in relation to COVID-19 mortality, most research has not considered sex as a critical factor in data analysis. Here, we highlight fundamental biological differences that exist between males and females, and how these may make significant contributions to the male-biased COVID-19 mortality. We present preclinical evidence identifying the influence of biological sex on the expression and regulation of angiotensin-converting enzyme 2 (ACE2), which is the main receptor used by SARS-CoV-2 to enter cells. However, we note that there is a lack of reports showing that sexual dimorphism of ACE2 expression exists and is of functional relevance in humans. In contrast, there is strong evidence, especially in the context of viral infections, that sexual dimorphism plays a central role in the genetic and hormonal regulation of immune responses, both of the innate and the adaptive immune system. We review evidence supporting that ineffective anti-SARS-CoV-2 responses, coupled with a predisposition for inappropriate hyperinflammatory responses, could provide a biological explanation for the male bias in COVID-19 mortality. A prominent finding in COVID-19 is the increased risk of death with pre-existing cardiovascular comorbidities, such as hypertension, obesity, and age. We contextualize how important features of sexual dimorphism and inflammation in COVID-19 may exhibit a reciprocal relationship with comorbidities, and explain their increased mortality risk. Ultimately, we demonstrate that biological sex is a fundamental variable of critical relevance to our mechanistic understanding of SARS-CoV-2 infection and the pursuit of effective COVID-19 preventative and therapeutic strategies.
Article
Introduction: Immunopathogenesis is the process of the development of a disease that involves immune response or immune system. The focus of the study is basically on the understanding of control and management along with the role-play of host immune responses and the pathogenesis of the microorganism. Coronavirus (CoV) is a large family of viruses which will lead to mild to severe illness, with symptoms most commonly cough, cold, and fever. The severe forms include Middle East Respiratory Syndrome (MERS-CoV) and Severe Acute Respiratory Syndrome (SARS-CoV). The worldwide pandemic COVID-19 was started in Wuhan city. Wuhan city, located in the Hubei province has a huge market called Huanan Seafood market housing numerous non-vegetarian foods which has now become the place where the pandemic started. A new strain of the virus that had not been earlier detected in humans is the Novel coronavirus (nCoV). These ranges of viruses are known to be zoonotic in nature, which means they can be easily transmitted between animals and human patients. The significant challenge faced with these viruses is their evolutionary nature. Materials and Methods: This is a review article wherein various articles were searched through search engines like Google Scholar and Pubmed using keywords like pathogenesis of COVID-19, immune mechanism of COVID-19, the role of TNF-Alpha, interferons, and diagnosis of COVID-19. Over 50 articles were collected and reviewed thoroughly. Results and Discussion: The nCoV is made up of a special protein called the glycosylated protein, which is also called the Spike protein or S protein. It plays a vital role in allowing the entry of viruses into the host cell. The spike protein is a large type 1 transmembrane protein that contains a range of 1160 amino acids. Spike protein is a trimeric class one fusion protein that exists in a metadata level perfusion. The ectodomain of all coronaviruses is known to share the same organization in two domains: an N-terminal domain called S1 that is responsible for receptor binding action and a C terminal S2 domain which is responsible for fusion. This review will give details of the molecular immunopathogenesis of COVID-19.
Article
Significance Public health response to COVID-19 requires behavior changes—isolation at home, wearing masks. Its effectiveness depends on generalized compliance. Original data from two waves of a survey conducted in March−April 2020 in eight Organisation for Economic Co-operation and Development countries ( n = 21,649) show large gender differences in COVID-19−related beliefs and behaviors. Women are more likely to perceive the pandemic as a very serious health problem and to agree and comply with restraining measures. These differences are only partially mitigated for individuals cohabiting or directly exposed to COVID-19. This behavioral factor contributes to substantial gender differences in mortality and is consistent with women-led countries responding more effectively to the pandemic. It calls for gender-based public health policies and communication.
Preprint
COVID-19, caused by SARS-CoV-2, has recently affected over 300,000 people and killed more than 10,000. The manner in which the key immune cell subsets change and their states during the course of COVID-19 remain unclear. Here, we applied single-cell technology to comprehensively characterize transcriptional changes in peripheral blood mononuclear cells during the recovery stage of COVID-19. Compared with healthy controls, in patients in the early recovery stage (ERS) of COVID-19, T cells decreased remarkably, whereas monocytes increased. A detailed analysis of the monocytes revealed that there was an increased ratio of classical CD14 ⁺⁺ monocytes with high inflammatory gene expression as well as a greater abundance of CD14 ⁺⁺ IL1B ⁺ monocytes in the ERS. CD4 ⁺ and CD8 ⁺ T cells decreased significantly and expressed high levels of inflammatory genes in the ERS. Among the B cells, the plasma cells increased remarkably, whereas the naïve B cells decreased. Our study identified several novel B cell-receptor (BCR) changes, such as IGHV3-23 and IGHV3-7, and confirmed isotypes (IGHV3-15, IGHV3-30, and IGKV3-11) previously used for virus vaccine development. The strongest pairing frequencies, IGHV3-23-IGHJ4, indicated a monoclonal state associated with SARS-CoV-2 specificity. Furthermore, integrated analysis predicted that IL-1β and M-CSF may be novel candidate target genes for inflammatory storm and that TNFSF13, IL-18, IL-2 and IL-4 may be beneficial for the recovery of COVID-19 patients. Our study provides the first evidence of an inflammatory immune signature in the ERS, suggesting that COVID-19 patients are still vulnerable after hospital discharge. Our identification of novel BCR signaling may lead to the development of vaccines and antibodies for the treatment of COVID-19. Highlights - The immune response was sustained for more than 7 days in the early recovery stage of COVID-19, suggesting that COVID-19 patients are still vulnerable after hospital discharge. - Single-cell analysis revealed a predominant subset of CD14 ⁺⁺ IL1β ⁺ monocytes in patients in the ERS of COVID-19. - Newly identified virus-specific B cell-receptor changes, such as IGHV3-23, IGHV3-7, IGHV3-15, IGHV3-30, and IGKV3-11, could be helpful in the development of vaccines and antibodies against SARS-CoV-2. - IL-1β and M-CSF were discovered as novel mediators of inflammatory cytokine storm, and TNFSF13, IL-2, IL-4, and IL-18 may be beneficial for recovery.