Next Article in Journal
Lip-Reading Advancements: A 3D Convolutional Neural Network/Long Short-Term Memory Fusion for Precise Word Recognition
Previous Article in Journal
Toward Cancer Chemoprevention: Mathematical Modeling of Chemically Induced Carcinogenesis and Chemoprevention
Previous Article in Special Issue
Deep Learning and Federated Learning for Screening COVID-19: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Unravelling Insights into the Evolution and Management of SARS-CoV-2

by
Aganze Gloire-Aimé Mushebenge
1,2,3,*,
Samuel Chima Ugbaja
2,4,
Nonkululeko Avril Mbatha
4,5,
Rene B. Khan
2 and
Hezekiel M. Kumalo
2,*
1
Discipline of Pharmaceutical Sciences, University of KwaZulu-Natal, Westville, Durban 4000, South Africa
2
Drug Research and Innovation Unit, Discipline of Medical Biochemistry, School of Laboratory Medicine and Medical Science, University of KwaZulu-Natal, Durban 4000, South Africa
3
Faculty of Pharmaceutical Sciences, University of Lubumbashi, Lubumbashi 1825, Democratic Republic of the Congo
4
Africa Health Research Institute, University of KwaZulu-Natal, Durban 4000, South Africa
5
Department of Human Physiology, School of Laboratory Medicine and Medical Science, University of KwaZulu-Natal, Durban 4000, South Africa
*
Authors to whom correspondence should be addressed.
BioMedInformatics 2024, 4(1), 385-409; https://doi.org/10.3390/biomedinformatics4010022
Submission received: 28 October 2023 / Revised: 23 January 2024 / Accepted: 1 February 2024 / Published: 3 February 2024

Abstract

:
Worldwide, the COVID-19 pandemic, caused by the brand-new coronavirus SARS-CoV-2, has claimed a sizable number of lives. The virus’ rapid spread and impact on every facet of human existence necessitate a continuous and dynamic examination of its biology and management. Despite this urgency, COVID-19 does not currently have any particular antiviral treatments. As a result, scientists are concentrating on repurposing existing antiviral medications or creating brand-new ones. This comprehensive review seeks to provide an in-depth exploration of our current understanding of SARS-CoV-2, starting with an analysis of its prevalence, pathology, and evolutionary trends. In doing so, the review aims to clarify the complex network of factors that have contributed to the varying case fatality rates observed in different geographic areas. In this work, we explore the complex world of SARS-CoV-2 mutations and their implications for vaccine efficacy and therapeutic interventions. The dynamic viral landscape of the pandemic poses a significant challenge, leading scientists to investigate the genetic foundations of the virus and the mechanisms underlying these genetic alterations. Numerous hypotheses have been proposed as the pandemic has developed, covering various subjects like the selection pressures driving mutation, the possibility of vaccine escape, and the consequences for clinical therapy. Furthermore, this review will shed light on current clinical trials investigating novel medicines and vaccine development, including the promising field of drug repurposing, providing a window into the changing field of treatment approaches. This study provides a comprehensive understanding of the virus by compiling the huge and evolving body of knowledge on SARS-CoV-2, highlighting its complexities and implications for public health, and igniting additional investigation into the control of this unprecedented global health disaster.

1. Introduction

The pandemic caused by the outbreak of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) in late 2019 caused a considerable amount of illness and high mortality rates globally [1,2]. Coronaviruses, a group of RNA viruses, cause various diseases in humans and animals and exhibit a characteristic crown-like shape under an electron microscope [3]. These viruses are zoonotic and are transmitted from animals to humans [4]. Notable among them are SARS-CoV, Middle East respiratory syndrome (MERS-CoV), and SARS-CoV-2 [3,5]. SARS-CoV initiated a global outbreak in 2003, with over 8000 cases and significant mortality [6]. MERS-CoV, identified in 2012, primarily affects the Middle East with a higher mortality rate [7]. SARS-CoV-2, which causes COVID-19, emerged in Wuhan, China, in late 2019, leading to a widespread global pandemic [8]. All three viruses share zoonotic origins, posing significant challenges to public health. SARS-CoV and MERS-CoV have been managed through measures such as quarantine, contact tracing, and travel restrictions, while no specific treatment or vaccine exists for MERS-CoV, making it an ongoing concern, particularly in the Middle East [9]. SARS-CoV-2, the most recent coronavirus, demonstrates a zoonotic origin, possibly from bats and an intermediate host like a pangolin, before infecting humans [10].
SARS-CoV-2’s fast spread and mutation have posed a serious challenge to global public health and medical communities [11]. Since its appearance, extensive efforts have been made to understand how SARS-CoV-2 is transmitted, its clinical presentation, and treatment strategies [12]. Understanding the evolution and epidemiology of the virus has laid the groundwork for the creation of solutions to stop its spread and lessen the pandemic’s negative effects on public health and the global economy [13].
The continuing evolution of the SARS-CoV-2 virus hampered significantly the creation of efficient prevention and management plans [14]. To provide efficient treatments and vaccines, scientists and researchers from all over the world have made enormous efforts to understand the mechanics of viral reproduction and mutation [15]. A number of vaccines with high success rates against SARS-CoV-2 have been developed as a result of these efforts, and several clinical trials are ongoing to investigate additional therapeutic possibilities [16].
Due to the disease’s variable clinical symptoms and uncertain course, SARS-CoV-2 care still faces significant hurdles [17]. Fever, coughing, and shortness of breath are the most typical signs of COVID-19. However, the virus can also cause severe respiratory distress, pneumonia, and multi-organ failure [18,19]. Management of COVID-19 requires a multidisciplinary approach, including supportive care, oxygen therapy, and antiviral treatments [20].
With severe morbidity and mortality, and extensive interruptions to everyday life and economic activities, SARS-CoV-2 has had an unprecedented worldwide impact [21]. The pandemic has made it clear how crucial it is for nations to work together and have a coordinated public health response for facing new infectious illnesses [22]. A thorough knowledge of the virus’s evolution, transmission, and clinical symptoms is necessary for the creation of efficient SARS-CoV-2 preventive and management measures.
This work intends to offer a summary of the current knowledge about SARS-CoV-2, including its evolution (see Figure 1), transmission, clinical manifestation, and management, in this in-depth review. This study reviews the most recent findings on the processes of viral replication, mutation, and transmission and the status of available therapies and vaccine research. In addition, we will explore the global impact of the pandemic and the public health response to mitigate the spread of the virus. This paper contributes to the existing literature on SARS-CoV-2 and is organised as follows: clinical description of SARS-CoV-2, SARS-CoV-2 prevalence and pathology, discussion on postulated hypothesis on COVID-19 mutations, COVID-19 therapies, vaccines, and other ongoing clinical trials therapies, drug repurposing for COVID-19, non-pharmacological interventions, and finally a conclusion and future perspectives for COVID-19 research.

2. Clinical Description of SARS-CoV-2

Despite containment efforts, the COVID-19 pandemic, which started in Wuhan, China, in December 2019, spread rapidly throughout the world [24,25]. The respiratory system is the primary target of COVID-19 symptoms, which frequently include a dry cough, acute respiratory discomfort, pneumonia, and tiredness. A fever is present in virtually all cases [26]. As with its predecessors SARS-CoV and MERS-CoV, COVID-19 is brought on by SARS-CoV-2, which has caused a global health emergency and high rates of morbidity and mortality [27]. As of 2 February 2024, COVID-19 has had more than 702 million confirmed cases and 6.9 million fatalities with a regular daily updated status on the Worldometer coronavirus monitoring instrument [28,29]. The World Health Organization classified COVID-19 as a Public Health Emergency of International Concern because of the pandemic’s intensity [30].
In addition to the above typical clinical manifestations of the virus, some patients may also exhibit unusual symptoms such as diarrhoea, headaches, and myalgia [26,31,32]. The condition can be mild or catastrophic in severity, and some patients require mechanical breathing and intensive care [32,33]. Secondary infections such as bacterial pneumonia, which possibly worsen patient outcomes, can affect the clinical course of COVID-19 [34].
Understanding the clinical characteristics of the illness is essential for providing appropriate medical care and creating successful preventative and treatment plans as the COVID-19 pandemic spreads. Healthcare specialists throughout the world are working non-stop to contain the pandemic because of the severity of the illness and its potential for spread, which have made it a global health emergency. Critical elements in the fight against COVID-19 include quick diagnostic tests, sufficient personal protective equipment for healthcare workers, and efficient vaccinations and antiviral treatments. To guide public health policy and stop the spread of the disease, more studies are needed on the clinical manifestation and pathogenesis of SARS-CoV-2.

3. SARS-CoV-2 Prevalence and Pathology

Coronaviruses such as SARS-CoV, MERS-CoV, and SARS-CoV-2 were first discovered in bats and spread to humans via intermediate hosts such as civet cats and camels [35]. The first instances of human-to-animal transmission were recorded in Hong Kong in February 2020, but the COVID-19 pandemic has shown that companion animals such as cats and dogs can also be susceptible to the virus [36]. In addition, SARS-CoV-2 has been detected in monkeys, white-tailed deer, and minks, indicating a wider spectrum of possible hosts. Numerous animal species, including ferrets, hamsters, macaques, and baboons, have been shown susceptibility to SARS-CoV or SARS-CoV-2 infection in experimental infection studies [37,38].

3.1. SARS-CoV-2 Transmission, Clinical Presentation, and Risk Factors for Severe Disease and Fatality

There have been reports of gastrointestinal involvement, and the discovery of viral RNA in faecal samples raises the possibility of faecal–oral transmission [39]. Isolation and medical monitoring are essential preventive measures because asymptomatic carriers can unintentionally spread the infection to others [39,40]. A median hospital stay of 12 days is required for 20% of confirmed COVID-19 patients, and 25% of patients in hospitals require acute critical care [40]. Adults 55 years and older are the main demographic for severe COVID-19 cases. Mortality risk increases gradually, with a mortality rate of 1.4–4.9% in the 55–74 age group, 4.3–10.5% in the 75–84 age group, and 10.4–27.3% in the 85-plus age group [41].
According to the available data, SARS-CoV-2 is a naturally occurring virus that is mainly spread by inhaling cough droplets [42]. Another important method of transmission occurs when hands that have come into contact with droplet-contaminated surfaces touch the face, eyes, or nose [42,43]. Despite the lack of certainty regarding the seasonality of SARS-CoV-2, mounting evidence points to a potential role for climate factors in the spread of the virus [44].
There are three stages of SARS-CoV-2 clinical pathology: mild, severe, and critical, with critical being the stage that results in mortality [45]. Adults who are infected typically show no symptoms or only mild, temporary symptoms, whereas those who show symptoms are most contagious the day before symptoms appear [46,47].
With a typical incubation time of 4–6 days, COVID-19 clinical signs include respiratory and intestinal problems [48]. It is difficult to detect transmission chains and conduct subsequent tracing because the clinical symptoms are less severe than those associated with SARS and MERS infections [49]. Severe COVID-19 symptoms are more likely to develop in older, immune-compromised people with pre-existing diseases such as cardiovascular disease, hypertension, asthma, and diabetes [37,50].
There have been reports of gastrointestinal involvement, and the discovery of viral RNA in faecal samples raises the possibility of faecal–oral transmission [51,52]. Isolation and medical monitoring are essential preventative measures because asymptomatic carriers can unintentionally spread the infection to others [51]. To stop the virus’s cycle of spread, preventive measures were implemented, including limiting population movements, preventing large gatherings, and closing educational institutions [53].
Some individuals who recovered from COVID-19 had severe lung damage by the 10th day after the onset of symptoms [54]. In addition, pneumonia linked to SARS-CoV-2 infection was observed to include a significant portion of the lower respiratory tract [55]. Contrary to SARS-CoV and MERS-CoV, studies indicate that SARS-CoV-2 infection during pregnancy does not result in maternal fatalities, and there is no proof that the virus is transmitted to unborn children either intrauterinally or transplacentally [56].
Furthermore, long COVID, or post-acute sequelae of SARS-CoV-2 infection (PCC), is a prevalent condition that occurs months after COVID-19 infection and is characterised by symptoms such as fatigue, dyspnoea, memory loss, diffuse pain, and orthostasis [57]. PCC has significant medical, psychosocial, and economic impacts, leading to widespread unemployment and substantial lost wages [57,58]. Pathophysiological mechanisms involve central nervous system inflammation, viral reservoirs, persistent spike protein, cell receptor dysregulation, and autoimmunity [59]. In long-term COVID, symptoms persisted for up to two years post-infection [60]. This symptomatology is heterogeneous, and potential mechanisms include viral persistence, inflammation, immune dysregulation, autoimmune reactions, latent infections, endothelial dysfunction, and alterations in gut microbiota [60,61]. Healthcare systems are grappling with the social significance of post-COVID syndrome, emphasising the need for dynamic patient follow-up and rehabilitation programmes [62]. The U.S. CDC acknowledges symptoms lasting over four weeks as “Post-COVID Condition,” but the precise pathogenesis remains multifactorial. It involves organ dysfunction, immune system responses, and the effects of severe illness, causing microvascular injury and abnormalities in organ functioning [63]. Long-term symptoms, affecting approximately 10% of COVID survivors, necessitate a comprehensive rehabilitation approach to improve various body systems’ functions [64].

3.2. Profile Characteristics and Prognostic Markers of COVID-19/SARS-CoV-2

Aspartate aminotransferase and hypersensitive troponin I levels are high in COVID-19 individuals, and they also exhibit unique blood laboratory profile traits such as lymphopenia, leukopoenia, thrombocytopenia, and RNAaemia [65]. Procalcitonin levels begin at a normal range but gradually increase as the disease progresses, indicating an elevated risk of secondary infections [65,66]. Erythrocyte sedimentation rate (ESR) and C-reactive protein (CRP) levels have increased in COVID-19, whereas platelet count and procalcitonin levels are normally within the range. Aspartate aminotransferase (AST), alanine transaminase (ALT), lactate dehydrogenase (LDH), creatine phosphokinase (CPK), creatinine, and prothrombin time, which can be employed as diagnostic markers, are associated with higher levels in severe instances (see Figure 2) [65,66,67].
Prognostic indicators for COVID-19 include lymphopenia, neutrophilia, increased levels of LDH, C-reactive protein, D-dimer, total bilirubin, hepatic transaminases, ferritin, and troponins [68]. The numbers of T cells, B cells, and natural killer (NK) cells are also lowered in severe instances, as are the numbers of helpers, regulatory, and memory T cells. Severe instances are marked by higher levels of proinflammatory cytokines and chemokines such as IFN-, IL-1, IP-10, MCP-1, TNF-, G-CSF, IL-8, IL-10, and MIP-1A [67,68].
Unlike SARS and MERS, in which only Th1 cytokines are elevated, COVID-19 shows an elevation of both Th1 and Th2 cytokines [69]. The release of pro-IL-1b and subsequent synthesis of mature IL-1b, which regulates fever, pulmonary inflammation, and fibrosis, are triggered by the interaction of SARS-CoV-2 with Toll-like receptors (TLR) [70]. Therefore, IL-37 and IL-38 could be considered as appropriate therapeutic agents and may be highly beneficial in COVID-19 patients to reduce pulmonary inflammation by suppressing IL-1b and other proinflammatory IL-family members [71]. Different HLA types and varying epitope binding affinities influenced various immunopathological effects induced by the novel human coronavirus COVID-19 in humans [72].
Genomic and microsatellite analyses identified 1191 protein targets related to ACE2, revealing 305 host factors associated with COVID-19/asthma comorbidity [73]. Enrichment analyses highlighted the significance of metabolic processes, Th1 and Th2 cell differentiation, and PPAR signalling pathways in this comorbidity. Key host factors, including HRAS, IFNG, CAT, CDH1, FASN, ACLY, CCL5, VCAM1, SCD, and HMGCR, were identified, with some highly expressed in lung tissues [74,75,76]. In the exploration of genetic variations in the ACE2 and TMPRSS2 receptor genes in COVID-19 patients, there were associations between specific genotypes and expressions of these genes with SARS-CoV-2 positivity, including clinical outcomes [77]. The identification of haplotypes and variants/mutations contributes to the understanding of host genetic factors influencing COVID-19 susceptibility, offering insights for future vaccine development and therapeutic approaches [78,79].

4. Discussion on the Postulated Hypothesis on COVID-19 Mutations

Periods of relative stability followed by the emergence of variants marked the evolution of COVID-19. In an unprecedented effort to stop the COVID-19 epidemic, the scientific community from around the world has teamed up [80]. Similar to other RNA viruses, COVID-19 exhibits a high mutation rate that can be caused by copying errors during viral replication, recombination, or contact with agents that can neutralise the virus, such as host antibodies [81,82]. The transport and load of the virus to ACE2 target cells may be improved by molecular changes that lessen the damage to the viral capsid, which is crucial for safeguarding the viral genome and replication [82]. This would increase infectivity without necessarily changing the virus’ inherent pathogenicity or virulence. Mutations that increase the affinity of spike S-protein to receptors on ACE2 cells could also increase viral load and infectivity [83].
In the variants of concern, Omicron (B.1.1.529) played a pivotal role in influencing infectivity with numerous mutations in the receptor-binding domain (RBD). The Omicron lineage, characterised by numerous sublineages, including BA.1, BA.2, BA.4, and BA.5, rapidly became globally dominant, with variations in the receptor-binding domain (RBD) of the spike protein. These variants raise concerns regarding vaccine effectiveness and potential reinfections [84]. Despite these mutations, the Omicron RBD binds to the human ACE2 (hACE2) receptor with a similar affinity as the prototype RBD, possibly due to compensatory mutations [85]. Key residues identified, including Q493, Q498, N501, F486, K417, and F456, are crucial for tight binding to hACE2 [86]. In the Omicron variant, mutated residues R493, S496, and R498 formed new interactions with ACE2, compensating for the reduced ACE2 binding affinity, leading to similar biochemical ACE2 binding affinities for Delta and Omicron [87,88]. These mutations contribute to increased infectivity and transmissibility, accompanied by a notable ability to evade neutralising antibodies, highlighting the complex interplay between spike protein mutations and viral characteristics [87,88,89,90,91]. This molecular understanding could inform the development of therapeutic and prophylactic agents targeting these variants.
Furthermore, the D614G mutation in the SARS-CoV-2 spike protein was discovered early in the epidemic and quickly became the predominant form everywhere [54,55]. The D614G mutation in the SARS-CoV-2 spike protein enhances viral transmission and infectivity [92,93]. This mutation is associated with increased binding to the human cell-surface receptor ACE2, leading to heightened replication in primary human airway epithelial cultures, human ACE2 knock-in mice, and enhanced replication and transmissibility in hamster and ferret models [92,94]. The variant containing S(D614G) exhibited a very competitive advantage during the transmission bottleneck, explaining its global predominance [94,95,96]. Population genetic analysis indicates a selective advantage for the 614G variant, reflected in increased frequency, higher viral load, and a younger age of patients. This supports its role in improved transmission without indicating higher mortality or clinical severity [97,98].
The mutation improves the virus’s capacity to infect cells and may increase its transmissibility [99]. Other changes to the spike protein could make the virus more resistant to antibodies produced by the host immune system or vaccines, thus decreasing their effectiveness [100]. Therefore, it is essential to track SARS-CoV-2 mutations to spot any potentially dangerous variants and create workable defences (see Figure 3).
Studying and comprehending the underlying principles of viral evolution is essential given the rapid evolution of SARS-CoV-2 and the high mutation rates observed [98,99,100,102]. This information will be useful in creating efficient vaccinations and therapies that can be quickly modified to address newly discovered virus strains. The progress achieved thus far emphasises the need for ongoing multidisciplinary research and collaboration to contain the ongoing COVID-19 pandemic [101]. The scientific community has made tremendous progress in understanding the biology of SARS-CoV-2.

5. Advancements in SARS-CoV-2/COVID-19 Management

5.1. COVID-19 Therapies, Vaccines, and Other Ongoing Clinical Trial Therapies

Effective treatments and vaccines are urgently needed to combat the COVID-19 pandemic (see Table 1). It is crucial to create decoy ligands with structure-based designs that can obstruct crucial infectivity-mediating activities such as docking, binding, entrance, and replication [103]. While natural supplements such as vitamin D and zinc are being investigated for their potential prophylactic benefits, supercomputers have been used to screen the library of already licenced medications for potential therapeutics against COVID-19 [104]. Daily vitamin D supplements were found to protect against acute respiratory infections in a recent meta-analysis [105], whereas increased intracellular zinc concentrations have been demonstrated to hinder the reproduction of several RNA viruses [106]. However, excessive intake of zinc can lead to undesirable sequelae, and it is not currently recommended to give elemental zinc supplementation above the recommended dietary allowance for the prevention of COVID-19 [106,107].
Previous studies have shown reduced neutralisation of Omicron by postvaccination serum, indicating the need for ongoing surveillance of genetic and antigenic changes [96]. Monoclonal antibodies, a promising therapeutic for COVID-19, face challenges in effectiveness due to evolving spike mutations [108]. The Omicron variant, with 37 amino acid substitutions in the spike protein, exhibits enhanced affinity for the ACE2 receptor, causing marked reductions in neutralising activity against Omicron compared with the ancestral virus [109]. The emergence of variants, including Alpha, Beta, Gamma, and Delta, has further complicated the pandemic landscape, necessitating additional research avenues to understand their impact on transmission, reinfection risk, and vaccine protection [110,111]. The challenges of long-lasting symptoms in a young patient emphasise the need for therapeutic approaches such as heparin-induced extracorporeal LDL/fibrinogen precipitation (H.E.L.P.) apheresis [112].
Moreover, the COVA trial revealed that 20-hydroxyecdysone (BIO101) demonstrated statistically significant efficacy, showing a 43.8% reduction in the risk of death or respiratory failure (p = 0.0426), with nine patients needing treatment. The results confirm BIO101′s good safety profile and support its clinical relevance in modulating the renin– angiotensin system through MASR activation for treating COVID-19 [113,114]. These findings indicate the need for further investigation of BIO101 as a potential treatment for hospitalised patients with severe pneumonia due to COVID-19 [113].
Table 1. Summary of current potential anti-SARS-CoV-2 agents.
Table 1. Summary of current potential anti-SARS-CoV-2 agents.
Approved SARS-CoV-2 Vaccines
Vaccine NameManufacturerTypeDosage/Post-dosageEfficacyTargetRef.
BNT162b2Pfizer, BioNTechmRNATwo doses, 4–8 weeks apart. A booster (4–6 months after).95%Spike protein[115,116]
mRNA-1273ModernamRNATwo doses, 4–8 weeks apart. Dose 3 at least 4 weeks after Dose 2. A single booster dose (0.25 mL) may be administered at least 5 months after completing a primary series.94.1%Spike protein[115,117]
Ad26.COV2.SJohnson and JohnsonViral vectorOne dose (preferred). Administration of the second dose to increase level of protection against symptomatic infection. W.H.O recommends interval of 2 to 6 months. Booster dose after 90 days.72% (in the U.S.)Spike protein[115,118]
ChAdOx1-S [recombinant]AstraZeneca, University of OxfordViral vectorTwo doses, with an interval of 8 to 12 weeks. A booster dose may be considered 4–6 months after completion of the primary vaccination series.70.4% (average)Spike protein[115,119]
SARS-CoV-2 therapeutics
Drug nameManufacturerTypeTargetAntiviral AgentStatusRef.
RemdesivirGilead SciencesAntiviralRNA polymeraseNucleotide analogueFDA-approved for emergency use in hospitalised patients[120]
BaricitinibEli Lilly and CompanyAnti-inflammatoryAP-1Janus kinase inhibitorFDA-approved for emergency use along with remdesivir[121,122]
TocilizumabRocheAnti-inflammatoryIL-6The monoclonal antibody (mAb)FDA-approved for emergency use in hospitalised patients[123,124]
SotrovimabGlaxoSmithKline and Vir BiotechnologyThe monoclonal antibody (mAb)Spike proteinThe monoclonal antibody (mAb) FDA-approved for emergency use in high-risk individuals[125,126]
MolnupiravirMerck & Co.AntiviralRNA polymeraseNucleotide analogueCurrently under review for emergency use authorisation[127,128]
Ongoing Clinical Trials for SARS-CoV-2
Study nameSponsorTypePhaseTargetAntiviral Agent/StatusRef.
ACTIV-6NIHTherapeutics3VariousVarious/Ongoing[129,130]
COMET-ICENIAID, LillyTherapeutics3VariousVarious/Ongoing[126,131]
REGN-COV2RegeneronTherapeutics3Spike proteinThe monoclonal antibody (mAb)/Ongoing[132,133]
COV-BOOSTUniversity of OxfordVaccine2Spike proteinN/A/Ongoing[134]
COV-FLUNovavaxVaccine3Influenza virusN/A/Ongoing[135]

5.2. Drug Repurposing for COVID-19

Drug repurposing has become a vital tactic in the pursuit of successful COVID-19 treatment [93,136,137]. It includes looking at old antiviral medications and substances that are either approved or being researched for use against other viral infections [136]. The World Health Organization conducted the multinational clinical trial SOLIDARITY to investigate the efficacy of repurposed medications in treating COVID-19 [138]. Drugs such as lopinavir–ritonavir, chloroquine, remdesivir, and favipiravir are among those being investigated. The possible applications of these medications for treating COVID-19 will be discussed in this article [139].
The FDA has approved the medication combination Lopinavir–Ritonavir for the treatment of HIV-1. Ritonavir boosts the efficacy of lopinavir by delaying the rate at which it is metabolised in the liver, whereas lopinavir is a protease inhibitor that prevents virus particle formation [140]. The medication has shown some promise in treating COVID-19, but more clinical trial findings are needed to confirm its effectiveness [136,140].
An RNA-dependent RNA polymerase inhibitor called favipiravir has shown promise in combating influenza and other viral diseases [141]. Initial clinical trials conducted in Shenzhen and Wuhan have demonstrated its efficacy against SARS-CoV-2 [142]. Favipiravir-treated patients had a stronger therapeutic response, especially in terms of faster viral clearance and a higher rate of improvement in chest imaging [143]. The National Medical Products Administration of China has approved favipiravir as the first anti-COVID-19 medication in the nation considering these encouraging findings [142,143].
Moderately priced medications such as chloroquine and hydroxychloroquine, which are used to treat autoimmune disorders and malaria, are thought to reduce endosomal pH, preventing SARS-CoV-2 replication [43,144,145]. Although chloroquine (CQ) and hydroxychloroquine (HCQ) have shown in vitro inhibition of coronaviruses such as SARS-CoV and SARS-CoV-2, clinical study results have been inconsistent, indicating an incomplete understanding of their antiviral mechanisms [146,147]. Research indicates that HCQ dose dependently inhibits the ACE2 enzyme, but structural changes from the ACE2–S protein interaction may affect HCQ binding [148,149,150,151]. Despite FDA emergency use authorisation for COVID-19 treatment, it was revoked on 15 June 2020 because clinical studies demonstrated ineffectiveness in reducing death rates or shortening recovery time for hospitalised patients. These findings align with those of other studies questioning the ability of HCQ to inhibit or eliminate SARS-CoV-2 [152,153].
The nucleotide analogue prodrug Remdesivir provides broad-spectrum antiviral action against various RNA viruses [102]. In contrast to protease inhibitors, which focus on the late steps of virus reproduction, RNA-dependent RNA polymerase inhibits the early stage of viral replication [103]. It has been used as an experimental medicine for the treatment of Ebola, MERS-CoV, and SARS-CoV-2 and has been demonstrated to limit the replication of SARS-CoV-2 in animal models [104]. Clinical improvement was observed in 68% of patients treated with Remdesivir in a sample of critically ill patients hospitalised for severe COVID-19 [105,106]. A study revealed that remdesivir-treated patients, mostly men in their fifth decade with comorbidities, had a lower mortality risk, reduced incidence of ARDS, and lower ventilator dependency compared with the control group [154].The drug demonstrated effectiveness in mitigating kidney involvement caused by SARS-CoV-2, with no significant adverse effects reported [154,155].
Furthermore, a study explored drug repurposing as a cost-effective strategy for SARS-CoV-2 treatment by docking 16 antiviral approved drugs against the protease protein (6M03) responsible for viral replication [156]. Delavirdine, fosamprenavir, imiquimod, stavudine, and zanamivir exhibited excellent results, indicating their potentially strong drug profiles [156]. The isoquinoline derivative SLL-0197800, known for its anticoronaviral activities, was tested for its inhibitory effects on RNA-dependent RNA polymerase (RdRp) and 3′-to-5′ exoribonuclease (ExoN) proteins [157]. In vitro assays demonstrated potent inhibitory activities with low EC50 values (0.16 and 0.27 μM, respectively). In silico results supported these findings, indicating the potential of SLL-0197800 against various coronaviral strains, including future versions, such as SARS emphasising, emphasising its specific anticoronaviral qualities. This study underscores the potential of drug repurposing and highlights SLL-0197800 as a promising candidate for broader coronaviral infections [157].
In addition, therapeutic options have been sought in natural products (terpenoids, alkaloids, saponins, and phenolics) with promising in vitro and in silico results for use in COVID-19 disease [158,159,160]. Among these, the most studied are resveratrol, quercetin, hesperidin, curcumin, myricetin, and betulinic acid, which were proposed as SARS-CoV-2 inhibitors [159]. Regarding natural products, resveratrol, curcumin, and quercetin have demonstrated in vitro antiviral activity against SARS-CoV-2, and in vivo, a nebulised formulation has been demonstrated to alleviate the respiratory symptoms of COVID-19 [158,159,161].

5.2.1. Vaccine Development and Challenges

Scientists worldwide are engaged in a race to develop effective vaccines against SARS-CoV-2, using traditional and next-generation platforms. Current COVID-19 vaccines induce immune responses against the viral spike protein or its subunits, thereby preventing cellular entry [162]. The emergence of SARS-CoV-2 variants poses challenges, prompting the design of a peptide-based vaccine with epitopes targeting pathogenic proteins [45]. Computational methods, reverse vaccinology, and immunoinformatics guide the creation of a multi-epitope-based peptide vaccine by considering mutations in spike glycoprotein. Immune profiling, codon optimisation, and in silico simulations contribute to vaccine development [163].
Inactivated and attenuated vaccines, protein subunit and virus-like particle vaccines, viral vector-based vaccinations, and more recent DNA- and RNA-based vaccines are a few of the techniques for vaccine creation that are being studied [164]. All approaches are being developed simultaneously to create an effective vaccine, and each has advantages and disadvantages (Figure 4). Because it is present in all coronaviruses encountered and is exposed to the immune system of a person, the spike protein is thought to be the most promising vaccine among the structured proteins of the virus because it allows the body to mount an immune response against it and retain it for future defence [165].
Notably, all FDA-approved vaccines effectively generate neutralising antibodies. The Pfizer–BioNTech and Moderna mRNA vaccines, among others, have received global approval, reflecting ongoing efforts to combat the evolving virus [163,166,167]. Several vaccines have been approved in different parts of the world, such as CoronaVac, BBIBP-CorV, CoviVac, Covaxin, Oxford–AstraZeneca vaccine (ChAdOx1 nCoV-19), Sputnik V, the Johnson & Johnson vaccine, Convidicea, RBD-Dimer, and EpiVacCorona [168,169].
Figure 4. An overview of the main platforms for technology employed in the creation of the COVID-19 vaccine, the SARS-CoV-2 variants of concern, the corresponding changes in their spike proteins, and the variables that could affect the efficacy of the vaccines already on the market. Created with BioRender.com as adapted from source [170].
Figure 4. An overview of the main platforms for technology employed in the creation of the COVID-19 vaccine, the SARS-CoV-2 variants of concern, the corresponding changes in their spike proteins, and the variables that could affect the efficacy of the vaccines already on the market. Created with BioRender.com as adapted from source [170].
Biomedinformatics 04 00022 g004
The mRNA-1273 vaccine, a revolutionary RNA-based vaccine created by Moderna Therapeutics, was the first to start clinical testing [171]. To inject nanoparticles into the body, it exploits a portion of the spike protein genetic code [171,172]. The first phase I clinical trial of this vaccine, which had shown promise in animal research, began on March 16, 2020, in partnership with the National Institutes of Health (NIH), and involved 45 healthy people between the ages of 18 and 55 [172].
There are numerous additional mRNA-based vaccines in various stages of development [173]. To address viral infection, further cutting-edge vaccine strategies and treatment interventions are being explored. For instance, Codagenix developed a live-attenuated vaccine employing a reverse technique by changing the viral sequences by substituting its optimised codons with non-optimised ones to weaken the virus [174], and Inovio Pharmaceuticals’ INO-4800 is a DNA-based vaccination using the spike gene [175]. The Milken Institute COVID-19 Treatment and Vaccine Tracker provides information on several vaccinations currently being developed, along with their present progress [176]. Significant international vaccine funding organisations supported numerous innovative initiatives to select the most promising vaccines for future mass production.
Subsequently, a study examined the impact of different treatments on SARS-CoV-2 antibody levels in patients with autoimmune rheumatic disease undergoing Pfizer/BioNTech mRNA vaccination [177]. Original adalimumab treatment resulted in stable antibody levels, whereas biosimilar adalimumab showed a significant decrease [177]. This study underscores the dominant influence of treatment type on antibody changes over time, with negligible contributions from other variables [178]. In the context of mRNA vaccines, a booster administered 3–4 weeks after the initial vaccination substantially increased protective antibody levels [179]. Ipsilateral boost of the SARS-CoV-2 mRNA vaccine induced more germinal centre B cells specific to the receptor-binding domain (RBD), generating increased bone marrow plasma cells compared with the contralateral boost. An ipsilateral boost also rapidly produced high-affinity RBD-specific antibodies with enhanced cross-reactivity to the Omicron variant [179]. Evaluating T-cell responses in patients with late-stage lung cancer receiving immune-modulating agents, including anti-PD-1/PD-L1, showed distinct CD8+ T-cell responses with only marginal CD4+ T-cell responses to the Omicron variant. This emphasises the need for heightened protective measures for cancer patients because of qualitative deviations in T-cell responses [180].

5.2.2. Experimental Therapeutic Interventions

Convalescent Plasma (CP) Therapy

Since the start of the pandemic, there has been active research into experimental therapeutic interventions for COVID-19. Convalescent plasma (CP) therapy, a conventional method, has been used successfully for more than a century to treat infectious diseases such as SARS, MERS, and H1N1 [181]. To provide this treatment, neutralising antibody-rich plasma from a recovered patient is extracted and given to the infected patient [182]. Significant improvement has been observed in preliminary tests on patients with severe COVID-19, and further clinical trials are still being conducted. Researchers are also profiling specific antibodies from recovered patients to create functional antibodies as COVID-19 treatment, in addition to CP therapy [182,183].
Subsequently, more than 500 distinct antibodies were found in the serum of a recovering COVID-19 patient, and companies such as AbCellera and Eli Lilly are collaborating to create medicines based only on human IgG1 monoclonal antibodies (mAbs) [184]. mAbs targeting the SARS-CoV-2 spike protein, isolated from memory B cells of an individual infected with SARS-CoV in 2003, exhibited potent neutralisation of SARS-CoV-2 and SARS-CoV [185]. The antibody S309, which recognises a conserved epitope containing glycan, demonstrates strong neutralisation without competing with receptor attachment. Antibody cocktails including S309 enhance SARS-CoV-2 neutralisation, potentially limiting the emergence of escape mutants [185,186]. These findings highlight the therapeutic potential of these antibodies, especially S309, for prophylaxis or post-exposure therapy to mitigate severe COVID-19. In addition, a study identified 11 potent neutralising antibodies against SARS-CoV-2, with antibody 414–1 showing the best IC50 of 1.75 nM, providing promising therapeutic candidates for COVID-19 treatment [185,186,187].
In addition, cutting-edge methods for treating COVID-19, such as aerosolised siRNAs and nanoviricides, are being developed [188]. A technique created by Alnylam Pharmaceuticals that delivers aerosolised siRNAs directly to the lungs is undergoing in vitro and in vivo testing [189]. On the other hand, “virucidal nanomicelles” are being chemically attached to the S protein to form nanoviricides [190]. Because complement factor 5a is the primary contributor to tissue damage in patients, InflaRx and Beijing Defengrei Biotechnology are developing human IgG1 mAbs against it [189,190]. These antibodies have already received Chinese government approval for clinical trials. These cutting-edge treatments may be able to effectively cure COVID-19 and help to contain the current pandemic [190].

Soluble Human Angiotensin-Converting Enzyme 2

Because of its capacity to prevent SARS-CoV-2 replication, soluble human angiotensin-converting enzyme 2 (ACE2) has become recognised as a potential COVID-19 treatment option [191]. The virus enters human cells through the cellular receptor ACE2; hence, inhibiting the binding between the Spike protein and ACE2 may be a useful therapeutic approach [192]. Recent in vitro investigations have demonstrated the therapeutic potential of human recombinant soluble ACE2 (hrsACE2), which can considerably lower viral loads in Vero cells and block viral infection in constructed human blood arteries and kidney organoids [193,194]. These results indicate that hrsACE2 has the potential to safeguard patients from lung injury and SARS-CoV-2 infection by preventing viral entry into target cells (see Figure 5) [195].
In the fight against COVID-19, the use of soluble ACE2 as a therapeutic intervention shows promise [197]. hrsACE2 can prevent viral replication and lower viral loads by preventing communication between the virus and its host receptor [198]. Moreover, hrsACE2 has shown efficacy against SARS-CoV-2 in human blood vessels and kidney organoids, indicating its potential to protect patients from the virus’s severe lung damage [108]. Therefore, COVID-19-patient therapy may benefit from the use of soluble ACE2. However, more investigations are required to establish its security and effectiveness in clinical trials [199].

6. Non-Pharmacological Interventions

Implementing non-pharmacological interventions (NPIs) was, therefore, the most efficient public health response to the outbreak [38,199,200]. NPIs included early case detection and isolation, in-depth contact tracing of suspected secondary cases, travel prohibitions, tight contact reductions, physical segregation, increased cleanliness, and routine hand washing [201]. Closing non-essential public areas, services, and facilities was one of these strategies. Another is for educational institutions to switch to digital learning modalities and for enterprises to implement self-isolation/work from home programmes [38]. According to modelling projections, integrated NPIs are anticipated to have the biggest and fastest impact on reducing the reproductive number and slowing the rate of viral transmission if they are adopted early in the outbreak [199,202,203]. The creation of efficient treatment interventions and vaccines is made possible using the knowledge gained from these NPIs, which are interim measures while the effort to better understand viral genomes continues [38,201].
To reduce transmission, such measures include early detection of infected persons, seclusion, and tracking of their close connexions. Travel restrictions, social withdrawal, and enhanced hygiene helped to stop the virus’s spread [204]. NPIs can reduce viral reproduction and delay viral transmission if implemented early, which would reduce the impact of the outbreak. However, while NPIs are effective in the interim, efforts must continue to develop effective therapeutic interventions and vaccines to address the ongoing crisis [205,206].
Despite the challenges posed by the COVID-19 pandemic, maintaining a regular physical activity practise was proven to be beneficial for reducing the likelihood of Musculoskeletal (MSK) pain in infected individuals [187]. These findings are crucial for healthcare professionals managing long-term COVID patients experiencing MSK pain [207,208]. In addition, a study that supports the effectiveness of non-pharmacological strategies and a proposed control system grounded in control system theory aims to guide public decision policies in managing COVID-19 spread and preventing healthcare system overload by anticipating peak infection rates and implementing timely interventions [207,208,209].

7. Conclusions and Future Perspectives for COVID-19 Research

Reducing infections, lowering the strain on healthcare systems, and lessening the pandemic’s social and economic effects are the three main goals of efforts to limit the COVID-19 pandemic. Non-pharmacological therapies will continue to serve as the main line of protection while we wait for viable vaccinations. Therefore, projections and planning for anticipated healthcare capacity can be informed by accurate and current data on the daily number of new cases and the case characteristics [210]. The bacille Calmette—Guerin childhood vaccine as well as national immunisation programmes may have an impact on the pandemic intensity. COVID-19 will undoubtedly have a large worldwide impact that could take a long time to reverse [211]. To combat upcoming pandemics, healthcare systems must consider including efficient regulatory mechanisms. The approach to the present pandemic has already been influenced by lessons learned from the earlier SARS-CoV outbreaks in Hong Kong, Singapore, and Taiwan [212,213]. With regard to the pathogenicity, transmissibility, and therapeutic response of the viral isolates, genomic characterisation will also affect regional and global populations.
Additionally, artificial intelligence (AI) plays a crucial role in various aspects of addressing the challenges posed by COVID-19, including modelling and simulation, employing AI robotics for medical quarantine, and predicting the spread of the virus [214] AI systems, encompassing machine learning, deep learning, convolutional neural networks, and cognitive computing, offer significant utility in virus detection, comprehensive screening, continuous monitoring, and alleviating the workload for healthcare providers [199,214,215,216]. Furthermore, they proved instrumental in anticipating potential interactions with novel treatments. Considering these advancements, we recommend that computer scientists focus their efforts on refining and advancing these AI-driven methods, as they hold promise for effective responses to future outbreaks [38,214,215].
The international community must cooperate to make the greatest technology resources available to combat the current pandemic and prepare for potential future epidemics [217]. For the development of efficient vaccinations and the discovery of new drugs, it is essential to understand the genetic makeup of viral strains [172,218]. Data-driven initiatives should guide planning and predictions for expected healthcare capacity [219]. To combat upcoming pandemics, effective regulatory measures and national immunisation strategies should be considered [38,219]. To anticipate and track infections before an outbreak occurs, AI should be tested. The social, cultural, and economic infrastructures will be significantly impacted by the COVID-19 pandemic over the long term; thus, it is crucial to draw lessons from this experience and improve readiness for future breakouts.

Author Contributions

Conceptualization, A.G.-A.M., S.C.U. and H.M.K.; writing—original draft preparation, A.G.-A.M. and S.C.U.; writing—review and editing, A.G.-A.M., S.C.U. and N.A.M.; supervision, H.M.K. and R.B.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the College of Health Sciences, University of KwaZulu-Natal for the support.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Acter, T.; Uddin, N.; Das, J.; Akhter, A.; Choudhury, T.R.; Kim, S. Evolution of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) as coronavirus disease 2019 (COVID-19) pandemic: A global health emergency. Sci. Total Environ. 2020, 730, 138996. [Google Scholar] [CrossRef]
  2. Kamel Boulos, M.N.; Geraghty, E.M. Geographical tracking and mapping of coronavirus disease COVID-19/severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) epidemic and associated events around the world: How 21st century GIS technologies are supporting the global fight against outbreaks and epidemics. Int. J. Health Geogr. 2020, 19, 8. [Google Scholar] [PubMed]
  3. Nassar, A.; Ibrahim, I.M.; Amin, F.G.; Magdy, M.; Elgharib, A.M.; Azzam, E.B.; Nasser, F.; Yousry, K.; Shamkh, I.M.; Mahdy, S.M.; et al. A review of human coronaviruses’ receptors: The host-cell targets for the crown bearing viruses. Molecules 2021, 26, 6455. [Google Scholar] [CrossRef]
  4. Ye, Z.-W.; Yuan, S.; Yuen, K.-S.; Fung, S.-Y.; Chan, C.-P.; Jin, D.-Y. Zoonotic origins of human coronaviruses. Int. J. Biol. Sci. 2020, 16, 1686. [Google Scholar] [CrossRef]
  5. Chikowe, I.; Mtewa, A.G.; Tembo, D.; Smith, D.; Ibrahim, E.; Mwamatope, B.; Nkhungulu, J.; Kumpalume, P.; Maroyi, A. Potential of Malawi’s medicinal plants in COVID-19 disease management: A review. Malawi Med. J. 2021, 33, 85–107. [Google Scholar] [CrossRef] [PubMed]
  6. Singh, A.; Prasad, R.; Gupta, A.; Das, K.; Gupta, N. Severe acute respiratory syndrome coronavirus-2 and pulmonary tuberculosis: Convergence can be fatal. Monaldi Arch. Chest Dis. 2020, 90. [Google Scholar] [CrossRef]
  7. Bleibtreu, A.; Bertine, M.; Bertin, C.; Houhou-Fidouh, N.; Visseaux, B. Focus on Middle East respiratory syndrome coronavirus (MERS-CoV). Med. Mal. Infect. 2020, 50, 243–251. [Google Scholar] [CrossRef] [PubMed]
  8. Hu, B.; Guo, H.; Zhou, P.; Shi, Z.-L. Characteristics of SARS-CoV-2 and COVID-19. Nat. Rev. Microbiol. 2021, 19, 141–154. [Google Scholar] [CrossRef]
  9. Chamola, V.; Hassija, V.; Gupta, V.; Guizani, M. A comprehensive review of the COVID-19 pandemic and the role of IoT, drones, AI, blockchain, and 5G in managing its impact. IEEE Access 2020, 8, 90225–90265. [Google Scholar] [CrossRef]
  10. Liu, P.; Jiang, J.-Z.; Wan, X.-F.; Hua, Y.; Li, L.; Zhou, J.; Wang, X.; Hou, F.; Chen, J.; Zou, J. Are pangolins the intermediate host of the 2019 novel coronavirus (SARS-CoV-2)? PLoS Pathog. 2020, 16, e1008421. [Google Scholar] [CrossRef]
  11. Machado, D.J.; Scott, R.; Guirales, S.; Janies, D.A. Fundamental evolution of all Orthocoronavirinae including three deadly lineages descendent from Chiroptera-hosted coronaviruses: SARS-CoV, MERS-CoV and SARS-CoV-2. Cladistics 2021, 37, 461. [Google Scholar] [CrossRef]
  12. Dhama, K.; Patel, S.K.; Pathak, M.; Yatoo, M.I.; Tiwari, R.; Malik, Y.S.; Singh, R.; Sah, R.; Rabaan, A.A.; Bonilla-Aldana, D.K. An update on SARS-CoV-2/COVID-19 with particular reference to its clinical pathology, pathogenesis, immunopathology and mitigation strategies. Travel Med. Infect. Dis. 2020, 37, 101755. [Google Scholar] [CrossRef]
  13. Seitz, B.M.; Aktipis, A.; Buss, D.M.; Alcock, J.; Bloom, P.; Gelfand, M.; Harris, S.; Lieberman, D.; Horowitz, B.N.; Pinker, S.; et al. The pandemic exposes human nature: 10 evolutionary insights. Proc. Natl. Acad. Sci. USA 2020, 117, 27767–27776. [Google Scholar] [CrossRef] [PubMed]
  14. Villa, A.; Brunialti, E.; Dellavedova, J.; Meda, C.; Rebecchi, M.; Conti, M.; Donnici, L.; De Francesco, R.; Reggiani, A.; Lionetti, V.; et al. DNA aptamers masking angiotensin converting enzyme 2 as an innovative way to treat SARS-CoV-2 pandemic. Pharmacol. Res. 2022, 175, 105982. [Google Scholar] [CrossRef] [PubMed]
  15. Chen, Y.; Liu, Q.; Guo, D. Emerging coronaviruses: Genome structure, replication, and pathogenesis. J. Med. Virol. 2020, 92, 418–423. [Google Scholar] [CrossRef]
  16. Awadasseid, A.; Wu, Y.; Tanaka, Y.; Zhang, W. Effective drugs used to combat SARS-CoV-2 infection and the current status of vaccines. Biomed. Pharmacother. 2021, 137, 111330. [Google Scholar] [CrossRef]
  17. Hodgson, S.H.; Mansatta, K.; Mallett, G.; Harris, V.; Emary, K.R.; Pollard, A.J. What defines an efficacious COVID-19 vaccine? A review of the challenges assessing the clinical efficacy of vaccines against SARS-CoV-2. Lancet Infect. Dis. 2021, 21, e26–e35. [Google Scholar] [CrossRef]
  18. Sanyaolu, A.; Okorie, C.; Marinkovic, A.; Patidar, R.; Younis, K.; Desai, P.; Hosein, Z.; Padda, I.; Mangat, J.; Altaf, M. Comorbidity and its impact on patients with COVID-19. SN Compr. Clin. Med. 2020, 2, 1069–1076. [Google Scholar] [CrossRef] [PubMed]
  19. Singhal, T. A review of coronavirus disease-2019 (COVID-19). Indian J. Pediatr. 2020, 87, 281–286. [Google Scholar] [CrossRef] [PubMed]
  20. Abu Haleeqa, M.; Alshamsi, I.; Al Habib, A.; Noshi, M.; Abdullah, S.; Kamour, A.; Ibrahim, H. Optimizing supportive care in COVID-19 patients: A multidisciplinary approach. J. Multidiscip. Healthc. 2020, 13, 877–880. [Google Scholar] [CrossRef]
  21. Formenti, B.; Gregori, N.; Crosato, V.; Marchese, V.; Tomasoni, L.R.; Castelli, F. The impact of COVID-19 on communicable and non-communicable diseases in Africa: A narrative review. Le Infez. Med. 2022, 30, 30. [Google Scholar]
  22. Haldane, V.; De Foo, C.; Abdalla, S.M.; Jung, A.-S.; Tan, M.; Wu, S.; Chua, A.; Verma, M.; Shrestha, P.; Singh, S. Health systems resilience in managing the COVID-19 pandemic: Lessons from 28 countries. Nat. Med. 2021, 27, 964–980. [Google Scholar] [CrossRef]
  23. Jeong, G.U.; Song, H.; Yoon, G.Y.; Kim, D.; Kwon, Y.-C. Therapeutic strategies against COVID-19 and structural characterization of SARS-CoV-2: A review. Front. Microbiol. 2020, 11, 1723. [Google Scholar] [CrossRef]
  24. Chauhan, S. Comprehensive review of coronavirus disease 2019 (COVID-19). Biomed. J. 2020, 43, 334–340. [Google Scholar] [CrossRef]
  25. Liu, N.-N.; Tan, J.-C.; Li, J.; Li, S.; Cai, Y.; Wang, H. COVID-19 pandemic: Experiences in China and implications for its prevention and treatment worldwide. Curr. Cancer Drug Targets 2020, 20, 410–416. [Google Scholar]
  26. Pascarella, G.; Strumia, A.; Piliego, C.; Bruno, F.; Del Buono, R.; Costa, F.; Scarlata, S.; Agrò, F.E. COVID-19 diagnosis and management: A comprehensive review. J. Intern. Med. 2020, 288, 192–206. [Google Scholar] [CrossRef] [PubMed]
  27. Pal, M.; Berhanu, G.; Desalegn, C.; Kandi, V. Severe acute respiratory syndrome coronavirus-2 (SARS-CoV-2): An update. Cureus 2020, 12, e7423. [Google Scholar] [CrossRef] [PubMed]
  28. Loeffler-Wirth, H.; Schmidt, M.; Binder, H. COVID-19 transmission trajectories–monitoring the pandemic in the worldwide context. Viruses 2020, 12, 777. [Google Scholar] [CrossRef] [PubMed]
  29. Worldometer. COVID-19 Coronavirus Pandemic. Available online: https://www.worldometers.info/coronavirus/ (accessed on 27 October 2023).
  30. Xiang, M.; Zhang, Z.; Kuwahara, K. Impact of COVID-19 pandemic on children and adolescents’ lifestyle behavior larger than expected. Prog. Cardiovasc. Dis. 2020, 63, 531. [Google Scholar] [CrossRef] [PubMed]
  31. Petrosillo, N.; Viceconte, G.; Ergonul, O.; Ippolito, G.; Petersen, E. COVID-19, SARS and MERS: Are they closely related? Clin. Microbiol. Infect. 2020, 26, 729–734. [Google Scholar] [CrossRef] [PubMed]
  32. Pustake, M.; Tambolkar, I.; Giri, P.; Gandhi, C. SARS, MERS and COVID-19: An overview and comparison of clinical, laboratory and radiological features. J. Fam. Med. Prim. Care 2022, 11, 10. [Google Scholar] [CrossRef] [PubMed]
  33. Shang, Y.; Pan, C.; Yang, X.; Zhong, M.; Shang, X.; Wu, Z.; Yu, Z.; Zhang, W.; Zhong, Q.; Zheng, X. Management of critically ill patients with COVID-19 in ICU: Statement from front-line intensive care experts in Wuhan, China. Ann. Intensive Care 2020, 10, 73. [Google Scholar] [CrossRef] [PubMed]
  34. Ghosh, S.; Bornman, C.; Zafer, M.M. Antimicrobial Resistance Threats in the emerging COVID-19 pandemic: Where do we stand? J. Infect. Public Health 2021, 14, 555–560. [Google Scholar] [CrossRef]
  35. Rabi, F.A.; Al Zoubi, M.S.; Kasasbeh, G.A.; Salameh, D.M.; Al-Nasser, A.D. SARS-CoV-2 and coronavirus disease 2019: What we know so far. Pathogens 2020, 9, 231. [Google Scholar] [CrossRef] [PubMed]
  36. Meekins, D.A.; Gaudreault, N.N.; Richt, J.A. Natural and experimental SARS-CoV-2 infection in domestic and wild animals. Viruses 2021, 13, 1993. [Google Scholar] [CrossRef] [PubMed]
  37. Damialis, A.; Gilles, S.; Sofiev, M.; Sofieva, V.; Kolek, F.; Bayr, D.; Plaza, M.P.; Leier-Wirtz, V.; Kaschuba, S.; Ziska, L.H.; et al. Higher airborne pollen concentrations correlated with increased SARS-CoV-2 infection rates, as evidenced from 31 countries across the globe. Proc. Natl. Acad. Sci. USA 2021, 118, e2019034118. [Google Scholar] [CrossRef] [PubMed]
  38. Uddin, M.; Mustafa, F.; Rizvi, T.A.; Loney, T.; Al Suwaidi, H.; Al-Marzouqi, A.H.H.; Kamal Eldin, A.; Alsabeeha, N.; Adrian, T.E.; Stefanini, C.; et al. SARS-CoV-2/COVID-19: Viral genomics, epidemiology, vaccines, and therapeutic interventions. Viruses 2020, 12, 526. [Google Scholar] [CrossRef]
  39. Johnson, K.D.; Harris, C.; Cain, J.K.; Hummer, C.; Goyal, H.; Perisetti, A. Pulmonary and extra-pulmonary clinical manifestations of COVID-19. Front. Med. 2020, 7, 526. [Google Scholar] [CrossRef]
  40. Bhatraju, P.K.; Ghassemieh, B.J.; Nichols, M.; Kim, R.; Jerome, K.R.; Nalla, A.K.; Greninger, A.L.; Pipavath, S.; Wurfel, M.M.; Evans, L.; et al. COVID-19 in critically ill patients in the Seattle region—Case series. N. Engl. J. Med. 2020, 382, 2012–2022. [Google Scholar] [CrossRef]
  41. Callow, M.A.; Callow, D.D.; Smith, C. Older adults’ intention to socially isolate once COVID-19 stay-at-home orders are replaced with “safer-at-home” public health advisories: A survey of respondents in Maryland. J. Appl. Gerontol. 2020, 39, 1175–1183. [Google Scholar] [CrossRef]
  42. Azuma, K.; Yanagi, U.; Kagi, N.; Kim, H.; Ogata, M.; Hayashi, M. Environmental factors involved in SARS-CoV-2 transmission: Effect and role of indoor environmental quality in the strategy for COVID-19 infection control. Environ. Health Prev. Med. 2020, 25, 1–16. [Google Scholar] [CrossRef]
  43. Li, H.; Liu, Z.; Ge, J. Scientific research progress of COVID-19/SARS-CoV-2 in the first five months. J. Cell. Mol. Med. 2020, 24, 6558–6570. [Google Scholar] [CrossRef] [PubMed]
  44. Jit, M.; Cook, A.R. Informing Public Health Policies with Models for Disease Burden, Impact Evaluation, and Economic Evaluation. Annu. Rev. Public Health 2023, 45. [Google Scholar] [CrossRef] [PubMed]
  45. Rabaan, A.A.; Al-Ahmed, S.H.; Albayat, H.; Alwarthan, S.; Alhajri, M.; Najim, M.A.; AlShehail, B.M.; Al-Adsani, W.; Alghadeer, A.; Abduljabbar, W.A.; et al. Variants of SARS-CoV-2: Influences on the Vaccines’ Effectiveness and Possible Strategies to Overcome Their Consequences. Medicina 2023, 59, 507. [Google Scholar] [CrossRef] [PubMed]
  46. Killingley, B.; Mann, A.J.; Kalinova, M.; Boyers, A.; Goonawardane, N.; Zhou, J.; Lindsell, K.; Hare, S.S.; Brown, J.; Frise, R.; et al. Safety, tolerability and viral kinetics during SARS-CoV-2 human challenge in young adults. Nat. Med. 2022, 28, 1031–1041. [Google Scholar] [CrossRef] [PubMed]
  47. Ramadori, G.P. SARS-CoV-2-infection (COVID-19): Clinical course, viral acute respiratory distress syndrome (ARDS) and cause(s) of death. Med. Sci. 2022, 10, 58. [Google Scholar] [CrossRef]
  48. Walsh, K.A.; Spillane, S.; Comber, L.; Cardwell, K.; Harrington, P.; Connell, J.; Teljeur, C.; Broderick, N.; De Gascun, C.F.; Smith, S.M.; et al. The duration of infectiousness of individuals infected with SARS-CoV-2. J. Infect. 2020, 81, 847–856. [Google Scholar] [CrossRef] [PubMed]
  49. de Araujo, W.R.; Lukas, H.; Torres, M.D.; Gao, W.; de la Fuente-Nunez, C. Low-Cost Biosensor Technologies for Rapid Detection of COVID-19 and Future Pandemics. ACS Nano 2024, 18, 1757–1777. [Google Scholar] [CrossRef] [PubMed]
  50. Yang, W.; Cao, Q.; Qin, L.; Wang, X.; Cheng, Z.; Pan, A.; Dai, J.; Sun, Q.; Zhao, F.; Qu, J. Clinical characteristics and imaging manifestations of the 2019 novel coronavirus disease (COVID-19): A multi-center study in Wenzhou city, Zhejiang, China. J. Infect. 2020, 80, 388–393. [Google Scholar] [CrossRef]
  51. Baj, J.; Karakuła-Juchnowicz, H.; Teresiński, G.; Buszewicz, G.; Ciesielka, M.; Sitarz, R.; Forma, A.; Karakuła, K.; Flieger, W.; Portincasa, P.; et al. COVID-19: Specific and non-specific clinical manifestations and symptoms: The current state of knowledge. J. Clin. Med. 2020, 9, 1753. [Google Scholar] [CrossRef]
  52. Banerjee, D. The impact of COVID-19 pandemic on elderly mental health. Int. J. Geriatr. Psychiatry 2020, 35, 1466. [Google Scholar] [CrossRef]
  53. Suzuki, H.; Miyamoto, T.; Hamada, A.; Nakano, A.; Okoshi, H.; Yamasawa, F. A guide for businesses and employers responding to novel coronavirus disease (COVID-19). J. Occup. Health 2021, 63, e12225. [Google Scholar]
  54. Yuen, K.-S.; Ye, Z.-W.; Fung, S.-Y.; Chan, C.-P.; Jin, D.-Y. SARS-CoV-2 and COVID-19: The most important research questions. Cell Biosci. 2020, 10, 40. [Google Scholar] [CrossRef] [PubMed]
  55. Yong, S.J. Long COVID or post-COVID-19 syndrome: Putative pathophysiology, risk factors, and treatments. Infect. Dis. 2021, 53, 737–754. [Google Scholar] [CrossRef] [PubMed]
  56. Zhang, Y.; Geng, X.; Tan, Y.; Li, Q.; Xu, C.; Xu, J.; Hao, L.; Zeng, Z.; Luo, X.; Liu, F. New understanding of the damage of SARS-CoV-2 infection outside the respiratory system. Biomed. Pharmacother. 2020, 127, 110195. [Google Scholar] [CrossRef]
  57. Mueller, M.R.; Ganesh, R.; Hurt, R.T.; Beckman, T.J. Post-COVID conditions. Mayo Clin. Proc. 2023, 98, 1071–1078. [Google Scholar] [CrossRef] [PubMed]
  58. Dryden, M.; Mudara, C.; Vika, C.; Blumberg, L.; Mayet, N.; Cohen, C.; Tempia, S.; Parker, A.; Nel, J.; Perumal, R. Post-COVID-19 condition 3 months after hospitalisation with SARS-CoV-2 in South Africa: A prospective cohort study. Lancet Glob. Health 2022, 10, e1247–e1256. [Google Scholar] [CrossRef] [PubMed]
  59. Haidar, M.A.; Shakkour, Z.; Reslan, M.A.; Al-Haj, N.; Chamoun, P.; Habashy, K.; Kaafarani, H.; Shahjouei, S.; Farran, S.H.; Shaito, A. SARS-CoV-2 involvement in central nervous system tissue damage. Neural Regen. Res. 2022, 17, 1228. [Google Scholar] [PubMed]
  60. Marjenberg, Z.; Leng, S.; Tascini, C.; Garg, M.; Misso, K.; El Guerche Seblain, C.; Shaikh, N. Risk of long COVID main symptoms after SARS-CoV-2 infection: A systematic review and meta-analysis. Sci. Rep. 2023, 13, 15332. [Google Scholar] [CrossRef]
  61. Kenny, G.; Townsend, L.; Savinelli, S.; Mallon, P.W. Long COVID: Clinical characteristics, proposed pathogenesis and potential therapeutic targets. Front. Mol. Biosci. 2023, 10, 1157651. [Google Scholar] [CrossRef]
  62. Derksen, C.; Rinn, R.; Gao, L.; Dahmen, A.; Cordes, C.; Kolb, C.; Becker, P.; Lippke, S. Longitudinal Evaluation of an Integrated Post–COVID-19/Long COVID Management Program Consisting of Digital Interventions and Personal Support: Randomized Controlled Trial. J. Med. Internet Res. 2023, 25, e49342. [Google Scholar] [CrossRef] [PubMed]
  63. Sherif, Z.A.; Gomez, C.R.; Connors, T.J.; Henrich, T.J.; Reeves, W.B. Pathogenic mechanisms of post-acute sequelae of SARS-CoV-2 infection (PASC). eLife 2023, 12, e86002. [Google Scholar] [CrossRef]
  64. Singh, S.J.; Baldwin, M.M.; Daynes, E.; Evans, R.A.; Greening, N.J.; Jenkins, R.G.; Lone, N.I.; McAuley, H.; Mehta, P.; Newman, J.; et al. Respiratory sequelae of COVID-19: Pulmonary and extrapulmonary origins, and approaches to clinical care and rehabilitation. Lancet Respir. Med. 2023, 11, 709–725. [Google Scholar] [CrossRef] [PubMed]
  65. Leveringhaus, E.S. Investigations of Cellular Determinants Involved in the Entry Process of Bovine and Porcine Pestiviruses. Ph.D. Thesis, Stiftung Tierärztliche Hochschule Hannover, Hannover, Germany, 2022. [Google Scholar]
  66. Samprathi, M.; Jayashree, M. Biomarkers in COVID-19: An up-to-date review. Front. Pediatr. 2021, 8, 607647. [Google Scholar] [CrossRef]
  67. Siavoshi, F.; Safavi-Naini, S.A.A.; Shirzadeh Barough, S.; Azizmohammad Looha, M.; Hatamabadi, H.; Ommi, D.; Jalili Khoshnoud, R.; Fatemi, A.; Pourhoseingholi, M.A. On-admission and dynamic trend of laboratory profiles as prognostic biomarkers in COVID-19 inpatients. Sci. Rep. 2023, 13, 6993. [Google Scholar] [CrossRef]
  68. Hachim, M.Y.; Hachim, I.Y.; Naeem, K.B.; Hannawi, H.; Salmi, I.A.; Hannawi, S. D-dimer, troponin, and urea level at presentation with COVID-19 can predict ICU admission: A single centered study. Front. Med. 2020, 7, 585003. [Google Scholar] [CrossRef]
  69. Ryabkova, V.A.; Churilov, L.P.; Shoenfeld, Y. Influenza infection, SARS, MERS and COVID-19: Cytokine storm–the common denominator and the lessons to be learned. Clin. Immunol. 2021, 223, 108652. [Google Scholar] [CrossRef]
  70. Campbell, G.R.; To, R.K.; Hanna, J.; Spector, S.A. SARS-CoV-2, SARS-CoV-1, and HIV-1 derived ssRNA sequences activate the NLRP3 inflammasome in human macrophages through a non-classical pathway. Iscience 2021, 24, 102295. [Google Scholar] [CrossRef]
  71. Alijotas-Reig, J.; Esteve-Valverde, E.; Belizna, C.; Selva-O’Callaghan, A.; Pardos-Gea, J.; Quintana, A.; Mekinian, A.; Anunciacion-Llunell, A.; Miró-Mur, F. Immunomodulatory therapy for the management of severe COVID-19. Beyond the anti-viral therapy: A comprehensive review. Autoimmun. Rev. 2020, 19, 102569. [Google Scholar] [CrossRef] [PubMed]
  72. Domingo, E.; García-Crespo, C.; Lobo-Vega, R.; Perales, C. Mutation rates, mutation frequencies, and proofreading-repair activities in RNA virus genetics. Viruses 2021, 13, 1882. [Google Scholar] [CrossRef] [PubMed]
  73. Yan, Q.; Lin, X.-Y.; Peng, C.-W.; Zheng, W.-J.; Liu, X.-H.; Wen, W.-J.; Jiang, Y.; Zhan, S.-F.; Huang, X.-F. Network-based analysis between SARS-CoV-2 receptor ACE2 and common host factors in COVID-19 and asthma: Potential mechanistic insights. Biomed. Signal Process. Control 2024, 87, 105502. [Google Scholar] [CrossRef]
  74. Chen, P.; Wu, M.; He, Y.; Jiang, B.; He, M.-L. Metabolic alterations upon SARS-CoV-2 infection and potential therapeutic targets against coronavirus infection. Signal Transduct. Target. Ther. 2023, 8, 237. [Google Scholar] [CrossRef]
  75. Lipman, D.; Safo, S.E.; Chekouo, T. Multi-omic analysis reveals enriched pathways associated with COVID-19 and COVID-19 severity. PLoS ONE 2022, 17, e0267047. [Google Scholar] [CrossRef]
  76. Sawalha, A.H.; Zhao, M.; Coit, P.; Lu, Q. Epigenetic dysregulation of ACE2 and interferon-regulated genes might suggest increased COVID-19 susceptibility and severity in lupus patients. Clin. Immunol. 2020, 215, 108410. [Google Scholar] [CrossRef]
  77. Yildirim, Z.; Sahin, O.S.; Yazar, S.; Bozok Cetintas, V. Genetic and epigenetic factors associated with increased severity of COVID-19. Cell Biol. Int. 2021, 45, 1158–1174. [Google Scholar] [CrossRef]
  78. Gemmati, D.; Longo, G.; Gallo, I.; Silva, J.A.; Secchiero, P.; Zauli, G.; Hanau, S.; Passaro, A.; Pellegatti, P.; Pizzicotti, S. Host genetics impact on SARS-CoV-2 vaccine-induced immunoglobulin levels and dynamics: The role of TP53, ABO, APOE, ACE2, HLA-A, and CRP genes. Front. Genet. 2022, 13, 1028081. [Google Scholar] [CrossRef]
  79. Ovsyannikova, I.G.; Haralambieva, I.H.; Crooke, S.N.; Poland, G.A.; Kennedy, R.B. The role of host genetics in the immune response to SARS-CoV-2 and COVID-19 susceptibility and severity. Immunol. Rev. 2020, 296, 205–219. [Google Scholar] [CrossRef]
  80. Van Damme, W.; Dahake, R.; Delamou, A.; Ingelbeen, B.; Wouters, E.; Vanham, G.; Van De Pas, R.; Dossou, J.-P.; Ir, P.; Abimbola, S.; et al. The COVID-19 pandemic: Diverse contexts; different epidemics—How and why? BMJ Glob. Health 2020, 5, e003098. [Google Scholar] [CrossRef] [PubMed]
  81. Hassanpour, M.; Rezaie, J.; Nouri, M.; Panahi, Y. The role of extracellular vesicles in COVID-19 virus infection. Infect. Genet. Evol. 2020, 85, 104422. [Google Scholar] [CrossRef] [PubMed]
  82. Poduri, R.; Joshi, G.; Jagadeesh, G. Drugs targeting various stages of the SARS-CoV-2 life cycle: Exploring promising drugs for the treatment of COVID-19. Cell. Signal. 2020, 74, 109721. [Google Scholar] [CrossRef] [PubMed]
  83. Jamwal, S.; Gautam, A.; Elsworth, J.; Kumar, M.; Chawla, R.; Kumar, P. An updated insight into the molecular pathogenesis, secondary complications and potential therapeutics of COVID-19 pandemic. Life Sci. 2020, 257, 118105. [Google Scholar] [CrossRef]
  84. Qu, P.; Faraone, J.N.; Evans, J.P.; Zheng, Y.-M.; Carlin, C.; Anghelina, M.; Stevens, P.; Fernandez, S.; Jones, D.; Panchal, A.R.; et al. Enhanced evasion of neutralizing antibody response by Omicron XBB.1.5, CH.1.1, and CA.3.1 variants. Cell Rep. 2023, 42, 112443. [Google Scholar] [CrossRef] [PubMed]
  85. Zhou, H.; Møhlenberg, M.; Thakor, J.C.; Tuli, H.S.; Wang, P.; Assaraf, Y.G.; Dhama, K.; Jiang, S. Sensitivity to vaccines, therapeutic antibodies, and viral entry inhibitors and advances to counter the SARS-CoV-2 Omicron variant. Clin. Microbiol. Rev. 2022, 35, e0001422. [Google Scholar] [CrossRef] [PubMed]
  86. Lima Neto, J.X.; Vieira, D.S.; de Andrade, J.; Fulco, U.L. Exploring the Spike-hACE 2 Residue–Residue Interaction in Human Coronaviruses SARS-CoV-2, SARS-CoV, and HCoV-NL63. J. Chem. Inf. Model. 2022, 62, 2857–2868. [Google Scholar] [PubMed]
  87. Mannar, D.; Saville, J.W.; Zhu, X.; Srivastava, S.S.; Berezuk, A.M.; Tuttle, K.S.; Marquez, A.C.; Sekirov, I.; Subramaniam, S. SARS-CoV-2 Omicron variant: Antibody evasion and cryo-EM structure of spike protein–ACE2 complex. Science 2022, 375, 760–764. [Google Scholar] [CrossRef] [PubMed]
  88. Silva, S.J.R.D.; Kohl, A.; Pena, L.; Pardee, K. Recent insights into SARS-CoV-2 omicron variant. Rev. Med. Virol. 2023, 33, e2373. [Google Scholar] [PubMed]
  89. Rees-Spear, C.; Muir, L.; Griffith, S.A.; Heaney, J.; Aldon, Y.; Snitselaar, J.L.; Thomas, P.; Graham, C.; Seow, J.; Lee, N.; et al. The effect of spike mutations on SARS-CoV-2 neutralization. Cell Rep. 2021, 34, 108890. [Google Scholar] [CrossRef] [PubMed]
  90. Thomson, E.C.; Rosen, L.E.; Shepherd, J.G.; Spreafico, R.; da Silva Filipe, A.; Wojcechowskyj, J.A.; Davis, C.; Piccoli, L.; Pascall, D.J.; Dillen, J.; et al. Circulating SARS-CoV-2 spike N439K variants maintain fitness while evading antibody-mediated immunity. Cell 2021, 184, 1171–1187.e20. [Google Scholar] [CrossRef] [PubMed]
  91. Wang, R.; Zhang, Q.; Ge, J.; Ren, W.; Zhang, R.; Lan, J.; Ju, B.; Su, B.; Yu, F.; Chen, P. Analysis of SARS-CoV-2 variant mutations reveals neutralization escape mechanisms and the ability to use ACE2 receptors from additional species. Immunity 2021, 54, 1611–1621.e15. [Google Scholar] [CrossRef]
  92. Zhang, L.; Jackson, C.B.; Mou, H.; Ojha, A.; Peng, H.; Quinlan, B.D.; Rangarajan, E.S.; Pan, A.; Vanderheiden, A.; Suthar, M.S.; et al. SARS-CoV-2 spike-protein D614G mutation increases virion spike density and infectivity. Nat. Commun. 2020, 11, 6013. [Google Scholar] [CrossRef]
  93. Mushebenge, A.G.-A.; Ugbaja, S.C.; Mbatha, N.A.; Khan, R.B.; Kumalo, H.M. A Comprehensive Analysis of Structural and Functional Changes Induced by SARS-CoV-2 Spike Protein Mutations. COVID 2023, 3, 1454–1472. [Google Scholar] [CrossRef]
  94. Zhou, B.; Thao, T.T.N.; Hoffmann, D.; Taddeo, A.; Ebert, N.; Labroussaa, F.; Pohlmann, A.; King, J.; Steiner, S.; Kelly, J.N.; et al. SARS-CoV-2 spike D614G change enhances replication and transmission. Nature 2021, 592, 122–127. [Google Scholar] [CrossRef] [PubMed]
  95. Escalera, A.; Gonzalez-Reiche, A.S.; Aslam, S.; Mena, I.; Laporte, M.; Pearl, R.L.; Fossati, A.; Rathnasinghe, R.; Alshammary, H.; van de Guchte, A.; et al. Mutations in SARS-CoV-2 variants of concern link to increased spike cleavage and virus transmission. Cell Host Microbe 2022, 30, 373–387.e377. [Google Scholar] [CrossRef] [PubMed]
  96. Sun, F.; Wang, X.; Tan, S.; Dan, Y.; Lu, Y.; Zhang, J.; Xu, J.; Tan, Z.; Xiang, X.; Zhou, Y. SARS-CoV-2 quasispecies provides an advantage mutation pool for the epidemic variants. Microbiol. Spectr. 2021, 9, e0026121. [Google Scholar] [CrossRef] [PubMed]
  97. Dubey, A.; Choudhary, S.; Kumar, P.; Tomar, S. Emerging SARS-CoV-2 variants: Genetic variability and clinical implications. Curr. Microbiol. 2022, 79, 1–18. [Google Scholar] [CrossRef]
  98. Volz, E.; Hill, V.; McCrone, J.T.; Price, A.; Jorgensen, D.; O’Toole, Á.; Southgate, J.; Johnson, R.; Jackson, B.; Nascimento, F.F.; et al. Evaluating the effects of SARS-CoV-2 spike mutation D614G on transmissibility and pathogenicity. Cell 2021, 184, 64–75.e11. [Google Scholar] [CrossRef]
  99. Mengist, H.M.; Kombe Kombe, A.J.; Mekonnen, D.; Abebaw, A.; Getachew, M.; Jin, T. Mutations of SARS-CoV-2 spike protein: Implications on immune evasion and vaccine-induced immunity. Semin. Immunol. 2021, 55, 101533. [Google Scholar] [CrossRef]
  100. De Maio, N.; Walker, C.R.; Turakhia, Y.; Lanfear, R.; Corbett-Detig, R.; Goldman, N. Mutation rates and selection on synonymous mutations in SARS-CoV-2. Genome Biol. Evol. 2021, 13, evab087. [Google Scholar] [CrossRef]
  101. Karakose, T.; Polat, H.; Papadakis, S. Examining teachers’ perspectives on school principals’ digital leadership roles and technology capabilities during the COVID-19 pandemic. Sustainability 2021, 13, 13448. [Google Scholar] [CrossRef]
  102. Korber, B.; Fischer, W.M.; Gnanakaran, S.; Yoon, H.; Theiler, J.; Abfalterer, W.; Hengartner, N.; Giorgi, E.E.; Bhattacharya, T.; Foley, B.; et al. Tracking changes in SARS-CoV-2 spike: Evidence that D614G increases infectivity of the COVID-19 virus. Cell 2020, 182, 812–827. [Google Scholar] [CrossRef] [PubMed]
  103. Khan, S.; Fakhar, Z.; Hussain, A.; Ahmad, A.; Jairajpuri, D.S.; Alajmi, M.F.; Hassan, M.I. Structure-based identification of potential SARS-CoV-2 main protease inhibitors. J. Biomol. Struct. Dyn. 2022, 40, 3595–3608. [Google Scholar] [CrossRef] [PubMed]
  104. Ismail, A.A. SARS-CoV-2 (COVID-19): A short update on molecular biochemistry, pathology, diagnosis and therapeutic strategies. Ann. Clin. Biochem. 2022, 59, 59–64. [Google Scholar] [CrossRef] [PubMed]
  105. Jolliffe, D.A.; Camargo, C.A.; Sluyter, J.D.; Aglipay, M.; Aloia, J.F.; Ganmaa, D.; Bergman, P.; Bischoff-Ferrari, H.A.; Borzutzky, A.; Damsgaard, C.T.; et al. Vitamin D supplementation to prevent acute respiratory infections: A systematic review and meta-analysis of aggregate data from randomised controlled trials. Lancet Diabetes Endocrinol. 2021, 9, 276–292. [Google Scholar] [CrossRef]
  106. Te Velthuis, A.J.; van den Worm, S.H.; Sims, A.C.; Baric, R.S.; Snijder, E.J.; van Hemert, M.J. Zn2+ inhibits coronavirus and arterivirus RNA polymerase activity in vitro and zinc ionophores block the replication of these viruses in cell culture. PLoS Pathog. 2010, 6, e1001176. [Google Scholar] [CrossRef]
  107. Razzaque, M.S. COVID-19 pandemic: Can zinc supplementation provide an additional shield against the infection? Comput. Struct. Biotechnol. J. 2021, 19, 1371–1378. [Google Scholar] [CrossRef]
  108. Robinson, P.C.; Liew, D.F.; Tanner, H.L.; Grainger, J.R.; Dwek, R.A.; Reisler, R.B.; Steinman, L.; Feldmann, M.; Ho, L.-P.; Hussell, T. COVID-19 therapeutics: Challenges and directions for the future. Proc. Natl. Acad. Sci. USA 2022, 119, e2119893119. [Google Scholar] [CrossRef]
  109. Gerardi, V.; Rohaim, M.A.; Naggar, R.F.E.; Atasoy, M.O.; Munir, M. Deep Structural Analysis of Myriads of Omicron Sub-Variants Revealed Hotspot for Vaccine Escape Immunity. Vaccines 2023, 11, 668. [Google Scholar] [CrossRef] [PubMed]
  110. Gong, W.; Parkkila, S.; Wu, X.; Aspatwar, A. SARS-CoV-2 variants and COVID-19 vaccines: Current challenges and future strategies. Int. Rev. Immunol. 2023, 42, 393–414. [Google Scholar] [CrossRef]
  111. Lawrence, H.Y. Vaccine Rhetorics; The Ohio State University Press: Columbus, OH, USA, 2020. [Google Scholar]
  112. Jaeger, B.R.; Arron, H.E.; Kalka-Moll, W.M.; Seidel, D. The potential of heparin-induced extracorporeal LDL/fibrinogen precipitation (HELP)-apheresis for patients with severe acute or chronic COVID-19. Front. Cardiovasc. Med. 2022, 9, 1007636. [Google Scholar] [CrossRef]
  113. Dioh, W.; Chabane, M.; Tourette, C.; Azbekyan, A.; Morelot-Panzini, C.; Hajjar, L.; Lins, M.; Nair, G.; Whitehouse, T.; Mariani, J.; et al. Testing the efficacy and safety of BIO101, for the prevention of respiratory deterioration, in patients with COVID-19 pneumonia (COVA study): A structured summary of a study protocol for a randomised controlled trial. Trials 2021, 22, 42. [Google Scholar] [CrossRef]
  114. Lobo, S.M.; Plantefève, G.; Nair, G.; Cavalcante, A.J.; de Moraes, N.F.; Nunes, E.; Barnum, O.; Stadnik, C.M.B.; Lima, M.P.; Lins, M.; et al. Efficacy of oral 20-hydroxyecdysone (BIO101), a MAS receptor activator, in adults with severe COVID-19 (COVA): A randomized, placebo-controlled, phase 2/3 trial. eClinicalMedicine 2024, 102383. [Google Scholar] [CrossRef]
  115. Mascellino, M.T.; Di Timoteo, F.; De Angelis, M.; Oliva, A. Overview of the main anti-SARS-CoV-2 vaccines: Mechanism of action, efficacy and safety. Infect. Drug Resist. 2021, 14, 3459–3476. [Google Scholar] [CrossRef] [PubMed]
  116. Noor, R. Developmental Status of the Potential Vaccines for the Mitigation of the COVID-19 Pandemic and a Focus on the Effectiveness of the Pfizer-BioNTech and Moderna mRNA Vaccines. Curr. Clin. Microbiol. Rep. 2021, 8, 178–185. [Google Scholar] [CrossRef]
  117. Costanzo, M.; De Giglio, M.A.; Roviello, G.N. Anti-coronavirus vaccines: Past investigations on SARS-CoV-1 and MERS-CoV, the approved vaccines from BioNTech/Pfizer, Moderna, Oxford/AstraZeneca and others under Development Against SARSCoV-2 Infection. Curr. Med. Chem. 2022, 29, 4–18. [Google Scholar] [CrossRef] [PubMed]
  118. Chiang, T.P.-Y.; Connolly, C.M.; Ruddy, J.A.; Boyarsky, B.J.; Alejo, J.L.; Werbel, W.A.; Massie, A.; Christopher-Stine, L.; Garonzik-Wang, J.; Segev, D.L.; et al. Antibody response to the Janssen/Johnson & Johnson SARS-CoV-2 vaccine in patients with rheumatic and musculoskeletal diseases. Ann. Rheum. Dis. 2021, 80, 1365–1366. [Google Scholar]
  119. Gillion, V.; Jadoul, M.; Demoulin, N.; Aydin, S.; Devresse, A. Granulomatous vasculitis after the AstraZeneca anti–SARS-CoV-2 vaccine. Kidney Int. 2021, 100, 706. [Google Scholar] [CrossRef]
  120. Lo, H.S.; Hui, K.P.Y.; Lai, H.-M.; He, X.; Khan, K.S.; Kaur, S.; Huang, J.; Li, Z.; Chan, A.K.; Cheung, H.H.-Y.; et al. Simeprevir potently suppresses SARS-CoV-2 replication and synergizes with remdesivir. ACS Cent. Sci. 2021, 7, 792–802. [Google Scholar] [CrossRef]
  121. Drożdżal, S.; Rosik, J.; Lechowicz, K.; Machaj, F.; Szostak, B.; Przybyciński, J.; Lorzadeh, S.; Kotfis, K.; Ghavami, S.; Łos, M.J. An update on drugs with therapeutic potential for SARS-CoV-2 (COVID-19) treatment. Drug Resist. Updates 2021, 59, 100794. [Google Scholar] [CrossRef]
  122. Naik, R.R.; Shakya, A.K.; Aladwan, S.M.; El-Tanani, M. Kinase inhibitors as potential therapeutic agents in the treatment of COVID-19. Front. Pharmacol. 2022, 13, 806568. [Google Scholar] [CrossRef] [PubMed]
  123. Masiá, M.; Fernández-González, M.; Padilla, S.; Ortega, P.; García, J.A.; Agulló, V.; García-Abellán, J.; Telenti, G.; Guillén, L.; Gutiérrez, F. Impact of interleukin-6 blockade with tocilizumab on SARS-CoV-2 viral kinetics and antibody responses in patients with COVID-19: A prospective cohort study. eBioMedicine 2020, 60, 102999. [Google Scholar] [CrossRef] [PubMed]
  124. Pelaia, C.; Calabrese, C.; Garofalo, E.; Bruni, A.; Vatrella, A.; Pelaia, G. Therapeutic role of tocilizumab in SARS-CoV-2-induced cytokine storm: Rationale and current evidence. Int. J. Mol. Sci. 2021, 22, 3059. [Google Scholar]
  125. Ashoor, D.; Marzouq, M.; Fathallah, M.-D. In silico evaluation of anti SARS-CoV-2 antibodies neutralization power: A blueprint with monoclonal antibody Sotrovimab. Res. Sq. 2023, PREPRINT. [Google Scholar] [CrossRef]
  126. Gupta, A.; Gonzalez-Rojas, Y.; Juarez, E.; Casal, M.C.; Moya, J.; Falci, D.R.; Sarkis, E.; Solis, J.; Zheng, H.; Scott, N.; et al. Effect of sotrovimab on hospitalization or death among high-risk patients with mild to moderate COVID-19: A randomized clinical trial. JAMA 2022, 327, 1236–1246. [Google Scholar] [CrossRef]
  127. Ashour, N.A.; Abo Elmaaty, A.; Sarhan, A.A.; Elkaeed, E.B.; Moussa, A.M.; Erfan, I.A.; Al-Karmalawy, A.A. A systematic review of the global intervention for SARS-CoV-2 combating: From drugs repurposing to molnupiravir approval. Drug Des. Dev. Ther. 2023, 16, 685–715. [Google Scholar] [CrossRef] [PubMed]
  128. Lee, C.-C.; Hsieh, C.-C.; Ko, W.-C. Molnupiravir—A novel oral anti-SARS-CoV-2 agent. Antibiotics 2021, 10, 1294. [Google Scholar] [CrossRef] [PubMed]
  129. Bhimraj, A.; Gallagher, J.C. Lack of Benefit of Fluvoxamine for COVID-19. JAMA 2023, 329, 291–292. [Google Scholar] [CrossRef] [PubMed]
  130. Zaidi, A.K.; Dehgani-Mobaraki, P. The mechanisms of action of ivermectin against SARS-CoV-2—An extensive review. J. Antibiot. 2022, 75, 60–71. [Google Scholar] [CrossRef]
  131. Tuccori, M.; Ferraro, S.; Convertino, I.; Cappello, E.; Valdiserra, G.; Blandizzi, C.; Maggi, F.; Focosi, D. Anti-SARS-CoV-2 neutralizing monoclonal antibodies: Clinical pipeline. mAbs 2020, 12, 1854149. [Google Scholar] [CrossRef] [PubMed]
  132. Baral, P.K.; Yin, J.; James, M.N. Treatment and prevention strategies for the COVID 19 pandemic: A review of immunotherapeutic approaches for neutralizing SARS-CoV-2. Int. J. Biol. Macromol. 2021, 186, 490–500. [Google Scholar] [CrossRef]
  133. Taha, Y.; Wardle, H.; Evans, A.B.; Hunter, E.R.; Marr, H.; Osborne, W.; Bashton, M.; Smith, D.; Burton-Fanning, S.; Schmid, M.L.; et al. Persistent SARS-CoV-2 infection in patients with secondary antibody deficiency: Successful clearance following combination casirivimab and imdevimab (REGN-COV2) monoclonal antibody therapy. Ann. Clin. Microbiol. Antimicrob. 2021, 20, 85. [Google Scholar] [CrossRef]
  134. Liu, X.; Munro, A.P.; Feng, S.; Janani, L.; Aley, P.K.; Babbage, G.; Baxter, D.; Bula, M.; Cathie, K.; Chatterjee, K. Persistence of immunogenicity after seven COVID-19 vaccines given as third dose boosters following two doses of ChAdOx1 nCov-19 or BNT162b2 in the UK: Three month analyses of the COV-BOOST trial. J. Infect. 2022, 84, 795–813. [Google Scholar] [CrossRef] [PubMed]
  135. Li, Y.; Liu, P.; Hao, T.; Liu, S.; Wang, X.; Xie, Y.; Xu, K.; Lei, W.; Zhang, C.; Han, P. Rational design of an influenza-COVID-19 chimeric protective vaccine with HA-stalk and S-RBD. Emerg. Microbes Infect. 2023, 12, 2231573. [Google Scholar] [CrossRef] [PubMed]
  136. Martinez, M.A. Efficacy of repurposed antiviral drugs: Lessons from COVID-19. Drug Discov. Today 2022, 27, 1954–1960. [Google Scholar] [CrossRef]
  137. Mushebenge, A.G.-A.; Ugbaja, S.C.; Mbatha, N.A.; Khan, R.B.; Kumalo, H.M. Assessing the Potential Contribution of In Silico Studies in Discovering Drug Candidates That Interact with Various SARS-CoV-2 Receptors. Int. J. Mol. Sci. 2023, 24, 15518. [Google Scholar] [CrossRef] [PubMed]
  138. Consortium, W.S.T. Repurposed antiviral drugs for COVID-19—Interim WHO solidarity trial results. N. Engl. J. Med. 2021, 384, 497–511. [Google Scholar] [CrossRef] [PubMed]
  139. Costanzo, M.; De Giglio, M.A.; Roviello, G.N. SARS-CoV-2: Recent reports on antiviral therapies based on lopinavir/ritonavir, darunavir/umifenovir, hydroxychloroquine, remdesivir, favipiravir and other drugs for the treatment of the new coronavirus. Curr. Med. Chem. 2020, 27, 4536–4541. [Google Scholar] [CrossRef] [PubMed]
  140. Zeldin, R.K.; Petruschke, R.A. Pharmacological and therapeutic properties of ritonavir-boosted protease inhibitor therapy in HIV-infected patients. J. Antimicrob. Chemother. 2004, 53, 4–9. [Google Scholar] [CrossRef]
  141. Jain, M.S.; Barhate, S.D. Favipiravir has been investigated for the treatment of life-threatening pathogens such as Ebola virus, Lassa virus, and now COVID-19: A review. Asian J. Pharm. Res. 2021, 11, 39–42. [Google Scholar] [CrossRef]
  142. Hu, B.; Huang, S.; Yin, L. The cytokine storm and COVID-19. J. Med. Virol. 2021, 93, 250–256. [Google Scholar] [CrossRef]
  143. Shiraki, K.; Daikoku, T. Favipiravir, an anti-influenza drug against life-threatening RNA virus infections. Pharmacol. Ther. 2020, 209, 107512. [Google Scholar] [CrossRef]
  144. Pastick, K.A.; Okafor, E.C.; Wang, F.; Lofgren, S.M.; Skipper, C.P.; Nicol, M.R.; Pullen, M.F.; Rajasingham, R.; McDonald, E.G.; Lee, T.C.; et al. Review: Hydroxychloroquine and Chloroquine for Treatment of SARS-CoV-2 (COVID-19). Open Forum Infect. Dis. 2020, 7, ofaa130. [Google Scholar] [CrossRef] [PubMed]
  145. Singh, A.K.; Singh, A.; Shaikh, A.; Singh, R.; Misra, A. Chloroquine and hydroxychloroquine in the treatment of COVID-19 with or without diabetes: A systematic search and a narrative review with a special reference to India and other developing countries. Diabetes Metab. Syndr. Clin. Res. Rev. 2020, 14, 241–246. [Google Scholar] [CrossRef] [PubMed]
  146. Gbinigie, K.; Frie, K. Should chloroquine and hydroxychloroquine be used to treat COVID-19? A rapid review. BJGP Open 2020, 4, bjgpopen20X101069. [Google Scholar] [CrossRef] [PubMed]
  147. Leong, T.D.; Gray, A.L.; Kredo, T.; De Waal, R.; Cohen, K.; Parrish, A.G.; Dawood, H. Managing therapeutic uncertainty in the COVID-19 pandemic: Rapid evidence syntheses and transparent decision-making. S. Afr. Health Rev. 2021, 2021, 41–49. [Google Scholar] [CrossRef]
  148. Arshad, S.; Kilgore, P.; Chaudhry, Z.S.; Jacobsen, G.; Wang, D.D.; Huitsing, K.; Brar, I.; Alangaden, G.J.; Ramesh, M.S.; McKinnon, J.E.; et al. Treatment with hydroxychloroquine, azithromycin, and combination in patients hospitalized with COVID-19. Int. J. Infect. Dis. 2020, 97, 396–403. [Google Scholar] [CrossRef]
  149. Infante, M.; Ricordi, C.; Alejandro, R.; Caprio, M.; Fabbri, A. Hydroxychloroquine in the COVID-19 pandemic era: In pursuit of a rational use for prophylaxis of SARS-CoV-2 infection. Expert Rev. Anti-Infect. Ther. 2021, 19, 5–16. [Google Scholar] [CrossRef]
  150. Kalra, R.S.; Tomar, D.; Meena, A.S.; Kandimalla, R. SARS-CoV-2, ACE2, and hydroxychloroquine: Cardiovascular complications, therapeutics, and clinical readouts in the current settings. Pathogens 2020, 9, 546. [Google Scholar] [CrossRef]
  151. Yao, X.; Ye, F.; Zhang, M.; Cui, C.; Huang, B.; Niu, P.; Liu, X.; Zhao, L.; Dong, E.; Song, C. In vitro antiviral activity and projection of optimized dosing design of hydroxychloroquine for the treatment of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). Clin. Infect. Dis. 2020, 71, 732–739. [Google Scholar] [CrossRef]
  152. U.S. Food and Drug Administration (FDA). Memorandum Explaining Basis for Revocation of Emergency Use Authorization for Chloroquine Phosphate and Hydroxychloroquine Sulfate. Available online: https://www.fda.gov/news-events/press-announcements/coronavirus-COVID-19-update-fda-revokes-emergency-use-authorization-chloroquine-and (accessed on 27 October 2023).
  153. U.S. Food and Drug Administration (FDA). FDA Cautions against Use of Hydroxychloroquine or Chloroquine for COVID-19 Outside of the Hospital Setting or a Clinical Trial Due to Risk of Heart Rhythm Problems. Available online: https://www.fda.gov/drugs/drug-safety-and-availability/fda-cautions-against-use-hydroxychloroquine-or-chloroquine-COVID-19-outside-hospital-setting-or (accessed on 27 October 2023).
  154. Figueredo, J.; Lopez, L.F.; Leguizamon, B.F.; Samudio, M.; Pederzani, M.; Apelt, F.F.; Añazco, P.; Caballero, R.; Bianco, H. Clinical evolution and mortality of critically ill patients with SARS-CoV-2 pneumonia treated with remdesivir in an adult intensive care unit of Paraguay. BMC Infect. Dis. 2024, 24, 37. [Google Scholar] [CrossRef]
  155. Murmu, N.; Sarkar, M.; Dey, S.; Manna, R.; Roy, S.; Mondal, T.; Halder, S.; Bhattacharjee, N.; Dash, S.K.; Giri, B. Efficacy and limitations of repurposed drugs and vaccines for COVID-19. J. Med. Surg. Public Health 2023, 2, 100041. [Google Scholar] [CrossRef]
  156. Waseem, W.; Zafar, R.; Jan, M.S.; Alomar, T.S.; Almasoud, N.; Rauf, A.; Khattak, H. Drug repurposing of FDA-approved anti-viral drugs via computational screening against novel 6M03 SARS-COVID-19. Ir. J. Med. Sci. 2023, 193, 73–83. [Google Scholar] [CrossRef]
  157. Rabie, A.M.; Abdel-Dayem, M.A.; Abdalla, M. Promising experimental anti-SARS-CoV-2 agent “SLL-0197800”: The prospective universal inhibitory properties against the coming versions of the coronavirus. ACS Omega 2023, 8, 35538–35554. [Google Scholar] [CrossRef]
  158. Akter, R.; Rahman, M.R.; Ahmed, Z.S.; Afrose, A. Plausibility of natural immunomodulators in the treatment of COVID-19—A comprehensive analysis and future recommendations. Heliyon 2023, 9, e17478. [Google Scholar] [CrossRef]
  159. Velásquez, P.A.; Hernandez, J.C.; Galeano, E.; Hincapié-García, J.; Rugeles, M.T.; Zapata-Builes, W. Effectiveness of Drug Repurposing and Natural Products Against SARS-CoV-2: A Comprehensive Review. Clin. Pharmacol. Adv. Appl. 2024, 16, 1–25. [Google Scholar] [CrossRef]
  160. Mushebenge, A.G.; Ugbaja, S.C.; Mtambo, S.E.; Ntombela, T.; Metu, J.I.; Babayemi, O.; Chima, J.I.; Appiah-Kubi, P.; Odugbemi, A.I.; Ntuli, M.L.; et al. Unveiling the Inhibitory Potentials of Peptidomimetic Azanitriles and Pyridyl Esters towards SARS-CoV-2 Main Protease: A Molecular Modelling Investigation. Molecules 2023, 28, 2641. [Google Scholar] [CrossRef]
  161. Das, A.; Khan, S.; Roy, S.; Das, S. Phytochemicals for mitigating the COVID-19 crisis: Evidence from pre-clinical and clinical studies. Explor. Drug Sci. 2023, 1, 336–376. [Google Scholar] [CrossRef]
  162. Kyriakidis, N.C.; López-Cortés, A.; González, E.V.; Grimaldos, A.B.; Prado, E.O. SARS-CoV-2 vaccines strategies: A comprehensive review of phase 3 candidates. NPJ Vaccines 2021, 6, 28. [Google Scholar] [CrossRef] [PubMed]
  163. Raman, R.; Patel, K.J.; Ranjan, K. COVID-19: Unmasking emerging SARS-CoV-2 variants, vaccines and therapeutic strategies. Biomolecules 2021, 11, 993. [Google Scholar] [CrossRef]
  164. Sohag, A.A.M.; Hannan, M.A.; Rahman, S.; Hossain, M.; Hasan, M.; Khan, M.K.; Khatun, A.; Dash, R.; Uddin, M.J. Revisiting potential druggable targets against SARS-CoV-2 and repurposing therapeutics under preclinical study and clinical trials: A comprehensive review. Drug Dev. Res. 2020, 81, 919–941. [Google Scholar] [CrossRef] [PubMed]
  165. Velikova, T.; Georgiev, T. SARS-CoV-2 vaccines and autoimmune diseases amidst the COVID-19 crisis. Rheumatol. Int. 2021, 41, 509–518. [Google Scholar] [CrossRef] [PubMed]
  166. Kalinke, U.; Barouch, D.H.; Rizzi, R.; Lagkadinou, E.; Türeci, Ö.; Pather, S.; Neels, P. Clinical development and approval of COVID-19 vaccines. Expert Rev. Vaccines 2022, 21, 609–619. [Google Scholar] [CrossRef]
  167. Zhang, H.; Liu, Y.; Liu, Z. Nanomedicine approaches against SARS-CoV-2 and variants. J. Control. Release 2024, 365, 101–111. [Google Scholar] [CrossRef]
  168. D’Acunto, E.; Muzi, A.; Marchese, S.; Donnici, L.; Chiarini, V.; Bucci, F.; Pavoni, E.; Ferrara, F.F.; Cappelletti, M.; Arriga, R.; et al. Isolation and characterization of neutralizing monoclonal antibodies from a large panel of murine antibodies against RBD of the SARS-CoV-2 Spike protein. Antibodies 2024, 13, 5. [Google Scholar] [CrossRef]
  169. Hirotsu, Y.; Takatori, M.; Mochizuki, H.; Omata, M. Effectiveness of the severe acute respiratory syndrome coronavirus 2 Omicron BA. 5 bivalent vaccine on symptoms in healthcare workers with BA.5 infection. Vaccine X 2024, 17, 100433. [Google Scholar] [CrossRef]
  170. Aydillo, T.; Rombauts, A.; Stadlbauer, D.; Aslam, S.; Abelenda-Alonso, G.; Escalera, A.; Amanat, F.; Jiang, K.; Krammer, F.; Carratala, J.; et al. Immunological imprinting of the antibody response in COVID-19 patients. Nat. Commun. 2021, 12, 3781. [Google Scholar] [CrossRef]
  171. Machado, B.A.S.; Hodel, K.V.S.; Fonseca, L.M.D.S.; Pires, V.C.; Mascarenhas, L.A.B.; da Silva Andrade, L.P.C.; Moret, M.A.; Badaró, R. The importance of vaccination in the context of the COVID-19 pandemic: A brief update regarding the use of vaccines. Vaccines 2022, 10, 591. [Google Scholar] [CrossRef]
  172. Malik, J.A.; Mulla, A.H.; Farooqi, T.; Pottoo, F.H.; Anwar, S.; Rengasamy, K.R. Targets and strategies for vaccine development against SARS-CoV-2. Biomed. Pharmacother. 2021, 137, 111254. [Google Scholar] [CrossRef] [PubMed]
  173. Piperno, A.; Sciortino, M.T.; Giusto, E.; Montesi, M.; Panseri, S.; Scala, A. Recent advances and challenges in gene delivery mediated by polyester-based nanoparticles. Int. J. Nanomed. 2021, 16, 5981–6002. [Google Scholar] [CrossRef] [PubMed]
  174. Kumar, V.; Kumar, S.; Sharma, P.C. Recent advances in the vaccine development for the prophylaxis of SARS COVID-19. Int. Immunopharmacol. 2022, 111, 109175. [Google Scholar] [CrossRef] [PubMed]
  175. Khoshnood, S.; Arshadi, M.; Akrami, S.; Koupaei, M.; Ghahramanpour, H.; Shariati, A.; Sadeghifard, N.; Heidary, M. An overview on inactivated and live-attenuated SARS-CoV-2 vaccines. J. Clin. Lab. Anal. 2022, 36, e24418. [Google Scholar] [CrossRef] [PubMed]
  176. Tebas, P.; Yang, S.; Boyer, J.D.; Reuschel, E.L.; Patel, A.; Christensen-Quick, A.; Andrade, V.M.; Morrow, M.P.; Kraynyak, K.; Agnes, J.; et al. Safety and immunogenicity of INO-4800 DNA vaccine against SARS-CoV-2: A preliminary report of an open-label, Phase 1 clinical trial. eClinicalMedicine 2021, 31, 100689. [Google Scholar] [CrossRef]
  177. Dokoupilová, E.; Vetchý, D.; Pavloková, S.; Hanuštiaková, M. Effect of treatment with original or biosimilar adalimumab on SARS-CoV2 vaccination antibody titers. Int. J. Pharm. X 2024, 7, 100229. [Google Scholar] [CrossRef] [PubMed]
  178. Jung, W.; Yuan, D.; Kellman, B.; Gonzalez, I.G.d.S.; Clemens, R.; Milan, E.P.; Sprinz, E.; Cerbino Neto, J.; Smolenov, I.; Alter, G.; et al. Boosting with adjuvanted SCB-2019 elicits superior Fcγ-receptor engagement driven by IgG3 to SARS-CoV-2 spike. NPJ Vaccines 2024, 9, 7. [Google Scholar] [CrossRef] [PubMed]
  179. Jiang, W.; Maldeney, A.R.; Yuan, X.; Richer, M.J.; Renshaw, S.E.; Luo, W. Ipsilateral immunization after a prior SARS-CoV-2 mRNA vaccination elicits superior B cell responses compared to contralateral immunization. Cell Rep. 2024, 43, 113665. [Google Scholar] [CrossRef]
  180. Song, N.-J.; Chakravarthy, K.B.; Jeon, H.; Bolyard, C.; Reynolds, K.; Weller, K.P.; Reisinger, S.; Wang, Y.; Li, A.; Jiang, S. mRNA vaccines against SARS-CoV-2 induce divergent antigen-specific T-cell responses in patients with lung cancer. J. Immunother. Cancer 2024, 12, e007922. [Google Scholar] [CrossRef]
  181. Shrotri, M.; Swinnen, T.; Kampmann, B.; Parker, E.P. An interactive website tracking COVID-19 vaccine development. Lancet Glob. Health 2021, 9, e590–e592. [Google Scholar] [CrossRef]
  182. Wang, Y.; Huo, P.; Dai, R.; Lv, X.; Yuan, S.; Zhang, Y.; Guo, Y.; Li, R.; Yu, Q.; Zhu, K. Convalescent plasma may be a possible treatment for COVID-19: A systematic review. Int. Immunopharmacol. 2021, 91, 107262. [Google Scholar] [CrossRef]
  183. Baldeón, M.E.; Maldonado, A.; Ochoa-Andrade, M.; Largo, C.; Pesantez, M.; Herdoiza, M.; Granja, G.; Bonifaz, M.; Espejo, H.; Mora, F.; et al. Effect of convalescent plasma as complementary treatment in patients with moderate COVID-19 infection. Transfus. Med. 2022, 32, 153–161. [Google Scholar] [CrossRef]
  184. Ochani, R.; Asad, A.; Yasmin, F.; Shaikh, S.; Khalid, H.; Batra, S.; Sohail, M.R.; Mahmood, S.F.; Ochani, R.; Hussham Arshad, M.; et al. COVID-19 pandemic: From origins to outcomes. A comprehensive review of viral pathogenesis, clinical manifestations, diagnostic evaluation, and management. Infez. Med. 2021, 29, 20–36. [Google Scholar]
  185. Chadha, R.; Raghav, A.; Banerjee, B.; Sengar, A.; Sengar, M.; Raghav, P.K. Targets of SARS-CoV-2: Therapeutic implications for COVID-19. In Stem Cells; Elsevier: Amsterdam, The Netherlands, 2024; pp. 3–14. [Google Scholar]
  186. Wang, Y.; Zhang, Z.; Yang, M.; Xiong, X.; Yan, Q.; Cao, L.; Wei, P.; Zhang, Y.; Zhang, L.; Lv, K.; et al. Identification of a broad sarbecovirus neutralizing antibody targeting a conserved epitope on the receptor-binding domain. Cell Rep. 2024, 43, 113653. [Google Scholar] [CrossRef] [PubMed]
  187. Fenwick, C.; Turelli, P.; Perez, L.; Pellaton, C.; Esteves-Leuenberger, L.; Farina, A.; Campos, J.; Lana, E.; Fiscalini, F.; Raclot, C.; et al. A highly potent antibody effective against SARS-CoV-2 variants of concern. Cell Rep. 2021, 37, 109814. [Google Scholar] [CrossRef]
  188. Jones, B.E.; Brown-Augsburger, P.L.; Corbett, K.S.; Westendorf, K.; Davies, J.; Cujec, T.P.; Wiethoff, C.M.; Blackbourne, J.L.; Heinz, B.A.; Foster, D.; et al. The neutralizing antibody, LY-CoV555, protects against SARS-CoV-2 infection in nonhuman primates. Sci. Transl. Med. 2021, 13, eabf1906. [Google Scholar] [CrossRef]
  189. Sahoo, P.; Dey, J.; Mahapatra, S.R.; Ghosh, A.; Jaiswal, A.; Padhi, S.; Prabhuswamimath, S.C.; Misra, N.; Suar, M. Nanotechnology and COVID-19 convergence: Toward new planetary health interventions against the pandemic. OMICS J. Integr. Biol. 2022, 26, 473–488. [Google Scholar] [CrossRef] [PubMed]
  190. Mehta, A.; Michler, T.; Merkel, O.M. siRNA therapeutics against respiratory viral infections—What have we learned for potential COVID-19 therapies? Adv. Healthc. Mater. 2021, 10, 2001650. [Google Scholar] [CrossRef]
  191. Yang, Y.; Peng, F.; Wang, R.; Guan, K.; Jiang, T.; Xu, G.; Sun, J.; Chang, C. The deadly coronaviruses: The 2003 SARS pandemic and the 2020 novel coronavirus epidemic in China. J. Autoimmun. 2020, 109, 102434. [Google Scholar] [CrossRef] [PubMed]
  192. Zhang, H.; Penninger, J.M.; Li, Y.; Zhong, N.; Slutsky, A.S. Angiotensin-converting enzyme 2 (ACE2) as a SARS-CoV-2 receptor: Molecular mechanisms and potential therapeutic target. Intensive Care Med. 2020, 46, 586–590. [Google Scholar] [CrossRef]
  193. Bourgonje, A.R.; Abdulle, A.E.; Timens, W.; Hillebrands, J.L.; Navis, G.J.; Gordijn, S.J.; Bolling, M.C.; Dijkstra, G.; Voors, A.A.; Osterhaus, A.D.; et al. Angiotensin-converting enzyme 2 (ACE2), SARS-CoV-2 and the pathophysiology of coronavirus disease 2019 (COVID-19). J. Pathol. 2020, 251, 228–248. [Google Scholar] [CrossRef]
  194. Seyedpour, S.; Khodaei, B.; Loghman, A.H.; Seyedpour, N.; Kisomi, M.F.; Balibegloo, M.; Nezamabadi, S.S.; Gholami, B.; Saghazadeh, A.; Rezaei, N. Targeted therapy strategies against SARS-CoV-2 cell entry mechanisms: A systematic review of in vitro and in vivo studies. J. Cell. Physiol. 2021, 236, 2364–2392. [Google Scholar] [CrossRef]
  195. Ita, K. Coronavirus disease (COVID-19): Current status and prospects for drug and vaccine development. Arch. Med. Res. 2021, 52, 15–24. [Google Scholar] [CrossRef] [PubMed]
  196. Miners, S.; Kehoe, P.G.; Love, S. Cognitive impact of COVID-19: Looking beyond the short term. Alzheimer’s Res. Ther. 2020, 12, 1–16. [Google Scholar] [CrossRef]
  197. Abd El-Aziz, T.M.; Al-Sabi, A.; Stockand, J.D. Human recombinant soluble ACE2 (hrsACE2) shows promise for treating severe COVID19. Signal Transduct. Target. Ther. 2020, 5, 258. [Google Scholar] [CrossRef]
  198. Zanganeh, S.; Goodarzi, N.; Doroudian, M.; Movahed, E. Potential COVID-19 therapeutic approaches targeting angiotensin-converting enzyme 2; an updated review. Rev. Med. Virol. 2022, 32, e2321. [Google Scholar] [CrossRef]
  199. Yang, H.; Zhang, S.; Liu, R.; Krall, A.; Wang, Y.; Ventura, M.; Deflitch, C. Epidemic informatics and control: A holistic approach from system informatics to epidemic response and risk management in public health. In AI and Analytics for Public Health-Proceedings of the 2020 INFORMS International Conference on Service Science; Springer: Berlin/Heidelberg, Germany, 2021; pp. 1–46. [Google Scholar]
  200. Di Domenico, L.; Pullano, G.; Sabbatini, C.E.; Boëlle, P.-Y.; Colizza, V. Impact of lockdown on COVID-19 epidemic in Île-de-France and possible exit strategies. BMC Med. 2020, 18, 240. [Google Scholar] [CrossRef]
  201. Simandan, D.; Rinner, C.; Capurri, V. The academic left, human geography, and the rise of authoritarianism during the COVID-19 pandemic. Geogr. Ann. Ser. B Hum. Geogr. 2023, 1–21. [Google Scholar] [CrossRef]
  202. Behera, R.K.; Bala, P.K.; Rana, N.P.; Kayal, G. Self-promotion and online shaming during COVID-19: A toxic combination. Int. J. Inf. Manag. Data Insights 2022, 2, 100117. [Google Scholar] [CrossRef]
  203. Mendez-Brito, A.; El Bcheraoui, C.; Pozo-Martin, F. Systematic review of empirical studies comparing the effectiveness of non-pharmaceutical interventions against COVID-19. J. Infect. 2021, 83, 281–293. [Google Scholar] [CrossRef] [PubMed]
  204. Hidayat, A.M.; Choocharukul, K. Passengers’ Intentions to Use Public Transport during the COVID-19 Pandemic: A Case Study of Bangkok and Jakarta. Sustainability 2023, 15, 5273. [Google Scholar] [CrossRef]
  205. Hanson, K.E.; Caliendo, A.M.; Arias, C.A.; Englund, J.A.; Lee, M.J.; Loeb, M.; Patel, R.; El Alayli, A.; Kalot, M.A.; Falck-Ytter, Y. Infectious Diseases Society of America guidelines on the diagnosis of coronavirus disease 2019. Clin. Infect. Dis. 2020, ciaa760. [Google Scholar] [CrossRef] [PubMed]
  206. Sala, G.; Chakraborti, R.; Ota, A.; Miyakawa, T. Association of BCG vaccination policy and tuberculosis burden with incidence and mortality of COVID-19. medRxiv 2020, 3. [Google Scholar] [CrossRef]
  207. Aminu, J. The implications of misconceptions about coronavirus disease (COVID-19) pandemic in relation to its daily increases from Nigerian perspective. J. Infect. Dis. Epidemiol. 2020, 6, 156. [Google Scholar]
  208. Yang, J.; Li, X.; He, T.; Ju, F.; Qiu, Y.; Tian, Z. Impact of physical activity on COVID-19. Int. J. Environ. Res. Public Health 2022, 19, 14108. [Google Scholar] [CrossRef] [PubMed]
  209. Ayouni, I.; Maatoug, J.; Dhouib, W.; Zammit, N.; Fredj, S.B.; Ghammam, R.; Ghannem, H. Effective public health measures to mitigate the spread of COVID-19: A systematic review. BMC Public Health 2021, 21, 1015. [Google Scholar] [CrossRef]
  210. Cheng, V.C.-C.; Wong, S.-C.; Chuang, V.W.-M.; So, S.Y.-C.; Chen, J.H.-K.; Sridhar, S.; To, K.K.-W.; Chan, J.F.-W.; Hung, I.F.-N.; Ho, P.-L.; et al. The role of community-wide wearing of face mask for control of coronavirus disease 2019 (COVID-19) epidemic due to SARS-CoV-2. J. Infect. 2020, 81, 107–114. [Google Scholar] [CrossRef] [PubMed]
  211. To, K.K.-W.; Sridhar, S.; Chiu, K.H.-Y.; Hung, D.L.-L.; Li, X.; Hung, I.F.-N.; Tam, A.R.; Chung, T.W.-H.; Chan, J.F.-W.; Zhang, A.J.-X.; et al. Lessons learned 1 year after SARS-CoV-2 emergence leading to COVID-19 pandemic. Emerg. Microbes Infect. 2021, 10, 507–535. [Google Scholar] [CrossRef] [PubMed]
  212. Dumache, R.; Enache, A.; Macasoi, I.; Dehelean, C.A.; Dumitrascu, V.; Mihailescu, A.; Popescu, R.; Vlad, D.; Vlad, C.S.; Muresan, C. SARS-CoV-2: An overview of the genetic profile and vaccine effectiveness of the five variants of concern. Pathogens 2022, 11, 516. [Google Scholar] [CrossRef] [PubMed]
  213. World Health Organization. Genomic Sequencing of SARS-CoV-2: A Guide to Implementation for Maximum Impact on Public Health; World Health Organization: Geneva, Switzerland, 2021. [Google Scholar]
  214. Adly, A.S.; Adly, A.S.; Adly, M.S. Approaches based on artificial intelligence and the internet of intelligent things to prevent the spread of COVID-19: Scoping review. J. Med. Internet Res. 2020, 22, e19104. [Google Scholar] [CrossRef] [PubMed]
  215. McCall, B. COVID-19 and artificial intelligence: Protecting health-care workers and curbing the spread. Lancet Digit. Health 2020, 2, e166–e167. [Google Scholar] [CrossRef]
  216. Yu, K.-H.; Beam, A.L.; Kohane, I.S. Artificial intelligence in healthcare. Nat. Biomed. Eng. 2018, 2, 719–731. [Google Scholar] [CrossRef]
  217. Salvatore, M.; Purkayastha, S.; Ganapathi, L.; Bhattacharyya, R.; Kundu, R.; Zimmermann, L.; Ray, D.; Hazra, A.; Kleinsasser, M.; Solomon, S. Lessons from SARS-CoV-2 in India: A data-driven framework for pandemic resilience. Sci. Adv. 2022, 8, eabp8621. [Google Scholar] [CrossRef]
  218. Malik, J.A.; Ahmed, S.; Mir, A.; Shinde, M.; Bender, O.; Alshammari, F.; Ansari, M.; Anwar, S. The SARS-CoV-2 mutations versus vaccine effectiveness: New opportunities to new challenges. J. Infect. Public Health 2022, 15, 228–240. [Google Scholar] [CrossRef]
  219. Gulseven, O.; Al Harmoodi, F.; Al Falasi, M.; ALshomali, I. How the COVID-19 Pandemic Will Affect the UN Sustainable Development Goals? 2020. Available online: https://ssrn.com/abstract=3592933 (accessed on 27 October 2023).
Figure 1. Redrawn viral life cycle of SARS-CoV-2 via BioRender as adapted from the source: the S protein’s interaction with hACE2 initiates the SARS-CoV-2 viral life cycle and causes infection. Proteolytic cleavage occurring after receptor engagement enables the viral entrance and release of genomic RNA into the cytoplasm. Polyprotein translation is initiated by RNA and is broken into non-structural proteins by enzymes such as PLpro and Mpro. The replicase—transcriptase complex, which is responsible for genome synthesis—is composed of non-structural proteins. Once produced, structural proteins travel to the endoplasmic reticulum where they assemble with RNA encapsulated in N proteins. Once the newly created virions have broken through the cell membrane, they are exocytosis [23].
Figure 1. Redrawn viral life cycle of SARS-CoV-2 via BioRender as adapted from the source: the S protein’s interaction with hACE2 initiates the SARS-CoV-2 viral life cycle and causes infection. Proteolytic cleavage occurring after receptor engagement enables the viral entrance and release of genomic RNA into the cytoplasm. Polyprotein translation is initiated by RNA and is broken into non-structural proteins by enzymes such as PLpro and Mpro. The replicase—transcriptase complex, which is responsible for genome synthesis—is composed of non-structural proteins. Once produced, structural proteins travel to the endoplasmic reticulum where they assemble with RNA encapsulated in N proteins. Once the newly created virions have broken through the cell membrane, they are exocytosis [23].
Biomedinformatics 04 00022 g001
Figure 2. An overview of the clinical pathology, pathogenesis, and immunopathology of COVID-19 as adapted from the source: the virus impacts the immune system by modifying the production of biomolecules in immune cells during disease progression. The clinical signs of COVID-19 can range from digestive to pulmonary issues, and the disease has a 5.99% case fatality rate. Asymptomatic carriers can shed the virus for up to 21 days. Fever, coughing, exhaustion, and respiratory distress are examples of clinical symptoms. The cytokine storm, humoral and cellular processes, and lymphopenia are all components of the immune response. The virus primarily causes cytopathic effects on epithelial cells in the gastrointestinal and respiratory systems. COVID-19 patients have different lymphocyte subsets, cytokine upregulation, and immunopathologically altered leukocyte numbers [12].
Figure 2. An overview of the clinical pathology, pathogenesis, and immunopathology of COVID-19 as adapted from the source: the virus impacts the immune system by modifying the production of biomolecules in immune cells during disease progression. The clinical signs of COVID-19 can range from digestive to pulmonary issues, and the disease has a 5.99% case fatality rate. Asymptomatic carriers can shed the virus for up to 21 days. Fever, coughing, exhaustion, and respiratory distress are examples of clinical symptoms. The cytokine storm, humoral and cellular processes, and lymphopenia are all components of the immune response. The virus primarily causes cytopathic effects on epithelial cells in the gastrointestinal and respiratory systems. COVID-19 patients have different lymphocyte subsets, cytokine upregulation, and immunopathologically altered leukocyte numbers [12].
Biomedinformatics 04 00022 g002
Figure 3. Information of fast-spreading SARS-CoV-2 variants and major SARS-CoV-2 structures as adapted from the source: the illustration shows the precise amino acid mutations that are present in SARS-CoV-2 variations that are spreading quickly. (a) Highlights the important mutations in variants like B.1.1.7, B.1.351, and B.1.1.28.1 by highlighting them in red. (b) Spike protein, membrane protein, envelope protein, nucleocapsid protein, and RNA—the main structural elements of SARS-CoV-2—are shown. Angiotensin-converting enzyme 2 (ACE2) is emphasised as the main cellular entry point for SARS-CoV-2, highlighting its function as a transmembrane protein [101].
Figure 3. Information of fast-spreading SARS-CoV-2 variants and major SARS-CoV-2 structures as adapted from the source: the illustration shows the precise amino acid mutations that are present in SARS-CoV-2 variations that are spreading quickly. (a) Highlights the important mutations in variants like B.1.1.7, B.1.351, and B.1.1.28.1 by highlighting them in red. (b) Spike protein, membrane protein, envelope protein, nucleocapsid protein, and RNA—the main structural elements of SARS-CoV-2—are shown. Angiotensin-converting enzyme 2 (ACE2) is emphasised as the main cellular entry point for SARS-CoV-2, highlighting its function as a transmembrane protein [101].
Biomedinformatics 04 00022 g003
Figure 5. Schematic of the renin–angiotensin system, and the proposed treatment for COVID-19 targeting the SARS-CoV-2 viral-entry mechanism as adapted from source: The renin–angiotensin system and a suggested treatment strategy for COVID-19 that targets the SARS-CoV-2 viral-entry mechanism are depicted in this diagram. The receptor-binding domain (RBD) of the spike protein interacts with ACE2 on the left, enabling host cell entrance and infection. The physiological function of ACE2 in the renin–angiotensin system and its organ-protective effects are discussed in the middle. ACE2 hydrolyses Angiotensin II to produce angiotensins 1–7, imparting a protective effect. For good health, ACE/Ang II/AT1R and ACE2/Ang 1-6/MasR must be in equilibrium. According to recent research, increasing the amount of human recombinant soluble ACE2 (hrsACE2) at tissue sites may compete with endogenous ACE2, preventing SARS-CoV-2 from infecting host cells and lowering angiotensin II levels while maintaining the production of anti-SARS-CoV-2 antibodies [196].
Figure 5. Schematic of the renin–angiotensin system, and the proposed treatment for COVID-19 targeting the SARS-CoV-2 viral-entry mechanism as adapted from source: The renin–angiotensin system and a suggested treatment strategy for COVID-19 that targets the SARS-CoV-2 viral-entry mechanism are depicted in this diagram. The receptor-binding domain (RBD) of the spike protein interacts with ACE2 on the left, enabling host cell entrance and infection. The physiological function of ACE2 in the renin–angiotensin system and its organ-protective effects are discussed in the middle. ACE2 hydrolyses Angiotensin II to produce angiotensins 1–7, imparting a protective effect. For good health, ACE/Ang II/AT1R and ACE2/Ang 1-6/MasR must be in equilibrium. According to recent research, increasing the amount of human recombinant soluble ACE2 (hrsACE2) at tissue sites may compete with endogenous ACE2, preventing SARS-CoV-2 from infecting host cells and lowering angiotensin II levels while maintaining the production of anti-SARS-CoV-2 antibodies [196].
Biomedinformatics 04 00022 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mushebenge, A.G.-A.; Ugbaja, S.C.; Mbatha, N.A.; Khan, R.B.; Kumalo, H.M. Unravelling Insights into the Evolution and Management of SARS-CoV-2. BioMedInformatics 2024, 4, 385-409. https://doi.org/10.3390/biomedinformatics4010022

AMA Style

Mushebenge AG-A, Ugbaja SC, Mbatha NA, Khan RB, Kumalo HM. Unravelling Insights into the Evolution and Management of SARS-CoV-2. BioMedInformatics. 2024; 4(1):385-409. https://doi.org/10.3390/biomedinformatics4010022

Chicago/Turabian Style

Mushebenge, Aganze Gloire-Aimé, Samuel Chima Ugbaja, Nonkululeko Avril Mbatha, Rene B. Khan, and Hezekiel M. Kumalo. 2024. "Unravelling Insights into the Evolution and Management of SARS-CoV-2" BioMedInformatics 4, no. 1: 385-409. https://doi.org/10.3390/biomedinformatics4010022

Article Metrics

Back to TopTop