Next Article in Journal
Dysbiotic Gut Bacteria in Obesity: An Overview of the Metabolic Mechanisms and Therapeutic Perspectives of Next-Generation Probiotics
Next Article in Special Issue
Organ Involvement in COVID-19: A Molecular Investigation of Autopsied Patients
Previous Article in Journal
The Leaf Microbiome of Tobacco Plants across Eight Chinese Provinces
Previous Article in Special Issue
SARS-CoV-2 Serological and Biomolecular Analyses among Companion Animals in Campania Region (2020–2021)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Literature Review of Omicron: A Grim Reality Amidst COVID-19

1
Department of Restorative Dental Sciences, College of Dentistry, King Khalid University, Abha 61321, Saudi Arabia
2
Department of Periodontology and Oral Implantology, Dr. H.S.J. Institute of Dental Sciences, Panjab University, Chandigarh 160014, India
3
Department of Conservative Dentistry and Endodontics, JCD Dental College, Sirsa 125055, India
4
Department of Periodontics and Community Dental Sciences, College of Dentistry, King Khalid University, Abha 61321, Saudi Arabia
5
Department of Diagnostic Sciences and Oral Biology, College of Dentistry, King Khalid University, Abha 61321, Saudi Arabia
6
Biomedical Statistician, Office of Research, Kirk Kerkorian School of Medicine at University of Nevada, 2040 W. Charleston Blvd., Las Vegas, NV 89102, USA
7
Department of Clinical Laboratory Sciences, College of Applied Medical Sciences, King Khalid University, Abha 61321, Saudi Arabia
8
Department of Prosthodontics, College of Dentistry, King Khalid University, Abha 61321, Saudi Arabia
9
Department of Prosthodontics, Dr. H.S.J. Institute of Dental Sciences & Hospital, Punjab University, Chandigarh 160014, India
*
Authors to whom correspondence should be addressed.
Microorganisms 2022, 10(2), 451; https://doi.org/10.3390/microorganisms10020451
Submission received: 12 January 2022 / Revised: 26 January 2022 / Accepted: 8 February 2022 / Published: 16 February 2022
(This article belongs to the Special Issue The Ecology and Evolution of SARS-CoV-2)

Abstract

:
Coronavirus disease 2019 (COVID-19) first emerged in Wuhan city in December 2019, and became a grave global concern due to its highly infectious nature. The Severe Acute Respiratory Coronavirus-2, with its predecessors (i.e., MERS-CoV and SARS-CoV) belong to the family of Coronaviridae. Reportedly, COVID-19 has infected 344,710,576 people around the globe and killed nearly 5,598,511 persons in the short span of two years. On November 24, 2021, B.1.1.529 strain, later named Omicron, was classified as a Variant of Concern (VOC). SARS-CoV-2 has continuously undergone a series of unprecedented mutations and evolved to exhibit varying characteristics. These mutations have largely occurred in the spike (S) protein (site for antibody binding), which attribute high infectivity and transmissibility characteristics to the Omicron strain. Although many studies have attempted to understand this new challenge in the COVID-19 strains race, there is still a lot to be demystified. Therefore, the purpose of this review was to summarize the structural or virologic characteristics, burden, and epidemiology of the Omicron variant and its potential to evade the immune response.

1. Background

Coronavirus disease 2019 (COVID-19) first emerged in Wuhan city in December 2019, and became a grave global concern due to its highly infectious nature [1,2,3]. The Severe Acute Respiratory Coronavirus-2, with its predecessors (i.e., MERS-CoV and SARS-CoV) belong to the family of Coronaviridae. Reportedly, COVID-19 has infected 344,710,576 people around the globe and killed nearly 5,598,511 persons in the short span of two years [4]. As with other viruses, coronavirus constantly changes through genetic mutations, which have posed new challenges in the road to recovery. According to the United States (U.S.) government-led SARS-CoV-2 Interagency Group (SIG), SARS-CoV-2 variants can be categorized into four classes: Variant Being Monitored (VBM), Variant of Interest (VOI), Variant of Concern (VOC), and Variant of High Consequence (VOHC) [5]. Among the VOC class, variants such as Alpha (B.1.1.7), Beta (B.1.351), Gamma (P.1), and Delta (B.1.617.2 and AY lineages) evolved, and recently a new variant, B.1.1.529, was detected in several countries [5]. It quickly became a subject of discussion and exploration among the scientific community. The variant B.1.1.529 was first detected in Botswana, followed by South Africa between 11 November 2021 and 14 November 2021 [6]. It is a VOC due to its high transmissibility and less susceptibility to neutralization by antibodies produced either by previous viral exposure or vaccine administration [5,6]. According to the estimates, the cases of B.1.1.529 in South Africa grew by over fivefold in one week from 16 November to 25 November 2021 [7,8]. These upward trends remained for four weeks, followed by a rapid decline by 48% from 27 December 2021 to 2 January 2022 [7,8]. Other countries to report cases of infection by B.1.1.529 include, but are not limited to, France, The Netherlands, Germany, Portugal, Italy, the United Kingdom (U.K.), Canada, Hong Kong, Australia, and the United States. On 25 November, the B.1.1.529 variant was termed a Variant Under Monitoring by the United Kingdom Health Security Agency, and was considered as the maximally mutated variant amongst the other variants [7,8]. A day after, on 26 November 2021, this variant was designated as an official Variant of Concern by the World Health Organization (WHO) and named Omicron [9]. According to the epidemiological update provided by the European Surveillance System (TESSy), travel-related cases were identified in 13% of the confirmed infections, with nearly 70% of the remaining cases having arisen locally [10]. Omicron created chaos around the world and different studies are being conducted to study its presenting symptoms, transmission, risk of reinfection, severity, and its tendency to evade immune responses [11,12,13,14]. There are concerns related to its rampant transmission, which could hinder containment efforts, such as vaccine effectiveness. The surging trend in Omicron cases is worrisome, because this could cause an overwhelming demand on health care systems which have not yet completely recovered from the health and financial damages caused by the initial virus outbreak [14,15]. Although studies have attempted to understand this new challenge in the COVID-19 strains race, there is still a lot to be demystified. Therefore, the purpose of this review was to summarize the structural or virologic characteristics, burden, epidemiology of the Omicron variant and its potential to evade the immune response.

2. Structural Characteristics of the Omicron

The Omicron variant, which is a derivative of the Pango lineage B.1.1.529, exhibiting a variation in the 21 amino acid pertaining to the spike protein with the majority residing in the receptor binding domain (RBD) (residues 319–541) compared with the original strain [16,17,18,19,20,21]. SARS-CoV-2 has continuously undergone a series of unprecedented mutations and evolved to exhibit varying characteristics [16,17,18,19,20,21]. These mutations have largely occurred in the spike (S) protein (which is the site for antibody binding), which attribute high infectivity and transmissibility properties to Omicron variant [16,17,18,19,20]. According to the genomic reports, the S protein of Omicron has a total of 30 amino acid substitutions, 3 deletions, and 1 small insertion [21]. About 50% (n = 15) of amino acid substitutions occur exclusively in the receptor binding domain (RBD) [16,17,18,19,20,21]. Among the 15 RBD substitutes, N501Y and Q498R have a stronger affinity towards the angiotensin-converting enzyme (ACE-2 receptors), which explain the high transmissibility of the Omicron variant [16,17,18,19,20,21,22]. The ACE-2 receptors play a significant role in COVID-19 pathogenesis, which may involve serious organ failure [23].

3. Virologic Characteristics Explained (Omicron vs. Delta Variant)

Recent studies have reported that Omicron has a different mechanism for entering the host, and is capable of gaining cell entry independent of the transmembrane serine protease 2 (TMPRSS2) [24]. The entry pathway and viral replication of Omicron is through the endocytic pathway rather than TMPRS22 pathway, unlike Delta variant, which may have contributed to the differences in the disease presentation following exposure to Delta and Omicron variants [24]. TMPRS22 is highly expressed in alveolar cells of the lungs; thus, due to the lesser or no dependence of Omicron on the TMPRS22 pathway for its replication, lung involvement following exposure might be absent or limited [24,25]. In addition, the fusion capabilities to aid syncytia (structure that results following the fusion of multiple cells) formations are reduced in Omicron compared with the Delta variant. The reduced capacity of the syncytia formation translates to the lesser severity of the clinical manifestations and tissue tropism following the Omicron infection [24,25,26].

4. Epidemiology of Omicron and a Tropism Shift

Omicron spreads faster than the original virus; however, data related to its reproductive number remain limited. According to the European Center for Disease Prevention and Control, this variant could be more transmissible than the Delta variant [10]. Early reports from South Africa in November suggested the effective reproductive number (Re) of the variant to be in the range of 1.5–3; however, a recent estimate of Re as low as 0.75 was also reported [27]. These latest estimates should be interpreted with caution, because several factors, including changes in the testing measures or efforts and lag in case reporting, might have contributed to the change in the reproductive number [27]. The collective evidence suggests that Omicron has greater infectivity and potential to cause reinfection as compared with its predecessors [9,27]; however, data are insufficient to quantify its overall impact.
The symptoms of Omicron include a dry cough, scratchy throat, body aches, fatigue, runny nose, fever, and night symptoms [18,19]. According to recent evidence from South Africa, no peculiar symptoms associated with the variant have been reported, and some patients remained asymptomatic or only presented mild symptoms [10,20]. The patients infected with Omicron had fewer or no symptoms related to neurotropism (i.e., loss of taste and smell), which were peculiar with the earlier strains of the virus [18,28,29,30,31,32]. This tropism shift is attributed to the reduced dependence of Omicron on TMPRS22-expressing cells, such as the lower respiratory tract, brain, heart, kidney, and other extrapulmonary organs [33].
Among the initial Omicron VOC cases reported by the end of 2021 by EU/EEA countries to TESSy, 7 of every 10 cases were symptomatic [30]. Furthermore, the average age range for these reported cases was between 20 and 49 years [30]. The incidence of Omicron cases was higher among females compared with males [34]. The patients with preexisting conditions and those with the acute respiratory attacks were among the Omicron-associated hospitalizations [34]. Furthermore, it was reported that the risk of hospital and ICU admissions was lower compared with the preceding waves in South Africa [34]. In addition, a lower proportion of patients required oxygen therapy and mechanical ventilation [34]. The median length of stay was reduced by half in this current wave as compared with what was observed in the previous waves (3–4 days vs. 7–8 days) [34]. Preliminary data from South Africa suggested that Omicron causes less severe symptoms than the original SARS-CoV-2 virus, although rapidly increasing case loads and overall transmissibility remain the serious concerns. Omicron seems to weaken the association between cases and mortality, which is supported by the U.K.-based evidence, according to which the Omicron case rate in the United Kingdom was 35% per day; however, the death rate continuously declined, unlike the patterns which were observed in the first wave of the COVID-19 pandemic [30,35]. Data also suggested that every 3 in 10 cases which emerged in the existing cases were fully vaccinated, which points to the ability of Omicron to evade immune response [34].

5. Potential for Immune Escape

The population-level evidence from South Africa estimated that the hazard ratio for Omicron reinfection and primary infection was 2.39, suggestive of a possible evasion from natural immunity gained from previous infections [12,36]. This is consistent with the general premise that antibodies generated following the natural response have a lower titer and greater dissipation, thereby reducing immunity over time [12,36]. It is important to explore the dimensions of vaccine effectiveness, especially between countries which have used different types of vaccines. Given the genetic divergence of Omicron with its predecessors, its emergent genetic sequence coding the spike protein has been documented to sheath away from the respective immunoglobulins or humoral response [37]. Upon analyzing the titers of neutralizing antibodies of sera from vaccinated individuals, the neutralization capacity was lower for Omicron [37,38,39,40,41,42,43]. However, the neutralization capacity was maintained among vaccinated individuals, who also had a history of prior infection [38,39]. Interestingly, one study by Cele et al. [44] reported high levels of neutralizing antibodies in the plasma of subjects who had a history of previous infection and were fully vaccinated. According to an (unpublished) report by Andrew et al., the vaccine effectiveness was raised by 75–80% following the administration of a booster dose [45]. This might be relatively lower than the Delta variant, but still encouraging to deal with the unknowns of Omicron. Some studies have compared the immunogenic properties of homologous and heterologous vaccinations, and found that the heterologous prime-boost regimens exhibited higher neutralizing activity compared with the homologous vaccinations [46]. More data will be needed to confirm this finding. Existing data suggest a reduced efficiency of vaccines in antibody neutralization; however, the cell-mediated immunity remains resilient in offering protection from severe illness [37,38,39,40,41,42,43,47]. T cells offer cell-mediated immunity and have an ability to recognize mutated virus through multiple sites beyond the spike protein [37,38,39,40,41,42,43,44,45,46,47]. In addition, T cells provide long-lasting immunity which does not fades as quickly as natural antibodies [45,46,47,48].

6. Laboratory Testing

Although reverse transcription—polymerase chain reaction (RT-PCR) tests remain the gold standard in COVID-19 testing, many rapid antigen detection tests (RADTs) and self-RADTs have been introduced and widely adopted. With the emergence of new variants, reductions in the RADT sensitivity have not yet been reported [49]. Studies have reported the effectiveness of antigen tests in detecting the Omicron variant at the lowest dilutions [49]. It is important to note that studies which will provide more insight to the analytical sensitivity of RADTs to detect the Omicron variant in settings with high transmission are ongoing. RT-PCR-based S-gene target failure (SGTF) assays that do not detect the S-gene with Δ69–70 deletion may serve to screen for the Omicron VOC [49,50]. Screening for VOC-specific amino acid substitutions can also be performed using specific RT-PCR assays targeting single-nucleotide polymorphisms (SNPs) [49]. Some commercially available SNP test kits for the identification of T478K, N501Y and P681H fail to reliably identify these mutations for Omicron, even with S-gene mutations [50]. The U.S. FDA has identified and listed molecular tests that may be impacted by new Omicron variant mutations [51]. The currently widely used real-time RT-PCR tests can detect the new variant, but rapid antigen detection tests may need more confirmatory studies in this regard.

7. Therapeutics

The WHO established its Joint Advisory Group on COVID-19 with the aim of collecting and synthesizing evidence related to the potential impact of the emergence of Omicron on the effectiveness of drugs currently being used or under investigation [52,53,54]. At present, limited evidence is available regarding the effectiveness of bioequivalents of monoclonal antibody therapies, including Sotrovimab, Casirivimab, Imdevimab, Bamlanivimab, etc. [54] against Omicron. Some early reports have indicated that the cocktail mixture of Casirivimab and Imdevimab does not neutralize Omicron in vitro [53], whereas the neutralization capability of Sotrovimab was retained against Omicron [53,54]. There is presumptive thought based on the genetic analysis regarding the efficacy of Remedisivir that it may continue to be active against Omicron; however, laboratory confirmative study data are needed [53,54]. Clinical and laboratory data on the effectiveness of newer oral antivirals against Omicron are not yet available. Owing to genomic changes, the variant might be responsive to some available treatments and may not respond to other modalities effective against the original virus. To assess the effect on the three currently available monoclonal antibody treatments (Sotrovimab, Bamlanivimab and Etesevimab, and REGEN-COV) more data need to be evaluated [54].

8. Public Health Response to Minimize Omicron Transmission

Similarly to other infectious and highly transmissible diseases, efforts at macro (country level) and micro level (individual level) are vital. First and foremost, surveillance plays an integral role in the public health response in curbing the transmission of Omicron. Recently, the CDC established a genomic surveillance system to improve the understanding of predominant strains circulating in the community. The CDC has partnered with commercial laboratories and universities to supplement public health sequencing efforts and conduct genomic surveillance research [55]. According to the WHO recommendations, countries should extend their research infrastructure to develop a science-based approach to curb the spread of COVID-19. Ensuring equitable access to healthcare will be critical to overcoming the disparate effects of the pandemic and new challenges induced by the mutated strains. At individual level, physical distancing, the use of masks, avoiding public gatherings and getting vaccinated will remain useful in curbing the spread.

Author Contributions

Conceptualization, S.A. and V.G.; methodology, S.A.S, G.D.; software, Y.A.A., M.Y.A. and K.B.; validation, S.M.A., R.J. and A.O.; formal analysis, K.B.; investigation, S.A.; resources, P.S.; data curation, S.M.A.; writing—original draft preparation, S.A.; writing—review and editing, V.G.; visualization, K.B.; supervision, G.D.; project administration, S.A.S., M.Y.A. and Y.A.A.; funding acquisition, R.J. and A.O. All authors have read and agreed to the published version of the manuscript.

Funding

The research was funded by the Deanship of Scientific Research at King Khalid University through the Small Research Group Project under grant number (RGP.1/336/42).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Centers for Disease Control and Prevention Transmission of Coronavirus Disease 2019 (COVID-19). Available online: https://www.cdc.gov/coronavirus/2019-ncov/index.html (accessed on 21 January 2022).
  2. Coronaviridae Study Group of the International Committee on Taxonomy of Viruses. The species Severe acute respiratory syndrome-related coronavirus: Classifying 2019-nCoV and naming it SARS-CoV-2. Nat. Microbiol. 2020, 5, 536–544. [Google Scholar] [CrossRef] [Green Version]
  3. Wax, R.S.; Christian, M.D. Practical recommendations for critical care and anesthesiology teams caring for novel coronavirus (2019-nCoV) patients. Can. J. Anesth. 2020, 67, 568–576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. The Worldometers. COVID-19 Coronavirus Pandemic. 2022. Available online: https://www.worldometers.info/coronavirus/ (accessed on 21 January 2022).
  5. Centers for Disease Control and Prevention. SARS-CoV-2 Variant Classifications and Definitions. 2021. Available online: https://www.cdc.gov/coronavirus/2019-ncov/variants/variant-info.html (accessed on 6 December 2021).
  6. Centers for Disease Control and Prevention. Omicron Variant: What You Need to Know. 2021. Available online: https://www.cdc.gov/coronavirus/2019-ncov/variants/omicron-variant.html (accessed on 31 December 2021).
  7. Boynton, S. When Will the Omicron Wave End? Data Suggests It Could Be Soon, but Experts Are Wary. Available online: https://globalnews.ca/news/8494760/omicron-wave-end-covid/ (accessed on 10 January 2022).
  8. Government of South Africa. Cabinet Approves Several Changes to the Adjusted Alert Level 1 COVID19 Regulations [Internet]. 2021. Available online: https://www.gov.za/speeches/cabinet-approves-several-changes-adjusted-alertlevel-1-covid-19-regulations-30-dec-2021 (accessed on 10 January 2022).
  9. World Health Organization. Update on Omicron 2021. Available online: https://www.who.int/news/item/28-11-2021-update-on-omicron (accessed on 6 December 2021).
  10. European Centre for Disease Prevention and Control. Epidemiological Update: Omicron Variant of Concern (VOC)-Data As of 11 December 2021. Available online: https://www.ecdc.europa.eu/en/news-events/epidemiological-update-omicron-variant-concern-voc-data-11-december-2021 (accessed on 21 January 2022).
  11. Liu, L.; Iketani, S.; Guo, Y.; Chan, J.F.-W.; Wang, M.; Liu, L.; Luo, Y.; Chu, H.; Huang, Y.; Nair, M.S.; et al. Striking Antibody Evasion Manifested by the Omicron Variant of SARS-CoV-2. Nature 2021. [Google Scholar] [CrossRef] [PubMed]
  12. Pulliam, J.R.C.; van Schalkwyk, C.; Govender, N.; von Gottberg, A.; Cohen, C.; Groome, M.J.; Dushoff, J.; Mlisana, K.; Moultrie, H. Increased risk of SARS-CoV-2 reinfection associated with emergence of the Omicron variant in South Africa. Medrxiv 2021. [Google Scholar] [CrossRef]
  13. McCreesh, N.; Dlamini, V.; Edwards, A.; Oliver, S.; Dayi, N.; Dikgale, K.; Nxumalo, S.; Dreyer, J.; Baisley, K.; Siedner, M.J.; et al. Impact of the COVID-19 epidemic and related social distancing regulations on social contact and SARS-CoV-2 transmission potential in rural South Africa: Analysis of repeated cross-sectional surveys. BMC Infect. Dis. 2021, 21, 928. [Google Scholar] [CrossRef] [PubMed]
  14. Mohiuddin, M.; Kasahara, K. Investigating the aggressiveness of the COVID-19 Omicron variant and suggestions for possible treatment options. Respir. Med. 2021, 191, 106716. [Google Scholar] [CrossRef] [PubMed]
  15. French, G.; Hulse, M.; Nguyen, D.; Sobotka, K.; Wedster, K.; Corman, J.; Aboagye-Nyame, B.; Dion, M.; Johnson, M.; Zalinger, B.; et al. Impact of Hospital Strain on Excess Deaths During the COVID-19 Pandemic-United States, July 2020–July 2021. MMWR Morb. Mortal. Wkly. Rep. 2021, 70, 1613–1616. [Google Scholar] [CrossRef]
  16. Kannan, S.R.; Spratt, A.N.; Sharma, K.; Chand, H.S.; Byrareddy, S.N.; Singh, K. Omicron SARS-CoV-2 variant: Unique features and their impact on pre-existing antibodies. J. Autoimmun. 2022, 126, 102779. [Google Scholar] [CrossRef]
  17. Lubin, J.H.; Markosian, C.; Balamurugan, D.; Pasqualini, R.; Arap, W.; Burley, S.K.; Khare, S.D. Structural models of SARS-CoV-2 Omicron variant in complex with ACE2 receptor or antibodies suggest altered binding interfaces. bioRxiv 2021, 12, 472313. [Google Scholar] [CrossRef]
  18. Saxena, S.K.; Kumar, S.; Ansari, S.; Paweska, J.T.; Maurya, V.K.; Tripathi, A.K.; Abdel-Moneim, A.S. Characterization of the novel SARS-CoV-2 Omicron (B.1.1.529) variant of concern and its global perspective. J. Med. Virol. 2021. [Google Scholar] [CrossRef]
  19. 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. [Google Scholar] [CrossRef]
  20. Kumar, S.; Thambiraja, T.S.; Karuppanan, K.; Subramaniam, G. Omicron and Delta variant of SARS-CoV-2: A comparative computational study of spike protein. J. Med. Virol. 2021. [Google Scholar] [CrossRef]
  21. Centers for Disease Control And Prevention. Science Brief: Omicron (B.1.1.529) Variant. 2021. Available online: https://www.cdc.gov/coronavirus/2019-ncov/science/science-briefs/scientific-brief-omicron-variant.html (accessed on 31 December 2021).
  22. Scialo, F.; Daniele, A.; Amato, F.; Pastore, L.; Matera, M.G.; Cazzola, M.; Castaldo, G.; Bianco, A. ACE2: The Major Cell Entry Receptor for SARS-CoV-2. Lung 2020, 198, 867–877. [Google Scholar] [CrossRef]
  23. Joseph, S.; Nair, B.; Nath, L.R. The Ineluctable Role of ACE-2 Receptors in SARS-CoV-2 Infection and Drug Repurposing as a Plausible SARS-CoV-2 Therapy: A Concise Treatise. Curr. Mol. Med. 2021, 21, 888–913. [Google Scholar] [CrossRef]
  24. Zhao, H.; Lu, L.; Peng, Z.; Chen, L.-L.; Meng, X.; Zhang, C.; Chan, W.-M.; Chu, A.W.-H.; Chan, K.-H.; Jin, D.-Y.; et al. SARS-CoV-2 Omicron variant shows less efficient replication and fusion activity when compared with Delta variant in TMPRSS2-expressed cells. Emerg. Microbes Infect. 2022, 11, 277–283. [Google Scholar] [CrossRef]
  25. Saito, A.; Irie, T.; Suzuki, R.; Maemura, T.; Nsaaer, H.; Uriu, K.; Kosugi, Y.; Shirakawa, K.; Sadamasu, K.; Kimura, I.; et al. Enhanced fusogenicity and pathogenicity of SARS-CoV-2 Delta P681R mutation. Nature 2021. [Google Scholar] [CrossRef]
  26. Youk, J.; Kim, T.; Evans, K.V.; Jeong, Y.-I.; Hur, Y.; Hong, S.P.; Kim, J.H.; Yi, K.; Kim, S.Y.; Na, K.J.; et al. Three-Dimensional Human Alveolar Stem Cell Culture Models Reveal Infection Response to SARS-CoV-2. Cell Stem Cell 2020, 27, 905–919. [Google Scholar] [CrossRef]
  27. Abbott, S.; Hellewell, J.; Thompson, R.N.; Sherratt, K.; Gibbs, H.P.; Bosse, N.I.; Munday, J.D.; Meakin, S.; Doughty, E.L.; Chun, J.Y.; et al. Estimating the time-varying reproduction number of SARS-CoV-2 using national and subnational case counts. Wellcome Open Res. 2020, 5, 112. [Google Scholar] [CrossRef]
  28. Song, W.J.; Hui, C.K.M.; Hull, J.H.; Birring, S.S.; Mcgarver, L.; Mazzone, S.B.; Chung, K.F. Confronting COVID-19-associated cough and the post-COVID syndrome: Role of viral neurotropism, neuroinflammation, and neuroimmune responses. Lancet Respir. Med. 2021, 9, 533–544. [Google Scholar] [CrossRef]
  29. Report from Norway. Available online: https://www.eurosurveillance.org/content/10.2807/1560-7917.ES.2021.26.50.2101147#f1 (accessed on 10 January 2022).
  30. UK Health Security Agency (UKHSA). SARS-CoV-2 Variants of Concern and Variants Under Investigation in England-Technical Briefing 31. 2021. Available online: https://assets.publishing.service.gov.uk/government/uploads/system/uploads/attachment_data/file/1040076/Technical_Briefing_31.pdf (accessed on 10 January 2022).
  31. Ibekwe, T.S.; Fasunla, A.J.; Orimadegun, A.E. Systematic Review and Meta-analysis of Smell and Taste Disorders in COVID-19. OTO Open 2020, 4, 2473974X20957975. [Google Scholar] [CrossRef]
  32. Hoffmann, M.; Kleine-Weber, H.; Schroeder, S.; Krüger, N.; Herrler, T.; Erichsen, S.; Schiergens, T.S.; Herrler, G.; Wu, N.-H.; Nitsche, A.; et al. SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and Is Blocked by a Clinically Proven Protease Inhibitor. Cell 2020, 181, 271–280. [Google Scholar] [CrossRef]
  33. Dong, M.; Zhang, J.; Ma, X.; Tan, J.; Chen, L.; Liu, S.; Xin, Y.; Zhuang, L. ACE2, TMPRSS2 distribution and extrapulmonary organ injury in patients with COVID-19. Biomed. Pharmacother. 2020, 131, 110678. [Google Scholar] [CrossRef]
  34. Maslo, C.; Friedland, R.; Toubkin, M.; Laubscher, A.; Akaloo, T.; Kama, B. Characteristics and Outcomes of Hospitalized Patients in South Africa During the COVID-19 Omicron Wave Compared with Previous Waves. JAMA 2021, 327, 583–584. [Google Scholar] [CrossRef]
  35. Burki, T.K. Omicron variant and booster COVID-19 vaccines. Lancet Respir. Med. 2021, 10, e17. [Google Scholar] [CrossRef]
  36. Poudel, S.; Ishak, A.; Perez-Fernandez, J.; Garcia, E.; León-Figueroa, D.A.; Romaní, L.; Bonilla-Aldana, D.K.; Rodriguez-Morales, A.J. Highly mutated SARS-CoV-2 Omicron variant sparks significant concern among global experts—What is known so far? Travel Med. Infect. Dis. 2021. [Google Scholar] [CrossRef]
  37. Ahmed, S.F.; Quadeer, A.A.; McKay, M.R. SARS-CoV-2 T Cell Responses Elicited by COVID-19 Vaccines or Infection Are Expected to Remain Robust against Omicron. Viruses 2022, 14, 79. [Google Scholar] [CrossRef]
  38. Wilhelm, A.; Widera, M.; Grikscheit, K.; Toptan, T.; Schenk, B.; Pallas, C.; Metzler, M.; Kohmer, N.; Hoehl, S.; Helfritz, F.A.; et al. Reduced neutralization of SARS-CoV-2 Omicron variant by vaccine sera and monoclonal antibodies. medRxiv 2021. [Google Scholar] [CrossRef]
  39. Roessler, A.; Riepler, L.; Bante, D.; von Laer, D.; Kimpel, J. SARS-CoV-2 Omicron Variant Neutralization in Serum from Vaccinated and Convalescent Persons. N. Engl. J. Med. 2022. [Google Scholar] [CrossRef]
  40. Planas, D.; Saunders, N.; Maes, P.; Guivel-Benhassine, F.; Planchais, C.; Buchrieser, J.; Bolland, W.-H.; Porrot, F.; Staropoli, I.; Lemoine, F.; et al. Considerable escape of SARS-CoV-2 Omicron to antibody neutralization. Nature 2021. [Google Scholar] [CrossRef]
  41. Krammer, F. A correlate of protection for SARS-CoV-2 vaccines is urgently needed. Nat. Med. 2021, 27, 1147–1148. Available online: https://www.nature.com/articles/s41591-021-01432-4 (accessed on 10 January 2022). [CrossRef]
  42. Hartley, G.E.; Edwards, E.S.; Aui, P.M.; Varese, N.; Stojanovic, S.; McMahon, J.; Peleg, A.Y.; Boo, I.; Drummer, H.E.; Hogarth, P.M.; et al. Rapid generation of durable B cell memory to SARS-CoV-2 spike and nucleocapsid proteins in COVID-19 and convalescence. Sci. Immunol. 2020, 5, eabf8891. Available online: https://immunology.sciencemag.org/content/5/54/eabf8891/43 (accessed on 10 January 2022). [CrossRef] [PubMed]
  43. Lau, E.H.; Tsang, O.T.; Hui, D.S.; Kwan, M.Y.; Chan, W.-h.; Chiu, S.S.; Ko, R.; Chan, K.H.; Cheng, S.; Perera, R.; et al. Neutralizing antibody titres in SARS-CoV-2 infections. Nat. Commun. 2021, 12, 1–7. Available online: https://www.nature.com/articles/s41467-020-20247-4 (accessed on 10 January 2022). [CrossRef]
  44. Cele, S.; Jackson, L.; Khoury, D.S.; Khan, K.; Moyo-Gwete, T.; Tegally, H.; San, J.E.; Cromer, D.; Scheepers, C.; Amoako, D.; et al. SARS-CoV-2 Omicron has extensive but incomplete escape of Pfizer BNT162b2 elicited neutralization and requires ACE2 for infection. medRxiv 2021. [Google Scholar] [CrossRef]
  45. Andrews, N.; Stowe, J.; Kirsebom, F.; Toffa, S.; Rickeard, T.; Gallagher, E.; Gower, C.; Kall, M.; Groves, N.; Connell, A.M.O’.; et al. Effectiveness of COVID-19 vaccines against the Omicron (B.1.1.529) variant of concern. medRxiv 2021. Available online: https://khub.net/documents/135939561/430986542/Effectiveness+of+COVID-19+vaccines+against+Omicron+variant+of+concern.pdf/f423c9f4-91cb-0274-c8c5-70e8fad50074 (accessed on 10 January 2022). [CrossRef]
  46. Barros-Martins, J.; Hammerschmidt, S.I.; Cossmann, A.; Odak, I.; Stankov, M.V.; Morillas Ramos, G.; Dopfer-Jablonka, A.; Heidemann, A.; Ritter, C.; Friedrichsen, M.; et al. Immune responses against SARS-CoV-2 variants after heterologous and homologous ChAdOx1 nCoV-19/BNT162b2 vaccination. Nat. Med. 2021, 27, 1525–1529. [Google Scholar] [CrossRef] [PubMed]
  47. Hoffmann, M.; Krüger, N.; Schulz, S.; Cossmann, A.; Rocha, C.; Kempf, A.; Nehlmeier, I.; Graichen, L.; Moldenhauer, A.S.; Winkler, M.S.; et al. The Omicron variant is highly resistant against antibody-mediated neutralization: Implications for control of the COVID-19. Cell 2021. [Google Scholar] [CrossRef] [PubMed]
  48. De Marco, L.; Dorso, S.; Pirronello, M.; Verdiani, A.; Termine, A.; Fabrizio, C.; Capone, A.; Sabatini, A.; Guerrera, G.; Placido, R.; et al. Preserved T cell reactivity to the SARS-CoV-2 Omicron variant indicates continued protection in vaccinated individuals. bioRxiv 2021. [Google Scholar] [CrossRef]
  49. Deerain, J.; Druce, J.; Tran, T.; Batty, M.; Yoga, Y.; Fennell, M.; Dwyer, D.E.; Kok, J.; Williamson, D.A. Assessment of the analytical sensitivity of ten lateral flow devices against the SARS-CoV-2 omicron variant. J. Clin. Microbiol. 2021. [Google Scholar] [CrossRef] [PubMed]
  50. Korukluoglu, G.; Kolukirik, M.; Bayrakdar, F.; Ozgumus, G.G.; Altas, A.B.; Cosgun, Y.; Ketre Kolukirik, C.Z. 40 minutes RT-qPCR Assay for Screening Spike N501Y and HV69-70del Mutations. bioRxiv 2021. [Google Scholar] [CrossRef]
  51. US Food and Drug Administration (FDA). SARS-CoV-2 Viral Mutations: Impact on COVID-19 Tests. 2021. Available online: https://www.fda.gov/medical-devices/coronavirus-covid-19-and-medical-devices/sars-cov2-viral-mutations-impact-covid-19-tests (accessed on 10 January 2022).
  52. World Health Organization. WHO Joint Advisory Group on COVID-19 Therapeutics Prioritization. 2021. Available online: https://www.who.int/publications/m/item/who-joint-advisory-group-on-covid-19-therapeutics-prioritization—draft-statement-on-the-possible-effects-of-the-new-sars-cov-2-omicron-variant-on-treatment-of-hospitalized-covid-19-patients (accessed on 24 January 2022).
  53. Cameroni, E.; Saliba, C.; Bowen, J.E.; Rosen, L.E.; Culap, K.; Pinto, D.; VanBlargan, L.A.; Marco, A.D.; Zepeda, S.K.; Iulio, J.D.; et al. Broadly neutralizing antibodies overcome SARS-CoV-2 Omicron Antigenic Shift. Nature 2021. [Google Scholar] [CrossRef]
  54. Aggarwal, A.; Stella, A.O.; Walker, G.; Akerman, A.; Milogiannakis, V.; Brilot, F.; Amatayakul-Chantler, S.; Roth, N.; Coppola, G.; Schofield, P.; et al. SARS-CoV-2 Omicron: Evasion of potent humoral responses and resistance to clinical immunotherapeutics relative to viral strains of concer. medRxiv 2021. [Google Scholar] [CrossRef]
  55. Centers for Disease Control and Prevention. CDC’S role in Tracking Variants. 2021. Available online: https://www.cdc.gov/coronavirus/2019-ncov/variants/cdc-role-surveillance.html (accessed on 24 January 2022).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Arora, S.; Grover, V.; Saluja, P.; Algarni, Y.A.; Saquib, S.A.; Asif, S.M.; Batra, K.; Alshahrani, M.Y.; Das, G.; Jain, R.; et al. Literature Review of Omicron: A Grim Reality Amidst COVID-19. Microorganisms 2022, 10, 451. https://doi.org/10.3390/microorganisms10020451

AMA Style

Arora S, Grover V, Saluja P, Algarni YA, Saquib SA, Asif SM, Batra K, Alshahrani MY, Das G, Jain R, et al. Literature Review of Omicron: A Grim Reality Amidst COVID-19. Microorganisms. 2022; 10(2):451. https://doi.org/10.3390/microorganisms10020451

Chicago/Turabian Style

Arora, Suraj, Vishakha Grover, Priyanka Saluja, Youssef Abdullah Algarni, Shahabe Abullais Saquib, Shaik Mohammed Asif, Kavita Batra, Mohammed Y. Alshahrani, Gotam Das, Rajni Jain, and et al. 2022. "Literature Review of Omicron: A Grim Reality Amidst COVID-19" Microorganisms 10, no. 2: 451. https://doi.org/10.3390/microorganisms10020451

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

Article Metrics

Back to TopTop