SARS-CoV-2 and COVID-19: An Evolving Review of Diagnostics and Therapeutics

This manuscript (permalink) was automatically generated from greenelab/covid19-review@bd8f0ff on January 21, 2021.

Authors

COVID-19 Review Consortium: Vikas Bansal, John P. Barton, Simina M. Boca, Christian Brueffer, James Brian Byrd, Stephen Capone, Shikta Das, Anna Ada Dattoli, John J. Dziak, Jeffrey M. Field, Soumita Ghosh, Anthony Gitter, Rishi Raj Goel, Casey S. Greene, Marouen Ben Guebila, Fengling Hu, Nafisa M. Jadavji, Sergey Knyazev, Likhitha Kolla, Alexandra J. Lee, Ronan Lordan, Tiago Lubiana, Temitayo Lukan, Adam L. MacLean, David Mai, Serghei Mangul, David Manheim, Lucy D'Agostino McGowan, YoSon Park, Dimitri Perrin, Yanjun Qi, Diane N. Rafizadeh, Bharath Ramsundar, Halie M. Rando, Sandipan Ray, Michael P. Robson, Elizabeth Sell, Lamonica Shinholster, Ashwin N. Skelly, Yuchen Sun, Gregory L Szeto, Ryan Velazquez, Jinhui Wang, Nils Wellhausen

Abstract

Since late 2019, Coronavirus disease 2019 (COVID-19) has spread around the world, resulting in the declaration of a pandemic by the World Health Organization (WHO). This infectious disease is caused by the newly identified Severe acute respiratory syndrome-related coronavirus 2 (SARS-CoV-2). Research on the SARS-CoV-2 virus and the disease it causes is emerging rapidly through global scientific efforts. Short-term mitigation of viral impacts will require public health interventions, and long-term mitigation will require new diagnostic and therapeutic technologies. The urgency of the pandemic has led to a rapidly emerging scientific literature on SARS-CoV-2 and COVID-19. This manuscript represents a collaborative effort to organize and consolidate this body of literature. We present information about the virus in the context of what is known about related viruses, describe the pathogenesis of COVID-19, and synthesize studies emerging about the diagnosis and treatment of COVID-19 alongside literature about related illnesses. We summarize this emerging literature with an eye towards discussing elements of the disease that will be fundamental to efforts to develop interventions. Our review is a collaboratively-authored, evolving document into which we seek to incorporate the ever-expanding body of information on the topic. This document provides a snapshot as of October, 2020. We continue to accept new contributions and anticipate future snapshots until technologies to mitigate the pandemic are widely deployed.

How to Contribute

We invite potential contributors to introduce themselves through GitHub: https://github.com/greenelab/covid19-review/issues/17

We have established a community chat room on a service called Gitter: https://gitter.im/covid19-review/community

More information about how to contribute is available in a README document on GitHub: https://github.com/greenelab/covid19-review#sars-cov-2-and-covid-19-an-evolving-review-of-diagnostics-and-therapeutics

1 Introduction

On January 21, 2020, the World Health Organization (WHO) released its first report concerning what is now known as the Coronavirus Disease 2019 (COVID-19) [1]. This infectious disease came to international attention on December 31, 2019 following an announcement by national officials in China describing 44 cases of a respiratory infection of unknown cause. The first known cases were located in Wuhan City within the Hubei province of China, but the disease spread rapidly throughout China and subsequently around the world. At the time of the WHO’s first situation report [1], 282 confirmed cases had been identified. Most of these cases were in China, but one to two exported cases had also been identified in each of several neighboring countries (Thailand, Japan, and the Republic of Korea). One week later, 4,593 confirmed cases had been identified, spanning not only Asia, but also Australia, North America, and Europe [2]. On March 11, 2020, the WHO formally classified the situation as a pandemic [3]. On April 4, 2020, the WHO reported that the global number of confirmed cases had surpassed one million [4].

As international attention remains focused on the ongoing public health crisis, the scientific community has responded by mobilizing resources and turning much of its attention to the virus and disease. This rapid influx of information is disseminated by traditional publishing mechanisms, preprint servers, and press releases, which provide a venue for scientists to release findings without undergoing the formal publication process. While having information available is valuable to efforts to understand and combat COVID-19, many contributions come from researchers across a wide range of fields who have varying degrees of experience working on coronaviruses and related topics. The volume of information available, much of which has not gone through rigorous peer review, presents a significant challenge to individual efforts to keep abreast of the state of COVID-19 research [5]. However, research on these topics is proceeding so rapidly that any static review is likely to quickly become dated. Our goal as a community is to consolidate information about the virus in the context of related viruses and to synthesize rapidly emerging literature centered on the diagnosis and treatment of COVID-19. We used an open publishing framework, Manubot [6], to manage hundreds of contributions from the community to create a living, scholarly document. We designed software to generate figures, such as , that automatically update using external data sources. Our primary goal is to sort and distill informative content out of the overwhelming flood of information [5] and help the broader scientific community become more conversant on this critical subject. Thus, our approach has been to develop a real-time, collaborative effort that welcomes submissions from scientists worldwide into this ongoing effort. This document represents the first snapshot, which aims to reflect the state of the field as of October, 2020. We plan to refine and expand this document until technologies to mitigate the pandemic are widely available.

1.1 Interdisciplinary Context

Collaboration across several broad areas of research is critical, as different areas provide different information and context necessary to understanding the virus and disease. This review provides a biological perspective on the virus Severe acute respiratory syndrome-related coronavirus 2 (SARS-CoV-2) and efforts to develop diagnostic, prophylactic, and therapeutic responses to COVID-19. We provide only brief summaries of two other important perspectives on this pandemic: epidemiology and public health. Research in these areas often seeks to anticipate, model, and prevent outbreaks of infectious disease or to understand and manage human behavior relevant to health and disease. Their insights are critical to mounting a global response to the pandemic. Epidemiological analyses have investigated patterns of transmission within and between communities, the symptoms associated with and the duration of infection and/or contagiousness, and how the virus propagates, among other characteristics [7]. Epidemiology also has a close relationship to public policy because it provides model-based insights into how preventative measures and public response can shift outcomes [8]. Public health addresses social and human factors influencing individuals’ exposure and susceptibility to pathogens, such as resource availability, inequality, human behavior, and access to accurate information. Strategies from public health and epidemiology for managing the current epidemic have included the promotion of hand hygiene, social distancing, and personal protective equipment such as masks to mitigate spread, as well as containment approaches such as test, trace, and isolate, which depends on widespread testing, contact tracing, and quarantining. An effective public health management strategy involves response coordination, disease surveillance, intervention monitoring, risk communication, and health education (including the containment of “infodemics” of false information) [9]. Epidemiology and public health intersect with the topics addressed in this manuscript because they both inform and benefit from relevant biotechnological developments. For example, the development of accurate and fast diagnostic testing is relevant to test, trace, and isolate strategies for containment, and public education will be critical to deploying vaccines once they become available. The present analysis focuses less on human and social factors and more on the basic biology of infection, diagnosis, and recovery, but these areas are inextricable in understanding and responding to the COVID-19 pandemic.

1.2 Initial Characterization of SARS-CoV-2

The first genome sequence of the virus was released on January 3, 2020 and revealed that the cluster of pneumonia cases seen in Wuhan were caused by a novel coronavirus [10]. Multiple research groups have drafted the genome sequence of SARS-CoV-2 based on sequences developed from clinical samples collected from the lower respiratory tract, namely bronchoalveolar lavage fluid (BALF), and the upper respiratory tract, in the form of throat swabs [11,12,13]. Analysis of the SARS-CoV-2 genome revealed significant sequence homology with two coronaviruses known to infect humans, with about 79% identity to SARS-CoV-1 and 50% to MERS-CoV [13]. However, the highest degree of similarity was observed between SARS-CoV-2 and bat-derived SARS-like coronaviruses (bat-SL-CoVZC45 and bat-SL-CoVZXC21) [12,13], with identity between SARS-CoV-2 and RATG13 as high as 96.2% [12,14]. This evidence therefore suggests the SARS-CoV-2 virus is the result of zoonotic transfer of a virus from bats to humans. Nevertheless, some fragments between SARS-CoV-2 and RATG13 differ by up to 17%, suggesting a complex natural selection process during zoonotic transfer. While the S region is highly similar to that of viruses found in pangolins [15], there is no consensus about the origin of S in SARS-CoV-2, as it could potentially be the result either of recombination or coevolution [14,16]. Though the intermediate host serving as the source for the zoonotic introduction of SARS-CoV-2 to human populations has not yet been identified, the SARS-CoV-2 virus has been placed within the coronavirus phylogeny through comparative genomic analyses. Genomic analyses and comparisons to other known coronaviruses suggest that SARS-CoV-2 is unlikely to have originated in a laboratory – either purposely engineered and released, or escaped – and instead evolved naturally in an animal host [17]. While the position of the SARS-CoV-2 virus within the coronavirus phylogeny has been largely resolved, the functional consequences of molecular variation between this virus and other viruses, such as its bat and pangolin sister taxa or SARS-CoV-1, remain unknown [17]. Fortunately, the basic genome structure of coronaviruses is highly conserved, and insight into the mechanisms the virus uses to enter cells, replicate, and spread is likely to be available from prior research in coronaviruses.

1.3 Coronaviruses and Humans Hosts

Coronaviruses have long been known to infect animals and have been the subject of veterinary medical investigations and vaccine development efforts due to their effect on the health of companion and agricultural animals [18]. Most coronaviruses show little to no transmission in humans. However, today it is thought that approximately one-third of common cold infections are caused by four human coronaviruses (HCoV): Human coronavirus 229E (HCoV-229E), Human coronavirus NL63 (HCoV-NL63), Human coronavirus OC43 (HCoV-OC43), and Human coronavirus HKU1 (HCoV-HKU1) [19,20]. The first HCoV were identified in the 1960s: HCoV-229E in 1965 [21] and HCoV-OC43 in 1967 [22]. Both of these viruses cause cold-like symptoms [23,24]. Two additional HCoV were subsequently identified [25,26]. In 2003, HCoV-NL63 [25] was first identified in a 7-month-old infant and then in clinical specimens collected from seven additional patients, five of whom were infants younger than 1 year old and the remainder of whom were adults. CoV-HKU1 was identified in samples collected from a 71-year-old pneumonia patient in 2004 and then found in samples collected from a second adult patient [26]. These viruses are associated with respiratory diseases of varying severity, ranging from common cold to severe pneumonia, with severe symptoms mostly observed in immunocompromised individuals [27]. In addition to these relatively mild HCoV, however, highly pathogenic human coronaviruses have been identified, including Severe acute respiratory syndrome-related coronavirus (SARS-CoV or SARS-CoV-1) and Middle East respiratory syndrome-related coronavirus (MERS-CoV) [19,28,29].

At the time that SARS-CoV-1 emerged in the early 2000s, no HCoV had been identified in almost 40 years [28]. The first case of SARS was reported in November 2002 in the Guangdong Province of China, and over the following month, the disease spread more widely within China and then into several countries across multiple continents [28,30]. Unlike previously identified HCoV, SARS was much more severe, with an estimated death rate of 9.5% [30]. It was also highly contagious via droplet transmission, with a basic reproduction number (R0) of 4 (i.e., each person infected was estimated to infect four other people) [30]. However, the identity of the virus behind the infection remained unknown until April of 2003, when the SARS-CoV-1 virus was identified through a worldwide scientific effort spearheaded by the WHO [28]. SARS-CoV-1 belonged to a distinct lineage from the two other HCoV known at the time [30]. By July 2003, the SARS outbreak was officially determined to be under control, with the success credited to infection management practices [28]. A decade later, a second outbreak of severe respiratory illness associated with a coronavirus emerged, this time in the Arabian Peninsula. This disease, known as Middle East respiratory syndrome (MERS), was linked to another novel coronavirus, MERS-CoV. The fatality rate associated with MERS is much higher than that of SARS, at almost 35%, but the disease is much less easily transmitted, with an R0 of 1 [30]. Although MERS is still circulating, its low reproduction number has allowed for its spread to be contained [30]. The COVID-19 pandemic is thus associated with the seventh HCoV to be identified and the fifth since the turn of the millennium, though additional HCoVs may be in circulation but remain undetected.

SARS-CoV-1 and MERS-CoV were ultimately managed largely through infection management practices (e.g., mask wearing) and properties of the virus itself (i.e., low rate of transmission), respectively [28,30]. Vaccines were not used to control either virus, although vaccine development programs were established for SARS-CoV-1 [31]. In general, care for SARS and MERS patients focuses on supportive care and symptom management [30]. Clinical treatments for SARS and MERS developed during the outbreaks generally do not have strong evidence supporting their use. Common treatments included Ribavirin, an antiviral, often in combination with corticosteroids or sometimes interferon (IFN) medications, which would both be expected to have immunomodulatory effects [28]. However, retrospective and in vitro analyses have reported inconclusive results of these treatments on SARS and the SARS-CoV-1 virus, respectively [28]. IFNs and Ribavirin have shown promise in in vitro analyses of MERS, but their clinical effectiveness remains unknown [28]. Therefore, only limited strategy for the pharmaceutical management of COVID-19 can be adopted from previous severe HCoV infections. Research in response to prior outbreaks of HCoV-borne infections, such as SARS and MERS, have, however, provided a strong foundation for hypotheses about the pathogenesis of SARS-CoV-2 as well as potential diagnostic and therapeutic approaches.

1.3.1 Human Immune Response to Viral Threats

Understanding the fundamental organization of the human immune response to viral threats is critical to understanding the varied response to SARS-CoV-2. The human immune system utilizes a variety of innate and adaptive responses to protect against the pathogens it encounters. The innate immune system consists of barriers, such as the skin, mucous secretions, neutrophils, macrophages, and dendritic cells. It also includes cell-surface receptors that can recognize the molecular patterns of pathogens. The adaptive immune system utilizes antigen-specific receptors that are expressed on B and T lymphocytes. These components of the immune system typically act together; the innate response acts first, and the adaptive response begins to act several days after initial infection following the clonal expansion of T and B cells [32]. After a virus enters into a host cell, its antigen is presented by major histocompatibility complex 1 (MHC 1) molecules and is then recognized by cytotoxic T lymphocytes.

In the case of COVID-19, there is also concern about the immune system becoming over-active. One of the main immune responses contributing to the onset of acute respiratory distress syndrome (ARDS) in COVID-19 patients is cytokine storm syndrome (CSS), which causes an extreme inflammatory response due to a release of pro-inflammatory cytokines and chemokines by immune effector cells. In addition to respiratory distress, this mechanism can lead to organ failure and death in severe COVID-19 cases [33]. Details of how the human body responds to SARS-CoV-2, both in healthy and pathological ways, and how these mechanisms can inform the identification of diagnostic, prophylactic, and therapeutic responses are explored in detail throughout this manuscript.

1.3.2 Clinical Presentation of COVID-19

A great diversity of symptom profiles has been observed for COVID-19, although a large study from Wuhan, China suggests fever and cough as the two most common symptoms on admission [34]. One early retrospective study in China described the clinical presentations of patients infected with SARS-CoV-2 as including lower respiratory tract infection with fever, dry cough, and dyspnea [35]. This study [35] noted that upper respiratory tract symptoms were less common, which suggests that the virus targets cells located in the lower respiratory tract. However, data from the New York City region [36,37] showed variable rates of fever as a presenting symptom, suggesting that symptoms may not be consistent across samples. These differences are present when comparing both between institutions in similar locations and between different regions experiencing COVID-19 outbreaks, leading to conflicting reports of the frequency of fever as a presenting symptom for patients upon hospital admission. For example, even within New York City, one study [36] identified low oxygen saturation (<90% without the use of supplemental oxygen or ventilation support) in a significant percentage of patients upon presentation, while another study [37] reported cough, fever, and dyspnea as the most common presenting symptoms. The variability of both which symptoms present and their severity makes it difficult for public health agencies to provide clear recommendations for citizens regarding what symptoms indicate SARS-CoV-2 infection and should prompt isolation.

1.4 Role of the COVID-19 Review

Several review articles on aspects of COVID-19 have already been published. These have included reviews on the disease epidemiology [38], immunological response [39], diagnostics [40], and pharmacological treatments [39,41]. Others [42,43] provide narrative reviews of progress on some important ongoing COVID-19 research questions. With the worldwide scientific community uniting during 2020 to investigate SARS-CoV-2 and COVID-19 from a wide range of perspectives, findings from many disciplines are relevant on a rapid timescale to a broad scientific audience. Additionally, many findings are published as preprints, which are available prior to going through the peer review process. As a result, centralizing, summarizing, and critiquing new literature broadly relevant to COVID-19 can help to expedite the interdisciplinary scientific process that is currently happening at an advanced pace. We are particularly interested in providing background to the development of diagnostic, prophylactic, and therapeutic approaches to COVID-19. Two major concerns within diagnosis include the detection of current infections in individuals with and without symptoms, and the detection of past exposure without an active infection. In the latter category, identifying whether individuals can develop or have developed sustained immunity is also a major consideration. The development of high-throughput, affordable methods for detecting active infections and sustained immunity will be critical to understanding and controlling the disease. The identification of interventions that can mitigate the effect of the virus on exposed and infected individuals is a significant research priority. Some possible approaches include the identification of existing pharmaceuticals that reduce the severity of infection, either by reducing the virus’ virulence (e.g., antivirals) or managing the most severe symptoms of infection. Due to the long timeline for the development of novel pharmaceuticals, in most cases, research surrounding possible pharmaceutical interventions focuses on the identification and investigation of existing compounds whose mechanisms may be relevant to COVID-19. Other foci of current research include the identification of antibodies produced by survivors of COVID-19 and the development of vaccines. Understanding the mechanisms describing host-virus interactions between humans and SARS-CoV-2 is thus critical to identifying candidate therapeutics. An overview of the topics covered is visualized in Thus, in this review, we seek to consolidate information about efforts to develop strategies for diagnosis and therapeutics as new information is released by the scientific community. We include information from both traditional peer-reviewed scientific literature and from preprints, which typically have not undergone peer review but have been critically evaluated by the scientists involved in this effort. The goal of this manuscript is to present preliminary findings within the broader context of COVID-19 research and to identify the broad interpretations of new research, as well as limitations to interpretability.

2 Genomic and Viral Structure of SARS-CoV-2 in the Context of Pathogenesis, Symptomology, and Transmission

2.1 Abstract

The novel coronavirus SARS-CoV-2, which emerged in late 2019, has been a major force shaping the year 2020 as it spread around the world infecting tens of millions of people with coronavirus disease 2019 (COVID-19). While the viral species was unknown prior to January 2020, its similarity to other coronaviruses that infect humans has allowed for rapid insight into the mechanisms that it uses to infect humans, as well as the ways in which the human immune system can respond. Here, we contextualize SARS-CoV-2 among other coronaviruses and identify what is known and what can be inferred about its behavior once inside a human host. Because the genomic content of coronaviruses, which specifies the virus’s structure, is highly conserved, early genomic analysis provided a significant head start in predicting viral pathogenesis. The pathogenesis of the virus offers insights into symptomology, transmission, and individual susceptibility. Additionally, prior research into interactions between the human immune system and coronaviruses identified how these viruses can evade the immune system’s protective mechanisms. We also explore systems-level research into the regulatory and proteomic effects of SARS-CoV-2 infection and the immune response. Understanding the structure and behavior of the virus serves to contextualize the many facets of the COVID-19 pandemic and can influence efforts to control the virus and treat the disease.

2.2 Importance

COVID-19 involves a number of organ systems and can present with a wide range of symptoms. However, understanding how the virus infects epithelial cells can contextualize how these systems connect. Similarly, the modes of viral transmission have been under debate throughout much of 2020, yet the available research suggests that these patterns are very similar to those observed in closely related viruses like SARS-CoV-1 and possibly MERS-CoV. Exploring the structure, phylogeny, and pathogenesis of the virus therefore helps to guide interpretation of the broader impacts of the virus on the human body and on human populations. For this reason, an in-depth exploration of viral mechanisms is critical to a robust understanding of the COVID-19 pandemic.

2.3 Introduction

The current coronavirus disease 2019 (COVID-19) pandemic, caused by the Severe acute respiratory syndrome-related coronavirus 2 (SARS-CoV-2) virus, represents an acute global health crisis. Symptoms of the disease can range from mild to severe to fatal [44] and can affect a variety of organs and systems [45]. Outcomes of the disease can include acute respiratory distress (ARDS) and acute lung injury, as well as damage to other organ systems [45,46]. Understanding the progression of the disease, including these diverse symptoms, depends on understanding how the virus interacts with the host. Additionally, the fundamental biology of the virus can provide insights into how it is transmitted among people, which can, in turn, inform efforts to control its spread. As a result, a thorough understanding of the pathogenesis of SARS-CoV-2 is a critical foundation on which to build an understanding of COVID-19 and the pandemic as a whole.

The rapid identification and release of the genomic sequence of the virus in January 2020 [10] provided early insight into the virus in a comparative genomic context. The viral genomic sequence clusters with known coronaviruses (order Nidovirales, family Coronaviridae, subfamily Orthocoronavirinae). Phylogenetic analysis of the coronaviruses reveals four major subclades, each corresponding to a genus: the alpha, beta, delta, and gamma coronaviruses. Among them, alpha- and betacoronaviruses infect mammalian species, gammacoronaviruses infect avian species, and deltacoronaviruses infect both mammalian and avian species [47]. The novel virus now known as SARS-CoV-2 was identified as betacoronavirus belonging to the B lineage based on phylogenetic analysis of a polymerase chain reaction (PCR) amplicon fragment from five patients along with the full genomic sequence [48]. This lineage also includes the Severe acute respiratory syndrome-related coronavirus (SARS-CoV-1) that caused the 2002-2003 outbreak of Severe Acute Respiratory Syndrome (SARS) in humans [48]. Because viral structure and mechanisms of pathogenicity are highly conserved within the order, this phylogenetic analysis provides a basis for forming hypotheses about how the virus interacts with hosts, including which tissues, organs, and systems would be most susceptible to SARS-CoV-2 infection. Coronaviruses that infect humans (HCoV) are not common, but prior research into other HCoV such as SARS-CoV-1 and Middle East respiratory syndrome-related coronavirus (MERS-CoV), as well as other viruses infecting humans such as a variety of influenza species, established a strong foundation that accelerated the pace of SARS-CoV-2 research.

Coronaviruses are large viruses that can be identified by their distinctive “crown-like” shape (Figure 1). Their spherical virions are made from lipid envelopes ranging from 100 to 160 nanometers (nm) in which peplomers of two to three spike (S) glycoproteins are anchored, creating the crown [49,50]. These spikes, which are critical to both viral pathogenesis and to the response by the host immune response, have been visualized using cryo-electron microscopy [51]. Because they induce the human immune response, they are also the target of many proposed therapeutic agents. Viral pathogenesis is typically broken down into three major components: entry, replication, and spread [52]. However, in order to draw a more complete picture of pathogenesis, it is also necessary to examine how infection manifests clinically, identify systems-level interactions between the virus and the human body, and consider the possible effects of variation or evolutionary change on pathogenesis and virulence. In this way, both biomedicine and genomics are important pieces of the puzzle of SARS-CoV-2 presentation and pathogenesis.

2.4 Coronavirus Structure and Pathogenesis

2.4.1 Structure of Coronaviruses

Genome structure is highly conserved among coronaviruses, meaning that the relationship between the SARS-CoV-2 genome and its pathogenesis can be inferred from prior research in related viral species. The genomes of viruses in the Nidovirales order share several fundamental characteristics. They are non-segmented, which means the viral genome is contained in a single capsid, and are enveloped, which means that the genome and capsid are encased by a lipid bilayer. Coronaviruses have large positive-sense RNA (ssRNA+) genomes ranging from 27 to 32 kilobases in length [13,53]. The SARS-CoV-2 genome lies in the middle of this range at 29,903 bp [13]. Genome organization is highly conserved within the order [53]. There are three major regions: one containing the replicase gene and one containing the genes encoding structural proteins [53] (Figure 1). The replicase gene comprises about two-thirds of the genome and consists of two open reading frames that are translated with ribosomal frameshifting [53]. This polypeptide is then translated into 16 non-structural proteins (nsp), except in Gammacoronaviruses where nsp1 is absent, that form replication machinery used to synthesize viral RNA [54]. The remaining third of the genome encodes structural proteins, including the spike (S), membrane (M), envelope (E), and nucleocapsid (N) proteins. Additional accessory genes are sometimes present between these two regions, depending on the species or strain. Much attention has been focused on the S protein, which is a critical structure involved in cell entry.

Figure 1: Structure of SARS-CoV-2 capsid and genome. A) The genomic structure of coronaviruses is highly conserved, including the order and organization of genes such as the spike (S), envelope (E), membrane (M) and nucleocapsid (N) proteins. B) The physical structure of the coronavirus virion, including the components encoded by conserved genes S, E, M and N.

2.4.2 Pathogenic Mechanisms of Coronaviruses

While, like most viruses, it is possible that SARS-CoV-1 and SARS-CoV-2 can enter cells through endocytosis, coronaviruses are able to target cells for entry through fusion with the plasma membrane [55,56]. This process is conserved among coronaviruses and is closely associated with the content of their genomes. Cell entry proceeds in three steps: binding, cleavage, and fusion. First, the viral spike protein binds to a host cell via a recognized receptor. Coronaviruses can bind to a range of host receptors [57,58], with binding conserved only at the genus level [47]. Viruses in the betacoronavirus genus, to which SARS-CoV-2 belongs, are known to bind to the CEACAM1 protein, 5-N-acetyl-9-O-acetyl neuraminic acid (Neu 5,9 Ac2), and to the angiotensin-converting enzyme 2 (ACE2) [57]. SARS-CoV-2 has a high affinity for the human ACE2 receptor, which is expressed in the vascular epithelium, other epithelial cells, and cardiovascular and renal tissues [59,60]. The binding process is guided by the molecular structure of the spike protein, which is structured in three segments: an ectodomain, a transmembrane anchor, and an intracellular tail [61]. The ectodomain forms the crown-like structures on the viral membrane and contains two subdomains known as the S1 and S2 subunits [62]. The S1 (N-terminal) domain forms the head of the crown and contains the receptor binding motif, and the S2 (C-terminal) domain forms the stalk that supports the head [62]. The S1 subunit guides the binding of the virus to a host cell receptor, and the S2 subunit guides the fusion process [61].

After the binding of the S1 subunit to a receptor, the spike protein is often cleaved at the S1-S2 boundary by a host protease [63,64,65]. Similar to SARS-CoV, SARS-CoV-2 exhibits redundancy in which host proteases can cleave the S protein [66]. Specifically, both transmembrane protease serine protease-2 (TMPRSS2) and cathepsins B/L have been shown to mediate SARS-CoV-2 S protein proteolytic priming, and small molecule inhibition of these enzymes fully inhibited viral entry in vitro [66,67]. Proteolytic priming prepares the S protein for fusion [64,65]. The two subunits remain bound by van der Waals forces, with the S1 subunit stabilizing the S2 subunit during the membrane fusion process [63]. Electron microscopy suggests that in some coronaviruses, including SARS-CoV-1 and MERS-CoV, a six-helix bundle separates the two subunits in the postfusion conformation, and the unusual length of this bundle facilitates membrane fusion through the release of additional energy [47]. Cleavage at a second site within S2 by these same proteases activates S for fusion by inducing conformational changes [63]. The viral membrane can then fuse with the endosomal membrane to release the viral genome into the host cytoplasm. Once the virus enters a host cell, the replicase gene is translated and assembled into the viral replicase complex. This complex then synthesizes the double-stranded RNA (dsRNA) genome from the genomic ssRNA(+). The dsRNA genome is transcribed and replicated to create viral mRNAs and new ssRNA(+) genomes [53,68]. From there, the virus can spread into other cells. In this way, the genome of SARS-CoV-2 provides insight into the pathogenic behavior of the virus.

2.4.3 Immune Evasion Strategies

Research in other HCoV provides some indication of how SARS-CoV-2 infection proceeds in spite of the human immune response. By infecting the epithelium, viruses such as SARS-CoV-1 are known to bypass the physical barriers such as skin and mucus that comprise the immune system’s first line of defense [69]. Once the virus infiltrates host cells, it is adept at evading detection. CD163+ and CD68+ macrophage cells are especially crucial for the establishment of SARS-CoV-1 in the body [69]. These cells most likely serve as viral reservoirs that help shield SARS-CoV-1 from the innate immune response. According to a study on the viral dissemination of SARS-CoV-1 in Chinese macaques, viral RNA could be detected in some monocytes throughout the process of differentiation into dendritic cells [69]. Thus lack of active viral replication allows SARS-CoV-1 to escape innate immunity because reduced levels of detectable viral RNA allow the virus to avoid both natural killer (NK) cells and Toll-like receptors [69]. Even during replication, SARS-CoV-1 is able to mask its dsRNA from detection by the immune system. Although dsRNA is a pathogen-associated molecular pattern (PAMP) that would typically initiate a response from the innate immune system [70], in vitro analysis of nidoviruses including SARS-CoV-1 suggests that these viruses can induce the development of double-membrane vesicles that protect the dsRNA signature from being detected by the host immune system [71]. This protective envelope can therefore insulate these coronaviruses from the innate immune response’s detection mechanism [33].

HCoV are also known to interfere with the host immune response, rather than just evade it. For example, the virulence of SARS-CoV is increased by nsp1, which can suppress host gene expression by stalling mRNA translation and inducing endonucleolytic cleavage and mRNA degradation [72]. SARS-CoV-1 also evades the immune response by interfering with type I IFN induction signaling, which is a mechanism that leads to cellular resistance to viral infections. SARS-CoV-1 employs methods such as ubiquitination and degradation of RNA sensor adaptor molecules MAVS and TRAF3/6 [73]. Also, MERS-CoV downregulates antigen presentation via MHC class I and MHC class II, which leads to a reduction in T cell activation [73]. These evasion mechanisms, in turn, can lead to systemic infection. Coronaviruses such as SARS-CoV-1 are also able to evade the humoral immune response through other mechanisms, such as inhibiting certain cytokine pathways or down-regulating antigen presentation by the cells [71].

2.4.4 Host Cell Susceptibility

ACE2 and TMPRSS2 have been identified as the primary entry portal and as a critical protease, respectively, in facilitating the entry of SARS-CoV-1 and SARS-CoV-2 into a target cell [51,66,74,75,76]. This finding has led to a hypothesized role for ACE2 and TMPRSS2 expression in determining which cells, tissues, and organs are most likely to be infected by SARS-CoV-2. The ACE2 receptor is expressed in numerous organs, such as the heart, kidney, and intestine, but it is most prominently expressed in alveolar epithelial cells; this pattern of expression is expected to contribute to the virus’ association with lung pathology [59,77]. Clinical investigations of COVID-19 patients have detected SARS-CoV-2 transcripts in bronchoalveolar lavage fluid (BALF) (93% of specimens), sputum (72%), nasal swabs (63%), fibrobronchoscopy brush biopsies (46%), pharyngeal swabs (32%), feces (29%) and blood (1%) [78]. Two studies reported that SARS-CoV-2 could not be detected in the urine specimens [78,79]; however, a third study identified four urine samples (out of 58) that were positive for SARS-CoV-2 nucleic acids [80]. Although respiratory failure remains the leading cause of death for COVID-19 patients [81], SARS-CoV-2 infection can damage many other organ systems including the heart [82], kidneys [83,84], liver [85], and gastrointestinal tract [86,87]. As it becomes clear that SARS-CoV-2 infection can damage multiple organs, the scientific community is pursuing multiple avenues of investigation in order to build a consensus about how the virus affects the human body.

2.5 Clinical Presentation of COVID-19

Pathogenesis is closely linked with the clinical presentation of the disease caused by the virus. Reports have described diverse symptom profiles associated with COVID-19. Differences in the frequency of symptoms are found when comparing both between institutions in similar locations and between different regions. A large study from Wuhan, China conducted early in the pandemic identified fever and cough as the two most common symptoms that patients reported at hospital admission [34], while a retrospective study in China described the clinical presentations of patients infected with SARS-CoV-2 as including lower respiratory tract infection with fever, dry cough, and dyspnea [35]. This study [35] noted that upper respiratory tract symptoms were less common, suggesting that the virus preferentially targets cells located in the lower respiratory tract. However, data from the New York City region [36,37] revealed variable rates of fever as a presenting symptom, suggesting that symptoms may not be consistent across samples. For example, even within New York City, one study [36] identified low oxygen saturation (<90% without the use of supplemental oxygen or ventilation support) in a significant percentage of patients upon presentation, while another study [37] reported cough, fever, and dyspnea as the most common presenting symptoms. The variability of both which symptoms present and their severity has presented challenges for public health agencies to provide clear recommendations for citizens regarding what symptoms indicate SARS-CoV-2 infection and should prompt isolation. Patients may also experience loss of smell, myalgias (muscle aches), fatigue, or headache. Gastrointestinal symptoms can also present [88], and the CDC includes nausea and vomiting, as well congestion and runny nose, on its list of symptoms consistent with COVID-19 [44]. A recent preprint using data from an app-based survey of 500,000 individuals in the US found that among those tested for SARS-CoV-2, a loss of taste or smell, fever, and a cough were significant predictors of a positive test result [89]. It is important to note that in this study, the predictive value of symptoms may be underestimated if they are not specific to COVID-19 because the outcome measured was a positive, as opposed to a negative, COVID-19 test result. At the time the surveys were conducted, due to limits in US testing infrastructure, respondents typically needed to have some symptoms known to be specific to COVID-19 in order to qualify for testing in the first place. Widespread testing of asymptomatic individuals may therefore provide additional insight into the range of symptoms associated with COVID-19.

Consistent with the wide range of symptoms observed, COVID-19 can affect diverse body systems in addition to causing respiratory problems [90]. For example, COVID-19 can lead to acute kidney injury, especially in patients with severe respiratory symptoms or certain preexisting conditions [91]. It can also cause neurological complications [92,93], potentially including stroke, seizures or meningitis [94,95]. COVID-19 has also been associated with an increased incidence of large vessel stroke, particularly in patients under the age of 40 [96], and other thrombotic events including pulmonary embolism and deep vein thrombosis [97]. The mechanism behind these complications has been suggested to be related to coagulopathy, with reports indicating the presence of antiphospholipid antibodies [98] and elevated levels of d-dimer and fibrinogen degradation products in deceased patients [99]. Other viral infections have been associated with coagulation defects and changes to the coagulation cascade; notably SARS was also found to lead to disseminated intravascular coagulation and was associated with both pulmonary embolism and deep vein thrombosis [100]. The mechanism behind these insults has been suggested to be related to inflammation-induced increases in the von Willebrand factor clotting protein, leading to a pro-coagulative state [100]. Abnormal clotting (thromboinflammation or coagulopathy) has been increasingly discussed recently as a possible key mechanism in many cases of severe COVID-19, and may be associated with the high d-dimer levels often observed in severe cases [101,102,103]. This excessive clotting in lung capillaries has been suggested to be related to a dysregulated activation of the complement system, part of the innate immune system [104,105].

2.5.1 Cytokine Release Syndrome

Symptoms of a disease can be caused by a pathogen, but they can also be caused by the immune system’s reaction to the pathogen. A dysregulated immune response can cause significant damage to the host [106,107,108]. The inflammatory response has received particular attention for its role in both a healthy response to infection and a pathogenic one. Inflammation is one of the most visible components of the immune response, as it is responsible for the hallmarks of injury, such as pain, heat, and swelling [109]. In response to injury or to signaling by pattern recognition receptors (PRRs) indicating the detection of a molecular pattern associated with a pathogen or foreign body, the immune system stimulates leukocytes that travel to the site of the threat, where they then produce cytokines [109]. Cytokines are a diverse group of small proteins that play an important role in intercellular signaling [110]. Cytokines can be both pro- and anti-inflammatory, which means they can either stimulate or inhibit the production of additional cytokines [110,111]. Some notable pro-inflammatory cytokines include the interleukins: IL-1β and IL-6 and tumor necrosis factor α (TNF-α) [111]. Anti-inflammatory cytokines play an immunoregulatory role complementary to the cascading effect of pro-inflammatory cytokines [110,111]. A number of interleukins and interferons play anti-inflammatory roles, and receptors or receptor antagonists for inflammatory cytokines are also important for regulating inflammation [111]. IL-10 is an anti-inflammatory cytokine of particular note because it regulates the expression of TNF-α, IL-1, and IL-6 [111]. When the pro- and anti-inflammatory responses are both commensurate with the threat posed, the immune system drives a shift back to homeostasis [112]. However, when the responses are disproportionate, the cytokine response can become dysregulated. Too low of an inflammatory response will not eliminate the immune threat [112]. In contrast, if the response is dysregulated towards excessive pro-inflammatory cytokine activity, inflammation can cascade [113] and cause cell damage, among other problems [109]. Elevated levels of inflammation over the long-term are associated with many chronic health conditions, including type 2 diabetes, dementia and Alzheimer’s, and arthritis, among others [114]. On a shorter timescale, dysregulated systemic inflammation can cause sepsis, which can lead to multi-organ failure and death [110,115].

Cytokines have been investigated for their role in the immune response to lung infections long before the COVID-19 pandemic. Dysregulation of the inflammatory response, including elevated levels of pro-inflammatory cytokines, is found in patients with ARDS, which is a severe condition that can arise from pneumonia, SARS, and COVID-19 [113]. One study of patients with and at risk for ARDS, specifically those who were intubated for medical ventilation, found that shortly after the onset of ARDS, anti-inflammatory cytokine concentration in BALF increased relative to the concentration of pro-inflammatory cytokines [116]. The results suggest that an increase in pro-inflammatory cytokines such as IL-6 may signal the onset of ARDS, but recovery depends on an increased anti-inflammatory response [116]. However, patients with severe ARDS were excluded from this study. Acute phase response to an infection can also cause damage to the capillary endothelium, allowing leaks that disrupt the balance between pro-inflammatory cytokines and their regulators [116]. Hyperactivity of the pro-inflammatory response due to lung infection is commonly associated with acute lung injury and more rarely with the more severe manifestation, ARDS [110]. The heightened inflammatory response in the lungs can also serve as a source for systemic inflammation, or sepsis, which can lead to multi-organ failure [110]. The shift from local to systemic inflammation is a phenomenon often referred to broadly as a cytokine storm [110] or, more precisely, as cytokine release syndrome (CRS) [117]. Sepsis is a known possible complication of pneumonia, and in an analysis of over 1,400 US pneumonia patients, IL-6, tumor necrosis factor (TNF), and IL-10 were found to be elevated at intake in patients who developed severe sepsis and/or ultimately deceased [118]. IL-6 and TNF are pro-inflammatory cytokines, while IL-10; is anti-inflammatory [118]. However, this study reported that unbalanced pro-/anti-inflammatory cytokine profiles were rare, although they measured only the three cytokines listed above. Prior work therefore made it clear that pulmonary infection and injury were associated with systemic inflammation through sepsis. While IL-6 is a biomarker sometimes used to assess cytokine storm activity in sepsis [110], the relationship between cytokine profiles and the risks associated with sepsis may be more complex. IL-6 is a pleiotropic cytokine that plays an integral role in both the inflammatory and anti-inflammatory responses and is associated with both healthy and pathological responses to viral threat [119].

The inflammatory response was identified early on as a potential driver of COVID-19 outcomes due to existing research in SARS and emerging research in COVID-19. In addition to the known role of cytokines in ARDS and lung infection more broadly, immunohistological analysis at autopsy of patients who deceased from SARS revealed that ACE2-expressing cells that were infected by SARS-CoV-1 showed elevated expression of IL-6, IL-1β, and TNF-α [120]. Similarly, the introduction of the S protein from SARS-CoV-1 to mouse macrophages was found to increase production of IL-6 and TNF-α [121]. For SARS-CoV-2 infection leading to COVID-19, early reports described a cytokine storm syndrome-like response in patients with particularly severe infections [77,122,123]. Among patients hospitalized with COVID-19 in Wuhan, China, 112 out of 191 (59%) developed sepsis, including all 54 of the non-survivors [35]. However, the argument has been made that while the cytokine levels observed in COVID-19 patients fall outside of the normal range, they are not as high as typically found in patients with ARDS [124]. Regardless, inflammation has received significant interest both in regards to the pathology of COVID-19 as well as potential avenues for treatment, as the similarities between the cytokine storm and the pathophysiology of COVID-19 has led to the suggestion that a number of immunomodulatory pharmaceutical interventions could hold therapeutic value for the treatment of COVID-19 [125].

2.5.2 Pediatric Presentation

The presentation of COVID-19 infection can vary greatly among pediatric patients and, in some cases, manifests in distinct ways from COVID-19 in adults. Evidence suggests that while children and adolescents tend to have mostly asymptomatic infections, those that are symptomatic exhibit a mild illness [126,127,128,129]. A review examined symptoms reported in 17 studies of children infected with COVID-19 during the early months of the COVID-19 epidemic in China and one study from Singapore [130]. Of the more than a thousand cases described, the most common reports were for mild symptoms such as fever, dry cough, fatigue, nasal congestion and/or runny nose, while three children were reported to be asymptomatic. Severe lower respiratory infection was described in only one of the pediatric cases reviewed. Gastrointestinal symptoms such as vomiting or diarrhea were occasionally reported. Radiologic findings were not always reported in the case studies reviewed, but when they were mentioned they included bronchial thickening, ground-glass opacities, and/or inflammatory lesions [130]. Neurological symptoms have also been reported [131].

These analyses indicate that most pediatric cases of COVID-19 are not severe. Indeed, it is estimated that less than 1% of pediatric cases result in critical illness [128,132]. However, serious complications and, in rare cases, deaths have occurred [133]. Of particular interest, children have occasionally experienced a serious inflammatory syndrome, multisystem inflammatory syndrome in children (MIS-C), following COVID-19 infection. This syndrome is similar in some respects to Kawasaki disease or to Kawasaki disease shock syndrome [134,135,136] and is thought to be a distinct clinical manifestation of SARS-CoV-2 due to its distinct cytokine profile and the presence of burr cells in peripheral blood smears [137,138]. MIS-C has been associated with heart failure in some cases [139]. One case study [140] described an adult who appeared to show symptoms similar to MIS-C after exposure to COVID-19, but cautioned against broad conclusions; a second possible adult case has also been reported [141]. The presentation of SARS-CoV-2 infection is therefore likely to be largely distinct between adult and pediatric populations.

2.6 Systems-Level Effects

Systems biology provides a cross-disciplinary analytical paradigm through which the host response to an infection can be analyzed. This field integrates the “omics” (genomics, transcriptomics, proteomics, metabolomics, etc.) with bioinformatics and other computational approaches. These cutting-edge research paradigms hold enormous potential for the study of the complexity of biological systems and human diseases [142]. Over the last decade, systems biology approaches have been used widely to study the pathogenesis of diverse types of life-threatening acute and chronic infectious diseases [143]. Omics-based studies have also provided meaningful information regarding host immune responses and surrogate protein markers in several viral, bacterial and protozoan infections [144]. Though the complex pathogenesis and clinical manifestations of SARS-CoV-2 infection are not yet fully understood, omics technologies offer the opportunity for discovery-driven analysis of biological changes associated with SARS-CoV-2 infection. For example, previous studies suggest that infection by coronaviruses, such as SARS-CoV-1 and MERS-CoV, and other viruses is associated with the upregulation of ACE2. In several preliminary assays and an analysis of previous microarray data, ACE2 expression was reported to be significantly upregulated following infection of human embryonic kidney cells and human airway epithelial cells [77]. This study also reported that direct stimulation with inflammatory cytokines such as type I interferons (e.g., IFNβ) resulted in the upregulation of ACE2 in human bronchial epithelial cells, with treated groups showing four-fold higher ACE2 expression than control groups at 18 hours post-treatment [77]. Whether SARS-CoV-2 facilitates the positive regulation of its own transmission between host cells is still unclear, the host immune response itself likely plays a key role in mediating infection-associated pathologies. A systems-biology approach allows for analyses such as these to identify possible phenotypic and endophenotypic responses to SARS-CoV-2 infection and to develop new hypotheses about how pathogenesis proceeds.

2.6.1 Transcriptomics

In addition to the study described above, two other studies have profiled expression following SARS-CoV-2 infection using human cell lines. The first study [145] compared transcriptional responses to SARS-CoV-2 and to other respiratory viruses, including MERS-CoV, SARS-CoV, human parainfluenza virus 3 (HPIV3), respiratory syncytial virus (RSV), and influenza A virus (IAV). The authors analyzed the responses of three human cell lines: A549 (adenocarcinomic human alveolar basal epithelial cells), Calu-3 (human airway epithelial cells derived from human bronchial submucosal glands), and MRC-5 (human fetal lung fibroblast cells). As the viral receptor ACE2 has low expression in A549 cells, they supplemented the A549 cells with adenovirus (AdV)-based vectors expressing either mCherry (a fluorescent protein used as a control) or ACE2 (A549-ACE2). The authors also measured host transcriptional responses to SARS-CoV-2 in primary normal human bronchial epithelial cells (HBEC or NHBE cells), nasal washes from an animal model (ferret), and lung samples from two COVID-19 patients. Differential expression (DE) analysis was then carried out to compare infected cells with control cells that underwent only a mock treatment. In the hosts where SARS-CoV-2 was able to replicate efficiently, DE analysis revealed that the transcriptional response was significantly different from the response to all of the other viruses tested. A unique proinflammatory cytokine signature associated with SARS-CoV-2 was present in cells exposed to both high and low doses of the virus, with the cytokines IL-6 and IL1RA uniquely elevated in response to SARS-CoV-2 relative to other viruses. However, the A549-ACE2 cells showed significant IFN-I or IFN-III expression when exposed to high, but not low, doses of SARS-CoV-2. This finding suggests that IFN induction is dependent on the extent of exposure. Similarly, in cells from the NHBE line, ferrets, and COVID-19 patients, chemokine signaling was significantly enriched, but there was no significant induction of IFN-I or IFN-III. Together, these results suggest that SARS-CoV-2 induces a limited antiviral state with low IFN-I or IFN-III expression and a moderate IFN-stimulated gene response, in contrast to other viruses. However, in ACE2-expressing A549 cells, this state could be overcome by using a 10-fold increase in SARS-CoV-2 exposure. This finding suggests that the SARS-CoV-2 interferon antagonist is insufficient for large doses of the virus [145]. This hypothesis was further supported by a recent study [146] that showed that the SARS-CoV-2 ORF3b gene suppresses IFNB1 promoter activity (IFN-I induction) more efficiently than the SARS-CoV-1 ORF3b gene. Taken together, these findings suggest that a unique cytokine profile is associated with the response to the SARS-CoV-2 virus and that this response differs depending on the extent of exposure.

Another study [147] analyzed cells’ transcriptional response to SARS-CoV-2 and SARS-CoV-1 over time. They characterized the response of three human cell lines, H1299 (human non-small cell lung carcinoma cell line), Calu-3, and Caco-2 (human epithelial colorectal adenocarcinoma cell line), at 4 to 36 hours post infection (hpi). Using poly(A) bulk RNA-seq, the authors found negligible susceptibility of H1299 cells (< 0.08 viral read percentage of total reads) compared to Caco-2 and Calu-3 cells (>10% of viral reads). This finding suggests that the risk of infection varies among cell types, and that cell type could influence which hosts are more or less susceptible. Based on visual inspection of microscopy images alongside transcriptional profiling, the authors also showed distinct responses among the host cell lines evaluated. In contrast to Caco-2, Calu-3 cells infected with SARS-CoV-2 showed signs of impaired growth and cell death at 24 hpi, as well as moderate IFN induction with a strong up-regulation of IFN-stimulated genes. Interestingly, the results were similar to those reported in Calu-3 cells exposed to much higher levels of SARS-CoV-2 [145], as described above. This finding suggests that IFN induction in Calu-3 cells is not dependent on the level of exposure, in contrast to A549-ACE2 cells. The discrepancy could be explained by the observations that Calu-3 cells are highly susceptible to SARS-CoV-2 and show rapid viral replication [67], whereas A549 cells are incompatible with SARS-CoV-2 infection [148]. This discrepancy raises the concern that in vitro models may vary in their similarity to the human response, underscoring the importance of follow-up studies in additional models.

2.6.2 Proteomics

One early proteomics study investigated changes associated with in vitro SARS-CoV-2 infection using Caco-2 cells [149]. This study reported that SARS-CoV-2 induced alterations in multiple vital physiological pathways, including translation, splicing, carbon metabolism and nucleic acid metabolism in the host cells. Another area of interest is whether SARS-CoV-2 is likely to induce similar changes to other HCoV. For example, because of the high level of sequence homology between SARS-CoV-2 and SARS-CoV-1, it has been hypothesized that sera from convalescent SARS-CoV-1 patients might show some efficacy in cross-neutralizing SARS-CoV-2-S-driven entry [66]. However, despite the high level of sequence homology, certain protein structures might be immunologically distinct, which would be likely to prohibit effective cross-neutralization across different SARS species [150]. Consequently, proteomic analyses of SARS-CoV-1 might also provide some essential information regarding the new pathogen [151,152].

Considering the paucity of omics-level big data sets for SARS-CoV-2 currently available, existing data hubs that contain information for other coronaviruses such as UniProt, NCBI Genome Database, The Immune Epitope Database and Analysis Resource (IEDB), and The Virus Pathogen Resource (ViPR) will serve as useful resources for computational and bioinformatics research on SARS-CoV-2. Using such databases, the systems-level reconstruction of the protein-protein interaction (PPI) will enable the generation of hypotheses about the mechanism of action of SARS-CoV-2 and suggest potential drug targets. In an initial study [153], 26 of the 29 SARS-CoV-2 proteins were cloned and expressed in HEK293T kidney cells, allowing for the identification of 332 high-confidence human proteins interacting with them. Notably, this study suggested that SARS-CoV-2 interacts with innate immunity pathways. Ranking pathogens by the similarity between their interactomes and that of SARS-CoV-2 suggested West Nile virus, Mycobacterium tuberculosis, and human papillomavirus infections as the top three hits. Therefore, given the lung symptoms associated with COVID-19, the Mycobacterium tuberculosis host-pathogen interactome in particular might provide new insights to the mechanism of SARS-CoV-2 infection. Additionally, it was suggested that the envelope protein, E, could disrupt host bromodomain-containing proteins, i.e., BRD2 and BRD4, that bind to histones, and the spike protein could likely intervene in viral fusion by modulating the GOLGA7-ZDHHC5 acyl-transferase complex to increase palmitoylation, which is a post-translational modification that affects how proteins interact with membranes [154].

Another study [155] used patient-derived peripheral blood mononuclear cells (PBMCs) to identify 251 host proteins targeted by SARS-CoV-2. This study also reported that more than 200 host proteins were disrupted following infection. In particular, a network analysis showed that nsp9 and nsp10 interacted with NF-Kappa-B-Repressing Factor (NKRF), which encodes a transcriptional repressor that mediates repression of genes responsive to Nuclear Factor kappa-light-chain-enhancer of activated B-cells (NF-kB). These genes are important to pro-, and potentially also anti-, inflammatory signaling [156]. This finding could explain the exacerbation of the immune response that shapes the pathology and the high cytokine levels characteristic of COVID-19, possibly due to the chemotaxis of neutrophils mediated by IL-8 and IL-6. Finally, it was suggested [157] that the E protein of both SARS-CoV-1 and SARS-CoV-2 has a conserved Bcl-2 Homology 3 (BH3)-like motif, which could inhibit anti-apoptosis proteins, e.g., BCL2, and trigger the apoptosis of T cells. Several compounds are known to disrupt the host-pathogen protein interactome, largely through the inhibition of host proteins. Therefore, preliminary research in the proteomics of SARS-CoV-2 infection suggests that drugs modulating the protein-level interactions between virus and host might be worth investigating.

2.7 Viral Evolution and Virulence

Like that of SARS-CoV-1, the entry of SARS-CoV-2 into host cells is mediated by interactions between the viral spike glycoprotein, S, and human ACE2 (hACE2) [63,66,158,159,160,161,162,163]. Differences in how the S proteins of the two viruses interact with the hACE2 receptor could also partially account for the increased transmissibility of SARS-CoV-2. Recent studies have reported conflicting binding constants for the S-hACE2 interaction, though they have agreed that the SARS-CoV-2 S protein binds with equal, if not greater, affinity than the SARS-CoV-1 S protein does [51,63,161]. The C-terminal domain of the SARS-CoV-2 S protein in particular was identified as the key region of the virus that interacts with hACE2, and the crystal structure of the C-terminal domain of the SARS-CoV-2 S protein in complex with hACE2 reveals stronger interaction and a higher affinity for receptor binding than that of SARS-CoV-1 [162]. Among the 14 key binding residues identified in the SARS-CoV-1 S protein, eight are conserved in SARS-CoV-2, and the remaining six are semi-conservatively substituted, potentially explaining variation in binding affinity [63,161]. Recent crystal structures have shown that the receptor-binding domain (RBD) of the SARS-CoV-2 S protein, like that of other coronaviruses, undergoes stochastic hinge-like movement that flips it from a “closed” conformation, in which key binding residues are hidden at the interface between protomers, to an “open” one [51,63]. Because the RBD plays such a critical role in viral entry, blocking its interaction with ACE2 could represent a promising therapeutic approach. Nevertheless, despite the high structural homology between the SARS-CoV-2 RBD and that of SARS-CoV-1, monoclonal antibodies targeting SARS-CoV-1-RBD failed to bind to SARS-CoV-2-RBD [51]. Promisingly, though, sera from convalescent SARS patients inhibited SARS-CoV-2 viral entry in vitro, albeit with lower efficiency than it inhibited SARS-CoV-1 [66].

Comparative genomic analysis reveals that several regions of the coronavirus genome are likely critical to virulence. The S1 domain of the spike protein, which contains the receptor binding motif, evolves more rapidly than S’s S2 domain [57,58]. However, even within the S1 domain, some regions are more conserved than others, with the receptors in S1’s N-terminal domain (S1-NTD) evolving more rapidly than those in its C-terminal domain (S1-CTD) [58]. Both S1-NTD and S1-CTD are involved in receptor binding and can function as RBDs to bind proteins and sugars [57], but RBDs in the S1-NTD typically bind to sugars, while those in the S1-CTD recognize protein receptors [47]. Viral receptors show higher affinity with protein receptors than sugar receptors [47], which suggests that positive selection on or relaxed conservation of the S1-NTD might reduce the risk of a deleterious mutation that would prevent binding. The SARS-CoV-2 S protein also contains an RRAR furin recognition site at the S1/S2 junction [51,63], setting it apart from both bat coronavirus RaTG13, with which it shares 96% genome sequence identity, and SARS-CoV-1 [12]. Such furin cleavage sites are commonly found in highly virulent influenza viruses, and as such may contribute to the heightened pathogenicity of SARS-CoV-2 [164,165]. Effective cell entry is a critical component to pathogenesis and therefore an important process to understand when examining possible therapeutics.

2.8 Mechanism of Transmission

Once a human host is infected with a virus, person-to-person viral transmission of a virus can occur through several possible mechanisms. The primary mechanisms associated with respiratory viruses are contact, droplet, and aerosol transmission [166]. Contact transmission can occur through either direct contact with a contagious person or indirect contact with active viral particles on a contaminated surface [167]. This latter mode of transmission is also called fomite transmission [168]. Viral particles can enter the body if they then come in contact with the oral, nasal, eye, or other mucus membranes [167]. Droplet transmission occurs when a contagious individual sneezes, coughs, or exhales and produces respiratory droplets that can contain a large number of viral particles [167]. Contact with these droplets can occur either through direct exposure to the droplets, such as breathing in droplets produced by a sneeze, or exposure to particles that have settled on a surface [167]. Aerosol transmission refers to much smaller particles (less than 5 micrometers) that are also produced by sneezing, coughing, or exhaling [166,167]. The small size of these particles allows them to remain suspended over a longer period of time and potentially to be moved by air currents [167]. Additionally, viral particles deposited on surfaces via large respiratory droplets can also later be aerosolized [167]. Droplet and/or contact transmission are both well-accepted modes of transmission for many viruses associated with common human illnesses, including influenza and rhinovirus [167]. The extent to which aerosol transmission contributes to the spread of respiratory viruses is less clear. In influenza A, for example, viral particles can be detected in aerosols produced by infected individuals, but the extent to which these particles drive the spread of influenza A infection remains under debate [166,167,169,170,171]. Regardless of its role in the spread of influenza A, however, aerosol transmission likely played a role in outbreaks such as the 1918 Spanish Influenza (H1N1) and 2009 “swine flu” (pH1N1) [171]. Contact, droplet, and aerosol transmission are therefore all worth evaluating when considering possible modes of transmission for a respiratory virus like SARS-CoV-2.

2.8.1 Transmission of HCoV

All three of these mechanisms have been identified as contributors to the transmission of HCoV [167], including the highly pathogenic coronaviruses SARS-CoV-1 and MERS-CoV [28,172]. Transmission of SARS-CoV-1 is thought to proceed primarily through droplet transmission, but aerosol transmission is also considered possible [167], and fomite transmission may have also played an important role in some outbreaks [173]. Similarly, the primary mechanism of MERS transmission is thought to be droplets because inter-individual transmission appears to be associated with close interpersonal contact (e.g., household or healthcare settings), but aerosolized particles of the MERS virus have been reported to persist much more robustly than influenza A under a range of environmental conditions [174,175]. While droplet-based and contact transmission were initially considered to be the primary modes by which SARS-CoV-2 spread [176], as additional information has emerged, the possibility of aerosol transmission has also been raised [177,178,179]. For example, the detection of SARS-CoV-2 viral particles in air samples taken from hospitals treating COVID-19 patients led to the concern that the virus could be spreading via aerosols [180]. The stability of the virus both in aerosols and on a variety of surfaces appeared similar to that of SARS-CoV-1 [178], and fomite transmission could also play a role in transmission (e.g., [181]). However, while the possibility of aerosol transmission seems plausible, the evidence suggests that droplet and contact transmission are the dominant mechanisms driving the spread of the virus [182], and the risk of fomite transmission under real-world conditions is likely to be substantially lower than the conditions used for experimental analyses [183]. These mechanisms may differ in their relevance to different types of transmission events, such as transmission within households, nosocomial transmissions, and transmission in indoor versus outdoor spaces.

2.8.2 Symptoms and Viral Spread

Other aspects of pathogenesis are also important to understanding how the virus spreads, especially the relationship between symptoms, viral shedding, and contagiousness. Symptoms associated with reported cases of COVID-19 range from mild to severe [44], but some individuals who contract COVID-19 remain asymptomatic throughout the duration of the illness [184]. The incubation period, or time period between exposure and the onset of symptoms, has been estimated at five to eight days (4.91 and 7.54 in two different cities) [185,186], and estimates suggest that viral shedding may begin long before the onset of symptoms (12.3 days with a 95% CI of 5.9 - 17.0) and peak around the onset of symptoms [187]. As these trends became apparent, concerns arose due to the potential for individuals who did not or did not yet show symptoms to transmit the virus [188]. Recovered individuals may also be able to transmit the virus after their symptoms cease. Estimates of the communicable period based on twenty-four individuals who tested positive for SARS-CoV-2 prior to or without developing symptoms estimated that individuals may be contagious for one to twenty-one days, but they note that this estimate may be low [184]. Initially, viral nucleic acids were reported to remain at observable levels in the respiratory specimens of recovering hospitalized COVID-19 patients for a median of 20 days and with a maximum observed duration through 37 days, when data collection for the study ceased [35]. As more estimates of the duration of viral shedding are released, they are beginning to converge around approximately three weeks from first positive PCR test and/or onset of symptoms (which, if present, are usually identified within three days of the initial PCR test). For example, viral shedding was reported for up to 28 days following symptom onset by a second study [189] and for one to 24 days from first positive PCR test with a median of 12 days [79]. On the other hand, almost 70% of patients were reported to still have symptoms at the time that viral shedding ceased, although all symptoms reduced in prevalence between onset and cessation of viral shedding (CVS) [190]. They also reported that the median time that elapsed between the onset of symptoms and CVS was 23 days and between first positive PCR test and CVS was 17 days [190]. The fact that this study reported symptom onset to predate the first positive PCR test by an average of three days, however, suggests that there may be some methodological differences between it and related studies. Furthermore, an analysis of residents of a nursing home with a known SARS-CoV-2 case measured similar viral load in residents who were asymptomatic regardless of whether they later developed symptoms, and the load in the asymptomatic residents was comparable to that of residents who displayed either typical of atypical symptoms [191]. Taken together, these results suggest that the presence or absence of symptoms are not reliable predictors of viral shedding or of SARS-CoV-2 status (e.g, [192]). However, viral shedding is not necessarily indicative of contagiousness. The risk of spreading the infection was low after ten days from the onset of symptoms, as viral load in sputum was found to be unlikely to pose a significant risk based on their efforts to culture samples in vitro [189].

The extent to which asymptomatic (or presymptomatic) individuals are able to transmit SARS-CoV-2 has been a question of both scientific and community interest. Early reports (February and March 2020) described transmission from presymptomatic SARS-CoV-2-positive individuals to close family contacts [193,194]. One of these reports [194] also included a description of an individual who tested positive for SARS-CoV-2 but never developed symptoms. Later analyses also sought to estimate the proportion of infections that could be traced back to a presymptomatic or asymptomatic individual (e.g., [195]). Estimates of the proportion of individuals with asymptomatic infections have varied widely. The proportion of asymptomatic individuals on board the Diamond Princess cruise ship, which was the site of an early COVID-19 outbreak, was estimated at 17.9% [196]. In contrast, a model using the prevalence of antibodies among residents of Wuhan, China estimated a much higher of asymptomatic cases, at approximately 7 in 8, or 87.5% [197]. An estimate of the population of care homes in London found that, among the residents (median age 85), the rate of asymptomatic infection was 43.8%, and among the caretakers (median age 47), the rate was 49.1% [198]. The duration of viral shedding may also be longer in individuals with asymptomatic cases of COVID-19 compared to those who do show symptoms [199]. As a result, the potential for individuals who do not know they have COVID-19 to spread the virus raises significant concerns. In Singapore and Tianjin, two cities studied to estimate incubation period, an estimated 40-50% and 60-80% of cases, respectively, were estimated to be caused by contact with an asymptomatic individuals [185]. An analysis of viral spread in the Italian town of Vo’, which was the site of an early COVID-19 outbreak, revealed that 42.5% of cases were asymptomatic and that the rate was similar across age groups [200]. They argued that the town’s lockdown was imperative for controlling the spread of COVID-19 because it isolated the asymptomatic individuals. While more models are likely to emerge to better explore the effect of asymptomatic individuals on SARS-CoV-2 transmission, these results suggest that strategies for identifying and containing asymptomatic, but contagious individuals are important for managing community spread.

2.8.3 Estimating the Fatality Rate

Estimating the occurrence of asymptomatic and mild COVID-19 cases is important to identifying the mortality rate associated with COVID-19. The mortality rate of greatest interest would be the total number of fatalities as a fraction of the total number of people infected. One metric reported is often case fatality rate (CFR), which simply compares the number of COVID-19 related deaths to the number of confirmed or suspected cases. However, in locations without universal testing protocols, it is impossible to identify all exposed or all infected individuals because so many asymptomatic or mild cases go undetected. Therefore, a more informative metric is the infection fatality rate (IFR), which compares the known deaths to the estimated number of cases. Meta-analyses have produced estimates of global IFR ranging from as low as 0.1% to as high as 1.04% [201,201,202,203], and this estimate was also supported by a repeated cross-sectional serosurvey conducted in New York City that revealed an estimated IFR of 0.97% [204]. All of these estimates note that IFR varies widely around the world. Estimates of infection rates are becoming more feasible as more data becomes available for modeling and will be bolstered as serological testing becomes more common and available.

2.9 Dynamics of Transmission

Disease spread dynamics can be estimated using R0, the basic reproduction number, and Rt, the effective reproduction number. Thus, accurate estimates are crucial to understanding the dynamics of infection and to predicting the effects of different interventions. R0 and the timescale of infection (measured by the infectious period and the exposed period) govern population-level epidemic dynamics, with R0 being one of the most important epidemiological parameters [205]. R0 is the average number of new (secondary) infections caused by one infected person, assuming a wholly susceptible population [206]. A simple mechanistic model used to describe infectious disease dynamics is a susceptible-infected-recovered (SIR) compartmental model. In this formulation individuals move through three states: susceptible, infected, and recovered; two parameters, \(\gamma\) and \(\beta\), specify the rate at which the infectious recover, and the infection transmission rate, respectively. In this simple formulation, R0 is estimated as the ratio of \(\beta\) and \(\gamma\).[205,207]. A pathogen can invade a susceptible population only if R0 > 1 [205,208]. The spread of an infectious disease at a particular time t can be quantified by Rt, the effective reproduction number, which assumes that part of the population has already recovered (and thus gained immunity to reinfection) or that mitigating interventions were put into place. For example, if only a fraction St of the population is still susceptible, Rt = St x R0. When Rt is greater than 1, an epidemic grows (i.e., the proportion of the population that is infectious increases); when Rt is less than 1, the proportion of the population that is infectious decreases. R0 and Rt can be estimated directly from epidemiological data or inferred using mathematical modeling. Modeling approaches are typically based upon a classic epidemiological model structure: the SIR model and its extensions [209,210]. In the context of SARS-CoV-2, more complex modified susceptible-exposed-infectious-recovered (SEIR) models are commonly used.

Estimates of R0 for COVID-19 lie in the range R0=1.4-6.5 [211,212,213]. Variation in R0 is expected between different populations, and the estimated values of R0 discussed below are for specific populations in specific environments. The different estimates of R0 should not necessarily be interpreted as a range of estimates of the same underlying parameter. In one study of international cases, the predicted value was R0=1.7 [214]. In China (both Hubei province and nationwide), the value was predicted to lie in the range R0=2.0-3.6 [211,215,216]. Another estimate based on a cruise ship where an outbreak occurred predicted R0=2.28 [217]. SEIR model-derived estimates of R0 range from 2.0 - 6.5 in China [218,219,220,221] to R0=4.8 in France [222]. Using the same model as for the French population, a study estimated R0=2.6 in South Korea [222], which is consistent with other studies [223]. From a meta-analysis of studies estimating R0, [212] predict the median as R0=2.79.

Inference of the effective reproduction number can provide insight into how populations respond to an infection and the effectiveness of interventions. In China, Rt was predicted to lie in the range 1.6-2.6 in January 2020, before travel restrictions [224]. Rt decreased from 2.35 one week before travel restrictions were imposed (January 23, 2020), to 1.05 one week after. Using their model, the authors also estimated the probability of new outbreaks occurring: the probability of a single individual exporting the virus and causing a large outbreak is 17-25%, assuming MERS-like or SARS-like transmission, and the probability of a large outbreak occurring after ≥4 infections exist at a new location is greater than 50%. An independent study came to similar conclusions, finding Rt=2.38 in the two-week period before January 23 with a decrease to Rt = 1.34 (using data from January 24 to February 3) or Rt=0.98 (using data from January 24 to February 8) [213]. In South Korea, Rt was inferred for February through March 2020 in two cities, Daegu (the center of the outbreak) and Seoul [223]. Metro data was also analyzed to estimate the effects of social distancing measures. Rt decreased in Daegu from around 3 to <1 over the period that social distancing measures were introduced. In Seoul, Rt decreased slightly, but remained close to 1 (and larger than Rt in Daegu). These findings indicate that social distancing measures appeared to be effective in containing the infection in Daegu, but in Seoul, Rt remained above 1, meaning secondary outbreaks remained possible. The study also shows the importance of region-specific analysis: the large decline in case load nationwide was mainly due to the Daegu region and could mask persistence of the epidemic in other regions, such as Seoul and Gyeonggi-do. In Iran, estimates of Rt declined from 4.86 in the first week to 2.1 by the fourth week after the first cases were reported [225]. In Europe, analysis of 11 countries inferred the dynamics of Rt over a time range from the beginning of the outbreak until March 28, 2020, by which point most countries had implemented major interventions (such as school closures, public gathering bans, and stay-at-home orders) [226]. Across all countries, the mean Rt before interventions began was estimated as 3.87; Rt varied considerably, from below 3 in Norway to above 4.5 in Spain. After interventions, Rt decreased by an average of 64% across all countries, with mean Rt=1.43. The lowest predicted value was 0.97 for Norway and the highest was 2.64 for Sweden, which may be in part because Sweden did not implement social distancing measures on the same scale as other countries. The study concludes that while large changes in Rt are observed, it is too early to tell whether the interventions put into place are sufficient to decrease Rt below 1.

More generally, population-level epidemic dynamics can be both observed and modelled [207]. Data and empirically determined biological mechanisms inform models, while models can be used to try to understand data and systems of interest or to make predictions about possible future dynamics, such as the estimation of capacity needs [227] or the comparison of predicted outcomes among prevention and control strategies [228,229]. Many current efforts to model Rt have led to tools that assist the visualization of estimates in real time or over recent intervals [230,231]. While these may be valuable resources, it is important to note that the estimates arise from models containing many assumptions and are dependent on the quality of the data they use, which varies widely by region.

2.10 Molecular Signatures and Transmission

Genetic variation in SARS-CoV-2 has been used to elucidate patterns over time and space. Mutations observed in individual SARS-CoV-2 genome sequences can be used to trace transmission patterns and have provided insights during outbreak investigations [14,232,233]. Similar mutations observed in several patients may indicate that the patients belong to the same transmission group. The tracking of SARS-CoV-2 mutations is recognized as an essential tool for controlling outbreaks that may facilitate tracing the paths of SARS-CoV-2’s spread [234]. Several studies used phylogenetic analysis to determine the source of local COVID-19 outbreaks in Connecticut (USA), [235], the New York City area (USA) [236], and Iceland [237]. There is an ongoing effort to collect SARS-CoV-2 genomes throughout the COVID-19 outbreak, and as of January 18, 2021 more than 381,000 genome sequences have been collected from patients. The sequencing data can be found at GISAID [238], NCBI [239], and COVID-19 data portal [240].

2.11 Conclusions

The novel coronavirus SARS-CoV-2 is the third HCoV to emerge in the 21st century, and research into previous HCoVs has provided a strong foundation for characterizing the pathogenesis and transmission of SARS-CoV-2. Critical insights into how the virus interacts with human cells have been gained from previous research in HCoV and other viral infections. As with other HCoV, the immune response to SARS-CoV-2 is likely driven by detection of its spike protein, which allows it to enter cells through the ACE2 receptor. Epithelial cells have also emerged as the major cellular target of the virus, contextualizing the respiratory and gastrointestinal symptoms that are frequently observed in COVID-19. Many of the mechanisms that facilitate the pathogenesis of SARS-CoV-2 are currently under consideration as possible targets for the treatment or prevention of COVID-19. Research in other viruses also provides a foundation for understanding the transmission of SARS-CoV-2 among people and can therefore inform efforts to control the virus’s spread. Though it remains a question whether aerosol and fomite transmission contribute to the spread of SARS-CoV-2 in real-world settings, in general, much like SARS-CoV-1 and MERS-CoV, this virus seems to be spread primarily by droplet transmission. Asymptomatic transmission was also a concern in the SARS outbreak of 2002-03 and, as in current pandemic, presented challenges for estimating rates of infection [241]. However, in 2021, we are fortunate to be able to build on top of 18 years of SARS-CoV research in order to rapidly ascertain the identity and behavior of the virus.

COVID-19 is a complex disease with many possible presentations that appear to vary across the lifespan. Variability in presentation, including cases with no respiratory symptoms or with no symptoms altogether, were also reported during the SARS epidemic at the beginning of the 21st century [241]. The variability of both which symptoms present and their severity have presented challenges for public health agencies seeking to provide clear recommendations regarding which symptoms indicate SARS-CoV-2 infection and should prompt isolation. Additionally, asymptomatic cases add complexity both to efforts to estimate critical statistics such as R0 and Rt, which are critical to understanding the transmission of the virus, and IFR, which is an important component of understanding its impact on a given population. The development of diagnostic technologies over the course of the pandemic have facilitated more accurate identification, including of asymptomatic cases. As more cases have been diagnosed, the health conditions and patient characteristics associated with more severe infection have also become more clear, although there are likely to be significant sociocultural elements that also influence these outcomes. While many efforts have focused on adults, and especially older adults because of the susceptibility of this demographic, additional research is needed to understand the presentation of COVID-19 and MIS-C in pediatric patients. As more information is uncovered about the pathogenesis of HCoV and SARS-CoV-2 specifically, the diverse symptomology of COVID-19 has and likely will continue to conform with the ever-broadening understanding of how SARS-CoV-2 functions within a human host.

While the SARS-CoV-2 virus is very similar to other HCoV in several ways, including in its genomic structure and the structure of the virus itself, there are also some differences that may account for differences in the COVID-19 pandemic compared to the SARS and MERS of the past two decades. The R0 of SARS-CoV-2 has been estimated to be similar to SARS-CoV-1 but much higher than that of MERS-CoV [30,30]. While the structures of the viruses are very similar, evolution among these species may account for differences in their transmissibility and virulence. For example, the acquisition of a furin cleavage site the S1/S2 boundary within the SARS-CoV-2 S protein may be associated with increased virulence. Additionally, concerns have been raised about the accumulation of mutations within the SARS-CoV-2 species itself, and whether these could influence virulence. The coming of age of genomic technologies has made these types of analyses feasible, and genomics research characterizing changes in SARS-CoV-2 along with temporal and spatial movement is likely to provide additional insights into whether within-species evolution influences the effect of the virus on the human host.

Additionally, omics technologies have made it feasible to characterize the host response to the virus. Although at present, much of the omics research has utilized in vitro models, such systems-level approaches represent a promising opportunity to characterize the host response to the SARS-CoV-2 virus. For example, analysis of PPI using publicly available data identified some similarities between COVID-19 and tuberculosis infection [153]. This finding suggests that insights into COVID-19 may be gained by systems-level analysis of a wide array of infections, even those that are not causes by viruses, as is the case with the bacterium M. tuberculosis. As more data is collected, large-scale, omics-based analyses will become more feasible.

Thus, though the COVID-19 crisis is still evolving, the insights acquired over the past 20 years of HCoV research have provided a solid foundation for understanding the SARS-CoV-2 virus and the disease it causes. As the scientific community continues to respond to COVID-19 and to elucidate more of the relationships between viral pathogenesis, transmission, and symptomology, and as more data about the regulatory shifts associated with COVID-19 becomes available, this understanding will no doubt continue to develop to reveal additional interdisciplinary connections among virology, pathogenesis, and health. As additional information becomes available, this review will be updated to reflect the changing state of research in this area. At present, understanding the SARS-CoV-2 virus and its pathogenesis is critical to a holistic understanding of the COVID-19 pandemic.

3 Evolutionary and Genomic Analysis of SARS-CoV-2

3.1 Abstract

3.2 Introduction

3.3 Initial Characterization of SARS-CoV-2

The first genome sequence of the virus was released on January 3, 2020 and revealed that the cluster of pneumonia cases seen in Wuhan were caused by a novel coronavirus [10]. Multiple research groups have drafted the genome sequence of SARS-CoV-2 based on sequences developed from clinical samples collected from the lower respiratory tract, namely bronchoalveolar lavage fluid (BALF), and the upper respiratory tract, in the form of throat swabs [11,12,13]. Analysis of the SARS-CoV-2 genome revealed significant sequence homology with two coronaviruses known to infect humans, with about 79% identity to SARS-CoV-1 and 50% to MERS-CoV [13]. However, in this analysis, the highest degree of similarity was observed between SARS-CoV-2 and bat-derived SARS-like coronaviruses (bat-SL-CoVZC45 and bat-SL-CoVZXC21) [13]. Other analyses have reported even greater similarity between SARS-CoV-2 and the bat coronavirus BatCoV-RaTG13, with identity as high as 96.2% [12,14], and the closely related pangolin coronavirus [15]. This evidence therefore suggests the SARS-CoV-2 virus is the result of zoonotic transfer of a virus from bats to humans. Nevertheless, some fragments between SARS-CoV-2 and RATG13 differ by up to 17%, suggesting a complex natural selection process during zoonotic transfer. While the S region is highly similar to that of viruses found in pangolins [15], there is no consensus about the origin of S in SARS-CoV-2, as it could potentially be the result either of recombination or coevolution [14,16]. Though the intermediate host serving as the source for the zoonotic introduction of SARS-CoV-2 to human populations has not yet been identified, the SARS-CoV-2 virus has been placed within the coronavirus phylogeny through comparative genomic analyses. Genomic analyses and comparisons to other known coronaviruses suggest that SARS-CoV-2 is unlikely to have originated in a laboratory – either purposely engineered and released, or escaped – and instead evolved naturally in an animal host [17]. While the position of the SARS-CoV-2 virus within the coronavirus phylogeny has been largely resolved, the functional consequences of molecular variation between this virus and other viruses, such as its bat and pangolin sister taxa or SARS-CoV-1, remain unknown [17]. Fortunately, the basic genome structure of coronaviruses is highly conserved, and insight into the mechanisms the virus uses to enter cells, replicate, and spread is likely to be available from prior research in coronaviruses.

Additionally, worldwide sequencing of viral samples has provided some preliminary insights into possible mechanisms of adaptation in the virus, and omics-based analysis of patient samples has elucidated some of the biological changes the virus induces in its human hosts.

3.4 Coronaviruses and Animal Hosts

Coronaviruses have long been known to infect animals and have been the subject of veterinary medical investigations and vaccine development efforts due to their effect on the health of companion and agricultural animals [18].

Most coronaviruses show little to no transmission in humans. However, today it is thought that approximately one-third of common cold infections are caused by four human coronaviruses (HCoV): Human coronavirus 229E (HCoV-229E), Human coronavirus NL63 (HCoV-NL63), Human coronavirus OC43 (HCoV-OC43), and Human coronavirus HKU1 (HCoV-HKU1) [19,20]. The first HCoV were identified in the 1960s: HCoV-229E in 1965 [21] and HCoV-OC43 in 1967 [22]. Both of these viruses typically cause cold-like symptoms, including upper and lower respiratory infections [23,24,242], but they have also been associated with gastrointestinal symptoms [243]. Two additional HCoV were subsequently identified [25,26]. In 2003, HCoV-NL63 [25] was first identified in a 7-month-old infant and then in clinical specimens collected from seven additional patients, five of whom were infants younger than 1 year old and the remainder of whom were adults. CoV-HKU1 was identified in samples collected from a 71-year-old pneumonia patient in 2004 and then found in samples collected from a second adult patient [26]. These viruses are associated with respiratory diseases of varying severity, ranging from common cold to severe pneumonia, with severe symptoms mostly observed in immunocompromised individuals [27], and also have gastrointestinal involvement in some cases [243]. In addition to these relatively mild HCoV, however, highly pathogenic human coronaviruses have been identified, including Severe acute respiratory syndrome-related coronavirus (SARS-CoV or SARS-CoV-1) and Middle East respiratory syndrome-related coronavirus (MERS-CoV) [19,28,29].

At the time that SARS-CoV-1 emerged in the early 2000s, no HCoV had been identified in almost 40 years [28]. The first case of SARS was reported in November 2002 in the Guangdong Province of China, and over the following month, the disease spread more widely within China and then into several countries across multiple continents [28,30]. Unlike previously identified HCoV, SARS was much more severe, with an estimated death rate of 9.5% [30]. It was also highly contagious via droplet transmission, with a basic reproduction number (R0) of 4 (i.e., each person infected was estimated to infect four other people) [30]. However, the identity of the virus behind the infection remained unknown until April of 2003, when the SARS-CoV-1 virus was identified through a worldwide scientific effort spearheaded by the WHO [28]. SARS-CoV-1 belonged to a distinct lineage from the two other HCoV known at the time [30]. By July 2003, the SARS outbreak was officially determined to be under control, with the success credited to infection management practices [28]. A decade later, a second outbreak of severe respiratory illness associated with a coronavirus emerged, this time in the Arabian Peninsula. This disease, known as Middle East respiratory syndrome (MERS), was linked to another novel coronavirus, MERS-CoV. The fatality rate associated with MERS is much higher than that of SARS, at almost 35%, but the disease is much less easily transmitted, with an R0 of 1 [30]. Although MERS is still circulating, its low reproduction number has allowed for its spread to be contained [30]. The COVID-19 pandemic is thus associated with the seventh HCoV to be identified and the fifth since the turn of the millennium, though additional HCoVs may be in circulation but remain undetected.

SARS-CoV-1 and MERS-CoV were ultimately managed largely through infection management practices (e.g., mask wearing) and properties of the virus itself (i.e., low rate of transmission), respectively [28,30]. Vaccines were not used to control either virus, although vaccine development programs were established for SARS-CoV-1 [31]. In general, care for SARS and MERS patients focuses on supportive care and symptom management [30]. Clinical treatments for SARS and MERS developed during the outbreaks generally do not have strong evidence supporting their use. Common treatments included Ribavirin, an antiviral, often in combination with corticosteroids or sometimes interferon (IFN) medications, which would both be expected to have immunomodulatory effects [28]. However, retrospective and in vitro analyses have reported inconclusive results of these treatments on SARS and the SARS-CoV-1 virus, respectively [28]. IFNs and Ribavirin have shown promise in in vitro analyses of MERS, but their clinical effectiveness remains unknown [28]. Therefore, only limited strategy for the pharmaceutical management of COVID-19 can be adopted from previous severe HCoV infections. Research in response to prior outbreaks of HCoV-borne infections, such as SARS and MERS, have, however, provided a strong foundation for hypotheses about the pathogenesis of SARS-CoV-2 as well as potential diagnostic and therapeutic approaches.

3.5 Zoonotic Transfer of Coronaviruses

3.6 Evolution of the SARS-CoV-2 Virus

3.6.1 Emergence of SARS-CoV-2

3.6.2 Evolution of SARS-CoV-2 Variants

Evolution in SARS-CoV-2 has also been observed over a short timescale. After zoonotic transfer, SARS-CoV-2 continued evolving in the human population [232]. The SARS-CoV-2 mutation rate is moderate compared to other RNA viruses [233], which likely restricts the pace of evolution in SARS-CoV-2. Nevertheless, genomic analyses have yielded statistical evidence of ongoing evolution. There are two known variants of the spike protein that differ by a single amino acid at position 614 (G614 and D614), and there is evidence that G614 had become more prevalent than D614 by June 2020 [244]. While there is a hypothesis that this genomic change increased the SARS-CoV-2 infectivity and virulence, this hypothesis has not yet been tested due to a lack of data [245]. Another study [233] identified 198 recurrent mutations in a dataset of 7,666 curated sequences. This pattern of convergent evolution at some sites could indicate that certain mutations confer an adaptive advantage. While it is evident that SARS-CoV-2 exhibits moderate potential for ongoing and future evolution, the relationship between mutations and pathogenicity is not yet known. Additional data is needed in order to understand patterns of evolutionary change and whether they are likely to affect virulence.

3.7 Conclusions

As of October 2020 the SARS-CoV-2 virus remains a serious worldwide threat. The scientific community has responded by rapidly collecting and disseminating information about the SARS-CoV-2 virus and the associated illness, COVID-19. The rapid identification of the genomic sequence of the virus allowed for early contextualization of SARS-CoV-2 among other known respiratory viruses. The pathogen is a coronavirus that is closely related to SARS-CoV-1, which caused the SARS pandemics of the early 2000s. Knowing the phylogenetic context and genomic sequence of the virus then allowed for rapid insights into its structure and pathogenesis.

4 Diagnostics

4.1 Abstract

4.2 Introduction

Identifying individuals who have contracted COVID-19 is crucial to slowing down the global pandemic. Given the high transmissibility of SARS-CoV-2, the development of reliable assays to detect SARS-CoV-2 infection even in asymptomatic carriers is vitally important. For instance, the deployment of wide-scale diagnostic testing followed by the isolation of infected people has been a key factor in South Korea’s successful strategy for controlling the spread of the virus. Following the first release of the genetic sequence of the virus by Chinese officials on January 10, 2020, the first test was released about 13 days later [246]. Diagnostic approaches utilizing a variety of methods are currently or have been developed. There are two main classes of diagnostic tests: molecular tests, which can diagnose an active infection by identifying the presence of SARS-CoV-2, and serological tests, which can assess whether an individual was infected in the past via the presence or absence of antibodies against SARS-CoV-2. Molecular tests are essential for identifying individuals for treatment and alerting their contacts to quarantine and be alert for possible symptoms. While serological tests may be of interest to individuals who wish to confirm they were infected with SARS-CoV-2 in the past, their potential for false positives means that they are not currently recommended for this use. However, serological tests provide population-level information for epidemiological analysis, as they can be used to estimate the extent of the infection in a given area. Thus, they may be useful in efforts to better understand the percent of cases that manifest as severe versus mild and for guiding public health and economic decisions regarding resource allocation and counter-disease measures.

Understanding the fundamental organization of the human immune response to viral threats is critical to understanding the varied response to SARS-CoV-2. The human immune system utilizes a variety of innate and adaptive responses to protect against the pathogens it encounters. The innate immune system consists of barriers, such as the skin, mucous secretions, neutrophils, macrophages, and dendritic cells. It also includes cell-surface receptors that can recognize the molecular patterns of pathogens. The adaptive immune system utilizes antigen-specific receptors that are expressed on B and T lymphocytes. These components of the immune system typically act together; the innate response acts first, and the adaptive response begins to act several days after initial infection following the clonal expansion of T and B cells [32]. After a virus enters into a host cell, its antigen is presented by major histocompatibility complex 1 (MHC 1) molecules and is then recognized by cytotoxic T lymphocytes.

In the case of COVID-19, there is also concern about the immune system becoming over-active. One of the main immune responses contributing to the onset of acute respiratory distress syndrome (ARDS) in COVID-19 patients is cytokine storm syndrome (CSS), which causes an extreme inflammatory response due to a release of pro-inflammatory cytokines and chemokines by immune effector cells. In addition to respiratory distress, this mechanism can lead to organ failure and death in severe COVID-19 cases [33].

4.3 Molecular Tests

Molecular tests are used to identify distinct genomic subsequences of a viral molecule in a sample and thus to diagnose an active viral infection. This first requires identifying which biospecimens are likely to contain the virus during infection and then acquiring these samples from the patient(s) to be tested. Common sampling sources for molecular tests include nasopharyngeal cavity samples, such as throat wash and saliva [247], and stool samples [248]. Once a sample is acquired from a patient, molecular testing will utilize a number of steps, described below, to analyze a sample and identify whether evidence of SARS-CoV-2 is present. When testing for RNA viruses like SARS-CoV-2, pre-processing is necessary in order to create DNA from the RNA sample. The DNA can then be amplified with PCR. Some tests use the results of the PCR itself to determine whether the pathogen is present, but in other cases, it may be necessary to sequence the amplified DNA. Sequencing requires an additional pre-processing step: library preparation. Library preparation is the process of preparing the sample for sequencing, typically by fragmenting the sequences and adding adapters [249]. In some cases, library preparation can involve other modifications of the sample, such as adding “barcoding” to identify a particular sample in the sequence data, which is useful for pooling samples from multiple sources. There are different reagents used for library preparation that are specific to identifying one or more target sections with PCR [250]. Sequential pattern matching is then used to identify unique subsequences of the virus that identify it in specific. If sufficient subsequences are found, the test is considered positive.

4.3.1 RT-PCR

Real-Time Polymerase Chain Reaction (RT-PCR) tests determine whether a target is present by measuring the rate of amplification during PCR compared to a standard. When the target is RNA, such as in the case of RNA viruses, the RNA must be converted into complementary DNA during pre-processing. The Drosten Lab, from Germany, was the first lab to establish and validate a diagnostic test to detect SARS-CoV-2. This test uses RT-PCR with reverse transcription [246] to detect several regions of the viral genome: the ORF1b of the RNA-dependent RNA polymerase (RdRP), the Envelope protein gene (E), and the Nucleocapsid protein gene (N). In collaboration with several other labs in Europe and in China, the researchers confirmed the specificity of this test with respect to other coronaviruses against specimens from 297 patients infected with a broad range of respiratory agents. Specifically this test utilizes two probes against RdRP, one of which is specific to SARS-CoV-2 [246]. Importantly, this assay was not found to return false positive results.

4.3.2 qRT-PCR

Chinese researchers developed a quantitative real-time reverse transcription PCR (qRT-PCR) test to identify two gene regions of the viral genome, ORF1b and N [251]. This assay was tested on samples coming two COVID-19 patients and a panel of positive and negative controls consisting of RNA extracted from several cultured viruses. The assay uses the N gene to screen patients, while the ORF1b gene region is used to confirm the infection [251]. In this case the test was designed to detect sequences conserved across sarbecoviruses, or viruses within the same subgenus as SARS-CoV-2. Considering that SARS-CoV-1 and SARS-CoV-2 are the only sarbecoviruses currently known to infect humans, a positive test can be assumed to indicate that the patient is infected with SARS-CoV-2. However, this test is not able to discriminate the genetics of viruses within the sarbecovirus clade.

4.3.2.1 dPCR

Digital PCR (dPCR) is a new generation of PCR technologies offering an alternative to traditional real-time quantitative PCR. In dPCR, a sample is partitioned into thousands of compartments, such as nanodroplets (droplet dPCR or ddPCR) or nanowells, and a PCR reaction takes place in each compartment. This design allows for a digital read-out where each partition is either positive or negative for the nucleic acid sequence being tested for, allowing for much higher throughput. While dPCR equipment is not yet as common as that for RT-PCR, dPCR for DNA targets generally achieves higher sensitivity than other PCR technologies while maintaining high specificity, though sensitivity is slightly lower for RNA targets [252]. High sensitivity is particularly relevant for SARS-CoV-2 detection, since low viral load in clinical samples can lead to false negatives. Suo et al. [253] performed a double-blind evaluation of ddPCR for SARS-CoV-2 detection on 57 samples, comprised by 43 samples from suspected positive patients and 14 from supposed convalescents, that had all tested negative for SARS-CoV-2 using RT-PCR. Despite the initial negative results, 33 out of 35 (94.3%) patients were later clinically confirmed positive. All of these individuals tested positive using ddPCR. Additionally, of 14 supposed convalescents who had received two consecutive negative RT-PCR tests, nine (64.2%) tested positive for SARS-CoV-2 using ddPCR. Two symptomatic patients tested negative with both RT-PCR and ddPCR, but were later clinically diagnosed positive, and 5 of the 14 suspected convalescents tested negative by ddPCR. While this study did not provide a complete head-to-head comparison to RT-PCR in all aspects, e.g., no samples testing positive using RT-PCR were evaluated by ddPCR, the study shows the potential of dPCR for viral detection even in highly diluted samples. In a second study, Dong et al. [254] compared the results of qRT-PCR and ddPCR testing for SARS-CoV-2 in 194 samples, including 103 samples from suspected patients, 75 from contacts and close contacts, and 16 from suspected convalescents. Of the 103 suspected patient samples, 29 were reported as positive, 25 as negative, and 49 as suspected by qRT-PCR; all patients were later confirmed to be SARS-CoV-2 positive. Of the qRT-PCR negative or suspected samples, a total of 61 (17 negative and 44 suspected) were later confirmed to be positive by ddPCR, improving the overall detection rate among these patients from 28.2% to 87.4%. Of 75 patient samples from contacts and close contacts, 48 tested negative with both methods, and these patients were observed to remain healthy. Within the remaining 27 patient samples, 10 tested positive, 1 negative, and 16 suspect with qRT-PCR. Fifteen out of 16 suspect samples and the negative test results were overturned by ddPCR, decreasing the rate of suspect cases from 21% to 1%. Importantly, all samples that tested positive using qRT-PCR also tested positive using ddPCR. Among the 16 convalescent patients, qRT-PCR identified 12 as positive, three as suspect, and one as negative, but RT-dPCR identified all 16 as positive. This evidence further indicates that the lower limit of detection made possible by ddPCR may be useful for identifying when COVID-19 patients are cleared of the virus. Overall, these studies suggest that dPCR is a promising tool for overcoming the problem of false-negative SARS-CoV-2 testing.

4.3.3 Pooled and Automated PCR Testing

Due to limited supplies and the need for more tests, several labs have found ways to pool or otherwise strategically design tests to increase throughput. The first such result came from Yelin et al. [255], who found they could pool up to 32 samples in a single qPCR run. This was followed by larger-scale pooling with slightly different methods [256]. Although these approaches are also PCR based, they allow for more rapid scaling and higher efficiency for testing than the initial PCR-based methods developed.

4.3.3.1 CRISPR-based Detection

Two CRISPR-associated nucleases, Cas12 and Cas13, have been used for nucleic acid detection. Multiple assays exploiting these nucleases have emerged as potential diagnostic tools for the rapid detection of SARS-CoV-2 genetic material and therefore SARS-CoV-2 infection. The SHERLOCK method (Specific High-sensitivity Enzymatic Reporter unLOCKing) from Sherlock Biosciences relies on Cas13a to discriminate between inputs that differ by a single nucleotide at very low concentrations [257]. The target RNA is amplified by RT-RPA and T7 transcription, and the amplified product activates Cas13a. The nuclease then cleaves a reporter RNA, which liberates a fluorescent dye from a quencher. Several groups have used the SHERLOCK method to detect SARS-CoV-2 viral RNA. An early study reported that the method could detect 7.5 copies of viral RNA in all 10 replicates, 2.5 copies in 6 out of 10, and 1.25 copies in 2 out of 10 runs [258]. It also reported 100% specificity and sensitivity on 114 RNA samples from clinical respiratory samples (61 suspected cases, among which 52 were confirmed and nine were ruled out by metagenomic next-generation sequencing, 17 nCoV-/HCoV+ cases and 36 samples from healthy subjects), and a reaction turnaround time of 40 minutes. A separate study screened four designs of SHERLOCK and extensively tested the best-performing assay. They determined the limit of detection to be 10 copies/μl using both fluorescent and lateral flow detection [259]. Lateral flow test strips are simple to use and read, but there are limitations in terms of availability and cost per test. Another group therefore proposed the CREST protocol (Cas13-based, Rugged, Equitable, Scalable Testing), which uses a P51 cardboard fluorescence visualizer, powered by a 9V battery, for the detection of Cas13 activity instead of immunochromatography [260]. CREST can be run, from RNA sample to result, with no need for AC power or a dedicated facility, with minimal handling in approximately 2 hours. Testing was performed on 14 nasopharyngeal swabs. CREST picked up the same positives as the CDC-recommended TaqMan assay with the exception of one borderline sample that displayed low-quality RNA.

The DETECTR method (DNA Endonuclease-Targeted CRISPR Trans Reporter) from Mammoth Biosciences involves purification of RNA extracted from patient specimens, amplification of extracted RNAs by loop-mediated amplification, which is a rapid, isothermal nucleic acid amplification technique, and application of their Cas12-based technology. In their assay, guide RNAs were designed to recognize portions of sequences corresponding to the SARS-CoV-2 genome, specifically the N2 and E regions [261]. In the presence of SARS-CoV-2 genetic material, sequence recognition by the guide RNAs results in double-stranded DNA cleavage by Cas12, as well as cleavage of a single-stranded DNA molecular beacon. The cleavage of this molecular beacon acts as a colorimetric reporter that is subsequently read out in a lateral flow assay and indicates the positive presence of SARS-CoV-2 genetic material and therefore SARS-CoV-2 infection. The 40-minute assay is considered positive if there is detection of both the E and N genes or presumptive positive if there is detection of either of them. The assay had 95% positive predictive agreement and 100% negative predictive agreement with the US Centers for Disease Control and Prevention SARS-CoV-2 real-time RT–PCR assay. The estimated limit of detection was 10 copies per μl reaction, versus 1 copy per μl reaction for the CDC assay. These results have been confirmed by other DETECTR approaches. Using RTRPA for amplification, another group detected 10 copies of synthetic SARS-CoV-2 RNA per μl of input within 60 minutes of RNA sample preparation in a proof-of-principle evaluation [262]. The DETECTR protocol was improved by combining RT-RPA and CRISPR-based detection in a one-pot reaction that incubates at a single temperature, and by using dual crRNAs (which increases sensitivity). This new assay, known as All-In-One Dual CRISPR-Cas12a (AIOD-CRISPR), detected 4.6 copies of SARS-CoV-2 RNA per μl of input in 40 minutes [263]. Another single-tube, constant-temperature approach using Cas12b instead of Cas12a achieved a detection limit of 5 copies/μl in 40-60 minutes [264]. It was also reported that that electric field gradients can be used to control and accelerate CRISPR assays by co-focusing Cas12-gRNA, reporters, and target [265]. The authors generated an appropriate electric field gradient using a selective ionic focusing technique known as isotachophoresis (ITP) implemented on a microfluidic chip. They also used ITP for automated purification of target RNA from raw nasopharyngeal swab samples. Combining this ITP purification with loop-mediated isothermal amplification, their ITP-enhanced assay to achieved detection of SARS-CoV-2 RNA (from raw sample to result) in 30 minutes.

There is an increasing body of evidence that CRISPR-based assays offer a practical solution for rapid, low-barrier testing in areas that are at greater risk of infection, such as airports and local community hospitals. In the largest study to date, DETECTR was compared to qRT-PCR on 378 patient samples [266]. The authors reported a 95% reproducibility. Both techniques were equally sensitive in detecting SARS-CoV-2. Lateral flow strips showed a 100% correlation to the high-throughput DETECTR assay. Importantly, DETECTR was 100% specific for SARS-CoV-2 and did not detect other human coronaviruses.

4.3.4 Limitations of Molecular Tests

Tests that identify SARS-CoV-2 using nucleic-acid-based technologies will identify only individuals with current infections and are not appropriate for identifying individuals who have recovered from a previous infection. Within this category, different types of tests have different limitations. For example, PCR-based test can be highly sensitive, but in high-throughput settings they can show several problems:

  1. False-negative responses, which can present a significant problem to large-scale testing. To reduce occurrence of false negatives, correct execution of the analysis is crucial [267].
  2. Uncertainty surrounding the SARS-CoV-2 viral shedding kinetics, which could affect the result of a test depending on when it was taken [267].
  3. Type of specimen, as it is not clear which clinical samples are best for detecting the virus [267].
  4. Expensive machinery, which might be present in major hospitals and/or diagnostic centers but is often not available to smaller facilities [268].
  5. Timing of the test, which might take up to 4 days to return results [268].
  6. The availability of supplies for testing, including swabs and testing media, has been limited [269].
  7. Because the guide RNA can recognize other interspersed sequences on the patient’s genome, false positives and a loss of specificity can occur.

Similarly, in tests that use CRISPR, false positives can occur due to the specificity of the technique, as the guide RNA can recognize other interspersed sequences on the patient’s genome. As noted above, false negatives are a significant concern for several reason. Importantly, clinical reports indicate that it is imperative to exercise caution when interpreting the results of molecular tests for SARS-CoV-2 because negative results do not necessarily mean a patient is virus-free [270].

4.4 Serological Tests

Although diagnostic tests based on the detection of genetic material can be quite sensitive, they cannot provide information about the extent of the disease over time. Most importantly, they would not work on a patient who has fully recovered from the virus at the time of sample collection. In this context, serological tests, which use serum to test for the presence of antibodies against SARS-CoV-2, are significantly more informative. Additionally, serological tests can help scientists to understand why the disease has a different course among patients, as well as which strategies might work to manage the spread of the infection. Furthermore, serological tests hold significant interest because of the possibility that they could provide information relevant to advancing economic recovery and allowing reopenings. For instance, people that had developed antibodies might be able to return to work, assuming (still-unproven) protective immunity [271]. Some infectious agents can be controlled through “herd immunity”, which is when a critical mass within the population acquires immunity through vaccination and/or infection, preventing an infectious agent from spreading widely. A simple SIR model predicts that to achieve the required level of exposure for herd immunity to be effective, at least (1-(1/R0)) fraction of the population must be immune or, equivalently, less than (1/R0) fraction of the population susceptible [208]. However, for SARS-CoV-2 and COVID-19, the R0 and mortality rates that have been observed suggest that relying on herd immunity without some combination of vaccines, proven treatment options, and strong non-pharmaceutical measures of prevention and control would likely result in a significant loss of life.

4.4.1 Sustained Immunity to COVID-19

In the process of mounting a response to a pathogen, the immune system produces antibodies specific to the pathogen. Understanding the acquisition and retention of antibodies is important both to the diagnosis of prior (inactive) infections and to the development of vaccines. The two immunoglobulin classes most pertinent to these goals are immunoglobulin M (IgM), which are the first antibodies produced in response to an infection, and immunoglobulin G (IgG), which are the most abundant antibodies found in the blood. Following SARS infection, IgM and IgG antibodies were detected in the second week post-infection. IgM titers peaked by the first month post-infection, and then declined to undetectable levels after day 180. IgG titers peaked by day 60, and persisted in all donors through the two-year duration of study [272]. A two-year longitudinal study following convalesced SARS patients with a mean age of 29 found that IgG antibodies were detectable in all 56 patients surveyed for at least 16 months, and remained detectable in all but 4 (11.8%) of patients through the full two-year study period [273]. These results suggest that immunity to SARS-CoV-1 is sustained for at least a year.

The persistence of antibodies to SARS-CoV-2 remains under investigation. Circulating antibody titers to other coronaviruses have been reported to decline significantly after 1 year [274]. Autopsies of lymph nodes and spleens from severe acute COVID-19 patients showed a loss of T follicular helper cells and germinal centers that may explain some of the impaired development of antibody responses [275]. An early study (initially released on medRxiv on February 25, 2020) presented a chemiluminescence immunoassay to a synthetic peptide derived from the amino acid sequence of the SARS_CoV-2 S protein [276]. This method was highly specific to SARS-CoV-2 and detected IgM in 57.2% and IgG in 71.4% and 57.2% of sera samples from 276 confirmed COVID-19 patients. They reported that they could detect IgG within two days of the onset of fever and were not able to detect IgM any earlier, a pattern they compared to findings in MERS. The duration over which these antibodies persist remains unknown.

4.4.2 Current Approaches

Several countries are now focused on implementing antibody tests, and in the United States, the FDA recently approved a serological test by Cellex for use under emergency conditions [277]. Specifically, the Cellex qSARS-CoV-2 IgG/IgM Rapid Test is a chromatographic immunoassay designed to qualitatively detect IgM and IgG antibodies against SARS-CoV-2 in the plasma (from a blood sample) of patients suspected to have developed the SARS-CoV-2 infection [277]. Such tests illuminate the progression of viral disease, as IgM are the first antibodies produced by the body and indicate that the infection is active. Once the body has responded to the infection, IgG are produced and gradually replace IgM, indicating that the body has developed immunogenic memory [278]. The test cassette contains a pad of SARS-CoV-2 antigens and a nitrocellulose strip with lines for each of IgG and IgM, as well as a control (goat IgG) [277]. In a specimen that contains antibodies against the SARS-CoV-2 antigen, the antibodies will bind to the strip and be captured by the IgM and/or IgG line(s), resulting in a change of color [277]. With this particular assay, results can be read within 15-20 minutes [277]. Other research groups, such as the Krammer lab of the Icahn School of Medicine at Mount Sinai, proposed an ELISA test that detects IgG and IgM that react against the receptor binding domain (RBD) of the spike proteins (S) of the virus [279]. The authors are now working to get the assay into clinical use [280].

4.4.3 Limitations of Serological Tests

Importantly, false-positives can occur due to the cross-reactivity with other antibodies according to the clinical condition of the patient [277]. Therefore, this test should be used in combination with RNA detection tests [277]. Due to the long incubation times and delayed immune responses of infected patients, serological tests are insufficiently sensitive for a diagnosis in the early stages of an infection. The limitations due to timing make serological tests far less useful for enabling test-and-trace strategies.

4.5 Possible Alternatives to Current Practices for Identifying Active Cases

Clinical symptoms are too similar to other types of pneumonia to be sufficient as a sole diagnostics criterion. In addition, as noted above, identifying asymptomatic cases is critical. Even among mildly symptomatic patients, a predictive model based on clinical symptoms had a sensitivity of only 56% and a specificity of 91% [281]. More problematic is that clinical symptom-based tests are only able to identify already symptomatic cases, not presymptomatic or asymptomatic cases. They may still be important for clinical practice, and for reducing tests needed for patients deemed unlikely to have COVID-19.

X-ray diagnostics have been reported to have high sensitivity but low specificity in some studies [282]. Other studies have shown that specificity varies between radiologists [283], though the sensitivity reported here was lower than that published in the previous paper. However, preliminary machine-learning results have shown far higher sensitivity and specificity from analyzing chest X-rays than was possible with clinical examination [284]. X-ray tests with machine learning can potentially detect asymptomatic or presymptomatic infections that show lung manifestations. This approach would still not recognize entirely asymptomatic cases. Given the above, the widespread use of X-ray tests on otherwise healthy adults is likely inadvisable.

4.6 Strategies and Considerations for Determining Whom to Test

Early in the COVID-19 pandemic, testing was typically limited to individuals considered high risk for developing serious illness [285]. This approach often involved limiting testing to people with severe symptoms and people showing mild symptoms that had been in contact with a person who had tested positive. Individuals who are asymptomatic (i.e., potential spreaders) and individuals who are able to recover at home are therefore often unaware of their status. However, this method of testing administration misses a high proportion of infections and does not allow for test-and-trace methods to be used. For instance, a recent study from Imperial College estimates that in Italy, the true number of infections was around 5.9 million in a total population ~60 million, compared to the 70,000 detected as of March 28th [226]. Another analysis, which examined New York state, indicated that as of May, 2020, approximately 300,000 cases had been reported in a total population of approximately 20 million [286]. This corresponded to ~1.5% of the population, but ~12% of individuals sampled statewide were estimated as positive through antibody tests (along with indications of spatial heterogeneity at higher resolution) [286]. Technological advancements that facilitate widespread, rapid testing will therefore improve the potential to accurately assess the rate of infection and aid in controlling the virus’ spread.

4.7 Conclusions

Major advancements have been made in identifying diagnostic approaches. The development of diagnostic technologies have been rapid, beginning with the release of the SARS-CoV-2 viral genome sequence in January. As of October 2020, a range of diagnostic tests have become available. One class of tests uses PCR (RT-PCR or qRT-PCR) to assess the presence of SARS-CoV-2 RNA, while another typically uses ELISA to test for the presence of antibodies to SARS-CoV-2. The former approach is useful for identifying active infections, while the latter measures hallmarks of the immune response and therefore can detect either active infections or immunity gained from prior infection. Combining these tests leads to extremely accurate detection of SARS-CoV-2 infection (98.6%), but when used alone, PCR-based tests are recommended before 5.5 days after the onset of the illness and antibody tests after 5.5 days [287]. Other strategies for testing can also influence the tests’ accuracy, such as the use of nasopharyngeal swabs versus BALF [287], which allow for trade-offs between patient’s comfort and test sensitivity. Additionally, technologies such as digital PCR may allow for scale-up in the throughput of diagnostic testing, facilitating widespread testing. One major question that remains is whether people who recover from SARS-CoV-2 develop sustained immunity, and over what period this immunity is expected to last. Some reports have suggested that some patients may develop COVID-19 reinfections (e.g., [288]), but the rates of reinfection are currently unknown. Serologic testing combined with PCR testing will be critical to confirming purported cases of reinfection and to identifying the duration over which immunity is retained and to understanding reinfection risks.

5 Identification and Development of Prophylactics and Therapeutics for COVID-19

5.1 Abstract

The novel coronavirus SARS-CoV-2 emerged in late 2019 and went on to significantly impact most countries worldwide. Only a small number of coronaviruses are known to infect humans, and only two are associated with the severe outcomes associated with SARS-CoV-2: SARS-CoV-1, a sister species that emerged in 2002, and MERS-CoV, which emerged in 2012. Although both of these previous pandemics were rapidly controlled, no vaccines or robust therapeutic interventions were identified. However, previous insights into the immune response to coronaviruses, many made in the context of better understanding SARS and MERS, have been beneficial in identifying approaches to the treatment and prophylaxis of COVID-19. A number of potential therapeutics were rapidly identified, leading to a large number of clinical trials investigating a variety of possible therapeutic approaches being initiated early on in the pandemic. Similarly, previous advances in vaccine development allowed for the rapid design and testing of candidate prophylactics. As a result, by the end of 2020, a small number of therapeutics and vaccines have been authorized by regulatory agencies such as the Food and Drugs Administration in the United States, and many more continue to be under investigation. Here, we describe a range of approaches for the treatment and prevention of COVID-19 along with their proposed mechanisms of action and the current status of clinical investigation into each candidate. The status of these investigations will continue to evolve, and this review will be updated as progress is made.

5.2 Importance

The COVID-19 pandemic is a rapidly evolving crisis. When the worldwide scientific community shifted its focus onto the SARS-CoV-2 virus and the disease it causes, a large number of possible pharmaceutical approaches for treatment and prevention were proposed. What is known about each of these potential interventions has evolved rapidly throughout 2020. In March 2020, we began monitoring a range of candidates and have continued to update this manuscript as new information became available throughout 2020. Some have been supported, others have been revealed to be unlikely to have any therapeutic benefits, and most require more data before a conclusion can be drawn. This rapidly changing area of research provides important insight into how the ongoing pandemic can be managed and also demonstrates the power of interdisciplinary collaboration to rapidly understand a virus and match its characteristics with existing or novel pharmaceuticals.

5.3 Introduction

The novel coronavirus Severe acute respiratory syndrome-related coronavirus 2 (SARS-CoV-2) emerged in late 2019 and quickly precipitated the worldwide spread of novel coronavirus disease 2019 (COVID-19), the disease caused by its infection. COVID-19 is associated with symptoms ranging from none (asymptomatic) to mild to severe, with approximately 2% of patients dying from COVID-19-related complicates such as acute respiratory disease syndrome (ARDS) [cite pathogenesis/transmission]. The virus is likely spread between people primarily by droplets, with the role of contact and aerosol transmission still in question [182,183]. As a result, public health guidelines have been critical to efforts to control the spread of the virus. However, as of December 2020, COVID-19 remains a significant worldwide concern (Figure 2), with cases in some places surging far above the numbers reported in early 2020. Due to the continued threat of the virus and the severity of the disease, the identification and development of prophylactic and therapeutic interventions have emerged as significant international priorities. Both approaches hold valuable potential for controlling the impact of the disease. Prophylactics induce immunity to prevent an individual from contracting a disease whereas therapeutics treat a disease in individuals who have already been infected. For a highly infectious virus like SARS-CoV-2, a vaccine would hold particular value because it could bolster the immune response to the virus of the population broadly, thereby driving a lower rate of infection and likely significantly reducing fatalities. However, the vaccine development process has historically been slow, and vaccines fail to provide immediate prophylactic protection or treat ongoing infections [289]. Thus, there is also an immediate need for treatments that palliate symptoms and prevent the most severe outcomes from infection. Fortunately, prior developments during other recent pandemics, especially those caused by human coronaviruses (HCoV), has provided a number of hypotheses guiding a biomedical approach to the novel coronavirus infection.

2,057,215 COVID-19 deaths had been reported worldwide as of January 19, 2021 (Figure 2).

Figure 2: Cumulative global COVID-19 deaths since January 22, 2020. Data are from the COVID-19 Data Repository by the Center for Systems Science and Engineering at Johns Hopkins University [290].

5.3.1 Lessons from Prior HCoV Outbreaks

SARS-CoV-2’s rapid shift from an unknown virus to a significant worldwide threat closely parallels the emergence of Severe acute respiratory syndrome-related coronavirus (SARS-CoV-1). The first documented case of COVID-19 was reported in Wuhan, China in November 2019, and the disease quickly spread worldwide during the early months of 2020. Similarly, the first case of SARS was reported in November 2002 in the Guangdong Province of China and spread within China and then into several countries across continents over the following months [28,30]. In fact, the virus causing COVID-19 was quickly identified through genome sequencing as a novel betacoronavirus closely related to SARS-CoV-1 10]. There are other similarities but also some differences in characteristics of the two viruses relevant to their spread. SARS infection is severe, with an estimated death rate of 9.5% [30], while estimates of the death rate associated with COVID-19 are much lower, at approximately 2% [cite pathogenesis/transmission section] . SARS-CoV-1 is highly contagious via droplet transmission and has a basic reproduction number (R0) of 4 (i.e., each person infected was estimated to infect four other people) [30]. SARS-CoV-2 also appears to be spread primarily by droplet transmission [182,183], and most estimates of its R0 fall between 2.5 and 3 [cite transmission/pathogenesis]. Furthermore, the 17-year difference in the timing of these two outbreaks has led to some major differences in the tools available for the international community’s response. At the time that SARS-CoV-1 emerged, no new HCoV had been identified in almost 40 years [28]. The identity of the virus underlying the SARS disease remained unknown until April of 2003, when the SARS-CoV-1 virus was characterized through a worldwide scientific effort spearheaded by the WHO [28]. In contrast, the SARS-CoV-2 genomic sequence was released on January 3, 2020 [10], only days after the international community became aware of the novel pneumonia-like illness now known as COVID-19. While SARS-CoV-1 belonged to a distinct lineage from the two other HCoV known at the time of its discovery [30], SARS-CoV-2 is closely related to SARS-CoV-1 and a more distant relative of another HCoV characterized in 2012, Middle East respiratory syndrome-related coronavirus [13,291]. Thus, despite their phylogenetic similarity, the viruses have different characteristics relevant to their spread, and the SARS-CoV-2 virus emerged under very different circumstances than SARS-CoV-1 in terms of scientific knowledge about HCoV.

The trajectories of the pandemics associated with each of the viruses have also diverged significantly. By July 2003, the SARS outbreak was officially determined to be under control, with the success credited to infection management practices such as mask wearing [28]. In contrast, MERS is still circulating and remains a concern because the fatality rate is very high at almost 35%, but the disease is much less easily transmitted, with an R0 of 1 [30]. The low R0 in combination with public health practices allowed for its spread to be contained [30]. Neither of these trajectories are comparable to that of SARS-CoV-2, which remains a serious threat worldwide a full year after the first known cases of COVID-19. However, early results suggest that pharmaceutical interventions for COVID-19 may be more successful than previous efforts to develop prophylactics and therapeutics for SARS and MERS. Vaccines were not used to control either virus, although vaccine development programs were established for SARS-CoV-1 [31]. Care for SARS and MERS patients prioritized supportive care and symptom management [30]. To the extent that clinical treatments for SARS and MERS were explored, there is generally a lack of evidence supporting their efficacy. For example, Ribavirin is an antiviral that was often used in combination with corticosteroids and sometimes interferon (IFN) medications to treat SARS and MERS [28], but its effects have been found to be inconclusive in retrospective and in vitro analyses of SARS and the SARS-CoV-1 virus, respectively [28]. IFNs and Ribavirin have shown promise in in vitro analyses of MERS, but their clinical effectiveness remains unknown [28]. Therefore, only limited pharmaceutical advances from prior HCoV outbreaks can be adopted to COVID-19. Importantly, though, prior analyses of the virological and pathogenic properties of SARS-CoV-1 and MERS-CoV provide a strong foundation for the development of hypotheses about SARS-CoV-2 that have served to accelerated the development and identification of potential prophylactic and therapeutic approaches.

5.3.2 Therapeutic Approaches

Therapeutic approaches to the current pandemic can utilize two potential avenues: they can reduce the symptoms that are harmful to COVID-19 patients, or they can directly target the virus to hinder the spread of infection. The goal of the former is to reduce the severity and risks of an active infection while for the latter, it is to inhibit the development of the virus once an individual is infected. A variety of symptom profiles with a range of severity are associated with COVID-19, many of which are not life-threatening. A study of COVID-19 patients in a hospital in Berlin, Germany found that the symptoms associated with the highest risk of death included infection-related symptoms, such as sepsis, respiratory symptoms such as ARDS, and cardiovascular failure or pulmonary embolism [292]. Therefore, therapeutics that reduce risk associated with these severe outcomes are of particular interest to reducing the pandemic death toll. Therapeutics that directly target the virus itself typically fall into the broad category of antivirals. Antiviral therapies hinder the spread of a virus within the host, rather than destroying existing copies of the virus, and these drugs can vary in their specificity to a narrow or broad range of viral targets. For both categories, uncertainty often surrounds treatments’ exact mechanisms of action, as most therapies have secondary or off-target effects.

In this review, we describe some treatments and prophylactics that have been considered for COVID-19 and classify them according to their biological properties, specifically whether they are biologics (produced from components of organisms) or small molecules. Small molecule drugs include drugs targeted at viral particles, drugs targeted at host proteins, and broad spectrum pharmaceuticals, while biologics include antibodies, interferons, and vaccines. A large number of clinical trials investigating a wide range of possible therapeutics and prophylactics for COVID-19 are currently in progress or have already been completed (Figure 3). The purpose of this review is to critically appraise the literature surrounding a subset of clinical trials and to evaluate a range of approaches to repurpose existing or develop novel approaches to the prevention, mitigation, and treatment of coronavirus infections. The identity of the virus underlying the SARS disease remained unknown until April of 2003, when the SARS-CoV-1 virus was characterized through a worldwide scientific effort spearheaded by the WHO [28]. In contrast, the SARS-CoV-2 genomic sequence was released on January 3, 2020 [10], only days after the international community became aware of the novel pneumonia-like illness now known as COVID-19. While SARS-CoV-1 belonged to a distinct lineage from the two other HCoV known at the time of its discovery [30], SARS-CoV-2 is closely related to SARS-CoV-1 and a more distant relative of another HCoV characterized in 2012, Middle East respiratory syndrome-related coronavirus [13,291]. Thus, despite their phylogenetic similarity, the viruses have different characteristics relevant to their spread, and the SARS-CoV-2 virus emerged under very different circumstances than SARS-CoV-1 in terms of scientific knowledge about HCoV.

Figure 3: COVID-19 clinical trials. There are 6,417 COVID-19 clinical trials and 168 trials with results as of November 9, 2020. The recruitment statuses and trial phases are shown only for trials in which the status or phase is recorded. The study types include only types used in at least five trials. The common interventions are all interventions used in at least ten trials. Combinations of interventions, such as Hydroxychloroquine + Azithromycin, are tallied separately from the individual interventions. Trials data are from the University of Oxford Evidence-Based Medicine Data Lab’s COVID-19 TrialsTracker [293].

5.4 Small Molecule Drugs

Small molecules are synthesized compounds of low molecular weight, typically less than 1 kilodalton (kDa) [294]. Small-molecule pharmaceutical agents have been a backbone of drug development since the discovery of penicillin in the early twentieth century [295]. It and other antibiotics have long been among the best known applications of small molecules to therapeutics, but biotechnological developments such as the prediction of protein-protein interactions have facilitated advances in precise targeting of specific structures using small molecules [295]. Small molecule drugs today encompass a wide range of therapeutics beyond antibiotics, including antivirals, protein inhibitors, and many broad-spectrum pharmaceuticals.

5.4.1 Small Molecule Antivirals

Antiviral drugs against SARS-CoV-2 are designed to inhibit replication of a virus within an epithelial host cell. This process requires inhibiting the replication cycle of a virus by disrupting one of six fundamental steps [296]. In the first step, the virus attaches to the host cell, which it next penetrates through endocytosis. Then the virus undergoes uncoating, which is classically defined as the release of viral contents into the host cell, and next the viral genetic material enters the nucleus where it gets replicated during the biosynthesis stage. During the assembly stage, viral proteins are translated, allowing new viral particles to be assembled, and during the final step, release, the new viruses are released into the extracellular environment. Many antiviral drugs are designed to inhibit the replication of viral genetic material during the biosynthesis step. Unlike DNA viruses, which can use the host enzymes to propagate themselves, RNA viruses like SARS-CoV-2 depend on their own polymerase, the RNA-dependent RNA polymerase (RdRP), for replication [297,298]. Targeting RdRP is therefore an effective strategy for antivirals against RNA viruses and is the proposed mechanism underlying the treatment of SARS and MERS with Ribavirin [299]. However, although antivirals are designed to target a virus, they can also impact other processes in the host and may have unintended effects.

5.4.1.1 Nucleoside and Nucleotide Analogs

5.4.1.1.1 Favipiravir

Favipiravir (Avigan), also known as T-705, was discovered by Toyama Chemical Co., Ltd. [300]. The drug was found to be effective at blocking viral amplification in several influenza subtypes as well as other RNA viruses, such as Flaviviridae and Picornaviridae, through a reduction in plaque formation [301] and viral replication in Madin-Darby Canine Kidney (MDCK) cells [302]. Furthermore, inoculation of mice with favipiravir was shown to increase survival of influenza infections [301,302]. In 2014, the drug was approved in Japan for the treatment of patients infected with influenza that was resistant to conventional treatments like neuraminidase inhibitors [303]. Favipiravir (6-fluoro-3-hydroxy-2-pyrazinecarboxamide) acts as a purine and purine nucleoside analogue that inhibits viral RNA polymerase in a dose-dependent manner across a range of RNA viruses, including Influenza virus [304,305,306,307,308]. Nucleotides and nucleosides are the natural building blocks for RNA synthesis. Because of this, modifications to nucleotides and nucleosides can disrupt key processes including replication [309]. Biochemical experiments showed that favipiravir was recognized as a purine nucleoside analogue and incorporated into the viral RNA template. A single incorporation does not influence RNA transcription; however, multiple events of incorporation lead to the arrest of RNA synthesis [310]. Evidence for T-705 inhibiting viral RNA polymerase are based on time-of-drug addition studies that found that viral loads were reduced with the addition of favipiravir in early times post-infection [304,307,308].

The effectiveness of favipiravir for treating patients with COVID-19 is currently under investigation. An open-label, nonrandomized, before-after controlled study was recently conducted [311]. The study included 80 COVID-19 patients (35 treated with favipiravir, 45 control) from the isolation ward of the National Clinical Research Center for Infectious Diseases (The Third People’s Hospital of Shenzhen), Shenzhen, China. The patients in the control group were treated with other antivirals, such as lopinavir and ritonavir. It should be noted that subsequent, large-scale analyses, the WHO Solidarity trial and the RECOVERY trial, identified no effect of lopinavir or of a lopinavir-ritonavir combination, respectively, on the metrics of COVID-19-related mortality that each assessed [???,312,313]. Treatment was applied on days 2-14; treatment stopped either when viral clearance was confirmed or at day 14. The efficacy of the treatment was measured by 1) the time until viral clearance using Kaplan-Meier survival curves, and 2) the improvement rate of chest computed tomography (CT) scans on day 14 after treatment. The study found that favipiravir increased the speed of recovery, measured as viral clearance from the patient by RT-PCR, with patients receiving favipiravir recovering in 4 days compared to 11 days for patients receiving antivirals such as lopinavir and ritonavir. Additionally, the lung CT scans of patients treated with favipiravir showed significantly higher improvement rates (91%) on day 14 compared to control patients (62%). However, there were adverse side effects in 4 (11%) favipiravir-treated patients and 25 (56%) control patients. The adverse side effects included diarrhea, vomiting, nausea, rash, and liver and kidney injury. Overall, despite the study reporting clinical improvement in favipiravir-treated patients, due to some issues with study design, it cannot be determined whether treatment with favipiravir had an effect or whether these patients would have recovered regardless of any treatment. For example, despite the significant differences observed between the two treatment groups, follow-up analysis is necessary due to the small sample size. The selection of patients did not take into consideration important factors such as previous clinical conditions or sex, and there was no age categorization. The study was neither randomized nor blinded, and the baseline control group was another antiviral instead of a placebo. Therefore, randomized controlled trials are still required.

5.4.1.1.2 Remdesivir

Remdesivir (GS-5734) is an intravenous antiviral that was developed by Gilead Sciences to treat Ebola Virus Disease (EVD). At the outset of the COVID-19 pandemic, it did not have any have any FDA-approved use. However, on May 1, 2020, the FDA issued an Emergency Use Authorization (EUA) for remdesivir for the treatment of hospitalized COVID-19 patients [314]. The EUA was based on information from two clinical trials, NCT04280705 and NCT04292899 [315,316,317,318]. Remdesivir is metabolized to GS-441524, an adenosine analog that inhibits a broad range of polymerases and then evades exonuclease repair, causing chain termination [319,320,321]. A clinical trial in the Democratic Republic of Congo found some evidence of effectiveness against EVD, but two antibody preparations were found to be more effective, and remdesivir was not pursued [322]. Although it was developed against EVD, remdesivir also inhibits polymerase and replication of the coronaviruses MERS-CoV and SARS-CoV-1 in cell culture assays with submicromolar IC50s [323]. It has also been found to inhibit SARS-CoV-2, showing synergy with chloroquine in vitro [321]. The effectiveness of remdesivir for treating patients with COVID-19 is currently under investigation.

Remdesivir was first used on some COVID-19 patients under compassionate use guidelines [324,325]. All were in late stages of COVID-19 infection, and these reports were inconclusive about the drug’s efficacy. Gilead Sciences, the maker of remdesivir, led a recent publication that reported outcomes for compassionate use of the drug in 61 patients hospitalized with confirmed COVID-19. Here, 200 mg of remdesivir was administered intravenously on day 1, followed by a further 100 mg/day for 9 days [318]. There were significant issues with the study design, or lack thereof. There was no randomized control group. The inclusion criteria were variable: some patients only required low doses of oxygen, while others required ventilation. The study included many sites, potentially with variable inclusion criteria and treatment protocols. The patients analyzed had mixed demographics. There was a short follow-up period of investigation. Some patients worsened, some patients died, and eight were excluded from the analysis mainly due to missing post-baseline information; thus, their health is unaccounted for. Therefore, even though the study reported clinical improvement in 68% of the 53 patients ultimately evaluated, due to the significant issues with study design, it could not be determined whether treatment with remdesivir had an effect or whether these patients would have recovered regardless of treatment. Another study comparing 5- and 10-day treatment regimens reported similar results but was also limited because of the lack of a placebo control [327]. The study did not alter our understanding of the efficacy of remdesivir in treating COVID-19, but the encouraging results provided motivation for placebo-controlled studies.

Remdesivir was later tested in a double-blind placebo-controlled phase 3 clinical trial performed at 60 trial sites, 45 of which were in the United states [316,317]. The trial recruited 1,062 patients and randomly assigned them to placebo treatment or treatment with remdesivir. Patients were stratified for randomization based on site and the severity of disease presentation at baseline [317]. The treatment was 200 mg on day 1, followed by 100 mg on days 2 through 10. Data was analyzed from a total of 1,059 patients who completed the 29 day course of the trial (517 assigned to remdesivir and 508 to placebo) [317]. The two groups were well matched demographically and clinically at baseline. Those who received remdesivir had a median recovery time of 10 days, as compared with 15 days in those who received placebo (rate ratio for recovery, 1.29; 95% CI, 1.12 to 1.49; p < 0.001). The Kaplan-Meier estimates of mortality by 14 days were 6.7% with remdesivir and 11.9% with placebo, with a hazard ratio (HR) for death of 0.55 and a 95% CI of 0.36 to 0.83, and at day 29, remdesivir corresponded to 11.4% and the placebo to 15.2% (HR: 0.73; 95% CI: 0.52 to 1.03). Serious adverse events were reported in 131 of the 532 patients who received remdesivir (24.6%) and in 163 of the 516 patients in the placebo group (31.6%). This study also reported an association between remdesivir administration and both clinical improvement and a lack of progression to more invasive respiratory intervention in patients receiving non-invasive and invasive ventilation at randomization [317]. Largely on the results of this trial, the FDA issued an Emergency Use Authorization (EUA) for remdesivir for the treatment of hospitalized COVID-19 patients. Clinical trials are currently underway to evaluate the use of remdesivir to treat COVID-19 patients at both early and late stages of infection and in combinations with other drugs (Figure 3). These trials include [321,328,329,330,331]. As of October 22, 2020, remdesivir received FDA approval based on three clinical trials [332].

However, results suggesting no effect of remdesivir on survival were reported by the WHO Solidarity trial [???]. This large-scale, open-label trial enrolled 11,330 adult in-patients at 405 hospitals in 30 countries around the world [???]. Patients were randomized in equal proportions into four experimental and a control conditions, corresponding to four candidate treatments for COVID-19 and SOC, respectively; no placebo was administered. The 2750 patients in the remdesivir group were administered 200 mg intravenously on the first day and 100 mg on each subsequent day until day 10 and assess for in-hospital death (primary endpoint), duration of hospitalization, and progression to mechanical ventilation. There were also 2708 control patients who would have been eligible and able to receive remdesivir were they not assigned to the control group. A total of 604 patients among these two cohorts deceased during initial hospitalization. The rate ratio of death between these two groups was therefore not significant (p = 0.50), suggesting that the administration of remdesivir did not affect survival. The two secondary analyses similarly did not find any effect of remdesivir. Additionally, the authors compared data from their study with data from three other studies of remdesivir (including [317]) stratified by supplemental oxygen status. The adjusted rate ratio for death based on this meta-analysis was 0.91. These results thus do not support the previous findings that remdesivir reduced median recovery time and mortality risk in COVID-19 patients.

In response to the results of the Solidarity trial, Gilead, which manufactures remdesivir under the name Veklury^R, released a statement pointing to the fact that the Solidarity trial was not placebo-controlled or double-blind and at the time of the statement had not been peer reviewed [333]; these sentiments have been echoed elsewhere [334]. Other critiques of this study have noted that antivirals are not typically targeted at patients with severe illness, and therefore remdesivir could be more beneficial for patients with mild rather than severe cases [313,335]. However, the publication associated with the trial sponsored by Gilead did suggest an effect of remdesivir on patients with severe disease, identifying an 11 versus 18 day recovery period (rate ratio for recovery: 1.31, 95% CI 1.12 to 1.52), although the results of a significance test were not provided [317]. Additionally, a smaller analysis comparing treatment with remdesivir to standard of care in patients with moderate COVID-19 reported a small effect size, corresponding to minimal differences between these groups at day 11 [336]. Therefore, some of the findings from this trial do seem to contradict the findings of the Solidarity trial. On the other hand, only 62% of patients in the Solidarity trial were randomized on the day of admission or one day afterwards [???], and differences in disease progression may influence the effect of remdesivir [313], so there could be be additional confounding factors in one or both trials. The broad international nature of the Solidarity clinical trial, which included countries with a wide range of economic profiles and a variety of healthcare systems, provides a much-needed global perspective in a pandemic [313]. Despite the findings of the Solidarity trial, remdesivir remains available for the treatment of COVID-19 in many places. Follow-up studies are needed and in many cases are underway to further investigate remdesivir-related outcomes, with possibilities including combinations of remdesivir with other drugs such as baricitinib, which is an inhibitor of Janus kinase 1 and 2 [337]. Similarly, the extent to which the remdesivir dosing regimen could influence outcomes is still under consideration. A randomized, open-label trial compared the effect of remdesivir on 397 patients with severe COVID-19 over 5 versus 10 days [315,327]. Patients in the two groups were administered 200 mg of remdesivir intravenously on the first day, followed by 100 mg on the subsequent four or nine days, respectively. The two groups differed significantly in their clinical status, with patients assigned to the 10-day group having more severe illness. This study also differed from most because it included not only adults, but also pediatric patients as young as 12 years old. This study found no significant differences across several outcomes for patients receiving a 5-day or 10-day course, when correcting for baseline clinical status. They did find evidence suggesting that the 10-day course might reduce mortality in the most severe patients at day 14, but the representation of this group in the study population was too low to justify any conclusions [327]. Thus, additional research is also required to determine whether the dosage and duration of remdesivir administration influences outcomes.

In summary, remdesivir is a first in class drug due to its FDA approval. Early investigations of this drug establish proof of principle that drugs targeting the virus can benefit COVID-19 patients. It also shows proof of principle that SARS-CoV-2 can be targeted at the level of viral replication, since remdesivir targets the viral RNA polymerase at high potency. Moreover, one of the most successful strategies for developing therapeutics for viral diseases is to target the viral replication machinery, which are typically virally encoded polymerases. Small molecule drugs targeting viral polymerases are the backbones of treatments for other viral diseases including HIV and Herpes. Notably, the HIV and herpes polymerases are a reverse transcriptase and a DNA polymerase, respectively, whereas SARS-CoV-2 encodes an RdRP, so most of the commonly used polymerase inhibitors are not likely to be active against SARS-CoV-2. In clinical use, polymerase inhibitors show short term benefits for HIV patients but for long term benefits they must be part of combination regimens. They are typically combined with protease inhibitors, integrase inhibitors, and even other polymerase inhibitors. Additional clinical trials of remdesivir in different patient pools and in combination with other therapies will refine its use in the clinic.

5.4.1.2 Protease Inhibitors

Several studies showed that viral proteases play an important role in the life cycle of viruses, including coronaviruses, by modulating the cleavage of viral polyprotein precursors [338]. Several FDA-approved drugs target proteases, including lopinavir and ritonavir for HIV infection and simeprevir for hepatitis C virus infection. In particular, serine protease inhibitors were suggested for the treatment of SARS and MERS viruses [339]. Recently, a study [66] suggested that camostat mesylate, an FDA-approved protease inhibitor (PI) could block the entry of SARS-CoV-2 into lung cells in vitro. However, to test the efficacy of PIs in patients, randomized clinical trials will need to be conducted on patients and healthy volunteers. Investigation of possible PIs that would work against SARS-CoV-2 has been driven by computational predictions, leading to the computer-aided design of a Michael acceptor inhibitor, N3, to target a protease critical to SARS-CoV-2 replication.

Discovery of the N3 mechanism arose from interest in the two polyproteins encoded by the SARS-CoV-2 replicase gene, pp1a and pp1ab, that are critical for viral replication and transcription [340]. These polyproteins must undergo proteolytic processing. This processing is usually conducted by Mpro, a 33.8-kDa SARS-CoV-2 protease that is therefore fundamental to viral replication and transcription. N3 was designed computationally [341] to bind in the substrate binding pocket of the Mpro protease of SARS-like coronaviruses [342], therefore inhibiting proteolytic processing. Subsequently, the structure of N3-bound SARS-CoV-2 Mpro was solved [340], confirming the computational prediction. N3 was tested in vitro on SARS-CoV-2-infected Vero cells, which belong to a line of cells established from the kidney epithelial cells of an African green monkey, and was found to inhibit SARS-CoV-2 [340].

Although N3 is a strong inhibitor of SARS-CoV-2 in vitro, its safety and efficacy still need to be tested in healthy volunteers and patients. After the design and confirmation of N3 as a highly potent Michael acceptor inhibitor and the identification of Mpro’s structure [340,343], 10,000 compounds were screened for their in vitro anti-Mpro activity. The six leads that were identified were ebselen, disulfiram, tideglusib, carmofur, and PX-12. In vitro analysis revealed that Ebselen had the strongest potency in reducing the viral load in SARS-CoV-2-infected Vero cells [340]. Ebselen is an organoselenium compound with anti-inflammatory and antioxidant properties [344] It has been proposed as a possible treatment for conditions ranging from bipolar disorder to diabetes to heart disease [344], and a preliminary investigation of ebselen as a treatment for noise-induced hearing loss provided promising reports of its safety [345]. For COVID-19, the NSP5 in SARS-CoV-2 contains a cysteine at the active site of Mpro, and ebselen is able to inactivate the protease by bonding covalently with this cysteine to form a selenosulfide [344,346]. Interestingly there has been some argument that selenium deficiency may be associated with more severe COVID-19 outcomes [347,348,349], possibly indicating that its antioxidative properties are protective [346]. On the other hand, ebselen and the other compounds identified are likely to be promiscuous binders, which could diminish their therapeutic potential [340]. While there is clear computational and in vitro support for ebselen’s potential as a COVID-19 therapeutic, results from clinical trials are not yet available for this compound.

In summary, N3 is a computationally designed molecule that inhibits the viral transcription through inhibiting Mpro. Ebselen is both a strong Mpro inhibitor and strong inhibitor of viral replication in vitro that was found to reduce SARS-CoV-2 viral load even more effectively than N3. Ebselen is a very promising compound since its safety has been demonstrated in other indications. However, Ebselen is likely a false positive since it is a promiscuous compound that can have many targets [350]. Therefore, compounds with higher specificity may be required to effectively translate to clinical trials.

5.4.2 Broad-Spectrum Pharmaceuticals

When a virus enters a host, the host becomes the virus’ environment. Therefore, the state of the host can also influence the virus’s ability to replicate and spread. Traditionally, viral targets have been favored for pharmaceutical interventions because altering host processes is likely to be less specific than targeting the virus directly [351]. On the other hand, targeting the host offers potential for a complementary strategy to antivirals that could broadly limit the ability of viruses to replicate [351]. As a result, therapeutic approaches that target host proteins have become an area of interest for SARS-CoV-2. Viral entry receptors in particular have emerged as a target of interest. Entry of SARS-CoV-2 into the cell depends on binding to the Angiotensin-converting enzyme 2 (ACE2) receptor, which is catalyzed by the enzyme encoded by TMPRSS2 [66]. In principle, drugs that reduce the expression of these proteins or sterically hinder viral interactions with them might reduce viral entry into cells.

Due to the urgent nature of the COVID-19 pandemic, many of the pharmaceutical agents that have been widely publicized as having possible therapeutic or prophylactic effects are broad-spectrum pharmaceuticals that pre-date the COVID-19 pandemic. These treatments are not specifically targeted at the virus itself or at the host receptors it relies on, but rather induce broad shifts in host biology that are hypothesized to be potential inhibitors of the virus. In most cases, interest in particular candidate medications arises because they are already available for other purposes. However, the fact that the targets of these agents are non-specific means that the mechanism of action can appear to be relevant to COVID-19 without a therapeutic or prophylactic effect being observed in rigorous testing. This category of drugs has also received significant attention from the media and general public, often before rigorous testing has been able to determine their effectiveness against SARS-CoV-2.

5.4.2.1 ACEi and ARB

Angiotensin-converting enzyme (ACE) inhibitors and angiotensin II receptor blockers (ARBs) are among today’s most commonly prescribed medications [352,353]. In the United States, for example, they are prescribed well over 100,000,000 times annually. Data from some animal models suggest that several, but not all, ACE inhibitors and several ARBs increase ACE2 expression in the cells of some organs [354]. Clinical studies have not established whether plasma ACE2 expression is increased in humans treated with these medications [355]. While randomized clinical trials are ongoing, a variety of observational studies have examined the relationship between exposure to ACE inhibitors or ARBs and outcomes in patients with COVID-19. An observational study of the association of exposure to ACE inhibitors or ARB with outcomes in COVID-19 was retracted from the New England Journal of Medicine [356]. Moreover, because observational studies are subject to confounding, randomized controlled trials are the standard means of assessing the effects of medications, and the findings of the various observational studies bearing on this topic cannot be interpreted as indicating a protective effect of the drug [357,358]. Several clinical trials testing the effects of ACE inhibitors or ARBs on COVID-19 outcomes are ongoing [359,360,361,362,363,364]. These studies of randomized intervention will provide important data for understanding whether exposure to ACEis or ARBs is associated with COVID-19 outcomes. Additional information about ACE2, observational studies of ACE inhibitors and ARBs in COVID-19, and clinical trials on this topic have been summarized [365].

5.4.2.2 Hydroxychloroquine and Chloroquine

Chloroquine (CQ) and hydroxychloroquine (HCQ) are lysosomotropic agents, meaning they are weak bases that can pass through the plasma membrane. Both drugs increase cellular pH by accumulating in their protonated form inside lysosomes [366,367]. This shift in pH inhibits the breakdown of proteins and peptides by the lysosomes during the process of proteolysis [367]. A number of mechanisms have been proposed through which these drugs could influence the immune response to pathogen challenge. For example, CQ/HCQ can interfere with digestion of antigens within the lysosome and inhibit CD4 T-cell stimulation while promoting the stimulation of CD8 T-cells [367]. CQ/HCQ can also decrease the production of certain key cytokines involved in the immune response, including interleukin-6 (IL-6), and inhibit the stimulation of toll-like receptors (TLR) and TLR signaling [367]. The drugs also have anti-inflammatory and photoprotective effects and may also affect rates of cell death, blood clotting, glucose tolerance, and cholesterol levels [367].

Interest in CQ and HCQ for treating COVID-19 was catalyzed by a mechanism observed in in vitro studies of both SARS-CoV-1 and SARS-CoV-2. In one study, CQ inhibited viral entry of SARS-CoV-1 into Vero E6 cells, a cell line that was derived from Vero cells in 1968, through the elevation of endosomal pH and the terminal glycosylation of the cellular entry receptor, ACE2 [368]. Increased pH within the cell, as discussed above, inhibits proteolysis, and terminal glycosylation of ACE2 is thought to interfere with virus-receptor binding. An in vitro study of SARS-CoV-2 infection of Vero cells found both HCQ and CQ to be effective in inhibiting viral replication, with HCQ being more potent [369]. Additionally, an early case study of three COVID-19 patients reported the presence of antiphospholipid antibodies in all three patients [98]. Antiphospholipid antibodies are central to the diagnosis of the antiphospholipid syndrome, a disorder that HCQ has often been used to treat [370,371,372]. Because the 90% effective concentration (EC90) of CQ in Vero E6 cells (6.90 μM) can be achieved in and tolerated by rheumatoid arthritis patients, it was hypothesized that it might also be possible to achieve the effective concentration in COVID-19 patients [373]. Additionally, HCQ has been found to be effective in treating HIV [374] and chronic Hepatitis C [375]. Together, these studies triggered initial enthusiasm about the therapeutic potential for HCQ and CQ against COVID-19. HCQ/CQ has been proposed both as a treatment for COVID-19 and a prophylaxis against SARS-CoV-2 exposure. HCQ/CQ are sometimes administered with azithromycin (AZ) and/or zinc supplementation.

5.4.2.2.1 Therapeutic Administration of HCQ/CQ

The initial study evaluating HCQ as a treatment for COVID-19 patients was published on March 20, 2020 by Gautret et al. [376]. This non-randomized, non-blinded, non-placebo clinical trial compared HCQ to standard of care (SOC) in 42 hospitalized patients in southern France. They reported that patients who received HCQ showed higher rates of virological clearance by nasopharyngeal swab on Days 3-6 when compared to standard care. This study also treated six patients with both HCQ + AZ and found this combination therapy to be more effective than HCQ alone. However, the design and analyses used showed weaknesses that severely limit interpretability of results, including the lack of randomization, lack of blinding, lack of placebo, lack of Intention-To-Treat analysis, lack of correction for sequential multiple comparisons, trial arms entirely confounded by hospital, false negatives in outcome measurements, lack of trial pre-registration, and small sample size. Two of these weaknesses are due to inappropriate data analysis and can therefore be corrected post hoc by recalculating the p-values (lack of Intention-To-Treat analysis and multiple comparisons). However, all other weaknesses are fundamental design flaws and cannot be corrected for. Thus, conclusions cannot be generalized outside of the study. The International Society of Antimicrobial Chemotherapy, the scientific organization that publishes International Journal of Antimicrobial Agents where the article appeared, has announced that the article does not meet its expected standard for publications [377], although it has not been officially retracted. Because of the preliminary data presented in this study, the use of HCQ in COVID-19 treatment was subsequently explored by other researchers. About one week later, a follow-up case study reported that 11 consecutive patients were treated with HCQ + AZ using the same dosing regimen [378]. One patient died, two were transferred to the ICU, and one developed a prolonged QT interval, leading to discontinuation of HCQ + AZ. As in the Gautret et al. study, the outcome assessed was virological clearance at Day 6 post-treatment, as measured in nasopharyngeal swabs. Of the ten living patients on Day 6, eight remained positive for SARS-CoV-2 RNA. Like in the original study, interpretability was severely limited by the lack of a comparison group and the small sample size. However, these results stand in contrast to the claims by Gautret et al. that all six patients treated with HCQ + AZ tested negative for SARS-CoV-2 RNA by Day 6 post-treatment. This case study illustrated the need for further investigation using robust study design to evaluate the efficacy of HCQ and/or CQ.

On April 10, 2020, a randomized, non-placebo trial of 62 COVID-19 patients at the Renmin Hospital of Wuhan University was released [379]. This study investigated whether HCQ decreased time to fever break or time to cough relief when compared to SOC [379]. This trial found HCQ decreased both average time to fever break and average time to cough relief, defined as mild or no cough. While this study improved on some of the methodological flaws in Gautret et al. by randomizing patients, it also had several flaws in trial design and data analysis that prevent generalization of the results. These weaknesses include the lack of placebo, lack of correction for multiple primary outcomes, inappropriate choice of outcomes, lack of sufficient detail to understand analysis, drastic disparities between pre-registration and published protocol, and small sample size. The choice of outcomes may be inappropriate as both fevers and cough may break periodically without resolution of illness. Additionally, for these outcomes, the authors reported that 23 of 62 patients did not have a fever and 25 of 62 patients did not have a cough at the start of the study, but the authors failed to describe how these patients were included in a study assessing time to fever break and time to cough relief. It is important to note here that the authors claimed “neither the research performers nor the patients were aware of the treatment assignments.” This blinding seems impossible in a non-placebo trial because at the very least, providers would know whether they were administering a medication or not, and this knowledge could lead to systematic differences in the administration of care. Correction for multiple primary outcomes can be adjusted post hoc by recalculating p-values, but all of the other issues were design and statistical weaknesses that cannot be corrected for. Additionally, the observation of drastic disparities between pre-registration and published protocol could indicate p-hacking [380]. The design limitations mean that the conclusions cannot be generalized outside of the study. A second randomized trial, conducted by the Shanghai Public Health Clinical Center, analyzed whether HCQ increased rates of virological clearance at day 7 in respiratory pharyngeal swabs compared to standard care [381]. This trial was published in Chinese along with an abstract in English, and only the English abstract was read and interpreted for this review. The trial found comparable outcomes in virological clearance rate, time to virological clearance, and time to body temperature normalization between the treatment and control groups. A known weakness is small sample size, with only 30 patients enrolled and 15 in each arm. This problem suggests the study is underpowered to detect potentially useful differences and precludes interpretation of results. Additionally, because only the abstract could be read, other design and analysis issues could be present. Thus, though these studies added randomization to their assessment of HCQ, their conclusions should be interpreted very cautiously. These two studies assessed different outcomes and reached differing conclusions about the efficacy of HCQ for treating COVID-19; the designs of both studies, especially with respect to sample size, meant that no general conclusions can be made about the efficacy of the drug.

Several widely reported studies on HCQ have issues with data integrity and/or provenance. A Letter to the Editor published in BioScience Trends on March 16, 2020 claimed that numerous clinical trials have shown that HCQ is superior to control treatment in inhibiting the exacerbation of COVID-19 pneumonia [382]. This letter has been cited by numerous primary literature, review articles, and media alike [383,384]. However, the letter referred to 15 pre-registration identifiers from the Chinese Clinical Trial Registry. When these identifiers are followed back to the registry, most trials claim they are not yet recruiting patients or are currently recruiting patients. For all of these 15 identifiers, no data uploads or links to publications could be located on the pre-registrations. At the very least, the lack of availability of the primary data means the claim that HCQ is efficacious against COVID-19 pneumonia cannot be verified. Similarly, a recent multinational registry analysis [385] analyzed the efficacy of CQ and HCQ with and without a macrolide, which is a class of antibiotics that includes Azithromycin, for the treatment of COVID-19. The study observed 96,032 patients split into a control and four treatment conditions (CQ with and without a macrolide; HCQ with and without a macrolide). They concluded that treatment with CQ and HCQ was associated with increased risk of de novo ventricular arrhythmia during hospitalization. However, this study has since been retracted by The Lancet due to an inability to validate the data used [386]. These studies demonstrate that increased skepticism in evaluation of the HCQ/CQ and COVID-19 literature may be warranted, possible because of the significant attention HCQ and CQ have received as possible treatments for COVID-19 and the politicization of these drugs.

Despite the fact that the study suggesting that CQ/HCQ increased risk of ventricular arrhythmia in COVID-19 patients has now been retracted, previous studies have identified risks associated with HCQ/CQ. A patient with systemic lupus erythematosus developed a prolonged QT interval that was likely exacerbated by use of HCQ in combination with renal failure [387]. A prolonged QT interval is associated with ventricular arrhythmia [388]. Furthermore, a separate study [389] investigated the safety associated with the use of HCQ with and without macrolides between 2000 and 2020. The study involved 900,000 cases treated with HCQ and 300,000 cases treated with HCQ + AZ. The results indicated that short-term use of HCQ was not associated with additional risk, but that HCQ + AZ was associated with an enhanced risk of cardiovascular complications (15-20% increased risk of chest pain) and a two-fold increased risk of mortality. Therefore, whether studies utilize HCQ alone or HCQ in combination with a macrolide may be an important consideration in assessing risk. As results from initial investigations of these drug combinations have emerged, concerns about the efficacy and risks of treating COVID-19 with HCQ and CQ has led to the removal of CQ/HCQ from SOC practices in several countries [390,391]. As of May 25, 2020, WHO had suspended administration of HCQ as part of the worldwide Solidarity Trial [392], and later the final results of this large-scale trial that compared 947 patients administered HCQ to 906 controls revealed no effect on the primary outcome, mortality during hospitalization (rate ratio: 1.19; P=0.23)

As additional research has emerged, HCQ/CQ have increasingly been demonstrated to be ineffective against COVID-19 and to carry a number of significant side effects. A randomized, open-label, non-placebo trial of 150 COVID-19 patients was conducted in parallel at 16 government-designated COVID-19 centers in China to assess the safety and efficacy of HCQ [393]. The trial compared treatment with HCQ in conjunction with SOC to SOC alone in 150 infected patients who were assigned randomly to the two groups (75 per group). The primary endpoint of the study was the negative conversion rate of SARS-CoV-2 in 28 days, and the investigators found no difference in this parameter between the groups. The secondary endpoints were an amelioration of the symptoms of the disease such as axillary temperature ≤36.6°C, SpO2 >94% on room air, and disappearance of symptoms like shortness of breath, cough, and sore throat. The median time to symptom alleviation was similar across different conditions (19 days in HCQ+SOC vs. 21 days in SOC). Additionally, 30% of the patients receiving SOC+HCQ reported adverse outcomes compared to 8.8% of patients receiving only SOC, with the most common adverse outcome in the SOC+HCQ group being diarrhea (10% vs. 0% in the SOC group, p=0.004). However, there are several factors that limit the interpretability of this study. Most of the enrolled patients had mild-to-moderate symptoms (98%), and the average age was 46. SOC in this study included the use of antivirals (Lopinavir-Ritonavir, Arbidol, Oseltamivir, Virazole, Entecavir, Ganciclovir, and Interferon alfa), which appeared to introduce confounding effects. Thus, to isolate the effect of HCQ, SOC would need to exclude the use of antivirals. In this trial, the samples used to test for the presence of the SARS-CoV-2 virus were collected from the upper respiratory tract, and the authors indicated that the use of upper respiratory samples may have introduced false negatives (e.g., [78]). Another limitation of the study that the authors acknowledge was that the HCQ treatment began, on average, at a 16-day delay from the symptom onset. The fact that this study was open-label and lacked a placebo limits interpretation, and additional analysis is required to determine whether HCQ reduces inflammatory response. Therefore, despite some potential areas of investigation identified in post hoc analysis, this study cannot be interpreted as providing support for HCQ as a therapeutic against COVID-19.

Additional evidence comes from a retrospective analysis [394] that examined data from 368 COVID-19 patients across all United States Veteran Health Administration medical centers. The study retrospectively investigated the effect of the administration of HCQ (n=97), HCQ + AZ (n=113), and no HCQ (n=158) on 368 patients. The primary outcomes assessed were death and the need for mechanical ventilation. Standard supportive care was rendered to all patients. Due to the low representation of women (N=17) in the available data, the study included only men, and the median age was 65 years. The rate of death in the HCQ-only treatment condition was 27.8% and in the HCQ + AZ treatment condition, it was 22.1%. In comparison to the 14.1% rate of death in the no-HCQ cohort, these data indicated a statistically significant elevation in the risk of death for the HCQ-only group compared to the no-HCQ group (adjusted HR: 2.61, p=0.03), but not for the HCQ + AZ group compared to the no-HCQ group (adjusted HR: 1.14; p=0.72). Further, the risk of ventilation was similar across all three groups (adjusted HR: 1.43, p=0.48 (HCQ) and 0.43, p=0.09 (HCQ + AZ) compared to no HCQ). The study thus showed evidence of an association between increased mortality and HCQ in this cohort of COVID-19 patients but no change in rates of mechanical ventilation among the treatment conditions. The study had a few limitations: it was not randomized, and the baseline vital signs, laboratory tests, and prescription drug use were significantly different among the three groups. All of these factors could potentially influence treatment outcome. Furthermore, the authors acknowledge that the effect of the drugs might be different in females and pediatric subjects, since these subjects were not part of the study. The reported result that HCQ + AZ is safer than HCQ contradicts the findings of the previous large-scale analysis of twenty years of records that found HCQ + AZ to be more frequently associated with cardiac arrhythmia than HCQ alone [389]; whether this discrepancy is caused by the pathology of COVID-19, is influenced by age or sex, or is a statistical artifact is not presently known.

Finally, findings from the Randomized Evaluation of COVID-19 Therapy (RECOVERY) trial were released on October 8, 2020. This study used a randomized, open-label design to study the effects of HCQ compared to standard care at 176 hospitals in the United Kingdom [395]. This large study enrolled 11,197 hospitalized patients whose physicians believed it would not harm them to participate. Patients were randomized into either the control group or one of the treatment arms, with twice as many patients enrolled in the control group as any treatment group. Of the patients eligible to receive HCQ, 1,561 were randomized into the HCQ arm while 3,155 were randomized into the control arm. The demographics of the HCQ and control groups were similar in terms of average age (65 years), proportion female (approximately 38%), ethnic make-up (73% versus 76% white), and prevalence of pre-existing conditions (56% versus 57% overall). In the HCQ arm of the study, patients received 800 mg at baseline and again after 6 hours, then 400 mg at 12 hours and every subsequent 12 hours. The primary outcome analyzed was all-cause mortality, and patient vital statistics were reported by physicians upon discharge or death, or else at 28 days following HCQ administration if they remained hospitalized. The secondary outcome assessed was the combined risk of progression to invasive mechanical ventilation or death within 28 days. By the advice of an external data monitoring committee, the HCQ arm of the study was reviewed early, leading to it being closed due a lack of support for HCQ as a treatment for COVID-19. Patients who received HCQ had a longer duration of hospitalization than patients receiving usual care, were less likely to be discharged alive within 28 days, and were more likely to progress to mechanical ventilation. This large-scale study thus builds upon studies in the United States and China to suggest that HCQ is not an effective treatment, and in fact may negatively impact COVID-19 patients due to its side effects. The rates of COVID-19-related reported in the RECOVERY trial did not differ between the control and HCQ arms, but patients receiving HCQ were more likely to die due to cardiac events. Therefore, though none of the studies have been blinded, taken together it becomes clear that all of the available evidence points to significant concerns about the administration of HCQ to hospitalized COVID-19 patients, without providing any support for its efficacy.

5.4.2.2.2 HCQ for the Treatment of Mild Cases

One additional possible therapeutic application considered was the treatment of mild COVID-19 cases in otherwise healthy individuals. This possibility was assessed in a randomized, open-label, multi-center analysis conducted in Catalonia (Spain) [396]. This analysis enrolled adults 18 and older who had been experiencing mild symptoms of COVID-19 for fewer than five days. Participants were randomized into an HCQ arm (N=136) and a control arm (N=157), and those in the treatment arm were administered 800 mg of HCQ on the first day of treatment followed by 400 mg on each of the subsequent six days. The primary outcome assessed was viral clearance at days 3 and 7 following the onset of treatment, and secondary outcomes were clinical progression and time to complete resolution of symptoms. They found no significant differences between the two groups. This study thus suggests that HCQ does not improve recovery from COVID-19, even in otherwise healthy adult patients with mild symptoms.

5.4.2.2.3 Prophylactic Administration of HCQ

An initial study of the possible prophylactic application of HCQ utilized a randomized, double-blind, placebo-controlled design to analyze the administration of HCQ prophylactically [397]. Asymptomatic adults in the United States and Canada who had been exposed to SARS-CoV-2 within the past four days were enrolled in an online study to see whether administration of HCQ over five days would influence the probability of developing COVID-19 symptoms over a 14-day period. Of the participants, 414 received HCQ and 407 received a placebo. The participants averaged 40 years of age and 51.6% were women. They found no significant difference in the rate of symptomatic illness between the two groups (11.8% HCQ, 14.3% placebo, p=0.35). The HCQ condition was associated with side effects, with 40.1% of patients reporting side effects compared to 16.8% in the control group (p<0.001). However, likely due to the high enrollment of healthcare workers (66% of participants) and the well-known side effects associated with HCQ, a large number of participants were able to correctly identify whether they were receiving HCQ or a placebo (46.5% and 35.7%, respectively). Furthermore, due to a lack of availability of diagnostic testing, only 20 of the 107 cases were confirmed with a PCR-based test to be positive for SARS-CoV-2. The rest were categorized as “probable” or “possible” cases by a panel of four physicians who were blind to the treatment status. One possible confounder is that a patient presenting one or more symptoms, which included diarrhea, was defined as a “possible” case, but diarrhea is also a common side effect of HCQ. Additionally, four of the twenty PCR-confirmed cases did not develop symptoms until after the observation period had completed, suggesting that the 14-day trial period may not have been long enough or that some participants also encountered secondary exposure events. Finally, in addition to the young age of the participants in this study, which ranged from 32 to 51, there were possible impediments to generalization introduced by the selection process, as 2,237 patients who were eligible but had already developed symptoms by day 4 were enrolled in a separate study. It is therefore likely that asymptomatic cases were over-represented in this sample, which would not have been detected based on the diagnostic criteria used. Therefore, while this study does represent the first effort to conduct a randomized, double-blind, placebo-controlled investigation of HCQ’s effect on COVID-19 symptoms in a large sample, the lack of PCR tests and several other design flaws significantly impede interpretation of the results. However, in line with the results from therapeutic studies, there is no evidence that HCQ provides protection against COVID-19.

A second study [398] examined the effect of administering HCQ to healthcare workers as a pre-exposure prophylactic. The primary outcome assessed was the conversion from SARS-CoV-2 negative to SARS-CoV-2 positive status over the 8 week study period. This study was also randomized, double-blind, and placebo-controlled and sought to address some of the limitations of the first prophylactic study. They aimed to enroll 200 healthcare workers, preferentially those working with COVID-19 patients, at two hospitals within the University of Pennsylvania hospital system in Philadelphia, PA. Participants were randomized 1:1 to receive either 600 mg of HCQ daily or a placebo, and their SARS-CoV-2 infection status and antibody status were assessed using RT-PCR and serological testing, respectively, at baseline, 4 weeks, and 8 weeks following the beginning of the treatment period. The statistical design of the study accounted for interim analyses at 50 and 100 participants in case efficacy or futility of HCQ for prophylaxis became clear earlier than completion of enrollment. The 139 individuals enrolled comprised a study population that was fairly young (average age 33) and made of largely of people who were white, women, and without pre-existing conditions. At the second interim analysis, more individuals in the treatment group than the control group had contracted COVID-19 (4 versus 3), causing the estimate z-score to fall below the pre-established threshold for futility. As a result, the trial was terminated early, offering additional evidence against the use of HCQ for prophylaxis.

5.4.2.2.4 Summary of HCQ/CQ Research Findings

Early in vitro evidence indicated that HCQ could be an effective therapeutic against SARS-CoV-2 and COVID-19, leading to early interest in its potential as both a therapeutic and prophylaxis. Initially it was hypothesized that CQ/HCQ might be effective against SARS-CoV-2, especially since CQ and HCQ have both been found to inhibit the expression of CD154 in T-cells and to reduce TLR signaling that leads to the production of pro-inflammatory cytokines [399]. Clinical trials for COVID-19 have more often used HCQ rather than CQ because it offers the advantages of being cheaper and having fewer side effects than CQ. However, a large analysis of patients receiving HCQ from January 2000 through March 2020 reported that the combination of HCQ and azithromycin, but not other macrolides, was associated with an elevated risk of cardiovascular complications and mortality. Multiple clinical studies have already been carried out to assess HCQ as a therapeutic agent for COVID-19, and many more are in progress. To date, none of these studies have used randomized, double-blind, placebo-controlled designs with a large sample size, which would be the gold standard. Despite the design limitations (which would be more likely to produce false positives than false negatives), initial optimism about HCQ has not been substantiated by these studies. The most methodologically rigorous HCQ-focused analysis to date [397] evaluated the prophylactic potential of HCQ and found no significant differences between the treatment and control groups, and the WHO’s global Solidarity trial similarly reported no effect of HCQ on mortality. Thus, HCQ/CQ are not likely to be effective therapeutic or prophylactic agents against COVID-19. Additionally, one study identified an increased risk of mortality in older men receiving HCQ, and administration of HCQ and HCQ+AZ did not decrease the use of mechanical ventilation in these patients [394]. HCQ use for COVID-19 could also lead to shortages for anti-malarial or anti-rheumatic use, where it has been definitively proven to be effective. Despite significant attention, these drugs appear to be ineffective against COVID-19. Several countries have now removed CQ from their SOC for COVID-19 due to the lack of evidence of efficacy and the frequency of adverse effects.

5.4.2.3 Dexamethasone

Dexamethasone (9α-fluoro-16α-methylprednisolone) is a synthetic corticosteroid that binds to glucocorticoid receptors [400,401]. It was first synthesized in the late 1950s as an anti-inflammatory and has been used to treat rheumatoid arthritis and other inflammatory conditions [402,403]. Steroids such as dexamethasone are widely available and affordable, and they are often used to treat community-acquired pneumonia [404]. A clinical trial that began in 2012 recently reported that dexamethasone may improve outcomes for patients with ARDS [405]. However, meta-analysis of a small amount of available data regarding the use of dexamethasone to treat SARS suggested that it may, in fact, be associated with patient harm [406], although these findings may have been biased by the fact that all of the studies examined were observational and a large number of inconclusive studies were not included [407]. Dexamethasone works as an anti-inflammatory by binding to glucocorticoid receptors with higher affinity than endogenous cortisol [408]. In order to understand how dexamethasone works to reduce inflammation, it is necessary to consider the stress response more broadly. In response to stress, corticotropin‐releasing hormone stimulates the release of neurotransmitters known as catecholamines, such as epinephrine, and steroid hormones known as glucocorticoids, such as cortisol [409,410]. While catecholamines are often associated with the fight-or-flight response, the specific role that glucocorticoids play is less clear, although they are thought to be important to restoring homeostasis [411]. Immune challenge is a stressor that is known to interact closely with the stress response. The immune system can therefore interact with the central nervous system; for example, macrophages can both respond to and produce catecholamines [409]. Additionally, the production of both catecholamines and glucocorticoids is associated with inhibition of proinflammatory cytokines such as IL-6, IL-12, and tumor necrosis factor-α (TNF‐α) and the stimulation of anti-inflammatory cytokines such as IL-10, meaning that the stress response can regulate inflammatory immune activity [410]. Administration of dexamethasone has been found to correspond to dose-dependent inhibition of IL-12 production, but not to affect IL-10 [412]; the fact that this relationship could be disrupted by administration of a glucocorticoid-receptor antagonist suggests that it is regulated by the receptor itself [412]. Thus, the administration of dexamethasone for COVID-19 is likely to simulate the release of glucocorticoids endogenously during stress, resulting in binding of the synthetic steroid to the glucocorticoid receptor and the associated inhibition of the production of proinflammatory cytokines. In this model, dexamethasone reduces inflammation by stimulating the mechanism that reduces inflammation following a threat such as immune challenge. Immunosuppressive drugs such as steroids are typically contraindicated in the setting of infection [413], but because COVID-19 results in hyperinflammation that appears to contribute to mortality via lung damage, immunosuppression may be a helpful approach to treatment [122]. The decision of whether and/or when to counter hyperinflammation with immunosuppression in the setting of COVID-19 was an area of intense debate, as the risks of inhibiting antiviral immunity needed to be weighed against the beneficial anti-inflammatory effects [414]. As a result, guidelines early in the pandemic typically recommended avoiding treating COVID-19 patients with corticosteroids such as dexamethasone [406].

The application of dexamethasone for the treatment of COVID-19 was evaluated as part of the multi-site RECOVERY trial in the United Kingdom [415]. Over 6,000 hospitalized COVID-19 patients were assigned into standard care or treatment (dexamethasone) arms of the trial with a 2:1 ratio. At the time of randomization, some patients were ventilated (16%), others were on non-invasive oxygen (60%), and others were breathing independently (24%). Patients in the treatment arm were administered dexamethasone either orally or intravenously at 6 mg per day for up to 10 days. The primary end-point was the patient’s status at 28-days post-randomization (mortality, discharge, or continued hospitalization), and secondary outcomes analyzed included the progression to invasive mechanical ventilation over the same period. The 28-day mortality rate was found to be lower in the treatment group than in the standard care group (21.6% vs 24.6%, p <0.001). However, this finding was driven by differences in mortality among patients who were receiving mechanical ventilation or supplementary oxygen at the start of the study. The report indicated that dexamethasone reduced 28-day mortality relative to standard care in patients who were ventilated (29.3% vs. 41.4%) and among those who were receiving oxygen supplementation (23.3% vs. 26.2%) at randomization, but not in patients who were breathing independently (17.8% vs. 14.0%). One possible confounder is that patients receiving mechanical ventilation tended to be younger than patients who were not receiving respiratory support (by 10 years on average) and to have had symptoms for a longer period. However, adjusting for age did not change their conclusions, although the duration of symptoms was found to be significantly associated with the effect of dexamethasone administration. These findings also suggested that dexamethasone may have reduced progression to mechanical ventilation, especially among patients who were receiving oxygen support at randomization. Thus, this large, randomized, and multi-site, albeit not placebo-controlled, study suggests that administration of dexamethasone to patients who are unable to breathe independently may significantly improve survival outcomes. Additionally, dexamethasone is a widely available and affordable medication, raising the hope that it could be made available to COVID-19 patients globally.

The results of the RECOVERY trial’s analysis of dexamethasone suggest that this therapeutic is effective primarily in patients who had been experiencing symptoms for at least seven days and patients who were not breathing independently [416]. Thus, it seems likely that dexamethasone is useful for treating inflammation associated with immunopathy or cytokine release syndrome. In fact, corticosteroids such as dexamethasone are sometimes used to treat CRS [417]. It is not surprising that administration of an immunosuppressant would be most beneficial when the immune system was dysregulated towards inflammation. However, it is also unsurprising that care must be taken in administering an immunosuppressant to patients fighting a viral infection. In particular, the concern has been raised that treatment with dexamethasone might increase patient susceptibility to concurrent (e.g., nosocomial) infections [418]. Additionally, the drug could potentially slow viral clearance and inhibit patients’ ability to develop antibodies to SARS-CoV-2 [406,418]. Furthermore, dexamethasone has been associated with side effects that include psychosis, glucocorticoid-induced diabetes, and avascular necrosis [406], and the RECOVERY trial did not report outcomes with enough detail to be able to determine whether they observed similar complications. However, since these results were released, strategies have been proposed for administering dexamethasone alongside more targeted treatments to minimize the likelihood of negative side effects [418]. Given the available evidence, dexamethasone is currently the most promising treatment for severe COVID-19.

5.5 Biologics

Biologics are produced from components of living organisms or viruses. They include treatments such as humanized monoclonal antibodies (mAb) tocilizumab (TCZ), and neutralizing antibodies (nAbs), as well as prophylactics such as vaccines. Historically produced from animal tissue, biologics have become increasingly feasible to produce as recombinant DNA technologies have advanced [419]. Often, they are glycoproteins or peptides [420], but whole viruses can also be used therapeutically or prophylactically, not only for vaccines but also as vectors for gene therapy or therapeutic proteins or for oncolytic virotherapy [421]. They are typically catabolized by the body to their amino acid components [420]. There are many differences on the development side between biologics and synthesized pharmaceuticals, such as small molecule drugs. Biologics are typically orders of magnitude larger than small molecule drugs and their physiochemical properties are often much less understood [420]. They are often heat sensitive, and their toxicity can vary, as it is not directly associated with the primary effects of the drug [420]. However, this class includes some extremely significant medical breakthroughs, including insulin for the management of diabetes and the smallpox vaccine. As a result, it is another possible avenue through which the pharmacological management of SARS-CoV-2 infection can be explored.

5.5.1 Tocilizumab

TCZ is a receptor antibody that was developed to manage chronic inflammation caused by the continuous synthesis of the cytokine IL-6 [422]. IL-6 is a pro-inflammatory cytokine belonging to the interleukin family, which is comprised by immune system regulators that are primarily responsible for immune cell differentiation. Often used to treat conditions such as rheumatoid arthritis [422], TCZ has become a pharmaceutical of interest for the treatment of COVID-19 because of the role IL-6 plays in this disease. While secretion of IL-6 can be associated with chronic conditions, it is a key player in the innate immune response and is secreted by macrophages in response to the detection of pathogen-associated molecular patterns (PAMPs) and damage-associated molecular patterns (DAMPs) [422]. An analysis of 191 in-patients at two Wuhan hospitals revealed that blood concentrations of IL-6 differed between patients who did and did not recover from COVID-19. Patients who ultimately deceased had higher IL-6 levels at admission than those who recovered [35]. Additionally, IL-6 levels remained higher throughout the course of hospitalization in the patients who ultimately deceased [35]. This finding provided some early evidence that COVID-19 deaths may be induced by the hyperactive immune response, often referred to as cytokine release syndrome (CRS) or cytokine storm syndrome (CSS), as IL-6 plays a key role in this response [110]. In this context, the observation of elevated IL-6 in patients who died may reflect an over-production of proinflammatory interleukins, suggesting that TCZ may palliate some of the most severe symptoms of COVID-19 associated with increased cytokine production.

Human IL-6 is a 26-kDa glycoprotein that consists of 184 amino acids and contains two potential N-glycosylation sites and four cysteine residues. It binds to a type I cytokine receptor (IL-6Rα or glycoprotein 80) that exists in both membrane-bound (IL-6Rα) and soluble (sIL-6Rα) forms [423]. It is not the binding of IL-6 to the receptor that initiates pro- and/or anti-inflammatory signaling, but rather the binding of the complex to another subunit, known as IL-6Rβ or glycoprotein 130 (gp13) [423,424]. Unlike membrane-bound IL-6Rα, which is only found on hepatocytes and some types of leukocytes, gp130 is found on most cells [425]. When IL-6 binds to sIL-6Rα, the complex can then bind to a gp130 protein on any cell [425]. The binding of IL-6 to IL-6Rα is termed classical signaling, while its binding to sIL-6Rα is termed trans-signaling [425,426,427]. These two signaling processes are thought to play different roles in health and illness. For example, trans-signaling may play a role in the proliferation of mucosal T-helper TH2 cells associated with asthma, while an earlier step in this proliferation process may be regulated by classical signaling [425]. Similarly, IL-6 is known to play a role in Crohn’s Disease via trans-, but not classical, signaling [425]. Both classical and trans-signaling can occur through three independent pathways: the Janus-activated kinase-STAT3 pathway, the Ras/Mitogen-Activated Protein Kinases (MAPK) pathway and the Phosphoinositol-3 Kinase/Akt pathway [423]. These signaling pathways are involved in a variety of different functions, including cell type differentiation, immunoglobulin synthesis, and cellular survival signaling pathways, respectively [423]. The ultimate result of the IL-6 cascade is to direct transcriptional activity of various promoters of pro-inflammatory cytokines, such as IL-1, TFN, and even IL-6 itself, through the activity of NF-κB [423]. IL-6 synthesis is tightly regulated both transcriptionally and post-transcriptionally, and it has been shown that viral proteins can enhance transcription of the IL-6 gene by strengthening the DNA-binding activity between several transcription factors and IL-6 gene-cis-regulatory elements [428]. Therefore, drugs inhibiting the binding of IL-6 to IL-6Rα or sIL-6Rα are of interest for combating the hyperactive inflammatory response characteristic of CRS and CSS. TCZ is a humanized monoclonal antibody that binds both to the insoluble and soluble receptor of IL-6, providing de facto inhibition of the IL-6 immune cascade.

Tocilizumab is being administered either as an intervention or as concomitant medication in 84 COVID-19 clinical trials (Figure 3). No randomized, placebo-controlled studies of TCZ have currently released results. Therefore, no conclusions can be drawn about its efficacy for the treatment of COVID-19. However, early interest in TCZ as a possible treatment for COVID-19 emerged from a very small retrospective study in China that examined 20 patients with severe symptoms in early February 2020 and reported rapid improvement in symptoms following treatment with TCZ [429]. Subsequently, a number of retrospective studies have been conducted in several countries. Many studies use a retrospective, observational design, where they compare outcomes for COVID-19 patients who received TCZ to those who did not over a set period of time. For example, one of the largest retrospective, observational analysis released to date [430] compared the rates at which patients who received TCZ deceased or progressed to invasive medical ventilation over a 14-day period compared to patients receiving only SOC. Under this definition, SOC could include other drugs such as HCQ, azithromycin, lopinavir-ritonavir or darunavir-cobicistat, or heparin. While this study was not randomized, a subset of patients who were eligible to receive TCZ were unable to obtain it due to shortages; however, these groups were not directly compared in the analysis. After adjusting for variables such as age, sex, and SOFA score, they found that patients treated with TCZ were less likely to progress to invasive medical ventilation and/or death (adjusted HR = 0.61, CI 0.40-0.92, p=0.020), although analysis of death and ventilation separately suggests that this effect may have been driven by differences in the death rate (20% of control versus 7% of TCZ-treated patients). They reported particular benefits for patients whose PaO2/FiO2 ratio, also known as the Horowitz Index for Lung Function, fell below a 150 mm Hg threshold. They found no differences between groups administered subcutaneous versus intravenous TCZ.

Another retrospective observational analysis of interest examined the charts of patients at a hospital in Connecticut, USA where 64% of all 239 COVID-19 patients in the study period were administered TCZ based on assignment by a standardized algorithm [431]. They found that TCZ administration was associated with more similar rates of survivorship in patients with severe versus nonsevere COVID-19 at intake, defined based on the amount of supplemental oxygen needed. They therefore proposed that their algorithm was able to identify patients presenting with or likely to develop CRS as good candidates for TCZ. This study also reported higher survivorship in Black and Hispanic patients compared to white patients when adjusted for age. The major limitation with interpretation for these studies is that there may be clinical characteristics that influenced medical practitioners decisions to administer TCZ to some patients and not others. One interesting example therefore comes from an analysis of patients at a single hospital in Brescia, Italy, where TCZ was not available for a period of time [432]. This study compared COVID-19 patients admitted to the hospital before and after March 13, 2020, when the hospital received TCZ. Therefore, patients who would have been eligible for TCZ prior to this arbitrary date did not receive it as treatment, making this retrospective analysis something of a natural experiment. Despite this design, demographic factors did not appear to be consistent between the two groups, and the average age of the control group was older than the TCZ group. The control group also had a higher percentage of males and a higher incidence of comorbidities such as diabetes and heart disease. All the same, the multivariate HR, which adjusted for these clinical and demographic factors, found a significant difference between survival in the two groups (HR=0.035, CI=0.004-0.347, p = 0.004). They reported improvement of survival outcomes after the addition of TCZ to their SOC regime, with 11 of 23 patients (47.8%) admitted prior to March 13th dying compared to 2 of 62 (3.2%) admitted afterwards. They also reported a reduced progression to mechanical ventilation in the TCZ group. However, this study also holds a significant limitation: the time delay between the two groups means that knowledge about how to treat the disease likely improved over this timeframe as well. All the same, the results of these observational retrospective studies provide support for TCZ as a pharmaceutical of interest for follow-up in clinical trials.

In addition to the retrospective observational studies, other analysis have utilized a retrospective case-control design to match pairs of patients with similar baseline characteristics, only one of whom received TCZ for COVID-19. In one such study, TCZ was significantly associated with a reduced risk of progression to ICU admission or death [433]. This study examined only 20 patients treated with TCZ (all but one of the patients treated with TCZ in the hospital during the study period) and compared them to 25 patients receiving SOC. For the combined primary endpoint of death and/or ICU admission, only 25% of patients receiving TCZ progressed to an endpoint compared to 72% in the SOC group (p = 0.002, presumably based on a chi-square test based on the information provided in the text). When the two endpoints were examined separately, progression to invasive medical ventilation remained significant (32% SOC compared to 0% TCZ, p = 0.006) but not for mortality (48% SOC compared to 25% TCZ, p = 0.066). In contrast, a study that compared 96 patients treated with TCZ to 97 patients treated with SOC only in New York City found that differences in mortality did not differ between the two groups, but that this difference did become significant when intubated patients were excluded from the analysis [434]. Taken together, these findings suggest that future clinical trials of TCZ may want to include intubation as an endpoint. However, these studies should be approached with caution, not only because of the small number of patients enrolled and the retrospective design, but also because they performed a large number of statistical tests and did not account for multiple hypothesis testing. These last findings highlight the need to search for a balance between impairing a harmful immune response, such as the one generated during CRS and CSS, and preventing the worsening of the clinical picture of the patients by potential new viral infections.

Though data about TCZ for COVID-19 is still only just emerging, some meta-analyses and systematic reviews have investigated the available data. One meta-analysis [435] evaluated 19 studies published or released as preprints prior to July 1, 2020 and found that the overall trends supportive of the frequent conclusion that TCZ does improve survivorship, with a significant HR of 0.41. This trend improved when they excluded studies that administered a steroid alongside TCZ, with a significant HR of 0.04. They also found some evidence for reduced IMV or ICU admission, but only when excluding all studies except a small number reporting estimates that were adjusted for possible bias introduced by the challenges of stringency during the enrollment process. A systematic analysis of nine case-control studies estimated a HR of 0.482, which was also significant [436]. Although these estimates are similar, it is important to note that they are drawing from the same literature and are therefore likely to be affected by the same biases in publication. A second systematic review of studies investigating TCZ treatment for COVID-19 analyzed 31 studies that had been published or released as pre-prints and reported that none carried a low risk of bias (RoB) [437]. Therefore, the present evidence is not likely to be sufficient for conclusions about the efficacy of TCZ.

Additionally, there are possible risks associated with the administration of TCZ for COVID-19. TCZ has been used for over a decade to treat rheumatoid arthritis [438], and a recent study found the drug to be safe for pregnant and breastfeeding women [439]. However, TCZ may increase the risk of developing infections [438], and rheumatoid arthritis patients with chronic hepatitis B infections had a high risk of hepatitis B virus reactivation when TCZ was administered in combination with other RA drugs [440]. As a result, TCZ is contraindicated in patients with active infections such as tuberculosis [441]. Previous studies have investigated, with varying results, a possible increased risk of infection in rheumatoid arthritis patients administered TCZ [442,443], although another study reported that the incidence rate of infections was higher in clinical practice rheumatoid arthritis patients treated with TCZ than in the rates reported by clinical trials [444]. In the investigation of 544 Italian COVID-19 patients, the group treated with TCZ was found to be more likely to develop secondary infections, with 24% compared to 4% in the control group [430]. Reactivation of hepatitis B and herpes simplex virus 1 was also reported in a small number of patients in this study, all of whom were receiving TCZ. A July 2020 case report described negative outcomes of two COVID-19 patients after receiving TCZ, including one death; however, both patients were intubated and had entered septic shock prior to receiving TCZ [445], likely indicating a severe level of cytokine production. Additionally, D-dimer and sIL2R levels were reported by one study to increase in patients treated with TCZ, which raised concerns because of the potential association between elevated D-dimer levels and thrombosis and between sIL2R and diseases where T-cell regulation is compromised [431]. An increased risk of bacterial infection was also identified in a systematic review of the literature, based on the unadjusted estimates reported [435]. In summary, TCZ administration to COVID-19 patients is not without risks, may introduce additional risk of developing secondary infections, and should be approached especially cautiously for patients who have latent viral infections.

In summary, approximately 25% of coronavirus patients develop ARDS, which is caused by an excessive early response of the immune system which can be a component of cytokine release syndrome [431] and cytokine storm syndrome [441]. This overwhelming inflammation is triggered by IL-6. TCZ is an inhibitor of IL-6 and therefore may be able to neutralize the inflammatory pathway that leads to the cytokine storm. While the mechanism suggests TCZ may be beneficial for the treatment of COVID-19 patients experiencing excessive immune activity, no randomized controlled trials are available assessing its effect. However, small initial studies have found preliminary indications that TCZ may reduce progression to invasive medical ventilation and/or death. It should be noted that SOC varied widely across retrospective studies, with one study administering HCQ, lopinavir-ritonavir, antibiotics, and/or heparin as part of standard care. Interest in TCZ as a treatment for COVID-19 was supported by two meta-analyses that converged on a HR estimate of approximately 0.45 [435,436], but a third meta-analysis found that all of the available literature carries a risk of bias, with even the largest available TCZ studies to date carrying a moderate risk of bias under the ROBINS-I criteria [437]. Additionally, different studies used different dosages, number of doses, and methods of administration; ongoing research may be needed to optimize administration of TCZ [446], although similar results were reported by one study for intravenous and subcutaneous administration [430]. Clinical trials that are in progress are likely to provide additional insight into the effectiveness of this drug for the treatment of COVID-19 along with how it should be administered.

5.5.2 Neutralizing Antibodies

Monoclonal antibodies have revolutionized the way we treat human diseases. They have become some of the best-selling drugs in the pharmaceutical market in recent years [447]. There are currently 79 FDA approved mAbs on the market including antibodies for viral infections (e.g. Ibalizumab for HIV and Palivizumab for RSV) [447,448]. Virus-specific neutralizing antibodies commonly target viral surface glycoproteins or host structures, thereby inhibiting viral entry through receptor binding interference [449,450]. This section discusses current efforts in developing neutralizing antibodies against SARS-CoV-2 and how expertise gained from previous approaches for MERS-CoV and SARS-CoV-1 may benefit antibody development.

5.5.2.1 Spike (S) Neutralizing Antibody

During the first SARS epidemic in 2002, nAbs were found in SARS-CoV-1-infected patients [451,452]. Several studies following up on these findings identified various S-glycoprotein epitopes as the major targets of neutralizing antibodies against SARS-CoV-1 [453]. The passive transfer of immune serum containing nAbs from SARS-CoV-1-infected mice resulted in protection of naïve mice from viral lower respiratory tract infection upon intranasal challenge [454]. Similarly, a meta-analysis suggested that administration of plasma from recovered SARS-CoV-1 patients reduced mortality upon SARS-CoV-1 infection [455]. Similar results were observed in MERS-CoV infection during the second coronavirus-related epidemic of the 21st century. In these cases, neutralizing antibodies were identified against various epitopes of the receptor binding domain (RBD) of the S glycoprotein [456,457]. Coronaviruses use trimeric spike (S) glycoproteins on their surface to bind to host cell receptors, such as ACE2, allowing for cell entry [63,66]. Each S glycoprotein protomer is comprised of an S1 domain, also called the RBD, and an S2 domain. The S1 domain binds to host cell receptors while the S2 domain facilitates the fusion between the viral envelope and host cell membranes [453]. Although targeting of the host cell receptor ACE2 shows efficacy in inhibiting SARS-CoV-2 infection [74], given the physiological relevance of ACE2 [458], it would be favorable to target virus-specific structures rather than host receptors. This forms the rationale of developing neutralizing antibodies against the S glycoprotein, disrupting its interaction with ACE2 and other receptors and thereby inhibiting viral entry.

The first human neutralizing antibody against SARS-CoV-2 targeting the trimeric spike (S) glycoproteins was developed using hybridoma technology [459], where antibody-producing B-cells developed by mice can be inserted into myeloma cells to produce a hybrid cell line (the hybridoma) that is grown in culture. The 47D11 clone was able to cross-neutralize SARS-CoV-1 and SARS-CoV-2 by a mechanism that is different from receptor binding interference. The exact mechanism of how this clone neutralizes SARS-CoV-2 and inhibits infection in vitro remains unknown, but a potential mechanism might be antibody-induced destabilization of the membrane prefusion structure [459,460]. The ability of this antibody to prevent infection at a feasible dose needs to be validated in vivo, especially since in vitro neutralization effects have been shown to not be reflective of in vivo efficacy [461]. Only a week later, a different group successfully isolated multiple nAbs targeting the RBD of the S glycoprotein from blood samples taken from COVID-19 patients in China [150]. Interestingly, the patient-isolated antibodies did not cross-react with RBDs from SARS-CoV-1 and MERS-CoV, although cross-reactivity to the trimeric spike proteins of SARS-CoV-1 and MERS-CoV was observed. This finding suggests that the RBDs between the three coronavirus species are immunologically distinct and that the isolated nAbs targeting the RBD of SARS-CoV-2 are species specific. While this specificity is desirable, it also raises the question of whether these antibodies are more susceptible to viral escape mechanisms. Viral escape is a common resistance mechanism to nAb therapy due to selective pressure from neutralizing antibodies [462,463]. For HIV, broadly neutralizing antibodies (bnAbs) targeting the CD4 binding site (CD4bs) show greater neutralization breadth than monoclonal antibodies, which target only specific HIV strains [464]. For MERS-CoV, a combination of multiple neutralizing antibodies targeting different antigenic sites prevented neutralization escape [465]. It was found that the different antibody isolates did not target the same epitopes, suggesting that using them in combination might produce a synergistic effect that prevents viral escape [150]. It was also demonstrated that binding affinity of the antibodies does not reflect their capability to compete with ACE2 binding. Furthermore, no conclusions about correlations between the severity of disease and the ability to produce neutralizing antibodies can be drawn at this point. Rather, higher neutralizing antibody titers were more frequently found in patients with severe disease. Correspondingly, higher levels of anti-spike IgG were observed in patients that deceased from infection compared to patient that recovered [466].

Results from the SARS and MERS epidemics thus provide valuable lessons for the design of neutralizing antibodies for the current outbreak. The findings for SARS-CoV-1 and MERS-CoV can aid in identifying which structures constitute suitable targets for nAbs, despite the fact that the RBD appears to be distinct between the three coronavirus species. These studies also suggest that a combination of nAbs targeting distinct antigens might be necessary to provide protection [465]. The biggest challenge remains identifying antibodies that not only bind to their target, but also prove to be beneficial for disease management. On that note, a recently published study indicates that anti-spike antibodies could make the disease worse rather than eliminating the virus [466]. These findings underscores our current lack of understanding the full immune response to SARS-CoV-2.

5.5.3 Interferons

IFNs are a family of cytokines critical to activating the innate immune system response against viral infections. Interferons are classified into three categories based on their receptor specificity: types I, II and III [110]. Specifically, IFNs I (IFN-𝛼 and 𝛽) and II (IFN-𝛾) induce the expression of antiviral proteins that bring the viral RNA to degradation [467]. Among these IFNs, IFN-𝛽 has already been found to strongly inhibit the replication of other coronaviruses, such as SARS-CoV-1, in cell culture, while IFN-𝛼 and 𝛾 were shown to be less effective in this context [467]. There is evidence that patients with higher susceptibility to ARDS indeed show deficiency in IFN-𝛽. For instance, infection with other coronaviruses impairs IFN-𝛽 expression and synthesis, allowing the virus to escape the innate immune response [468]. On March 18 2020, Synairgen plc received approval to start a phase II trial for SNG001, an IFN-𝛽-1a formulation to be delivered to the lungs via inhalation. SNG001 was already shown to be effective in reducing viral load in an in vivo model of swine flu and in vitro models of other coronavirus infections [469]. In July, a press release from Synairgen stated that SNG001 reduced progression to ventilation in a double-blind, placebo-controlled, multi-center study of 101 patients with an average age in the late 50s [470]. They also reported that patients in the treatment group showed greater recovery and lower breathlessness. However, given that this information was released in a press release rather than in a manuscript and thus cannot be thoroughly reviewed, these findings should be considered preliminary. Additionally, the WHO Solidarity trial reported no significant effect of IFN-𝛽1a on patient survival during hospitalization [???].

5.5.4 Vaccines

5.5.4.1 Vaccine Development

Flu-like illnesses caused by viruses are a common target of vaccine development programs, and influenza vaccine technology in particular has made many strides. During the H1N1 influenza outbreak, vaccine development was accelerated because of the existing infrastructure, along with the fact that regulatory agencies had already decided that vaccines produced using egg- and cell-based platforms could be licensed under the regulations used for a strain change. Critiques of the production and distribution of the H1N1 vaccine have stressed the need for alternative development-and-manufacturing platforms that can be readily adapted to new pathogens. Although a monovalent H1N1 vaccine was not available before the pandemic peaked in the United States and Europe, it was available soon afterward as a stand-alone vaccine that was eventually incorporated into commercially available seasonal influenza vaccines [471]. If H1N1 vaccine development provides any indication, considering developing and manufacturing platforms for promising COVID-19 vaccine trials early could hasten the emergence of an effective prophylactic vaccine against SARS-CoV-2.

The first critical step towards developing a vaccine against SARS-CoV-2 was characterizing the target, which fortunately happened early in the COVID-19 outbreak with the sequencing and dissemination of the viral genome [472]. The Coalition for Epidemic Preparedness Innovations (CEPI) is coordinating global health agencies and pharmaceutical companies to develop vaccines against SARS-CoV-2. As of September 2020, there were over 180 vaccine candidates against SARS-CoV-2 in development [473]. Unlike many global vaccine development programs previously, such as with H1N1, the vaccine development landscape for COVID-19 includes vaccines produced by a wide array of technologies. Experience in the field of oncology is encouraging COVID-19 vaccine developers to use next-generation approaches to vaccine development, which have led to the great diversity of vaccine development programs [474]. These diverse technology platforms include DNA, RNA, virus-like particle, recombinant protein, both replicating and non-replicating viral vectors, live attenuated virus, and inactivated virus approaches (Figure 4). Given the wide range of vaccines under development, it is possible that some vaccine products may eventually be shown to be more effective in certain subpopulations, such as children, pregnant women, immunocompromised patients, the elderly, etc. The requirements for a successful vaccine trial and deployment are complex and may require coordination between government, industry, academia, and philanthropic entities [475]. While little is currently known about immunity to SARS-CoV-2, vaccine development typically tests for serum neutralizing activity, as this has been established as a biomarker for adaptive immunity in other respiratory illnesses [476].

Figure 4: Vaccine Development Strategies. Several different strategies can and are being employed for the development of vaccines. Each approach capitalizes on different features of the SARS-CoV-2 virus and delivery through a different platform.
5.5.4.1.1 DNA Vaccines

This vaccination method involves the direct introduction of a plasmid containing a DNA sequence encoding the antigen(s) against which an immune response is sought into appropriate tissues [477]. This approach may offer several advantages over traditional vaccination approaches, such as the stimulation of both B- as well as T-cell responses and the absence of any infectious agent. Currently, a Phase I safety and immunogenicity clinical trial of INO-4800, a prophylactic vaccine against SARS-CoV-2, is underway [478]. The vaccine developer Inovio Pharmaceuticals Technology is overseeing administration of INO-4800 by intradermal injection followed by electroporation with the CELLECTRA® device to healthy volunteers. Electroporation is the application of brief electric pulses to tissues in order to permeabilize cell membranes in a transient and reversible manner. It has been shown that electroporation can enhance vaccine efficacy by up to 100-fold, as measured by increases in antigen-specific antibody titers [479]. The safety of the CELLECTRA® device has been studied for over seven years, and these studies support the further development of electroporation as a safe vaccine delivery method [480]. The temporary formation of pores through electroporation facilitates the successful transportation of macromolecules into cells, allowing cells to robustly take up INO-4800 for the production of an antibody response. Approved by the U.S. FDA on April 6, 2020, the Phase I study is enrolling up to 40 healthy adult volunteers in Philadelphia, PA at the Perelman School of Medicine and at the Center for Pharmaceutical Research in Kansas City, MO. The trial has two experimental arms corresponding to the two locations. Participants in Experimental Group 1 will receive one intradermal injection of 1.0 milligram (mg) of INO-4800 followed by electroporation using the CELLECTRA® 2000 device twice, administered at Day 0 and Week 4. Participants in Experimental Group 2 will receive two intradermal injections of 1.0 mg (total 2.0 mg per dosing visit) of INO-4800 followed by electroporation using the CELLECTRA® 2000 device, administered at Day 0 and Week 4. Safety data and the initial immune responses of participants from the trial are expected by the end of the summer of 2021. The development of a DNA vaccine against SARS-CoV-2 by Inovio could be an important step forward in the world’s search for a COVID-19 vaccine. Although exciting, the cost of vaccine manufacturing and electroporation may make scaling the use of this technology for prophylactic use for the general public difficult.

5.5.4.1.2 RNA Vaccines

RNA vaccines are nucleic-acid based modalities that code for viral antigens against which the human body elicits a humoral and cellular immune response. The mRNA technology is transcribed in vitro and delivered to cells via lipid nanoparticles (LNP) [481]. They are recognized by ribosomes in vivo and then translated and modified into functional proteins [482]. The resulting intracellular viral proteins are displayed on surface MHC proteins, provoking a strong CD8+ T cell response as well as a CD4+ T cell and B cell-associated antibody responses [482]. Naturally, mRNA is not very stable and can degrade quickly in the extracellular environment or the cytoplasm. The LNP covering protects the mRNA from enzymatic degradation outside of the cell [483]. Codon optimization to prevent secondary structure formation and modifications of the poly-A tail as well as the 5’ untranslated region to promote ribosomal complex binding can increase mRNA expression in cells. Furthermore, purifying out dsRNA and immature RNA with FPLC (fast performance liquid chromatography) and HPLC (high performance liquid chromatography) technology will improve translation of the mRNA in the cell [482,484]. Vaccines based on mRNA delivery confer many advantages over traditional viral vectored vaccines and DNA vaccines. In comparison to live attenuated viruses, mRNA vaccines are non-infectious and can be synthetically produced in an egg-free, cell-free environment, thereby reducing the risk of a detrimental immune response in the host [485]. Unlike DNA vaccines, mRNA technologies are naturally degradable and non-integrating, and they do not need to cross the nuclear membrane in addition to the plasma membrane for their effects to be seen [482]. Furthermore, mRNA vaccines are easily, affordably, and rapidly scalable.

Although mRNA vaccines have been developed for therapeutic and prophylactic purposes, none have previously been licensed or commercialized. Nevertheless, they have shown promise in animal models and preliminary clinical trials for several indications, including rabies, coronavirus, influenza, and cytomegalovirus [486]. Preclinical data previously identified effective antibody generation against full-length FPLC-purified influenza hemagglutinin stalk-encoding mRNA in mice, rabbits, and ferrets [487]. Similar immunological responses for mRNA vaccines were observed in humans in Phase I and II clinical trials operated by the pharmaceutical-development companies Curevac and Moderna for rabies, flu, and zika [484]. Positively charged bilayer LNPs carrying the mRNA attract negatively charged cell membranes, endocytose into the cytoplasm [483], and facilitate endosomal escape. LNPs can be coated with modalities recognized and engulfed by specific cell types. LNPs that are 150nm or less effectively enter into lymphatic vessels.

There are three types of RNA vaccines: non-replicating, in vivo self-replicating, and in vitro dendritic cell non-replicating [488]. Non-replicating mRNA vaccines consist of a simple open reading frame (ORF) for the viral antigen flanked by the 5’ UTR and 3’ poly-A tail. In vivo self-replicating vaccines encode a modified viral genome derived from single-stranded, positive sense RNA alphaviruses [482,484]. The RNA genome encodes the viral antigen along with proteins of the genome replication machinery, including an RNA polymerase. Structural proteins required for viral assembly are not included in the engineered genome [482]. Self-replicating vaccines produce more viral antigens over a longer period of time, thereby evoking a more robust immune response [488]. Finally, in vitro dendritic cell non-replicating RNA vaccines limit transfection to dendritic cells. Dendritic cells are potent antigen-presenting immune cells that easily take up mRNA and present fragments of the translated peptide on their MHC proteins, which can then interact with T cell receptors. Ultimately, primed T follicular helper cells can stimulate germinal center B cells that also present the viral antigen to produce antibodies against the virus [489]. These cells are isolated from the patient, grown and transfected ex vivo, and reintroduced to the patient [490].

mRNA-1273 was the first COVID-19 vaccine to enter a phase I clinical in the United States. ModernaTX, Inc. is currently spearheading an investigation on the immunogenicity and reactogenicity of mRNA-1273, a conventional lipid nanoparticle encapsulated RNA encoding a full-length prefusion stabilized spike (S) protein for SARS-CoV-2 [491]. In a study that is ongoing, forty-five participants were enrolled and given intramuscular injections of mRNA-1273 in their deltoid muscle on day 1 and day 29, and then followed for the next twelve months. Healthy males and non-pregnant females aged 18-55 years were recruited for this study and divided into three groups receiving 25, 100, or 250 micrograms (μg) of mRNA-1273. IgG ELISA assays on patient serology samples are being used to examine the immunogenicity of the vaccine [491].

A preliminary report describing initial safety and immunogenicity findings is now available [476]. Binding antibodies were observed at two weeks after the first dose at all concentrations. At the time point one week after the second dose was administered on day 29, the pseudotyped lentivirus reporter single-round-of-infection neutralization assay (PsVNA), which was used to assess neutralizing activity, reached a median level similar to the median observed in convalescent plasma samples. Participants reported mild and moderate systemic adverse events after the day 1 injection, and one severe local event was observed in each of the two highest dose levels. The second injection led to severe systemic adverse events for three of the participants at the highest dose levels, with one participant in the group being evaluated at an urgent care center on the day after the second dose. The reported localized adverse events from the second dose were similar to those from the first.

In summary, mRNA vaccines are promising tools in the prevention and control of pandemics because they may sidestep many challenges associated with traditional vaccine design and manufacturing. As of August 2020, placebo-blinded and randomized results for mRNA vaccines are not yet available, but the initial results are promising. In the dose ranging study, even the lowest dose of 25 μg of mRNA-1273 appeared to be sufficient to induce an immunogenicity profile comparable to that observed in convalescent plasma and appeared to be well tolerated among the population of healthy, non-pregnant participants from 18-55 years of age. Based on the preliminary results from mRNA-1273 [476], a new trial aims to enroll 30,000 participants who will be randomized for two injections of 100 μg of mRNA-1273 or placebo, again spaced 28 days apart [492].

5.5.4.2 Adjuvants for Vaccines

Adjuvants include a variety of molecules or larger microbial-related products that have an effect on the immune system or an immune response of interest. They can either be comprised of or contain immunostimulants or immunomodulators. Adjuvants are sometimes included within vaccines, especially vaccines other than live-attenuated and inactivated viruses, in order to enhance the immune response. A review on the development of SARS-CoV-2 vaccines [493] also included a brief summary of the potential of adjuvants for these vaccines, including a brief description of some already commonly used adjuvants. Different adjuvants can regulate different types of immune responses, so the type or combination of adjuvants used in a vaccine will depend on both the type of vaccine and concern related to efficacy and safety. A variety of possible mechanisms for adjuvants have been researched [494,495,496], including the following: induction of DAMPs that can be recognized by certain PRRs of the innate immune system; functioning as PAMP that can also be recognized by certain PRRs; and more generally enhancing the humoral or cellular immune responses. Selection of one or more adjuvants requires considering how to promote the advantageous effects of the components and/or immune response and, likewise, to inhibit possible deleterious effects. There are also considerations related to the method of delivering (or co-delivering) the adjuvant and antigen components of a vaccine.

5.5.4.3 Trained Immunity

Another approach that is being investigated explores the potential for vaccines that are not made from the SARS-CoV-2 virus to confer what has been termed trained immunity. In a recent review [497], trained immunity was defined as forms of memory that are temporary (e.g., months or years) and reversible. It is induced by exposure to whole-microorganism vaccines or other microbial stimuli that generates heterologous protective effects. Trained immunity can be displayed by innate immune cells or innate immune features of other cells, and it is characterized by alterations to immune responsiveness to future immune challenges due to epigenetic and metabolic mechanisms. These alterations can take the form of either an increased or decreased response to immune challenge by a pathogen. Trained immunity elicited by non-SARS-CoV-2 whole-microorganism vaccines could potentially improve SARS-CoV-2 susceptibility or severity [498].

One type of stimulus which research indicates can induce trained immunity is bacillus Calmette-Guerin (BCG) vaccination. BCG is an attenuated form of bacteria Mycobacterium bovis. The vaccine is most commonly administered for the prevention of tuberculosis in humans. Clinical trials in non-SARS-CoV-2-infected adults have been designed to assess whether BCG vaccination could have prophylactic effects against SARS-CoV-2 by reducing susceptibility, preventing infection, or reducing disease severity. A number of trials are now evaluating the effects of the BCG vaccine or the related vaccine VPM1002 [498,499,500,501,502,503,504,505,506,507,508,509,510,511,512].

The ongoing trials are using a number of different approaches. Some trials enroll healthcare workers, other trials hospitalized elderly adults without immunosuppression who get vaccinated with placebo or BCG at hospital discharge, and yet another set of trials older adults (>50 years) under chronic care for conditions like hypertension and diabetes. One set of trials, for example, uses time until first infection as the primary study endpoint; more generally, outcomes measured in some of these trials are related to incidence of disease and disease severity or symptoms. Some analyses have suggested a possible correlation at the country level between the frequency of BCG vaccination (or BCG vaccination policies) and the severity of COVID-19 [498]. Currently it is unclear whether this correlation has any connection to trained immunity. Many possible confounding factors are also likely to vary among countries, such as age distribution, detection efficiency, stochastic epidemic dynamic effects, differences in healthcare capacity over time in relation to epidemic dynamics, and these have not been adequately accounted for in current analyses. It is unclear whether there is an effect of the timing of BCG vaccination, both during an individual’s life cycle and relative to the COVID-19 pandemic. Additionally, given that severe SARS-CoV-2 may be associated with a dysregulated immune response, it is unclear what alterations to the immune response would be most likely to be protective versus pathogenic (e.g., [108,498,513,514]). The article [498] proposes that trained immunity might lead to an earlier and stronger response, which could in turn reduce viremia and the risk of later, detrimental immunopathology. While trained immunity is an interesting possible avenue to complement vaccine development efforts through the use of an existing vaccine, additional research is required to assess whether the BCG vaccine is likely to confer trained immunity in the case of SARS-CoV-2.

5.6 Discussion

With the emergence of the COVID-19 pandemic caused by the coronavirus SARS-CoV-2, the development and/or identification of therapeutic and prophylactic interventions became an issue of international urgency. In previous outbreaks of HCoV, namely SARS and MERS, the development of these interventions was very limited. As research has progressed, several potential approaches to treatment have emerged (Figure 5).. Most notably, remdesivir has been approved by the FDA for the treatment of COVID-19, and dexamethasone, which was approved by the FDA in 1958, has been found to improve outcomes for patients with severe COVID-19. Other potential therapies are being still being explored and require additional data (Figure 3). Additionally, major advances in vaccines using mRNA and adenoviruses that have led to three vaccines becoming available or close to becoming available in late 2020 (Figure 5). As more evidence becomes available, the potential for existing and novel therapies to improve outcomes for COVID-19 patients will become better understood.

Figure 5: Mechanism of Action for Potential Therapeutics Potential therapeutics currently being studied target the SARS-CoV-2 or modify the host environment through many different mechanisms. Here, the relationship between the virus and several therapeutics described above are visualized.

Insights into the pathogenesis and immune response to SARS-CoV-2 have also provided some insights into prophylactics [cite therapeutics]. Though some concerns remain about the duration of sustained immunity for convalescents, vaccine development efforts are ongoing and show initial promising results. The Moderna trial, for example, reported that the neutralizing activity in participants who received two doses of the vaccine was similar to that observed in convalescent plasma. Insights from the immune response to SARS-CoV-2 have also guided the identification of potential therapeutics. As cases have become better characterized, it has become evident that many patients experience an initial immune response to the virus that is typically characterized by fever, cough, dyspnea, and related symptoms. However, the most serious concern is cytokine release syndrome, when the body’s immune response becomes dysregulated, resulting in an extreme inflammatory response. The RECOVERY trial, a large-scale, multi-arm trial enrolling about 15% of all COVID-19 patients in the United Kingdom, was the first to identify that the widely available steroid dexamethasone seems to be beneficial for patients suffering from this immune dysregulation [415]. The results of efforts to identify therapeutic treatments to treat patients early in the course of infection have been more ambiguous. Early interest in the drugs hydroxychloroquine and chloroquine yielded no promising results from studies with robust experimental designs. On the other hand, the experimental drug remdesivir, which was developed for Ebola, has received enough support from early analyses to receive FDA approval, but results have been mixed. The potential for other drugs, such as tocilizumab, to reduce recovery time remains unclear, but some early results were promising.

One concern is that the presentation of COVID-19 appears to be heterogeneous across the lifespan. Many adult cases, especially in younger adults, present with mild symptoms or even asymptomatically, while others, especially in older adults, can be severe or fatal. In children, the SARS-CoV-2 virus can present as two distinct diseases, COVID-19 or MIS-C. The therapeutics and prophylactics discussed here were all tested in adults. One of the two mRNA vaccines, Pfizer and BioNTech’s BNT162b2, has been issued an EUA for patients as young as 16 [515], while ModernaTX has begun a clinical trial to assess its mRNA vaccine in adolescents ages 12 to 18 [516]. However, additional research is needed to identify therapeutics that address the symptoms characteristic of pediatric COVID-19 and MIS-C cases.

5.6.1 Potential Avenues of Interest for Therapeutic Development

Given what is currently known about these therapeutics for COVID-19, a number of related therapies beyond those explored above may also prove to be of interest. For example, the demonstrated benefit of dexamethasone and the ongoing potential of tocilizumab for treatment of COVID-19 suggests that other anti-inflammatory agents might also hold value for the treatment of COVID-19. Given that current evidence about treating COVID-19 with dexamethasone suggests that the need to curtail the cytokine storm inflammatory response to the virus can transcend the risks of immunosuppression, exploration of more anti-inflammatory agents may be warranted. While dexamethasone is considered widely available and generally affordable, the high costs of biologics such as tocilizumab therapy may present obstacles to wide-scale distribution of this drug if it proves of value. At the doses used for rheumatoid arthritis patients, the cost for tocilizumab ranges from $179.20 to $896 per dose for the IV form and $355 for the pre-filled syringe [517]. There are several anti-inflammatory agents used for the treatment of autoimmune diseases that may also be able to counter the effects of the cytokine storm induced by the virus, some of which, such as cyclosporine, are likely to be more cost-effective and readily available than biologics [518]. While tocilizumab targets IL-6, several other inflammatory markers could be potential targets, including TNF-alpha. Inhibition of TNF-alpha by an inhibitor such as Etanercept has been previously suggested for treatment of SARS-CoV-1 [519] and may be relevant for SARS-CoV-2 as well. Another anti-IL-6 antibody, sarilumab, is also being investigated [520,521]. Baricitinib and other small molecule inhibitors of the Janus-activated kinase pathway also curtail the inflammatory response and have been suggested as potential options for SARS-CoV-2 infections [522]. Baricitinib in particular may be able to reduce the ability of SARS-CoV-2 to infect lung cells [523]. Clinical trials studying baricitinib in COVID-19 have already begun in the US and in Italy [524,525]. Identification and targeting of further inflammatory markers that are relevant in SARS-CoV-2 infection may be of value for curtailing the inflammatory response and lung damage.

In addition to immunosuppressive treatments that are most beneficial late in disease progression, much research is focused on identifying treatments would be likely to benefit early-stage patients. For example, although studies of hydroxychloroquine have not supported the early theory-driven interest in this antiviral treatment, alternative compounds with related mechanisms may still have potential. Hydroxyferroquine derivatives of HCQ have been described as a class of bioorganometallic compounds that exert antiviral effects with some selectivity for SARS-CoV-1 in vitro [526]. Future work could explore whether such compounds exert antiviral effects against SARS-CoV-2 and whether they would be safer for use in COVID-19. Another potential approach is the development of antivirals, which could be broad-spectrum, specific to coronaviruses, or targeted to SARS-CoV-2. Development of new antivirals is complicated by the fact that none have yet been approved for human coronaviruses. Intriguing new options are emerging, however. Beta-D-N4-hydroxycytidine (NHC) is an orally bioavailable ribonucleotide analog showing broad-spectrum activity against RNA viruses, which may inhibit SARS-CoV-2 replication in vitro and in vivo in mouse models of HCoVs [527]. A range of other antivirals are also in development. Development of antivirals will be further facilitated as research reveals more information about the interaction of SARS-CoV-2 with the host cell and host cell genome, mechanisms of viral replication, mechanisms of viral assembly, and mechanisms of viral release to other cells; this can allow researchers to target specific stages and structures of the viral life cycle. Finally, antibodies against viruses, also known as antiviral monoclonal antibodies, could be an alternative as well and are described in detail in an above section. The goal of antiviral antibodies is to neutralize viruses through either cell-killing activity or blocking of viral replication [528]. They may also engage the host immune response, encouraging the immune system to hone in on the virus. Given the cytokine storm that results from immune system activation in response to the virus, which has been implicated in worsening of the disease, a neutralizing antibody (nAb) may be preferable. Upcoming work may explore the specificity of nAbs for their target, mechanisms by which the nAbs impede the virus, and improvements to antibody structure that may enhance the ability of the antibody to block viral activity.

Some research is also investigating potential therapeutics and prophylactics that would interact with components of the innate immune response. For example, there are a variety of TLRs, PRRs that recognize PAMPs and DAMPs. TLRs form a part of innate immune recognition and can more generally contribute to promoting both innate and adaptive responses [106]. In mouse models, poly(I:C) and CpG, which are agonists of toll-like receptors TLR3 and TLR9, respectively, showed protective effects when administered prior to SARS-CoV-1 infection [529]. Therefore, TLR agonists hold some potential for broad-spectrum prophylaxis.

Given that a large number of clinical trials are currently in progress, more information about the potential of these and other therapeutics should become available over time. This information, combined with advances in understanding the molecular structure and viral pathogenesis of SARS-CoV-2, may lead to a more complete understanding of how the virus affects the human host and what strategies can improve outcomes. To date, investigations of potential therapeutics for COVID-19 have focused primarily on repurposing existing drugs. This approach is necessary given the urgency of the situation as well as the extensive time required for developing and testing new therapies. However, in the long-term, new drugs specific for treatment of COVID-19 may also enter development. Development of novel drugs is likely to be guided by what is known about the pathogenesis and molecular structure of SARS-CoV-2. For example, understanding the various structural components of SARS-CoV-2 may allow for the development of small molecule inhibitors of those components. Currently, crystal structures of the SARS-CoV-2 main protease have recently been resolved [340,530], and efforts are already in place to perform screens for small molecule inhibitors of the main protease, which have yielded potential hits [340]. Much work remains to be done to determine further crystal structures of other viral components, understand the relative utility of targeting different viral components, perform additional small molecule inhibitor screens, and determine the safety and efficacy of the potential inhibitors. While still nascent, work in this area is promising. Over the longer term, this approach and others may lead to the development of novel therapeutics specifically for COVID-19 and SARS-CoV-2.

5.6.2 Conclusions of the Present Analysis

Due to the large number of clinical trials currently under examination (Figure 3), not all candidates are examined here. Instead, this review seeks to provide an overview of the range of mechanisms that have been explored and to examine some prominent candidates in the context of the pathogenesis of and immune response to SARS-CoV-2. As more research becomes available, this review will be updated to include additional therapeutics that emerge and to include new findings that are released about those discussed here. While no therapeutics or vaccines were developed for SARS-CoV-1 or MERS-CoV, the current state of COVID-19 research suggests that the body of literature produced before and after the emergence of these viruses has prepared the biomedical community for a rapid response to novel HCoV like SARS-CoV-2. At present, two pharmaceutical therapeutics and one vaccine have been authorized for the treatment and prevention, respectively, of COVID-19. As the COVID-19 pandemic continues to be a topic of significant worldwide concern, more information is expected to become available about pharmaceutical mechanisms that can be used to combat this, and possibly other, HCoV. These advances therefore not only benefit the international community’s ability to respond to the current crisis, but are also likely to shape responses to future viral threats.

6 Dietary Supplements and Nutraceuticals Under Investigation for COVID-19 Prevention and Treatment

6.1 Abstract

The coronavirus disease 2019 (COVID-19) caused by the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) has caused global disruption and a significant loss of life. Scientists and physicians around the world have worked to discover and develop novel treatments or repurpose existing treatments in order to identify potential prophylactic and therapeutic agents to stem the spread of SARS-CoV-2 and reduce the devastation of this pandemic. Dietary supplements and nutraceuticals are also being investigated in this context. Despite a lack of evidence to support their efficacy or safety for the treatment or prevention of COVID-19, the beginning of the pandemic led to increased sales of vitamins, supplements, nutraceuticals, and various other products purporting ‘immune-boosting’ effects. Considerable interest amongst the scientific community and emerging evidence of potential utility in other therapeutic contexts has led to further investigation of these products. In this review, we critically appraise the evidence surrounding dietary supplements and nutraceuticals for the prophylaxis and treatment of COVID-19. These include vitamin C, vitamin D, omega 3 polyunsaturated fatty acids (PUFA), probiotics, and zinc. These compounds were selected as they are currently under clinical investigation. For example, there is increasing evidence to suggest that vitamin D supplementation may be beneficial due to plausible associations between vitamin D deficiency with COVID-19 incidence and possibly disease severity. A growing number of scientific organizations are considering or advising for increased vitamin D supplementation to those at high risk of COVID-19. Overall, though, caution is warranted as further evidence and clinical trials are required before conclusive evidence-based recommendations can be formulated for any specific dietary supplements or nutraceuticals. Consumers should also be aware of misinformation and false promises surrounding some supplement products, which may lack safety and efficacy evidence due to poor regulation by authorities. However, considering that nutritional status plays a significant role in patient outcomes, it would be prudent to advise the general public to follow a healthy diet and lifestyle to prevent nutrient deficiencies and to maintain a healthy immune system.

6.2 Importance

Sales of dietary supplements and nutraceuticals have increased since the start of the pandemic due to their perceived health benefits. These include vitamin C, vitamin D, and zinc, which many consider synonymous with the treatment of respiratory infections and support of immune health. However, little is known about the efficacy of these dietary supplements and nutraceuticals against the novel coronavirus (SARS-CoV-2) or the disease it causes, COVID-19. This review provides a critical overview of the potential prophylactic and therapeutic value of various dietary supplements and nutraceuticals from the evidence available to date. Evidence indicates that vitamin D deficiency may be associated with greater incidence of infection and severe COVID-19 and that vitamin D supplementation may be of prophylactic and therapeutic value. However, considerably more research is required to determine whether dietary supplements and nutraceuticals exhibit prophylactic and therapeutic value against SARS-CoV-2 infection and COVID-19.

6.3 Introduction

Scientists and the medical community are scrambling to repurpose or discover novel host-directed therapies against the coronavirus diseases 2019 (COVID-19) pandemic caused by the spread of the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2). Some successful therapies such as remdesivir and dexamethasone have been identified for hospitalized patients. Furthermore, most societies have adopted the non-pharmacological preventative public health strategies that reduce the transmission of SARS-CoV-2. However, many individuals have sought additional protections via the consumption of various dietary supplements and nutraceuticals with purported beneficial effects. Indeed, a patient’s nutritional status seems to play a role in COVID-19 susceptibility and outcomes [531,532,533,534,535].

Nutraceuticals and dietary supplements are related but distinct non-pharmaceutical products. Nutraceuticals are classified as supplements with health benefits beyond their basic nutritional value [536,537]. The key difference between a dietary supplement and a nutraceutical is that nutraceuticals should not only supplement the diet, but also aid in the prophylaxis and/or treatment of a disorder or disease [538]. Due to the significant interest from the general public in these dietary additives, whether and to what extent nutraceuticals or dietary supplements can provide any prophylactic or therapeutic benefit remains a topic of interest for the scientific community. Despite the lack of any evidence to support their use, the beginning of the pandemic led sales of these products to soar. In the United States, dietary supplement and nutraceutical sales were worth $345 million for the entirety of 2019, whereas sales reached $435 million during the six-week period of the outbreak that ended on April 5 of 2020, which grew by a further $151 million by May 17 2020 [539]. In France, New Zealand, India, and China similar trends in sales were reported [540,541,542,543]. The increase in sales was driven by a consumer perception that dietary supplements and nutraceuticals would protect consumers from infection and/or mitigate the impact of infection due to the various “immune-boosting” claims of these products [544,545]. However, dietary supplements and nutraceuticals, unlike pharmaceuticals, are not subject to the same regulatory protocols that protect consumers of medicines. Indeed, nutraceuticals do not entirely fall under the responsibility of the Food and Drug Administration (FDA), but they are monitored as dietary supplements according to the Dietary Supplement, Health and Education Act 1994 (DSHEA) [546] and the Food and Drug Administration Modernization Act 1997 (FDAMA) [547]. Due to the increased sales of dietary supplements and nutraceuticals, in 1996 the FDA established the Office of Dietary Supplement Programs (ODSP) to increase surveillance. Novel products or nutraceuticals must now submit a new dietary ingredient notification to the ODSP for review.
There are significant concern that these legislations do not adequately protect the consumer as they ascribe responsibility to the manufacturers to ensure the safety of the product before manufacturing or marketing [548]. Manufacturers are not required to register or even seek approval from the FDA to produce or sell food supplements or nutraceuticals. Health or nutrient content claims for labeling purposes are approved based on an authoritative statement from the Academy of Sciences or relevant federal authorities once the FDA has been notified and on the basis that the information is known to be true and not deceptive [548].

In Europe, the safety of supplements are monitored by the European Union (EU) under Directive 2002/46/EC [549]. However, nutraceuticals are not directly mentioned. Consequently, nutraceuticals can be generally described as either a medicinal product under Directive 2004/27/EC [550] or as a ‘foodstuff’ under Directive 2002/46/EC of the European council. In order to synchronize the various existing legislations, Regulation EC 1924/2006 on nutrition and health claims was put into effect to assure customers of safety and efficacy of products and to deliver understandable information to consumers. However, specific legislation for nutraceuticals is still elusive. Health claims are permitted on a product label only following compliance and authorization according to the European Food Safety Authority (EFSA) guidelines on nutrition and health claims [551]. EFSA does not currently distinguish between food supplements and nutraceuticals for health claim applications of new products, as claim authorization is dependent on the availability of clinical data in order to substantiate efficacy [552]. However, these guidelines seem to provide more protection to consumers than the FDA regulations, but potentially at the cost of innovation in the sector [553]. Looking further afield, the situation becomes even more complicated at a global level where countries such as China and India have existing regulatory frameworks for traditional medicines and phytomedicines not commonly consumed by Western society [554]. Currently, there is a debate among scientists and regulatory authorities surrounding the development of a widespread regulatory framework to deal with the challenges of safety and health claim substantiation for nutraceuticals [548,552] as these products do not necessarily follow the same rigorous clinical trial frameworks used to approve the use of pharmaceuticals. Such regulatory disparities have been highlighted by the pandemic as many individuals and companies have attempted to profit from the vulnerabilities of others by overstating claims in relation to the treatment of COVID-19 using supplements and nutraceuticals. The FDA has written several letters to prevent companies marketing or selling products based on false hyperbolic promises to prevent SARS-CoV-2 infection or treat COVID-19 [555,556,557]. These letters came in response to efforts to market nutraceutical prophylactics against COVID-19, some of which charged the consumer as much as $23,000 [558]. There have even been some incidents highlighted in the media because of their potentially life threatening consequences; for example, the use of oleandrin was touted as a potential “cure” by individuals close to the former President of the United States despite its high toxicity [559]. Thus, heterogeneous and at times relaxed regulatory standards have permitted high-profile cases of the sale of nutraceuticals and dietary supplements that are purported to provide protection against COVID-19, despite a lack of research into these compounds.

Notwithstanding the issues of poor safety, efficacy, and regulatory oversight, some dietary supplements and nutraceuticals have exhibited therapeutic and prophylactic potential. Some have been linked with reduced immunopathology, antiviral and anti-inflammatory activities, or even the prevention of ARDS [544,560,561]. A host of potential candidates have been highlighted in the literature that target various aspects of the COVID-19 viral pathology, while others are thought to prime the host immune system. These candidates include vitamins and minerals along with extracts and omega-3 polyunsaturated fatty acids (n-3 PUFA) [562]. In vitro and in vivo studies suggest that nutraceuticals containing phycocyanobilin, N-acetylcysteine, glucosamine, selenium or phase 2 inductive nutraceuticals (e.g. ferulic acid, lipoic acid, or sulforaphane) can prevent or modulate RNA virus infections via amplification of the signaling activity of mitochondrial antiviral-signaling protein (MAVS) and activation of toll-like receptor 7 (TLR7) [563]. While promising, further animal and human studies are required to assess the therapeutic potential of these various nutrients and nutraceuticals against COVID-19. For the purpose of this review, we have highlighted some of the main dietary supplements and nutraceuticals that are currently under investigation for their potential prophylactic and therapeutic applications. These include n-3 PUFA, zinc, vitamins C and D, and probiotics.

6.4 n-3 PUFA

One supplement that has been explored for beneficial effects against various viral infections is n-3 PUFA [562], such as eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA). EPA and DHA intake can come from a diet high in fish or through dietary supplementation with fish oils or purified oils [564]. Other more sustainable sources of EPA and DHA include algae [565], which can also be exploited for their rich abundance of other bioactive compounds such as ACE inhibitor peptides and antiviral agents including phycobiliproteins, sulfated polysaccharides, and calcium-spirulan [566]. They can mediate inflammation and therefore may have the capacity to modulate the adaptive immune response [537,564,567]. Another potential mechanism that has led to interest in n-3 PUFA for viral infections is its potential as a precursor molecule for the biosynthesis of endogenous specialized proresolving mediators (SPM), such as protectins and resolvins, that actively resolve inflammation and infection [568]. Finally, some COVID-19 patients, particularly those with comorbidities, are at a significant risk of thrombotic complications including arterial and venous thrombosis [103,569]. Therefore, the use of prophylactic and therapeutic anticoagulants and antithrombotic agents is under consideration [570,571], which could potentially include n-3 PUFA.

SPM have exhibited beneficial effects against a variety of lung infections, including some caused by RNA viruses [572,573]. Indeed, protectin D1 has been shown to increase survival from H1N1 viral infection in mice by affecting the viral replication machinery [574]. Several mechanisms for SPM have been proposed, including preventing the release of pro-inflammatory cytokines and chemokines or increasing phagocytosis of cellular debris by macrophages [575]. In influenza, SPM promote antiviral B lymphocytic activities [576], and protectin D1 has been shown to increase survival from H1N1 viral infection in mice by affecting the viral replication machinery [574]. It is hypothesized that SPM could aid in the resolution of the cytokine storm and pulmonary inflammation associated with COVID-19 [577,578]. Another theory is that some comorbidities such as obesity could lead to deficiencies of SPM, which could in turn be related to the occurrence of adverse outcomes for COVID-19 [???]. However, not all studies are in agreement that n-3 PUFA or their resulting SPM are effective against infections [579]. At a minimum, the effectiveness of n-3 PUFA against infections would be dependent on the dosage, timing, and the specific pathogens responsible [580]. On another level, there is still the question of whether fish oils can raise the levels of SPM levels upon ingestion and in response to acute inflammation in humans [581]. Currently, Karolinska University Hospital is running a trial that will measure the levels of SPM as a secondary outcome following intravenous supplementation of n-3 PUFA in hospitalized COVID-19 patients to determine whether n-3 PUFA provides therapeutic value [582].

The increased risk of thrombotic complications in COVID-19 infected patients was reported relatively late in comparison to other manifestations of COVID-19 [97,569]. Considering that there is significant evidence that n-3 fatty acids and other fish oil-derived lipids possess antithrombotic properties and anti-inflammatory properties [583,584], they may have therapeutic value against the prothrombotic complications of COVID-19. Based on concern among the medical community that the use of investigational therapeutics on COVID-19 patients already on antiplatelet therapies due to a pre-existing comorbidities may lead to issues with dosing and drug choice, and/or negative drug-drug interactions [570], this supplementation holds particular interest for the treatment of patients already receiving pharmaceutical antiplatelet therapies. As a result, the use of other therapeutics, such as dietary sources of n-3 fatty acids or nutraceuticals with antiplatelet activities, may be beneficial and warrant further investigation. A new clinical trial [585] is currently recruiting COVID-19 positive patients to investigate the anti-inflammatory activity of a recently developed, highly purified nutraceutical derivative of EPA known as icosapent ethyl (VascepaTM) [586]. Other randomized controlled trials are in the preparatory stages with the intention of investigating the administration of EPA and other bioactive compounds to COVID-19 positive patients to determine whether anti-inflammatory effects or disease state improvements are observed [587,588]. Finally, while there have been studies investigating the therapeutic value of n-3 fatty acids against ARDS in humans, there is still limited evidence of their effectiveness [589]. It should be noted that the overall lack of human studies in this area means there is limited evidence as to whether these supplements could affect COVID-19 infection. Consequently, clinical trials have been proposed and many are in the preparatory stages. These trials will investigate whether the anti-inflammatory potential of n-3 PUFA and their derivatives is beneficial in the treatment of COVID-19. Therefore, the evidence is not present to draw conclusions about whether n-3 PUFA will be useful in treating COVID-19, but as there is likely little harm associated with a diet rich in fish oils, interest in these products by the general public is unlikely to have negative effects.

6.5 Zinc

There is evidence that certain nutrient supplements may exhibit some benefit against RNA viral infections. Zinc is a trace metal obtained from dietary sources or supplementation that is important for the maintenance of immune cells involved in adaptive and innate immunity [590]. Zinc supplements can be administered orally as a tablet or as a lozenge and they are available in many forms, such as zinc picolinate, zinc acetate, and zinc citrate. Zinc is also available from dietary sources including meat, seafood, nuts, seeds, legumes, and dairy. The role of zinc in immune function has been extensively reviewed [590]. Zinc is an important signaling molecule and zinc levels can alter host defense systems. In inflammatory situations such as an infection, zinc can regulate leukocyte immune responses and it can activate the nuclear factor kappa-light-chain-enhancer of activated B cells (NF-kB), thus altering cytokine production [591,592]. In particular, zinc supplementation can increase natural killer cell levels, which are important cells for host defense against viral infections [590,593].

Adequate zinc intake has been associated with reduced incidence of infection [594] and antiviral immunity [595]. A randomized, double-blind, placebo-controlled trial that administered zinc supplementation to elderly subjects over the course of a year found zinc deficiency to be associated with increased susceptibility to infection and that zinc deficiency could be prevented through supplementation [594]. Clinical trial data supports the utility of zinc to diminish the duration and severity of symptoms associated with common colds when it is provided within 24 hours of the onset of symptoms [596,597]. An observational study has shown that COVID-19 patients had significantly lower zinc levels in comparison to healthy controls, they also tended to have more complications (70.4% vs 30.0%, p = 0.009), and potentially prolonged hospital stays (7.9 vs 5.7 days, p = 0.048) [598].
In coronaviruses specifically, in vitro evidence demonstrates that the combination of zinc (Zn2+) and zinc ionophores (pyrithione) can interrupt the replication mechanisms of SARS-CoV-GFP (a fluorescently tagged SARS-CoV-1) and a variety of other RNA viruses [599,600]. Currently, there are over twenty clinical trials registered with the intention to use zinc in a preventative or therapeutic manner. However, many of these trials proposed the use of zinc in conjunction with hydroxychloroquine and azithromycin [601,602,603,604], and it is not known how the lack of evidence supporting the use of hydroxychloroquine will affect investigation of zinc. One retrospective observational study of New York University Langone hospitals in New York compared outcomes among hospitalized COVID-19 patients administered hydroxychloroquine and azithromycin with zinc sulfate (n = 411) versus hydroxychloroquine and azithromycin alone (n = 521). Notably, zinc is the only treatment that was used in this trial that is still under consideration as a therapeutic agent due to to the lack of efficacy and potential adverse events associated with hydroxychloroquine and azithromycin against COVID-19 [605,606,607]. While the addition of zinc sulfate did not affect the duration of hospitalization, the length of ICU stays or patient ventilation duration, univariate analyses demonstrated that zinc did increase the frequency of patients discharged and decreased the requirement for ventilation and referrals to the ICU, and reduced mortality [608]. However, a smaller retrospective study at Hoboken University Medical Center New Jersey failed to observe an association between zinc supplementation and survival of hospitalized patients [609].

Other trials, however, are investigating zinc in conjunction with other supplements such as vitamin C or n-3 PUFA [588,610]. Though there is, overall, encouraging data for zinc supplementation against the common cold and viral infections, there is currently limited evidence to suggest zinc supplementation has any beneficial effects against the current novel COVID-19; thus, the clinical trials that are currently underway will provide vital information on the efficacious use of zinc in COVID-19 prevention and/or treatment. However, given the limited risk, maintaining a healthy diet to ensure an adequate zinc status may be advisable for individuals seeking to reduce their likelihood of infection.

6.6 Vitamin C

Vitamins B, C, D, and E have also been suggested as potential nutrient supplement interventions for COVID-19 [562,611]. In particular vitamin C has been proposed as a potential therapeutic agent against COVID-19 due to its long history of use against the common cold and other respiratory infections [612,613]. Vitamin C can be obtained via dietary sources such as fruits and vegetables or via supplementation. Vitamin C plays a significant role in promoting immune function due to its effects on various immune cells. It affects inflammation by modulating cytokine production, decreasing histamine levels, enhancing the differentiation and proliferation of T- and B-lymphocytes, increasing antibody levels, and protecting against the negative effects of reactive oxygen species amongst other effects related to COVID-19 pathology [614,615,616]. Vitamin C is utilized by the body during viral infections, as evinced by lower concentrations in leukocytes and lower concentrations of urinary vitamin C. Post-infection, these levels return to baseline ranges [617,618,619,620,621]. Indeed, COVID-19 patients appear to be deficient in vitamin C [622,623]. COVID-19 is also associated with the formation of microthrombi and coagulopathy [105] that contributes to the observed lung pathology [624], which can be ameliorated by early infusions of vitamin C by inhibiting endothelial surface P-selectin expression and platelet-endothelial adhesion [625]. Indeed, intravenous vitamin C also reduced D-dimer levels in a case study of 17 COVID-19 patients [626], which are notably elevated in COVID-19 patients [???,102].

A recent meta-analysis found consistent support for regular vitamin C supplementation reducing the duration of the common cold, but that supplementation with vitamin C (> 200 mg) failed to reduce the incidence of colds [627]. Individual studies have found Vitamin C to reduce the susceptibility of patients to lower respiratory tract infections such as pneumonia [628]. Another meta-analysis has demonstrated in twelve trials that vitamin C supplementation reduced the length of stay of patients in intensive care units (ICUs) by 7.8% (95% CI: 4.2% to 11.2%; p = 0.00003). Furthermore, high doses (1-3 g/day) significantly reduced the length of an ICU stay by 8.6% in six trials (p = 0.003). Vitamin C also shortened the duration of mechanical ventilation by 18.2% in three trials in which patients required intervention for over 24 hours (95% CI 7.7% to 27%; p = 0.001) [629]. Despite these findings, the CITRUS ALI study failed to show a benefit of a 96-hour infusion of vitamin C to treat ARDS, which is a severe complication of COVID-19 infection [630]. Nevertheless, clinical trials have begun, as highlighted by Carr et al. [613], the first of which was initiated in Wuhan, China. Others will be conducted in Canada, China, Iran, and the USA. These trials intend to investigate the use of intravenous vitamin C in hospitalized COVID-19 patients. The first trial established in Wuhan, China [631] has reported initial results indicating that 12 g/12 hr of intravenous vitamin C for 7 days in 56 critically ill COVID-19 patients resulted in a promising reduction of 28-day mortality (p = 0.06) in univariate survival analysis [632]. Indeed, the same study reported a significant decrease in IL-6 levels by day 7 of vitamin C infusion (p = 0.04) [633].

Even though evidence supporting the use of vitamin C is beginning to emerge, we will not know how effective vitamin C is as a therapeutic for quite some time. Currently (as of December 2020) over fifteen trials are registered with clinicaltrials.gov that are either recruiting, active or are currently in preparation. When completed, these trials will provide crucial evidence on the efficacy of vitamin C as a therapeutic for COVID-19 infection. However, the majority of supplementation studies investigate the intravenous infusion of vitamin C in severe patients. Therefore, there is a lack of studies investigating the potential prophylactic administration of vitamin C via oral supplementation for healthy individuals or potentially asymptomatic SARS-CoV-2 positive patients. Once again, vitamin C intake is part of a healthy diet and the vitamin likely presents minimal risk, but its potential prophylactic or therapeutic effects against COVID-19 are yet to be determined.

6.7 Vitamin D

Of all of the supplements currently under investigation, vitamin D has become a leading prophylactic and therapeutic candidate against SARS-CoV-2. Vitamin D can modulate both the adaptive and innate immune system and is associated with various aspects of immune health and antiviral defense [634,635,636,637,638]. Vitamin D can be sourced through diet or supplementation, but it is mainly biosynthesized by the body on exposure to ultraviolet light (UVB) from sunlight. Vitamin D deficiency is associated with an increased susceptibility to infection [639]. In particular, vitamin D deficient patients are at risk of developing acute respiratory infections [640] and ARDS [640]. 1,25-dihydroxyvitamin D3 is the active form of vitamin D that is involved in adaptive and innate responses; however, due to its low concentration and a short half life of a few hours, it is measured by its longer lasting and more abundant precursor 25-hydroxyvitamin D. The vitamin D receptor is expressed in various immune cells, and vitamin D is an immunomodulator of antigen presenting cells, dendritic cells, macrophages, monocytes, and T- and B-lymphocytes [639,641]. Due to its potential immunomodulating properties, vitamin D supplementation may be advantageous to maintain a healthy immune system.

Early in the pandemic it was postulated that an individual’s vitamin D status may significantly affect their risk of developing COVID-19 [642]. This hypothesis was derived from the fact that the current pandemic emerged in winter in Wuhan China when 25-hydroxyvitamin D concentrations are at their lowest due to a lack of sunlight, whereas in the Southern Hemisphere, where it was nearing the end of the summer and higher 25-hydroxyvitamin D concentrations would be higher, the number of cases was low. This led researchers to question whether there was a seasonal component to the SARS-CoV-2 pandemic and whether vitamin D levels might play a role [642,643,644,645]. Though it is assumed that COVID-19 is seasonal, multiple other factors that can affect vitamin D levels should also be considered. These factors include an individual’s nutritional status, their age, their occupation, skin pigmentation, potential comorbidities, and the variation of exposure to sunlight due to latitude amongst others. Indeed, it has been estimated that each degree of latitude north of 28 degrees corresponded to a 4.4% increase of COVID-19 mortality, indirectly linking a persons vitamin D levels via exposure to UVB light to COVID-19 mortality [643].

As the pandemic evolves, further research of varying quality has investigated some of the potential links identified early in the pandemic [642] between vitamin D and COVID-19. Indeed, studies are beginning to investigate whether there is any prophylactic and/or therapeutic relationship between vitamin D and COVID-19. A study in Switzerland demonstrated that 27 SARS-CoV-2 positive patients exhibited 25-hydroxyvitamin D plasma concentrations that were significantly lower (11.1 ng/ml) than those of SARS-CoV-2 negative patients (24.6 ng/ml; p = 0.004), an association that held when stratifying for patients greater than 70 years old [646]. These findings seem to be supported by a Belgian observational study of 186 SARS-CoV-2 positive patients exhibiting symptoms of pneumonia, where 25-hydroxyvitamin D plasma concentrations were measured and a CT scan of the lungs was obtained upon hospitalization [647]. A significant difference in 25-hydroxyvitamin D levels was observed between the SARS-CoV-2 patients and 2,717 season-matched diseased controls. Both female and male patients possessed lower median 25-hydroxyvitamin D concentrations than the control group (18.6 ng/ml versus 21.5 ng/ml; p = 0.0016) and a higher rate of vitamin D deficiency (58.6% versus 42.5%). Evidence of sexual dimorphism was also apparent, as female patients had equivalent levels of 25-hydroxyvitamin D to the control group, whereas male patients were deficient in 25-hydroxyvitamin D relative to male controls (67% versus 49%; p = 0.0006). Notably, vitamin D deficiency was progressively lower in males with advancing radiological disease stages (p = 0.001). These studies are supported by several others that indicate that vitamin D status may be an independent risk factor for the severity of COVID-19 [648,649,650,651] and in COVID-19 patients relative to population-based controls [652]. Indeed, serum concentrations of 25-hydroxyvitamin D above 30 ng/ml, which indicate vitamin D sufficiency, seems to be associated with a reduction in serum C-reactive protein (CRP), an inflammatory marker, along with increased lymphocyte levels, which suggests that vitamin D levels may modulate the immune response by reducing risk for cytokine storm in response to SARS-CoV-2 infection [652]. A study in India has determined that COVID-19 fatality is higher in patients with severe COVID-19 and low serum 25-hydroxyvitamin D (mean level 6.2 ng/ml; 97% vitamin D deficient) levels versus asymptomatic non-severe patients with higher levels of vitamin D (mean level 27.9 ng/ml; 33% vitamin D deficient) [653]. In the same study, vitamin D deficiency was associated with higher levels of inflammatory markers including IL-6, ferritin, and tumor necrosis factor (TNF)-α. Collectively, these studies add to a multitude of observational studies reporting potential associations between low levels of 25-hydroxyvitamin D and COVID-19 incidence and severity [646,651,652,654,655,656,657,658,659,660].

Despite these studies establishing a link between vitamin D status and COVID-19 severity, an examination of data from the UK Biobank did not support this thesis [661,662]. These analyses examined 25-hydroxyvitamin D concentrations alongside SARS-CoV-2 positivity and COVID-19 mortality in over 340,000 UK Biobank participants. However, these studies have caused considerable debate that will likely be settled following further studies [663,664]. Overall, while the evidence suggest that there is likely an association between low serum 25-hydroxyvitamin D and COVID-19 incidence, these studies must be interpreted with caution as there is the potential for reverse causality, bias, and other confounding factors including that vitamin D deficiency is also associated with numerous pre-existing conditions and risk factors that can increase the risk for severe COVID-19 [531,643,665,666].

While these studies inform us of the potential importance of vitamin D sufficiency and the risk of SARS-CoV-2 infection and severe COVID-19, they fail to conclusively determine whether vitamin D supplementation can therapeutically affect the clinical course of COVID-19. In one study, 40 vitamin D deficient asymptomatic or mildly symptomatic participants patients were either randomized to receive 60,000 IU of cholecalciferol daily for at least 7 days (n = 16) or a placebo (n = 24) with a target serum 25-hydroxyvitamin D level >50 ng/ml. At day 7, 10 patients achieved >50 ng/ml, and another 2 by day 14. By the end of the study, the treatment group had a greater proportion of vitamin D-deficient participants that tested negative for SARS-CoV-2 RNA, and they had a significantly lower fibrinogen levels, potentially indicating a beneficial effect [667]. A pilot study in Spain determined that early high dose calcifediol (~21,000 IU days 1-2 and ~11,000 IU days 3-7 of hospital admission) with hydroxychloroquine and azithromycin to 50 hospitalized COVID-19 patients significantly reduced ICU admissions and may have reduced disease severity versus hydroxychloroquine and azithromycin alone [668].
While this study received significant criticism the National Institute for Health and Care Excellence (NICE) in the UK [669], an independent follow-up statistical analysis supported the findings of the study as a result of cholecalciferol treatment [670]. Another trial of 986 COVID-19 hospitalized patients in three UK hospitals administered cholecalciferol supplementation (≥ 280,000 IU in a time period of 7 weeks) to 151 patients and found an association with a reduced risk of COVID-19 mortality, regardless of baseline 25-hydroxyvitamin D levels [671]. Indeed, a double-blind, randomized, placebo-controlled trial of 240 hospitalized COVID-19 patients in Sao Paulo, Brazil administered a single 200,000 IU oral dose of vitamin D. While levels of 25-hydroxyvitamin D did increase from 21% to ~27% (p = 0.001) and were well tolerated, there was no reduction in the length of hospital stay, mortality, and no change to any other relevant secondary outcomes [672]. These early findings are still inconclusive with regards to the therapeutic value of vitamin D supplementation. However, other trials are underway including one trial that is investigating the utility of vitamin D as an immune-modulating agent by monitoring whether administration of vitamin D precipitates an improvement of health status in non-severe symptomatic COVID-19 patients and whether vitamin D prevents patient deterioration [673]. Other trials are examining various factors including mortality, symptom recovery, severity of disease, rates of ventilation, inflammatory markers such as CRP and IL-6, blood cell counts, and the prophylactic capacity of vitamin D administration [673,674,675,676]. Concomitant administration of vitamin D with pharmaceuticals such as aspirin [677] and bioactive molecules such as resveratrol [678] are also under investigation.

The effectiveness of vitamin D supplementation against COVID-19 remains open for debate. Yet, there is no doubt that vitamin D deficiency is a widespread issue and should be mitigated for not only because of its potential link to SARS-CoV-2 incidence [679], but also due to its importance for overall health. There is a possibility that safe exposure to sunlight could improve endogenous synthesis of vitamin D, potentially strengthening the immune system. However, sun exposure is not sufficient on its own, particularly in the winter months. Indeed, preemptive supplementation of vitamin D and encouraging people to maintain a healthy diet for optimum vitamin D status is likely to raise serum levels of 25-hydroxyvitamin D while being unlikely to carry major health risks while the possible link between vitamin D status and COVID-19 is further investigated. These principles seem to be the basis of a number of guidelines issued by some countries and scientific organizations that have advised supplementation of vitamin D during the pandemic.
The Académie Nationale de Médecine in France recommends rapid testing of 25-hydroxyvitamin D for people over 60 years old to identify those most at risk of vitamin D deficiency and advising them to obtain a bolus dose of 50,000 to 100,000 IU vitamin D to limit respiratory complications. They have also recommended that those under 60 years old should take 800 to 1,000 IU daily if they receive a SARS-CoV-2 positive test [680]. Both Public Health England and Public Health Scotland have advised members of the Black, Asian, and minority ethnic (BAME) communities to supplement for vitamin D in light of evidence that they may be at higher risk for vitamin D deficiency and other COVID-19 risk factors. Indeed, African Americans may have up to a 6-fold higher COVID-19 mortality rate than white populations [681,682,683]. However, other UK scientific bodies including the NICE recommends that individuals supplement for vitamin D as per usual UK government advice, but they warn that people should not supplement for vitamin D solely to prevent COVID-19. However, the NICE have provided guidelines for research to investigate the supplementation of vitamin D in the context of COVID-19 [684]. In Slovenia, doctors have been advised to provide nursing home patients with vitamin D [685]. Despite vitamin D deficiency being a widespread issue in the United States [686] the National Institutes of Health have stated that there is “insufficient data to recommend either for or against the use of vitamin D for the prevention or treatment of COVID-19” [687]. These are just some examples of how public health guidance has responded to the emerging evidence with regards to vitamin D and COVID-19. In contrast, some companies have used the emerging evidence surrounding vitamin D to sell products that claim to prevent and treat COVID-19, which in one incident required a federal court to intervene and issue an injunction barring the sale of vitamin-D-related products, due to a lack of clinical data to support their claims [688] It it clear that further studies and clinical trials are required to conclusively determine the prophylactic and therapeutic potential of vitamin D supplementation against COVID-19. Until such time that sufficient evidence emerges, individuals should follow their national guidelines surrounding vitamin D intake to achieve vitamin D sufficiency.

6.8 Probiotics

Probiotics are “live microorganisms that, when administered in adequate amounts, confer a health benefit on the host” [689]. Some studies suggest that probiotics are beneficial against common viral infections and there is modest evidence to suggest that they can modulate the immune response [690,691], and as a result it has been hypothesized that probiotics may have therapeutic value worthy of investigation against SARS-CoV-2 [692]. Probiotics and next-generation probiotics, which are more akin to pharmacological-grade supplements, have been associated with multiple potential beneficial effects for allergies, digestive tract disorders, and even metabolic diseases through their anti-inflammatory and immunomodulatory effects [693,694]. However, the mechanisms by which probiotics affect these various conditions would likely differ among strains, with the ultimate effect of the probiotic depending on the heterogeneous set of bacteria present [694]. Some of the beneficial effects of probiotics include reducing inflammation by promoting the expression of anti-inflammatory mediators, inhibiting toll-like receptors (TLR) 2 and 4, direct competition with pathogens, the synthesis of antimicrobial substances or other metabolites, improving intestinal barrier function, and/or favorably altering the gut microbiota and the brain-gut axis [694,695,696]. It is also thought that lactobacilli such as Lactobacillus paracasei, Lactobacillus plantarum and Lactobacillus rhamnosus have the capacity to bind to and inactivate some viruses via adsorptive and/or trapping mechanisms [697]. Other probiotic lactobacilli and even non-viable bacterium-like particles have been shown to reduce both viral attachment to host cells and viral titers, along with reducing cytokine synthesis, enhancing the antiviral IFN-α response, and inducing various other antiviral mechanisms [???,697,698,699,700,701,702,703,704,705]. These antiviral and immunobiotic mechanisms and others have been reviewed in detail elsewhere [561,692,706]. However, there is also a bi-directional relationship between the lungs and gut microbiota known as the gut-lung axis [707], whereby gut microbial metabolites and endotoxins may affect the lungs via the circulatory system and the lung microbiota in return may affect the gut [708]. Therefore, the gut-lung axis may play role in our future understanding of COVID-19 pathogenesis and become a target for probiotic treatments [709]. Moreover, as microbial dysbiosis of the respiratory tract and gut may play a role in some viral infections, it has been suggested that SARS-CoV-2 may interact with our commensal microbiota [561; 710; 10.3389/fmicb.2020.01840] and that the lung microbiome could play a role in developing immunity to viral infections [711]. These postulations, if correct, could lead to the development of novel probiotic and prebiotic treatments. However, significant research is required to confirm these associations and their relevance to patient care, if any.

Probiotics have tentatively been associated with the reduction of risk and duration of viral upper respiratory tract infections [712,713,714]. Some meta-analyses that have assessed the efficacy of probiotics in viral respiratory infections have reported moderate reductions in the incidence and duration of infection [713,715]. Indeed, randomized controlled trials have shown that administering Bacillus subtilis and Enterococcus faecalis [716], Lactobacillus rhamnosus GG [717], or Lactobacillus casei and Bifidobacterium breve with galactooligosaccharides [718] via the nasogastric tube to ventilated patients reduced the occurrence of ventilator-associated pneumonia in comparison to the respective control groups in studies of viral infections and sepsis. These findings are supported by a recently published meta-analysis [719]. There is a significant risk of ventilator-associated bacterial pneumonia in COVID-19 patients [720], but it can be challenging for clinicians to diagnose this infection due to the fact that severe COVID-19 infection presents with the symptoms of pneumonia [721]. Therefore, an effective prophylactic therapy for ventilator-associated pneumonia in severe COVID-19 patients would be of significant therapeutic value.

Probiotics are generally synonymous with the treatment of gastrointestinal issues due to their supposed anti-inflammatory and immunomodulatory effects [722]. Notably, gastrointestinal symptoms commonly occur in COVID-19 patients [723], and the ACE2 receptor is highly expressed in enterocytes of the ileum and colon, suggesting that these organs may be a potential route of infection [724,725]. Indeed, SARS-CoV-2 viral RNA has been detected in human feces [86,726], and fecal-oral transmission of the virus has not yet been ruled out [727]. Rectal swabs of some SARS-CoV-2 positive pediatric patients persistently tested positive for several days despite negative nasopharyngeal tests, indicating the potential for fecal viral shedding [728]. However, there is conflicting evidence for the therapeutic value of various probiotics against the incidence or severity of gastrointestinal symptoms in viral or bacterial infections such as gastroenteritis [729,730]. Nevertheless, it has been proposed that the administration of probiotics to COVID-19 patients and healthcare workers may prevent or ameliorate the gastrointestinal symptoms of COVID-19, a hypothesis that several clinical trials are now preparing to investigate [731,732]. Other studies are investigating whether probiotics may affect patient outcomes following SARS-CoV-2 infection [733].

Generally, the efficacy of probiotic use is a controversial topic among scientists. In Europe, EFSA has banned the term probiotics on products labels, which has elicited either criticism for EFSA or support for probiotics from researchers in the field [689,734,735]. This is due to the hyperbolic claims placed on the labels of various probiotic products, which lack rigorous scientific data to support their efficacy. Overall, the data supporting probiotics in the treatment or prevention of many different disorders and diseases is not conclusive as the quality of the evidence is generally considered low [712]. However, in the case of probiotics and respiratory infections, the evidence seems to be supportive of their potential therapeutic value. Consequently, several investigations are underway to investigate the prophylactic and therapeutic potential of probiotics for COVID-19. However, blind use of conventional probiotics for COVID-19 is cautioned against until the pathogenesis of SARS-CoV-2 is further established [736]. Until clinical trials investigating the prophylactic and therapeutic potential of probiotics for COVID-19 are complete, it is not possible to provide an evidence-based recommendation for their use. Despite these concerns, complementary use of probiotics as an adjuvant therapeutic has been proposed by the Chinese National Health Commission and National Administration of Traditional Chinese Medicine [87].

6.9 Discussion

The purported prophylactic and therapeutic benefits of dietary supplements and nutraceuticals for multiple disorders, diseases, and infections has been the subject of significant research and debate for the last few decades. Inevitably, scientists are also investigating the potential for these various products to treat or prevent COVID-19. This interest also extends to consumers, which led to a remarkable increase of sales of dietary supplements and nutraceuticals throughout the pandemic due to a desire to obtain additional protections from infection and disease. It is pertinent to clarify that the nutraceuticals discussed in this review, namely vitamin C, vitamin D, n-3 PUFA, zinc, and probiotics, have been selected because of their possible relevance to the biological mechanisms that can beneficially affect viral and respiratory infections and because they are currently under clinical investigation. However, there are various other products and molecules that have garnered scientific interest and require further investigation. These include polyphenols, lipid extracts, and tomato-based nutraceuticals among others, which have all been suggested for the prevention of potential cardiovascular complications of COVID-19 such as thrombosis [561,571]. Melatonin is another supplement that has been identified as a potential antiviral agent against SARS-CoV-2 using computational methods [737], and it has also been highlighted as a potential therapeutic agent for COVID-19 due to its documented antioxidant, anti-apoptotic, immunomodulatory, and anti-inflammatory effects [571,738,739]. Notably, melatonin, vitamin D and zinc were included in the treatment plan of the Former President of the United States upon hospitalization due to COVID-19 [740]. These are just some of the many substances and supplements under investigation that currently have limited evidence to support their use or supplementation with the aim to prevent or treat COVID-19. While there is plenty speculation surrounding the use of supplements, a survey of 327,720 users of the COVID Symptom Study App found that the consumption of n-3 PUFA supplements, probiotics, multivitamins, and vitamin D was associated with a lower risk of SARS-CoV-2 infection in women but not men after adjusting for potential confounders [741]. According to the authors, the sexual dimorphism observed may in part be because supplements may better support females due to known differences between the male and female immune systems, or it could be due to behavioral and health consciousness differences between the sexes [741]. Certainly, randomized controlled trials are required to investigate these findings further.

In this review, we report the findings to date for several dietary supplements and nutraceuticals. While existing evidence suggests potential benefits of n-3 PUFA and probiotic supplementation for COVID-19 treatment and prophylaxis, clinical data is lacking but trials are underway. Both zinc and vitamin C supplementation in hospitalized patients seem to be associated with positive outcomes; however, further clinical trials are required. In any case, both vitamin C and zinc intake are part of a healthy diet and both likely presents minimal risk when supplemented for, but their potential prophylactic or therapeutic effects against COVID-19 are yet to be determined. On the other hand, mounting evidence from observational studies indicates that there is an association between vitamin D deficiency and COVID-19 incidence has also been supported by meta-analysis [742]. Indeed, scientists are working to confirm these findings and to determine whether a patient’s serum 25-hydroxyvitamin D levels are also associated with COVID-19 severity. Clinical trials are required to determine whether preemptive vitamin D supplementation may mitigate against severe COVID-19. In terms of the therapeutic potential of vitamin D, initial evidence from clinical trials are conflicting, but seem to indicate that vitamin D supplementation may reduce COVID-19 severity [668]. The various clinical trials currently underway will be imperative to provide information on the efficacious use of vitamin D supplementation for COVID-19 prevention and/or treatment.

Finally, it is known that a patient’s nutritional status affects health outcomes in various infectious diseases [535], and COVID-19 is no different [533,743,744]. Some of the main risk factors for severe COVID-19, which also happen to be linked to poor nutritional status includes obesity, hypertension, cardiovascular diseases, type II diabetes mellitus, and indeed age-related malnutrition [531,533,745]. Although not the main focus of this review, it is important to consider the nutritional challenges associated with severe COVID-19 patients. Hospitalized COVID-19 patients tend to report an unusually high loss of appetite preceding admission, some suffer diarrhea and gastrointestinal symptoms that result in significantly lower food intake, and patients with poorer nutritional status were more likely to have worse outcomes and require nutrition therapy [746]. Dysphagia also seems to be a significant problem in pediatric patients that suffered multisystem inflammatory syndrome (MIS-C) [747] and rehabilitating COVID-19 patients, potentially contributing to poor nutritional status [748]. Almost two-thirds of discharged COVID-19 ICU patients exhibit significant weight loss, of which 26% had weight loss greater than 10% [744]. As investigated in this review, hospitalized patients also tend to exhibit vitamin D deficiency or insufficiency, which may be associated with greater disease severity [742]. Therefore, further research is required to determine how dietary supplements and nutraceuticals may contribute to the treatment of severely ill and rehabilitating patients who often rely on enteral nutrition.

6.10 Conclusions

Despite all the potential benefits of nutraceutical and dietary supplement interventions presented, currently there is a paucity of clinical evidence to support their use for the prevention or mitigation of COVID-19 infection. Nevertheless, optimal nutritional status can prime an individual’s immune system to protect against the effects of acute respiratory viral infections by supporting normal maintenance of the immune system [531,535]. Nutritional strategies can also play a role in the treatment of hospitalized patients, as malnutrition is a risk to COVID-19 patients [748]. Overall, supplementation of vitamin C, vitamin D, and zinc may be an effective method of ensuring their adequate intake to maintain optimal immune function, which may also convey beneficial effects against viral infections due to their immunomodulatory effects. Individuals should pay attention to their nutritional status, particularly their intake of vitamin D, considering its deficiency tends to be widespread. The prevailing evidence seems to indicate an association between vitamin D deficiency with COVID-19 incidence and, potentially, severity [643]. As a result, some international authorities have advised the general public, particularly those at high risk of infection, to consider vitamin supplementation. However, further well-controlled clinical trials are required to confirm the observations presented in this review. Many supplements and nutraceuticals designed for various ailments are available in the United States and beyond that are not strictly regulated [749]. Consequently, there can be safety and efficacy concerns associated with many of these products. Often, the vulnerable members of society can be exploited in this regard and unfortunately the COVID-19 pandemic is no different. As mentioned above, the FDA has issued warnings to several companies for advertising falsified claims in relation to the preventative and therapeutic capabilities of their products against COVID-19 [750]. Further intensive investigation is required to establish the effects of these nutraceuticals, if any, against COVID-19. Until more effective therapeutics are established, the most effective mitigation strategies consist of encouraging standard public health practices such as regular hand washing with soap, wearing a face mask, and covering a cough with your elbow [751], along with following social distancing measures, “shelter in place” guidelines, expansive testing, and contact tracing [752,753]. Indeed, in light of this review, it would also be pertinent to adopt a healthy diet and lifestyle following national guidelines in order to maintain optimal immune health.

7 Social Factors Influencing COVID-19 Exposure and Outcomes

7.1 Social Factors Influencing COVID-19 Outcomes

In addition to understanding the fundamental biology of the SARS-CoV-2 virus and COVID-19, it is critical to consider how the broader environment can influence both COVID-19 outcomes and efforts to develop and implement treatments for the disease. The evidence clearly indicates that social environmental factors are critical determinants of individuals’ and communities’ risks related to COVID-19. There are distinct components to COVID-19 susceptibility, and an individual’s risk can be elevated at one or all stages from exposure to recovery/mortality: an individual may be more likely to be exposed to the virus, more likely to get infected once exposed, more likely to have serious complications once infected, and be less likely to receive adequate care once they are seriously ill. The fact that differences in survival between Black and white patients were no longer significant after controlling for comorbidities and socioeconomic status (type of insurance, neighborhood deprivation score, and hospital where treatment was received) in addition to sex and age [754] underscores the relevance of social factors to understanding mortality differences between racial and ethnic groups. Moreover, the Black patients were younger and more likely to be female than white patients, yet still had a higher mortality rate without correction for the other variables [754]. Here, we outline a few systemic reasons that may exacerbate the COVID-19 pandemic in communities of color.

7.2 Factors Observed to be Associated with Susceptibility

As COVID-19 has spread into communities around the globe, it has become clear that the risks associated with this disease are not equally shared by all individuals or all communities. Significant disparities in outcomes have led to interest in the demographic, biomedical, and social factors that influence COVID-19 severity. Untangling the factors influencing COVID-19 susceptibility is a complex undertaking. Among patients who are admitted to the hospital, outcomes have generally been poor, with rates of admission to the intensive care unit (ICU) upwards of 15% in both Wuhan, China and Italy [34,755,756]. However, hospitalization rates vary by location [757]. This variation may be influenced by demographic (e.g., average age in the area), medical (e.g., the prevalence of comorbid conditions such as diabetes), and social (e.g., income or healthcare availability) factors that vary geographically. Additionally, some of the same factors may influence an individual’s probability of exposure to SARS-CoV-2, their risk of developing a more serious case of COVID-19 that would require hospitalization, and their access to medical support. As a result, quantifying or comparing susceptibility among individuals, communities, or other groups requires consideration of a number of complex phenomena that intersect across many disciplines of research. In this section, the term “risk factor” is used to refer to variables that are statistically associated with more severe COVID-19 outcomes. Some are intrinsic characteristics that have been observed to carry an association with variation in outcomes, whereas others may be more functionally linked to the pathophysiology of COVID-19.

7.2.1 Patient Traits Associated with Increased Risk

Two traits that have been consistently associated with more severe COVID-19 outcomes are male sex and advanced age (typically defined as 60 or older, with the greatest risk among those 85 and older [758]). In the United States, males and older individuals diagnosed with COVID-19 were found to be more likely to require hospitalization [759,760]. A retrospective study of hospitalized Chinese patients [35] found that a higher probability of mortality was associated with older age, and world-wide, population age structure has been found to be an important variable for explaining differences in outbreak severity [761]. The CFR for adults over 80 has been estimated upwards of 14% or even 20% [762]. Male sex has also been identified as a risk factor for severe COVID-19 outcomes, including death [763,764,765]. Early reports from China and Europe indicated that even though the case rates were similar across males and females, males were at elevated risk for hospital admission, ICU admission, and death [764], although data from some US states indicates more cases among females, potentially due to gender representation in care-taking professions [766]. In older age groups (e.g., age 60 and older), comparable absolute numbers of male and female cases actually suggests a higher rate of occurrence in males, due to increased skew in the sex ratio [764]. Current estimates based on worldwide data suggest that, compared to females, males may be 30% more likely to be hospitalized, 80% more likely to be admitted to the ICU, and 40% more likely to die as a result of COVID-19 [765]. There also may be a compounding effect of advanced age and male sex, with differences time to recovery worst for males over 60 years old relative to female members of their age cohort [767].

Both of these risk factors can be approached through the lens of biology. The biological basis for greater susceptibility with age is likely linked to the prevalence of extenuating health conditions such as heart failure or diabetes [762]. Several hypotheses have been proposed to account for differences in severity between males and females. For example, some evidence suggests that female sex hormones may be protective [764,766]. ACE2 expression in the kidneys of male mice was observed to be twice as high as that of females, and a regulatory effect of estradiol on ACE2 expression was demonstrated by removing the gonads and then supplementing with estradiol [766,768]. Other work in mice has shown an inverse association between mortality due to SARS-CoV-1 and estradiol, suggesting a protective role for the sex hormone [766]. Similarly, evidence suggests that similar patterns might be found in other tissues. A preliminary analysis identified higher levels of ACE2 expression in the myocardium of male patients with aortic valve stenosis showed than female patients, although this pattern was not found in controls [764]. Additionally, research has indicated that females respond to lower doses than males of heart medications that act on the Renin angiotensin aldosterone system (RAAS) pathway, which is shared with ACE2 [764]. Additionally, several components of the immune response, including the inflammatory response, may differ in intensity and timing between males and females [766,768]. This hypothesis is supported by some preliminary evidence showing that female patients who recovered from severe COVID-19 had higher antibody titers than males [766]. Sex steroids can also bind to immune cell receptors to influence cytokine production [764]. Additionally, social factors may influence risks related to both age and sex: for example, older adults are more likely to live in care facilities, which have been a source for a large number of outbreaks [769], and gender roles may also influence exposure and/or susceptibility due to differences in care-taking and/or risky behaviors (e.g., caring for elder relatives and smoking, respectively) [764] among men and women (however, it should be noted that both transgender men and women are suspected to be at heightened risk [770].)

7.2.2 Comorbid Health Conditions

A number of pre-existing or comorbid conditions have repeatedly been identified as risk factors for more severe COVID-19 outcomes. Several underlying health conditions were identified at high prevalence among hospitalized patients, including obesity, diabetes, hypertension, lung disease, and cardiovascular disease [757]. Higher Sequential Organ Failure Assessment (SOFA) scores have been associated with a higher probability of mortality [35], and comorbid conditions such as cardiovascular and lung disease as well as obesity were also associated with an increased risk of hospitalization and death, even when correcting for age and sex [763]. Diabetes may increase the risk of lengthy hospitalization [771] or of death [771,772]. [773] and [774] discuss possible ways in which COVID-19 and diabetes may interact. Obesity also appears to be associated with higher risk of severe outcomes from SARS-CoV-2 [775,776]. Obesity is considered an underlying risk factor for other health problems, and the mechanism for its contributions to COVID-19 hospitalization or mortality is not yet clear [777]. Dementia and cancer were also associated with the risk of death in an analysis of a large number (more than 20,000) COVID-19 patients in the United Kingdom [763]. It should be noted that comorbid conditions are inextricably tied to age, as conditions tend to be accumulated over time, but that the prevalence of individual comorbidities or of population health overall can vary regionally [778]. Several comorbidities that are highly prevalent in older adults, such as COPD, hypertension, cardiovascular disease, and diabetes, have been associated with CFRs upwards of 8% compared to an estimate of 1.4% in people without comorbidities [762,779]. Therefore, both age and health are important considerations when predicting the impact of COVID-19 on a population [778]. However, other associations may exist, such as patients with sepsis having higher SOFA scores – in fact, SOFA was developed for the assessment of organ failure in the context of sepsis, and the acronym originally stood for Sepsis-Related Organ Failure Assessment [780,781]. Additionally, certain conditions are likely to be more prevalent under or exacerbated by social conditions, especially poverty, as is discussed further below.

7.2.3 Ancestry

A number of studies have suggested associations between individuals’ racial and ethnic backgrounds and their COVID-19 risk. In particular, Black Americans are consistently identified as carrying a higher burden of COVID-19 than white Americans [759,760], with differences in the rates of kidney complications from COVID-19 particularly pronounced [91]. Statistics from a number of cities indicate significant discrepancies between the proportion of COVID-19 cases and deaths in Black Americans relative to their representation in the general population [681]. In addition to Black Americans, disproportionate harm and mortality from COVID-19 has also been noted in Latino/Hispanic Americans and in Native American and Alaskan Native communities, including the Navajo nation [782,783,784,785,786,787]. In Brazil, indigenous communities likewise carry an increased burden of COVID-19 [788]. In the United Kingdom, nonwhite ethnicity (principally Black or South Asian) was one of several factors found to be associated with a higher risk of death from COVID-19 [789].

From a genetic standpoint, it is highly unlikely that ancestry itself predisposes individuals to contracting COVID-19 or to experiencing severe COVID-19 outcomes. Examining human genetic diversity indicates variation over a geographic continuum, and that most human genetic variation is associated with the African continent [790]. African-Americans are also a more genetically diverse group relative to European-Americans, with a large number of rare alleles and a much smaller fraction of common alleles identified in African-Americans [791]. Therefore, the idea that African ancestry (at the continent level) might convey some sort of genetic risk for severe COVID-19 contrasts with what is known about worldwide human genetic diversity [792]. The possibility for genetic variants that confer some risk or some protection remains possible, but has not been widely explored, especially at a global level. Research in Beijing of a small number (n=80) hospitalized COVID-19 patients revealed an association between severe COVID-19 outcomes and homozygosity for an allele in the interferon-induced transmembrane protein 3 (IFITM3) gene, which was selected as a candidate because it was previously found to be associated with influenza outcomes in Chinese patients [793]. Genetic factors may also play a role in the risk of respiratory failure for COVID-19 [794,795,796]. However, genetic variants associated with outcomes within ancestral groups are far less surprising than genetic variants explaining outcomes between groups. Alleles in ACE2 and TMPRSS2 have been identified that vary in frequency among ancestral groups [797], but whether these variants are associated with COVID-19 susceptibility has not been explored.

Instead, examining patterns of COVID-19 susceptibility on a global scale that suggest that social factors are of primary importance in predicting mortality. Reports from several sub-Saharan African countries have indicated that the effects of the COVID-19 pandemic have been less severe than expected based on the outbreaks in China and Italy. In Kenya, for example, estimates of national prevalence based on testing blood donors for SARS-CoV-2 antibodies were consistent with 5% of Kenyan adults having recovered from COVID-19 [798]. This high seroprevalence of antibodies lies in sharp contrast to the low number of COVID-19 fatalities in Kenya, which at the time was 71 out of 2093 known cases [798]. Likewise, a serosurvey of health care workers in Blantyre City, Malawi reported an adjusted antibody prevalence of 12.3%, suggesting that the virus had been circulating more widely than thought and that the death rate was up eight times lower than models had predicted [799]. While several possible hypotheses for the apparent reduced impact of COVID-19 on the African continent are being explored, such as young demographics in many places [800], these reports present a stark contrast to the severity of COVID-19 in Americans and Europeans of African descent. Additionally, ethnic minorities in the United Kingdom also tend to be younger than white British living in the same areas, yet the burden of COVID-19 is still more serious for minorities, especially people of Black Caribbean ancestry, both in absolute numbers and when controlling for age and location [801]. Furthermore, the groups in the United States and United Kingdom that have been identified as carrying elevated COVID-19 burden, namely Black American, indigenous American, and Black and South Asian British, are quite distinct in their position on the human ancestral tree. What is shared across these groups is instead a history of disenfranchisement under colonialism and ongoing systematic racism. A large analysis of over 11,000 COVID-19 patients hospitalized in 92 hospitals across U.S. states revealed that Black patients were younger, more often female, more likely to be on Medicaid, more likely to have comorbidities, and came from neighborhoods identified as more economically deprived than white patients [754]. This study reported that when these factors were accounted for, the differences in mortality between Black and white patients were no longer significant. Thus, the current evidence suggests that the apparent correlations between ancestry and health outcomes must be examined in the appropriate social context.

7.3 Environmental Influences on Susceptibility

7.3.1 Exposure to COVID-19

Social distancing has emerged as one of the main social policies used to manage the COVID-19 epidemic in many countries. Many governments issued stay-at-home orders, especially in the initial months of the crisis. However, data clearly indicates that these orders impacted different socioeconomic groups differently. In U.S. counties with and without stay-at-home orders, smartphone tracking indicated a significant decrease in the general population’s mobility in April relative to February through March of 2020 (-52.3% and -60.8%, respectively) [802]. A linear relationship was observed between counties’ reduction in mobility and their wealth and health, as measured by access to health care, food security, income, space, and other factors [802]. Counties with greater reductions in mobility were also found to have much lower child poverty and household crowding and to be more racially segregated, and to have fewer youth and more elderly residents [802]. Similar associations between wealth and decreased mobility were observed in cellphone GPS data from Colombia, Indonesia, and Mexico collected between January and May 2020 [803], as well as in a very large data set from several US cities [804]. These disparities in mobility are likely to be related to the role that essential workers have played during the pandemic. Essential workers are disproportionately likely to be female, people of color, immigrants, and to have an income below 200% of the poverty line [805]. Black Americans in particular are over-represented among front-line workers and in professions where social distancing is infeasible [806]. Health care work in particular presents an increased risk of exposure to SARS-CoV-2 [806,807,808,809,810]. In the United Kingdom, (South) Asians are more likely than their white counterparts to be medical professionals [801], although BAME medical professionals are still disproportionately represented in the proportion of National Health Service staff deaths [811]. Similar trends have been reported for nurses, especially nurses of color, in the United States [812]. Furthermore, beyond the risks associated with work itself, use of public transportation may also impact COVID-19 risk [813]. The socioeconomic and racial/ethnic gaps in who is working on the front lines of the pandemic make it clear that socioeconomic privilege is likely to decrease the probability of exposure to SARS-CoV-2.

Increased risk of exposure can also arise outside the workplace. Nursing homes and skilled nursing facilities received attention early on as high-risk locations for COVID-19 outbreaks [814]. Prisons and detention centers also confer a high risk of exposure or infection [815,816]. Populations in care facilities are largely older adults, and in the United States, incarcerated people are more likely to be male and persons of color, especially Black [817]. Additionally, multi-generational households are less common among non-Hispanic white Americans than people of other racial and ethnic backgrounds [818], increasing the risk of exposure for more susceptible family members. Analysis suggests that household crowding may also be associated with increased risk of COVID-19 exposure [802], and household crowding is associated with poverty [819]. Forms of economic insecurity like housing insecurity, which is associated with poverty and more pronounced in communities subjected to racism [820,821], would be likely to increase household crowding and other possible sources of exposure. As a result, facets of systemic inequality such as mass incarceration of Black Americans and poverty are likely to increase the risk of exposure outside of the workplace.

7.3.2 Severity of COVID-19 Following Exposure

Following exposure to SARS-CoV-2, the likelihood that an individual develops COVID-19 and the severity of the disease presentation can be influenced by a number of social factors. As discussed above, a number of patient characteristics are associated with the likelihood of severe COVID-19 symptoms. In some cases, these trends run counter to those expected given rates of exposure: for example, although women are more likely to be exposed, men are more likely to be diagnosed with, hospitalized from, or die from COVID-19 [766]. In the case of comorbid conditions and racial/ethnic demographics, however, social factors are highly likely to modulate or at least influence the apparent association between these traits and the increased risk from COVID-19. In particular, the comorbidities and racial/ethnic correlates of severe COVID-19 outcomes suggest that poverty confers additional risk for COVID-19.

In order to explore the relationship between poverty and COVID-19 outcomes, it is necessary to consider how poverty impacts biology. In particular, we focus on the United States and the United Kingdom. Comorbidities that increase risk for COVID-19, including obesity, type II diabetes, hypertension, and cardiovascular disease, are known to be intercorrelated [822]. Metabolic conditions related to heightened inflammation, like obesity, type II diabetes, and hypertension, are more strongly associated with negative COVID-19 outcomes than other comorbid conditions, such as chronic heart disease [823]. As discussed above, dysregulated inflammation characteristic of cytokine release syndrome is one of the greatest concerns for COVID-19-related death. Therefore, it is possible that chronic inflammation characteristic of these metabolic conditions predisposes patients to COVID-19-related death [823]. The association between these diseases and severe COVID-19 outcomes is a concern from a health equity perspective because poverty exposes people to “obesogenic” conditions [824] and is therefore unsurprisingly associated with higher incidence of obesity and associated disorders [825]. Furthermore, cell phone GPS data suggests that lower socioeconomic status may also be associated with decreased access to healthy food choices during the COVID-19 pandemic [826,827], suggesting that health-related risk factors for COVID-19 may be exacerbated as the pandemic continues [828]. Chronic inflammation is a known outcome of chronic stress (e.g., [829,830,831,832]). Therefore, the chronic stress of poverty is likely to influence health broadly (as summarized in [833]) and especially during the stress of the ongoing pandemic.

A preprint [834] provided observational evidence that geographical areas in the United States that suffer from worse air pollution by fine particulate matter have also suffered more COVID-19 deaths per capita, after adjusting for demographic covariates. Although lack of individual-level exposure data and the impossibility of randomization make it difficult to elucidate the exact causal mechanism, this finding would be consistent with similar findings for all-cause mortality (e.g., [835]). Exposure to air pollution is associated with both poverty (e.g., [836]) and chronic inflammation [837]. Other outcomes of environmental racism, such as the proximity of abandoned uranium mines to Navajo land, can also cause respiratory illnesses and other health issues [787]. Similarly, preliminary findings indicate that nutritional status (e.g., vitamin D deficiency [652]) may be associated with COVID-19 outcomes, and reduced access to grocery stores and fresh food often co-occurs with environmental racism [787,838]. Taken together, the evidence suggests that low-income workers who face greater exposure to SARS-CoV-2 due to their home or work conditions are also more likely to face environmental and social stressors associated with increased inflammation, and therefore with increased risk from COVID-19. In particular, structural racism can play an important role on disease severity after SARS-CoV-2 exposure, due to consequences of racism which include an increased likelihood of poverty and its associated food and housing instability. COVID-19 can thus be considered a “syndemic”, or a synergistic interaction between several epidemics [839]. As a result, it is not surprising that people from minoritized backgrounds and/or with certain pre-existing conditions are more likely to suffer severe effects of COVID-19, but these “risk factors” are likely to be causally linked to poverty.

7.3.3 Access to Treatment

Finally, COVID-19 outcomes can be influenced by access to healthcare. Receiving care for COVID-19 can, but does not always, include receiving a positive test for the SARS-CoV-2 virus. For example, it is common to see treatment guidelines for suspected cases regardless of whether the presence of SARS-CoV-2 has been confirmed (e.g., [840]). Whether and where a patient is diagnosed can depend on their access to testing, which can vary both between and within countries. In the United States, it is not always clear whether an individual will have access to free testing [841,842]. The concern has been raised that more economic privilege is likely to correspond to increased access to testing, at least within the United States [843]. This is supported by the fact that African Americans seem to be more likely to be diagnosed in the hospital, while individuals from other groups were more likely to have been diagnosed in ambulatory settings in the community [759]. Any delays in treatment are a cause for concern [843], which could potentially be increased by an inability to acquire testing because in the United States, insurance coverage for care received can depend on a positive test [844].

Another important question is whether patients with moderate to severe cases are able to access hospital facilities and treatments, to the extent that they have been identified. Early findings from China as of February 2020 suggested the COVID-19 mortality rate to be much lower in the most developed regions of the country [845], although reported mortality is generally an estimate of CFR, which is dependent on rates of testing. Efforts to make treatment accessible for all confirmed and suspected cases of COVID-19 in China are credited with expanding care to people with fewer economic resources [846]. In the United States, access to healthcare varies widely, with certain sectors of the workforce less likely to have health insurance; many essential workers in transportation, food service, and other frontline fields are among those likely to be uninsured or underinsured [843]. As of 2018, Hispanic Americans of all races were much less likely to have health insurance than people from non-Hispanic backgrounds [847]. Therefore, access to diagnostics and care prior to the development of severe COVID-19 is likely to vary depending on socioeconomic and social factors, many of which overlap with the risks of exposure and of developing more severe COVID-19 symptoms. This discrepancy ties into concerns about broad infrastructural challenges imposed by COVID-19. A major concern in many countries has been the saturation of healthcare systems due to the volume of COVID-19 hospitalizations (e.g., [227]). Similarly, there have been shortages of supplies such as ventilators that are critical to the survival of many COVID-19 patients, leading to extensive ethical discussions about how to allocate limited resources among patients [848,849,850,851]. Although it is generally considered unethical to consider demographic factors such as age, sex, race, or ethnicity while making such decisions, and ideally this information would not be shared with triage teams tasked with allocating limited resources among patients [852], there are substantial concerns about implicit and explicit biases against older adults [853], premature infants [854], and people with disabilities or comorbidities [852,855,856]. Because of the greater burden of chronic disease in populations subjected to systemic racism, algorithms intended to be blind to race and ethnicity could, in fact, reinforce systemic inequalities caused by structural racism [857,858,859]. Because of this inequality, it has been argued that groups facing health disparities should be prioritized by these algorithms [860]. This approach would carry its own ethical concerns, including the fact that many resources that need to be distributed do not have well-established risks and benefits [860].

As the pandemic has progressed, it has become clear that ICU beds and ventilators are not the only limited resources that needs to be allocated, and, in fact, the survival rate for patients who receive mechanical ventilation is lower than these discussions would suggest [861]. Allocation of interventions that may reduce suffering, including palliative care, has become critically important [861,862]. The ambiguities surrounding the risks and benefits associated with therapeutics that have been approved under emergency use authorizations also present ethical concerns related to the distribution of resources [860]. For example, remdesivir, discussed above, is currently available for the treatment of COVID-19 under compassionate use guidelines and through expanded access programs, and in many cases has been donated to hospitals by Gilead [863,864]. Regulations guiding the distribution of drugs in situations like these typically do not address how to determine which patients receive them [864]. Prioritizing marginalized groups for treatment with a drug like remdesivir would also be unethical because it would entail disproportionately exposing these groups to a therapeutic that may or not be beneficial [860]. On the other hand, given that the drug is one of the most promising treatments available for many patients, using a framework that tacitly feeds into structural biases would also be unethical. At present, the report prepared for the Director of the CDC by Ethics Subcommittee of the CDC fails to address the complexity of this ethical question given the state of structural racism in the United States, instead stating that “prioritizing individuals according to their chances for short-term survival also avoids ethically irrelevant considerations, such as race or socioeconomic status” [865]. In many cases, experimental therapeutics are made available only through participation in clinical trials [866]. However, given the history of medical trials abusing minority communities, especially Black Americans, there is a history of unequal representation in clinical trial enrollment [866]. As a result, the standard practice of requiring enrollment in a clinical trial in order to receive experimental treatment may also reinforce patterns established by systemic racism.

7.3.4 Access to and Representation in Clinical Trials

Experimental treatments are often made available to patients primarily or even exclusively through clinical trials. The advantage of this approach is that clinical trials are designed to collect rigorous data about the effects of a treatment on patients. The disadvantage is that access to clinical trials is not equal among all people who suffer from a disease. Two important considerations that can impact an individual’s access to clinical trials are geography and social perceptions of clinical trials. For the first, the geographic distribution of trial recruitment efforts are typically bounded and can vary widely among difference locations, and for the second, the social context of medical interactions can impact strategies for and the success of outreach to different communities. Differential access to clinical trials raises concerns because it introduces biases that can influence scientific and medical research on therapeutics and prophylactics broadly. Concerns about bias in clinical trials need to address both trial recruitment and operation. In the present crisis, such biases are particularly salient because COVID-19 is a disease of global concern. Treatment is needed by people all over the world, and clinical research that characterizes treatment outcomes in a variety of populations is critically important.

Global representation in clinical trials is important to ensuring that experimental treatments are available equally to COVID-19 patients who may need them. The advantage to a patient of participation in a clinical trial is that they may receive an experimental treatment they would not have been able to access otherwise. The potential downsides of participation include that the efficacy and side effects of such treatments are often poorly characterized and that patients who enroll in clinical trials will in some cases run the risk of being assigned to a placebo condition where they do not receive the treatment but miss out on opportunities to receive other treatments. The benefits and burdens of clinical trials therefore need to be weighed carefully to ensure that they don’t reinforce existing health disparities. The WHO Director‐General Tedros Adhanom Ghebreyesus stated his condemnation of utilizing low and middle income countries as test subjects for clinical trials, yet having highly developed countries as the majority of clinical trial representation is also not the answer [867]. Figure 6 showcases two choropleths detailing COVID-19 clinical trial recruitment by country. China, the United States, and France are among the countries with the most clinical trial recruiting for trials with single-country enrollment. Many countries have little to no clinical trial recruiting, with the continents of Africa and South America much less represented than Asia, Europe, and North America. Trials that recruit across multiple countries do appear to broaden geographic representation, but these trials seem to be heavily dominated by the United States and European Union.

Figure 6: Geographic distribution of COVID-19 clinical trials. The density of clinical trials is reported at the country level. As of November 9, 2020, there are 6,417 trials in the University of Oxford Evidence-Based Medicine Data Lab’s COVID-19 TrialsTracker [293]. The top figure demonstrates the density of trials recruiting only from a singular country, while the bottom shows the distribution of recruitment for trials that involve more than one country.

A few different concerns arise from this skewed geographic representation in clinical trial recruitment. First, treatments such as remdesivir that are promising but primarily available to clinical trial participants are unlikely to be accessible by people in many countries. Second, it raises the concern that the findings of clinical trials will be based on participants from many of the wealthiest countries, which may lead to ambiguity in whether the findings can be extrapolated to COVID-19 patients elsewhere. Especially with the global nature of COVID-19, equitable access to therapeutics and vaccines has been a concern at the forefront of many discussions about policy (e.g., [868], yet data like that shown in Figure 6 demonstrates that accessibility is likely to be a significant issue. Another concern with the heterogeneous international distribution of clinical trials is that the governments of countries leading these clinical trials might prioritize their own populations once vaccines are developed, causing unequal health outcomes [869]. Additionally, even within a single state in the United States (Maryland), geography was found to influence the likelihood of being recruited into or enrolled in a clinical trial, with patients in under-served rural areas less likely to enroll [870]. Thus, geography both on the global and local levels may influence when treatments and vaccines are available and who is able to access them. Efforts such as the African Union’s efforts to coordinate and promote vaccine development [871] are therefore critical to promoting equity in the COVID-19 response.

Even when patients are located within the geographic recruitment area of clinical trials, however, there can still be demographic inequalities in enrollment. When efforts are made to ensure equal opportunity to participate in clinical trials, there is no significant difference in participation among racial/ethnic groups [872]. However, within the United States, real clinical trial recruitment numbers have indicated for many years that racial minorities, especially African-Americans, tend to be under-represented (e.g., [873,874,875,876]). This trend is especially concerning given the disproportionate impact of COVID-19 on African-Americans. Early evidence suggests that the proportion of Black, Latinx, and Native American participants in clinical trials for drugs such as remdesivir is much lower than the representation of these groups among COVID-19 patients [877].

One proposed explanation for differences among racial and ethnic groups in clinical trial enrollment refers to different experiences in healthcare settings. While some plausible reasons for the disparity in communication between physicians and patients could be a lack of awareness and education, mistrust in healthcare professionals, and a lack of health insurance [872], a major concern is that patients from certain racial and ethnic groups are marginalized even while seeking healthcare. In the United States, many patients experience “othering” from physicians and other medical professionals due to their race or other external characteristics such as gender (e.g., [878]). Many studies have sought to characterize implicit biases in healthcare providers and whether they affect their perceptions or treatment of patients. A systematic review that examined 37 such studies reported that most (31) identified racial and/or ethnic biases in healthcare providers in many different roles, although the evidence about whether these biases translated to different attitudes towards patients was mixed [879], with similar findings reported by a second systematic review [880]. However, data about real-world patient outcomes are very limited, with most studies relying on clinical vignette-based exercises [879], and other analyses suggest that physician implicit bias could impact the patient’s perception of the negativity/positivity of the interaction regardless of the physician’s explicit behavior towards the patient [881]. Because racism is a common factor in both, negative patient experiences with medical professionals are likely to compound other issues of systemic inequality, such as a lack of access to adequate care, a lack of insurance, or increased exposure to SARS-CoV-2 [882]. Furthermore, the experience of being othered is not only expected to impact patients’ trust in and comfort with their provider, but also may directly impact whether or not the patient is offered the opportunity to participate in a clinical trial at all. Some studies suggest communication between physicians and patients impacts whether or not a physician offers a patient participation in a clinical trial. For example, researchers utilized a linguistic analysis to assess mean word count of phrases related to clinical trial enrollment, such as voluntary participation, clinical trial, etc. [872]. The data indicated that the mean word count of the entire visit was 1.5 times more for white patients in comparison to Black patients. In addition, the greatest disparity between white and Black patients’ experience was the discussion of risks, with over 2 times as many risk-related words spoken with white patients than Black patients [872]. The trends observed for other clinical trials raise the concern that COVID-19 clinical trial information may not be discussed as thoroughly or as often with Black patients compared to white patients.

These discrepancies are especially concerning given that many COVID-19 treatments are being or are considered being made available to patients prior to FDA approval through Emergency Use Authorizations. In the past, African-Americans have been over-represented relative to national demographics in use of the FDA’s Exception From Informed Consent (EFIC) pathway [883]. Through this pathway, people who are incapacitated can receive an experimental treatment even if they are not able to consent and there is not sufficient time to seek approval from an authorized representative. This pathway presents concerns, however, when it is considered in the context of a long history of systematic abuses in medical experimentation where informed consent was not obtained from people of color, such as the Tuskegee syphilis experiments [884]. While the goal of EFIC approval is to provide treatment to patients who urgently need it, the combination of the ongoing legacy of racism in medicine renders this trend concerning. With COVID-19, efforts to prioritize people who suffer from systemic racism are often designed with the goal of righting some of these inequalities (e.g., [885]), but particular attention to informed consent will be imperative in ensuring these trials are ethical given that the benefits and risks of emerging treatments are still poorly characterized. Making a substantial effort to run inclusive clinical trials is also important because of the possibility that racism could impact how a patient responds to a treatment. For example, as discussed above, dexamethasone has been identified as a promising treatment for patients experiencing cytokine release syndrome, but the mechanism of action is tied to the stress response. A study from 2005 reported that Black asthma patients showed reduced responsiveness to dexamethasone in comparison to white patients and suggested Black patients might therefore require higher doses of the drug [886]. In the context of chronic stress caused by systemic racism, this result is not surprising: chronic stress is associated with dysregulated production of glucocorticoids [887] and glucocorticoid receptor resistance [888]. However, it underscores the critical need for treatment guidelines to take into account differences in life experience, which would be facilitated by the recruitment of patients from a wide range of backgrounds. Attention to the social aspects of clinical trial enrollment must therefore be an essential component of the medical research community’s response to COVID-19.

7.4 Conclusions and Future Directions

As the COVID-19 pandemic evolves, the scientific community’s response will be critical for identifying potential pharmacological and biotechnological developments that may aid in combating the virus and the disease it causes. However, this global crisis highlights the importance of mounting a response based on collaboration among a wide variety of disciplines. Understanding the basic science of the virus and its pathogenesis is imperative for identifying and envisioning possible diagnostic and therapeutic approaches; understanding how social factors can influence outcomes and shape implementation of a response is critical to disseminating any scientific advancements. Summarizing such a complex and ever-changing topic presents a number of challenges. This review represents the effort of over 50 contributors to distill and interpret the available information. However, this text represents a dynamic and evolving document, and we welcome continued contributions from all researchers who have insights into how these topics intersect. A multidisciplinary perspective is critical to understanding this evolving crisis, and in this review we seek to use open science tools to coordinate a response among a variety of researchers. We intend to publish additional updates as the situation evolves.

8 Discussion

As of October 2020 the SARS-CoV-2 virus remains a serious worldwide threat. The scientific community has responded by rapidly collecting and disseminating information about the SARS-CoV-2 virus and the associated illness, COVID-19. The rapid identification of the genomic sequence of the virus allowed for early contextualization of SARS-CoV-2 among other known respiratory viruses. The pathogen is a coronavirus that is closely related to SARS-CoV-1, which caused the SARS pandemics of the early 2000s. Knowing the phylogenetic context and genomic sequence of the virus then allowed for rapid insights into its structure and pathogenesis. As with other HCoV, the immune response to SARS-CoV-2 is likely driven by detection of its spike protein, which allows it to enter cells through the ACE2 receptor. Epithelial cells have also emerged as the major cellular target of the virus, contextualizing the respiratory and gastrointestinal symptoms that are frequently observed in COVID-19. However, as COVID-19 cases have been more widely characterized, it has become clear that the disease presentation is highly heterogeneous. Many cases, especially in younger adults, present with mild symptoms or even asymptomatically, while others, especially in older adults, can be severe or fatal. In children, the SARS-CoV-2 virus can present as two distinct diseases, COVID-19 or MIS-C. While the overall infection fatality rate remains unknown, estimates suggest that it is not more than 1%; however, the severity of many non-lethal cases makes COVID-19 an ongoing, significant concern.

Characterizing the rate of infection and fatality rates hinges on the availability of rapid and accurate diagnostic testing. Major advancements have been made in identifying diagnostic approaches. The development of diagnostic technologies have been rapid, beginning with the release of the SARS-CoV-2 viral genome sequence in January. As of October 2020, a range of diagnostic tests have become available. One class of tests uses PCR (RT-PCR or qRT-PCR) to assess the presence of SARS-CoV-2 RNA, while another typically uses ELISA to test for the presence of antibodies to SARS-CoV-2. The former approach is useful for identifying active infections, while the latter measures hallmarks of the immune response and therefore can detect either active infections or immunity gained from prior infection. Combining these tests leads to extremely accurate detection of SARS-CoV-2 infection (98.6%), but when used alone, PCR-based tests are recommended before 5.5 days after the onset of the illness and antibody tests after 5.5 days [287]. Other strategies for testing can also influence the tests’ accuracy, such as the use of nasopharyngeal swabs versus BALF [287], which allow for trade-offs between patient’s comfort and test sensitivity. Additionally, technologies such as digital PCR may allow for scale-up in the throughput of diagnostic testing, facilitating widespread testing. One major question that remains is whether people who recover from SARS-CoV-2 develop sustained immunity, and over what period this immunity is expected to last. Some reports have suggested that some patients may develop COVID-19 reinfections (e.g., [288]), but the rates of reinfection are currently unknown. Serologic testing combined with PCR testing will be critical to confirming purported cases of reinfection and to identifying the duration over which immunity is retained and to understanding reinfection risks.

One of the goals of characterizing the immune response is to identify strategies for the prophylactic enhancements of immunity. Though some concerns remain about the duration of sustained immunity for convalescents, vaccine development efforts are ongoing and show initial promising results. The Moderna trial, for example, reported that the neutralizing activity in participants who received two doses of the vaccine was similar to that observed in convalescent plasma. Vaccine development for COVID-19 is progressing rapidly compared to typical timelines, but vaccine development is still a lengthy process. In the meantime, some advances have also been made in the treatment of patients with COVID-19. As cases have become better characterized, it has become evident that many patients experience an initial immune response to the virus that is typically characterized by fever, cough, dyspnea, and related symptoms. However, the most serious concern is cytokine release syndrome, when the body’s immune response becomes dysregulated, resulting in an extreme inflammatory response. The RECOVERY trial, a large-scale, multi-arm trial enrolling about 15% of all COVID-19 patients in the United Kingdom, was the first to identify that the widely available steroid dexamethasone seems to be beneficial for patients suffering from this immune dysregulation [415]. Efforts to identify therapeutic treatments to treat patients early in the course of infection have been more ambiguous. Early interest in the drugs hydroxychloroquine and chloroquine yielded no promising results from studies with robust experimental designs. The experimental drug remdesivir, which was developed for Ebola, has received enough support from early analyses to receive FDA approval, but results have been mixed. The potential for other drugs, such as tocilizumab, to reduce recovery time remains unclear, but some early results were promising.

8.1 Additional Therapeutics of Interest

Given what is currently known about these therapeutics for COVID-19, a number of related therapies beyond those explored above may also prove to be of interest. For example, the demonstrated benefit of dexamethasone and the ongoing potential of tocilizumab for treatment of COVID-19 suggests that other anti-inflammatory agents might also hold value for the treatment of COVID-19. Given that current evidence about treating COVID-19 with dexamethasone suggests that the need to curtail the cytokine storm inflammatory response to the virus can transcend the risks of immunosuppression, exploration of more anti-inflammatory agents may be warranted. While dexamethasone is considered widely available and generally affordable, the high costs of biologics such as tocilizumab therapy may present obstacles to wide-scale distribution of this drug if it proves of value. At the doses used for rheumatoid arthritis patients, the cost for tocilizumab ranges from $179.20 to $896 per dose for the IV form and $355 for the pre-filled syringe [517]. There are several anti-inflammatory agents used for the treatment of autoimmune diseases that may also be able to counter the effects of the cytokine storm induced by the virus, some of which, such as cyclosporine, are likely to be more cost-effective and readily available than biologics [518]. While tocilizumab targets IL-6, several other inflammatory markers could be potential targets, including TNF-alpha. Inhibition of TNF-alpha by an inhibitor such as Etanercept has been previously suggested for treatment of SARS-CoV-1 [519] and may be relevant for SARS-CoV-2 as well. Another anti-IL-6 antibody, sarilumab, is also being investigated [520,521]. Baricitinib and other small molecule inhibitors of the Janus-activated kinase pathway also curtail the inflammatory response and have been suggested as potential options for SARS-CoV-2 infections [522]. Baricitinib in particular may be able to reduce the ability of SARS-CoV-2 to infect lung cells [523]. Clinical trials studying baricitinib in COVID-19 have already begun in the US and in Italy [524,525]. Identification and targeting of further inflammatory markers that are relevant in SARS-CoV-2 infection may be of value for curtailing the inflammatory response and lung damage.

In addition to immunosuppressive treatments that are most beneficial late in disease progression, much research is focused on identifying treatments would be likely to benefit early-stage patients. For example, although studies of hydroxychloroquine have not supported the early theory-driven interest in this antiviral treatment, alternative compounds with related mechanisms may still have potential. Hydroxyferroquine derivatives of HCQ have been described as a class of bioorganometallic compounds that exert antiviral effects with some selectivity for SARS-CoV-1 in vitro [526]. Future work could explore whether such compounds exert antiviral effects against SARS-CoV-2 and whether they would be safer for use in COVID-19. Another potential approach is the development of antivirals, which could be broad-spectrum, specific to coronaviruses, or targeted to SARS-CoV-2. Development of new antivirals is complicated by the fact that none have yet been approved for human coronaviruses. Intriguing new options are emerging, however. Beta-D-N4-hydroxycytidine (NHC) is an orally bioavailable ribonucleotide analog showing broad-spectrum activity against RNA viruses, which may inhibit SARS-CoV-2 replication in vitro and in vivo in mouse models of HCoVs [527]. A range of other antivirals are also in development. Development of antivirals will be further facilitated as research reveals more information about the interaction of SARS-CoV-2 with the host cell and host cell genome, mechanisms of viral replication, mechanisms of viral assembly, and mechanisms of viral release to other cells; this can allow researchers to target specific stages and structures of the viral life cycle. Many researchers have also focused their attention to the potential use of dietary supplements and nutraceuticals. Indeed, there has been recent interest for the potential use of vitamin D as a prophylactic and therapeutic agent against COVID-19 as several observational studies have linked low vitamin D status to its incidence [643,651]. These associations have yet to be confirmed and rigorous trials are required before considering supplementation recommendations. However, the nutritional status and general health of a patient can affect their outcomes in various diseases, thus it would be pertinent to advise people to follow a healthy diet and life style to the best of their ability to prevent nutrient deficiencies and insufficiencies and to maintain a healthy immune system [533]. Finally, antibodies against viruses, also known as antiviral monoclonal antibodies, could be an alternative as well and are described in detail in an above section. The goal of antiviral antibodies is to neutralize viruses through either cell-killing activity or blocking of viral replication [528]. They may also engage the host immune response, encouraging the immune system to hone in on the virus. Given the cytokine storm that results from immune system activation in response to the virus, which has been implicated in worsening of the disease, a neutralizing antibody (nAb) may be preferable. Upcoming work may explore the specificity of nAbs for their target, mechanisms by which the nAbs impede the virus, and improvements to antibody structure that may enhance the ability of the antibody to block viral activity.

Some research is also investigating potential therapeutics and prophylactics that would interact with components of the innate immune response. For example, there are a variety of TLRs, PRRs that recognize PAMPs and DAMPs. TLRs form a part of innate immune recognition and can more generally contribute to promoting both innate and adaptive responses [106]. In mouse models, poly(I:C) and CpG, which are agonists of toll-like receptors TLR3 and TLR9, respectively, showed protective effects when administered prior to SARS-CoV-1 infection [529]. Therefore, TLR agonists hold some potential for broad-spectrum prophylaxis.

Given that a large number of clinical trials are currently in progress, more information about the potential of these and other therapeutics should become available over time. This information, combined with advances in understanding the molecular structure and viral pathogenesis of SARS-CoV-2, may lead to a more complete understanding of how the virus affects the human host and what strategies can improve outcomes. To date, investigations of potential therapeutics for COVID-19 have focused primarily on repurposing existing drugs. This approach is necessary given the urgency of the situation as well as the extensive time required for developing and testing new therapies. However, in the long-term, new drugs specific for treatment of COVID-19 may also enter development. Development of novel drugs is likely to be guided by what is known about the pathogenesis and molecular structure of SARS-CoV-2. For example, understanding the various structural components of SARS-CoV-2 may allow for the development of small molecule inhibitors of those components. Currently, crystal structures of the SARS-CoV-2 main protease have recently been resolved [340,530], and efforts are already in place to perform screens for small molecule inhibitors of the main protease, which have yielded potential hits [340]. Much work remains to be done to determine further crystal structures of other viral components, understand the relative utility of targeting different viral components, perform additional small molecule inhibitor screens, and determine the safety and efficacy of the potential inhibitors. While still nascent, work in this area is promising. Over the longer term, this approach and others may lead to the development of novel therapeutics specifically for COVID-19 and SARS-CoV-2.

9 Application of an Open Publishing Framework to an Emerging Public health Crisis

9.1 Abstract

9.2 Introduction

As international attention remains focused on the ongoing public health crisis, the scientific community has responded by mobilizing resources and turning much of its attention to the virus and disease. This rapid influx of information is disseminated by traditional publishing mechanisms, preprint servers, and press releases, which provide a venue for scientists to release findings without undergoing the formal publication process. While having information available is valuable to efforts to understand and combat COVID-19, many contributions come from researchers across a wide range of fields who have varying degrees of experience working on coronaviruses and related topics. The volume of information available, much of which has not gone through rigorous peer review, presents a significant challenge to individual efforts to keep abreast of the state of COVID-19 research [5]. However, research on these topics is proceeding so rapidly that any static review is likely to quickly become dated. Our goal as a community is to consolidate information about the virus in the context of related viruses and to synthesize rapidly emerging literature centered on the diagnosis and treatment of COVID-19. We used an open publishing framework, Manubot [6], to manage hundreds of contributions from the community to create a living, scholarly document. We designed software to generate figures that automatically update using external data sources. Our primary goal is to sort and distill informative content out of the overwhelming flood of information [5] and help the broader scientific community become more conversant on this critical subject. Thus, our approach has been to develop a real-time, collaborative effort that welcomes submissions from scientists worldwide into this ongoing effort. This document represents the first snapshot, which aims to reflect the state of the field as of October, 2020. We plan to refine and expand this document until technologies to mitigate the pandemic are widely available.

In an effort to keep pace as new information about COVID-19 and SARS-CoV-2 becomes available, this project is an open, collaborative effort that invited contributions from the scientific community broadly, similar to previous efforts to develop collaborative reviews [889,890]. Contributors were recruited by word of mouth and on Twitter. Existing efforts to train early-career scientists were also integrated: Appendix A contains summaries written by the students, post-docs, and faculty of the Immunology Institute at the Mount Sinai School of Medicine [891,892], and two of the authors were recruited through the American Physician Scientist Association’s Virtual Summer Research Program [893]. The project was managed through GitHub [894] using Manubot [6] to continuously generate a version of the manuscript online [895]. Contributors developed text that was proposed through GitHub’s pull request system and then reviewed and approved by at least one other author. While this document reflects the current version of record, the online version will continue to be developed as information about the pandemic emerges. Below, we will describe the processes used to synthesize the literature.

9.3 Technical Infrastructure

9.3.1 Collaborative Writing and Manuscript Generation

Manubot [6] is a collaborative framework developed to adapt open-source software development techniques and version control for manuscript writing. Here, Manubot was used to generate a manuscript from text maintained using GitHub, a popular, online version control interface. The GitHub implementation allowed users to contribute either using git on the command line or using the GitHub user interface, and we developed documentation for users with less experience with this platform. Manubot also provides a functionality to create a bibliography using digital object identifiers (DOIs), website URLs, or other identifiers such as PubMed identifiers and arXiv IDs. Due to the needs of this project, project contributors also implemented new features in Manubot and Zotero, which Manubot uses to extract metadata for some types of citations. These features support directly citing clinical trial identifiers such as clinicaltrials:NCT04292899 [315] and generating the complete review manuscript along with the individual manuscripts reviewing specific topics. Finally, Manubot and GitHub Actions continuous integration allowed for scripted updates to be run each time the manuscript was generated. These scripts were used to check that the manuscript was built correctly, run spellchecking, and cross-reference the manuscripts cited in this review, summarized in Appendix A, and discussed in the project’s issues and pull requests.

9.3.2 Data Analysis and Visualization

The combination of Manubot and GitHub Actions also made it possible to dynamically update information such as statistics and visualizations in the manuscript. Data about worldwide cases and deaths from the COVID-19 Data Repository by the Center for Systems Science and Engineering at Johns Hopkins University [290] were read using a Python script. Similarly, the clinical trials statistics and figure were generated based on data from the University of Oxford Evidence-Based Medicine Data Lab’s COVID-19 TrialsTracker [293]. In both cases, frequency data were plotted using Matplotlib [896] in Python. The figure showing the geographic distribution of COVID-19 clinical trials was generated using the countries associated with the trials listed in the COVID-19 TrialsTracker, converting the country names to 3-letter ISO codes using pycountry or manual adjustment when necessary, and visualizing the geographic distribution of trial recruitment using geopandas.

GitHub Actions runs a nightly workflow to update these external data and regenerate the statistics and figures for the manuscript. The workflow uses the GitHub API to detect and save the latest commit of the external data sources, which are both GitHub repositories. It then downloads versioned data from that snapshot of the external repositories and runs bash and Python scripts to calculate the desired statistics and produce the summary figures. The statistics are stored in JSON files that are accessed by Manubot to populate the values of placeholder template variables dynamically every time the manuscript is built. For instance, the template variable {{ebm_trials_results}} in the manuscript is replaced by the actual number of clinical trials with results, 168. The template variables also include versioned URLs to the dynamically updated figures. The JSON files and figures are stored in the external-resources branch of the manuscript’s GitHub repository, which acts as versioned storage. The GitHub Actions workflow automatically adds and commits the new JSON files and figures to the external-resources branch every time it runs, and Manubot uses the latest version of these resources when it builds the manuscript.

The workflow file is available from https://github.com/greenelab/covid19-review/blob/master/.github/workflows/update-external-resources.yaml and the scripts are available from https://github.com/greenelab/covid19-review/tree/external-resources. The Python package versions are available in https://github.com/greenelab/covid19-review/blob/external-resources/environment.yml.

9.4 Article Selection and Evaluation

Relevant articles were identified and submitted as issues on GitHub for review. Articles were classified as diagnostic, therapeutic, or other, and a template was developed to guide the review of papers and preprints in each category. Following a framework often used for assessing medical literature, the review consisted of examining methods used in each relevant article, assignment (whether the study was observational or randomized), assessment, results, interpretation, and how well the study extrapolates [897]. For examples of each template, please see Appendices B-D.

9.4.1 Diagnostic Papers

9.4.1.1 Methods

Reviewers began by describing the study question(s) being investigated by the article. They then described the study population, the sample size, the prevalence of the disease in the study population, countries / regions considered in case of human subjects, demographics of participants, the setting, and any remaining inclusion / exclusion criteria considered. They then described the reference test or “gold standard,” if one was utilized.

9.4.1.2 Assignment

Reviewers described how new and reference tests were assigned, including additional relevant details about the study design. For example, reviewers were asked whether the diagnostic test resulted in rigorous assignments of case status or was biased towards sicker or healthier individuals.

9.4.1.3 Assessment

Reviewers described how the test was performed. For example, for both standard and reference tests, reviewers described technical details of assays used, when measurements were taken and by whom. Subsequently, they described how individuals were classified as positive or negative cases and whether results were precise and reproducible with repeated tests. Reviewers described whether there were any missing data, whether some participants underwent only one test, or whether there were individuals with inconclusive results.

9.4.1.4 Results

Reviewers reported the estimated sensitivity, specificity, positive predictive value (PPV), and negative predicted value (NPV), as well as confidence bounds around these measures, if provided.

9.4.1.5 Interpretation

Reviewers reported how well the test ruled in or ruled out disease based on the population, if there were identified side effects, and patient adherence.

9.4.1.6 Extrapolation

Reviewers described how well this test will extrapolate outside the measured population.

9.4.2 Therapeutic Papers

9.4.2.1 Methods

Reviewers began by describing the study question(s) being investigated by the article. They then described the study population, the sample size, the prevalence of the disease in the study population, countries / regions considered in case of human subjects, demographics of participants, the setting, and any remaining inclusion / exclusion criteria considered.

9.4.2.2 Assignment

Reviewers described how the treatment is assigned, whether it was an interventional or observational study, whether randomization took place, etc.

9.4.2.3 Assessment

9.4.2.3.1 Outcome Assessment

Reviewers described the outcome that was assessed and evaluated whether it was appropriate given the underlying study question. They described whether there were any missing data such as whether there were individuals lost to follow up. They then describe whether there were any potential sources of bias such as lack of blinding in a randomized controlled trial.

9.4.2.3.2 Statistical Methods Assessment

Reviewers described which statistical methods were used for inference and whether applied methods were appropriate for the study. They then described whether adjustments were made for possible confounders.

9.4.2.4 Results

Reviewers described the estimated association between the treatment and outcome. They described measures of confidence or statistical significance, if provided.

9.4.2.5 Interpretation

Reviewers described whether a causal claim could be made. They described whether any side effects or interactions with other drugs were identified, as well as any subgroup findings.

9.4.2.6 Extrapolation

Reviewers describe how the study may extrapolate to a different species or population.

9.5 Conclusions

Figure 7: Summary of the relationships among topics covered in this review.

Several review articles on aspects of COVID-19 have already been published. These have included reviews on the disease epidemiology [38], immunological response [39], diagnostics [40], and pharmacological treatments [39,41]. Others [42,43] provide narrative reviews of progress on some important ongoing COVID-19 research questions. With the worldwide scientific community uniting during 2020 to investigate SARS-CoV-2 and COVID-19 from a wide range of perspectives, findings from many disciplines are relevant on a rapid timescale to a broad scientific audience. Additionally, many findings are published as preprints, which are available prior to going through the peer review process. As a result, centralizing, summarizing, and critiquing new literature broadly relevant to COVID-19 can help to expedite the interdisciplinary scientific process that is currently happening at an advanced pace. We are particularly interested in providing background to the development of diagnostic, prophylactic, and therapeutic approaches to COVID-19. Two major concerns within diagnosis include the detection of current infections in individuals with and without symptoms, and the detection of past exposure without an active infection. In the latter category, identifying whether individuals can develop or have developed sustained immunity is also a major consideration. The development of high-throughput, affordable methods for detecting active infections and sustained immunity will be critical to understanding and controlling the disease. The identification of interventions that can mitigate the effect of the virus on exposed and infected individuals is a significant research priority. Some possible approaches include the identification of existing pharmaceuticals that reduce the severity of infection, either by reducing the virus’ virulence (e.g., antivirals) or managing the most severe symptoms of infection. Due to the long timeline for the development of novel pharmaceuticals, in most cases, research surrounding possible pharmaceutical interventions focuses on the identification and investigation of existing compounds whose mechanisms may be relevant to COVID-19. Other foci of current research include the identification of antibodies produced by survivors of COVID-19 and the development of vaccines. Understanding the mechanisms describing host-virus interactions between humans and SARS-CoV-2 is thus critical to identifying candidate therapeutics. An overview of the topics covered is visualized in Figure 7. Thus, in this review, we seek to consolidate information about efforts to develop strategies for diagnosis and therapeutics as new information is released by the scientific community. We include information from both traditional peer-reviewed scientific literature and from preprints, which typically have not undergone peer review but have been critically evaluated by the scientists involved in this effort. The goal of this manuscript is to present preliminary findings within the broader context of COVID-19 research and to identify the broad interpretations of new research, as well as limitations to interpretability.

10 Additional Items

10.1 Competing Interests

Author Competing Interests Last Reviewed
Halie M. Rando None 2020-03-22
Casey S. Greene None 2020-01-20
Michael P. Robson None 2020-11-12
Simina M. Boca None 2020-11-07
Nils Wellhausen None 2020-11-03
Ronan Lordan None 2020-11-03
Christian Brueffer Employee and shareholder of SAGA Diagnostics AB. 2020-11-11
Sandipan Ray None 2020-11-11
Lucy D'Agostino McGowan Received consulting fees from Acelity and Sanofi in the past five years 2020-11-10
Anthony Gitter Filed a patent application with the Wisconsin Alumni Research Foundation related to classifying activated T cells 2020-11-10
Anna Ada Dattoli None 2020-03-26
Ryan Velazquez None 2020-11-10
John P. Barton None 2020-11-11
Jeffrey M. Field None 2020-11-12
Bharath Ramsundar None 2020-11-11
Adam L. MacLean None 2020-11-11
Alexandra J. Lee None 2020-11-09
Immunology Institute of the Icahn School of Medicine None 2020-04-07
Fengling Hu None 2020-04-08
Nafisa M. Jadavji None 2020-11-11
Elizabeth Sell None 2020-11-11
Jinhui Wang None 2020-04-13
Diane N. Rafizadeh None 2020-11-11
Ashwin N. Skelly None 2020-11-11
Marouen Ben Guebila None 2020-11-11
Likhitha Kolla None 2020-11-16
David Manheim None 2020-11-11
Soumita Ghosh None 2020-11-09
James Brian Byrd Funded by FastGrants to conduct a COVID-19-related clinical trial 2020-11-12
YoSon Park None 2020-11-11
Vikas Bansal None 2020-05-26
Stephen Capone None 2020-11-11
John J. Dziak None 2020-11-11
Yuchen Sun None 2020-11-11
Yanjun Qi None 2020-07-09
Lamonica Shinholster None 2020-11-11
Temitayo Lukan None 2020-11-10
Sergey Knyazev None 2020-11-11
Dimitri Perrin None 2020-11-11
Serghei Mangul None 2020-11-11
Shikta Das None 2020-08-13
Gregory L Szeto None 2020-11-16
Tiago Lubiana None 2020-11-11
David Mai None 2021-01-08
COVID-19 Review Consortium None 2021-01-16
Rishi Raj Goel None 2021-01-20

10.2 Author Contributions

Author Contributions
Halie M. Rando A, B, C, D, E
Casey S. Greene X, Y, Z
Michael P. Robson Software
Simina M. Boca MISSING
Nils Wellhausen MISSING
Ronan Lordan MISSING
Christian Brueffer MISSING
Sandipan Ray MISSING
Lucy D'Agostino McGowan MISSING
Anthony Gitter MISSING
Anna Ada Dattoli Writing - Original Draft
Ryan Velazquez MISSING
John P. Barton MISSING
Jeffrey M. Field Writing - Original Draft
Bharath Ramsundar Investigation, Writing - Review & Editing
Adam L. MacLean MISSING
Alexandra J. Lee MISSING
Immunology Institute of the Icahn School of Medicine Data Curation
Fengling Hu Writing - Original Draft, Writing - Review & Editing
Nafisa M. Jadavji Writing - Original Draft, Writing - Review & Editing
Elizabeth Sell MISSING
Jinhui Wang MISSING
Diane N. Rafizadeh Writing - Original Draft, Writing - Review & Editing
Ashwin N. Skelly MISSING
Marouen Ben Guebila MISSING
Likhitha Kolla Writing - Original Draft
David Manheim Investigation, Writing - Original Draft
Soumita Ghosh Writing - Original Draft
James Brian Byrd MISSING
YoSon Park MISSING
Vikas Bansal MISSING
Stephen Capone Writing - Original Draft, Writing - Review & Editing
John J. Dziak MISSING
Yuchen Sun Visualization
Yanjun Qi Visualization
Lamonica Shinholster MISSING
Temitayo Lukan Investigation, Writing - Original Draft
Sergey Knyazev MISSING
Dimitri Perrin Writing - Original Draft, Writing - Review & Editing
Serghei Mangul MISSING
Shikta Das Writing - Review & Editing
Gregory L Szeto MISSING
Tiago Lubiana MISSING
David Mai MISSING
COVID-19 Review Consortium Project Administration
Rishi Raj Goel Writing - Original Draft, Writing - Review & Editing

10.3 Acknowledgements

We thank Nick DeVito for assistance with the Evidence-Based Medicine Data Lab COVID-19 TrialsTracker data. We thank Yael Evelyn Marshall who contributed writing (original draft) as well as reviewing and editing of pieces of the text but who did not formally approve the manuscript, as well as Ronnie Russell, who contributed text to and helped develop the structure of the manuscript early in the writing process and Matthias Fax who helped with writing and editing text related to diagnostics. We are grateful to the following contributors for reviewing pieces of the text: Nadia Danilova, James Eberwine and Ipsita Krishnan.

11 References

1. Novel Coronavirus (2019-nCoV) SITUATION REPORT - 1
World Health Organization
(2020-01-21) https://www.who.int/docs/default-source/coronaviruse/situation-reports/20200121-sitrep-1-2019-ncov.pdf

2. Novel Coronavirus (2019-nCoV) Situation Report - 8
World Health Organization
(2020-01-28) https://www.who.int/docs/default-source/coronaviruse/situation-reports/20200128-sitrep-8-ncov-cleared.pdf

3. Novel Coronavirus (2019-nCoV) Situation Report - 51
World Health Organization
(2020-03-11) https://www.who.int/docs/default-source/coronaviruse/situation-reports/20200311-sitrep-51-covid-19.pdf

4. Novel Coronavirus (2019-nCoV) Situation Report - 75
World Health Organization
(2020-04-04) https://www.who.int/docs/default-source/coronaviruse/situation-reports/20200404-sitrep-75-covid-19.pdf

5. How you can help with COVID-19 modelling
Julia R. Gog
Nature Reviews Physics (2020-04-08) https://doi.org/ggrsq3
DOI: 10.1038/s42254-020-0175-7 · PMCID: PMC7144181

6. Open collaborative writing with Manubot
Daniel S. Himmelstein, Vincent Rubinetti, David R. Slochower, Dongbo Hu, Venkat S. Malladi, Casey S. Greene, Anthony Gitter
PLOS Computational Biology (2019-06-24) https://doi.org/c7np
DOI: 10.1371/journal.pcbi.1007128 · PMID: 31233491 · PMCID: PMC6611653

7. Cases, Data, and Surveillance
CDC
Centers for Disease Control and Prevention (2020-02-11) https://www.cdc.gov/coronavirus/2019-ncov/cases-updates/about-epidemiology/index.html

8. IHME | COVID-19 Projections
Institute for Health Metrics and Evaluation
https://covid19.healthdata.org/

9. Managing epidemics: key facts about major deadly diseases.
World Health Organization
(2018)
ISBN: 9789241565530

10. A Novel Coronavirus Genome Identified in a Cluster of Pneumonia Cases — Wuhan, China 2019−2020
Wenjie Tan, Xiang Zhao, Xuejun Ma, Wenling Wang, Peihua Niu, Wenbo Xu, George F. Gao, Guizhen Wu, MHC Key Laboratory of Biosafety, National Institute for Viral Disease Control and Prevention, China CDC, Beijing, China, Center for Biosafety Mega-Science, Chinese Academy of Sciences, Beijing, China
China CDC Weekly (2020) https://doi.org/gg8z47
DOI: 10.46234/ccdcw2020.017

11. A new coronavirus associated with human respiratory disease in China
Fan Wu, Su Zhao, Bin Yu, Yan-Mei Chen, Wen Wang, Zhi-Gang Song, Yi Hu, Zhao-Wu Tao, Jun-Hua Tian, Yuan-Yuan Pei, … Yong-Zhen Zhang
Nature (2020-02-03) https://doi.org/dk2w
DOI: 10.1038/s41586-020-2008-3 · PMID: 32015508 · PMCID: PMC7094943

12. A pneumonia outbreak associated with a new coronavirus of probable bat origin
Peng Zhou, Xing-Lou Yang, Xian-Guang Wang, Ben Hu, Lei Zhang, Wei Zhang, Hao-Rui Si, Yan Zhu, Bei Li, Chao-Lin Huang, … Zheng-Li Shi
Nature (2020-02-03) https://doi.org/ggj5cg
DOI: 10.1038/s41586-020-2012-7 · PMID: 32015507 · PMCID: PMC7095418

13. Genomic characterisation and epidemiology of 2019 novel coronavirus: implications for virus origins and receptor binding
Roujian Lu, Xiang Zhao, Juan Li, Peihua Niu, Bo Yang, Honglong Wu, Wenling Wang, Hao Song, Baoying Huang, Na Zhu, … Wenjie Tan
The Lancet (2020-02) https://doi.org/ggjr43
DOI: 10.1016/s0140-6736(20)30251-8 · PMID: 32007145 · PMCID: PMC7159086

14. On the origin and continuing evolution of SARS-CoV-2
Xiaolu Tang, Changcheng Wu, Xiang Li, Yuhe Song, Xinmin Yao, Xinkai Wu, Yuange Duan, Hong Zhang, Yirong Wang, Zhaohui Qian, … Jian Lu
National Science Review (2020-06) https://doi.org/ggndzn
DOI: 10.1093/nsr/nwaa036 · PMCID: PMC7107875

15. Pangolin homology associated with 2019-nCoV
Tao Zhang, Qunfu Wu, Zhigang Zhang
Cold Spring Harbor Laboratory (2020-02-20) https://doi.org/ggpvpt
DOI: 10.1101/2020.02.19.950253

16. Emergence of SARS-CoV-2 through recombination and strong purifying selection
Xiaojun Li, Elena E. Giorgi, Manukumar Honnayakanahalli Marichannegowda, Brian Foley, Chuan Xiao, Xiang-Peng Kong, Yue Chen, S. Gnanakaran, Bette Korber, Feng Gao
Science Advances (2020-07) https://doi.org/gg6r93
DOI: 10.1126/sciadv.abb9153 · PMID: 32937441 · PMCID: PMC7458444

17. The proximal origin of SARS-CoV-2
Kristian G. Andersen, Andrew Rambaut, W. Ian Lipkin, Edward C. Holmes, Robert F. Garry
Nature Medicine (2020-03-17) https://doi.org/ggn4dn
DOI: 10.1038/s41591-020-0820-9 · PMID: 32284615 · PMCID: PMC7095063

18. Coronavirus immunogens
Linda J. Saif
Veterinary Microbiology (1993-11) https://doi.org/ckfn8b
DOI: 10.1016/0378-1135(93)90030-b · PMID: 8116187 · PMCID: PMC7117163

19. Origin and evolution of pathogenic coronaviruses
Jie Cui, Fang Li, Zheng-Li Shi
Nature Reviews Microbiology (2018-12-10) https://doi.org/ggh4vb
DOI: 10.1038/s41579-018-0118-9 · PMID: 30531947 · PMCID: PMC7097006

20. Human Coronaviruses: A Review of Virus–Host Interactions
Yvonne Lim, Yan Ng, James Tam, Ding Liu
Diseases (2016-07-25) https://doi.org/ggjs23
DOI: 10.3390/diseases4030026 · PMID: 28933406 · PMCID: PMC5456285

21. A New Virus Isolated from the Human Respiratory Tract.
D. Hamre, J. J. Procknow
Experimental Biology and Medicine (1966-01-01) https://doi.org/gg84fc
DOI: 10.3181/00379727-121-30734 · PMID: 4285768

22. Recovery in tracheal organ cultures of novel viruses from patients with respiratory disease.
K. McIntosh, J. H. Dees, W. B. Becker, A. Z. Kapikian, R. M. Chanock
Proceedings of the National Academy of Sciences (1967-04-01) https://doi.org/bhfd6w
DOI: 10.1073/pnas.57.4.933 · PMID: 5231356 · PMCID: PMC224637

23.:(unav)
Krzysztof Pyrc, Maarten F Jebbink, Ben Berkhout, Lia van der Hoek
Virology Journal (2004) https://doi.org/dnmj8m
DOI: 10.1186/1743-422x-1-7 · PMID: 15548333 · PMCID: PMC538260

24. Understanding Human Coronavirus HCoV-NL63.
Sahar Abdul-Rasool, Burtram C Fielding
The open virology journal (2010-05-25) https://www.ncbi.nlm.nih.gov/pubmed/20700397
DOI: 10.2174/1874357901004010076 · PMID: 20700397 · PMCID: PMC2918871

25. Identification of a new human coronavirus
Lia van der Hoek, Krzysztof Pyrc, Maarten F Jebbink, Wilma Vermeulen-Oost, Ron JM Berkhout, Katja C Wolthers, Pauline ME Wertheim-van Dillen, Jos Kaandorp, Joke Spaargaren, Ben Berkhout
Nature Medicine (2004-03-21) https://doi.org/b5wtsn
DOI: 10.1038/nm1024 · PMID: 15034574 · PMCID: PMC7095789

26. Characterization and Complete Genome Sequence of a Novel Coronavirus, Coronavirus HKU1, from Patients with Pneumonia
Patrick C. Y. Woo, Susanna K. P. Lau, Chung-ming Chu, Kwok-hung Chan, Hoi-wah Tsoi, Yi Huang, Beatrice H. L. Wong, Rosana W. S. Poon, James J. Cai, Wei-kwang Luk, … Kwok-yung Yuen
Journal of Virology (2005-01-15) https://doi.org/bk7m7h
DOI: 10.1128/jvi.79.2.884-895.2005 · PMID: 15613317 · PMCID: PMC538593

27. Coronavirus 229E-Related Pneumonia in Immunocompromised Patients
F. Pene, A. Merlat, A. Vabret, F. Rozenberg, A. Buzyn, F. Dreyfus, A. Cariou, F. Freymuth, P. Lebon
Clinical Infectious Diseases (2003-10-01) https://doi.org/dcjk64
DOI: 10.1086/377612 · PMID: 13130404 · PMCID: PMC7107892

28. SARS and MERS: recent insights into emerging coronaviruses
Emmie de Wit, Neeltje van Doremalen, Darryl Falzarano, Vincent J. Munster
Nature Reviews Microbiology (2016-06-27) https://doi.org/f8v5cv
DOI: 10.1038/nrmicro.2016.81 · PMID: 27344959 · PMCID: PMC7097822

29. Hosts and Sources of Endemic Human Coronaviruses
Victor M. Corman, Doreen Muth, Daniela Niemeyer, Christian Drosten
Advances in Virus Research (2018) https://doi.org/ggwx4j
DOI: 10.1016/bs.aivir.2018.01.001 · PMID: 29551135 · PMCID: PMC7112090

30. Three Emerging Coronaviruses in Two Decades
Jeannette Guarner
American Journal of Clinical Pathology (2020-04) https://doi.org/ggppq3
DOI: 10.1093/ajcp/aqaa029 · PMID: 32053148 · PMCID: PMC7109697

31. From SARS to MERS, Thrusting Coronaviruses into the Spotlight
Zhiqi Song, Yanfeng Xu, Linlin Bao, Ling Zhang, Pin Yu, Yajin Qu, Hua Zhu, Wenjie Zhao, Yunlin Han, Chuan Qin
Viruses (2019-01-14) https://doi.org/ggqp7h
DOI: 10.3390/v11010059 · PMID: 30646565 · PMCID: PMC6357155

32. 1. Overview of the human immune response
D CHAPLIN
Journal of Allergy and Clinical Immunology (2006-02) https://doi.org/b6zghf
DOI: 10.1016/j.jaci.2005.09.034 · PMID: 16455341

33. Molecular immune pathogenesis and diagnosis of COVID-19
Xiaowei Li, Manman Geng, Yizhao Peng, Liesu Meng, Shemin Lu
Journal of Pharmaceutical Analysis (2020-04) https://doi.org/ggppqg
DOI: 10.1016/j.jpha.2020.03.001 · PMID: 32282863 · PMCID: PMC7104082

34. Clinical Characteristics of Coronavirus Disease 2019 in China
Wei-jie Guan, Zheng-yi Ni, Yu Hu, Wen-hua Liang, Chun-quan Ou, Jian-xing He, Lei Liu, Hong Shan, Chun-liang Lei, David S. C. Hui, … Nan-shan Zhong
New England Journal of Medicine (2020-04-30) https://doi.org/ggm6dh
DOI: 10.1056/nejmoa2002032 · PMID: 32109013 · PMCID: PMC7092819

35. Clinical course and risk factors for mortality of adult inpatients with COVID-19 in Wuhan, China: a retrospective cohort study
Fei Zhou, Ting Yu, Ronghui Du, Guohui Fan, Ying Liu, Zhibo Liu, Jie Xiang, Yeming Wang, Bin Song, Xiaoying Gu, … Bin Cao
The Lancet (2020-03) https://doi.org/ggnxb3
DOI: 10.1016/s0140-6736(20)30566-3 · PMID: 32171076 · PMCID: PMC7270627

36. Presenting Characteristics, Comorbidities, and Outcomes Among 5700 Patients Hospitalized With COVID-19 in the New York City Area
Safiya Richardson, Jamie S. Hirsch, Mangala Narasimhan, James M. Crawford, Thomas McGinn, Karina W. Davidson, Douglas P. Barnaby, Lance B. Becker, John D. Chelico, Stuart L. Cohen, … and the Northwell COVID-19 Research Consortium
JAMA (2020-05-26) https://doi.org/ggsrkd
DOI: 10.1001/jama.2020.6775 · PMID: 32320003 · PMCID: PMC7177629

37. Clinical Characteristics of Covid-19 in New York City
Parag Goyal, Justin J. Choi, Laura C. Pinheiro, Edward J. Schenck, Ruijun Chen, Assem Jabri, Michael J. Satlin, Thomas R. Campion, Musarrat Nahid, Joanna B. Ringel, … Monika M. Safford
New England Journal of Medicine (2020-06-11) https://doi.org/ggtsjc
DOI: 10.1056/nejmc2010419 · PMID: 32302078 · PMCID: PMC7182018

38. COVID-19: Epidemiology, Evolution, and Cross-Disciplinary Perspectives
Jiumeng Sun, Wan-Ting He, Lifang Wang, Alexander Lai, Xiang Ji, Xiaofeng Zhai, Gairu Li, Marc A. Suchard, Jin Tian, Jiyong Zhou, … Shuo Su
Trends in Molecular Medicine (2020-05) https://doi.org/ggqgwq
DOI: 10.1016/j.molmed.2020.02.008 · PMID: 32359479 · PMCID: PMC7118693

39. Immunology of COVID-19: Current State of the Science
Nicolas Vabret, Graham J. Britton, Conor Gruber, Samarth Hegde, Joel Kim, Maria Kuksin, Rachel Levantovsky, Louise Malle, Alvaro Moreira, Matthew D. Park, … Uri Laserson
Immunity (2020-06) https://doi.org/ggt54g
DOI: 10.1016/j.immuni.2020.05.002 · PMID: 32505227 · PMCID: PMC7200337

40. COVID-19 diagnostics in context
Ralph Weissleder, Hakho Lee, Jina Ko, Mikael J. Pittet
Science Translational Medicine (2020-06-03) https://doi.org/gg339m
DOI: 10.1126/scitranslmed.abc1931 · PMID: 32493791

41. Pharmacologic Treatments for Coronavirus Disease 2019 (COVID-19)
James M. Sanders, Marguerite L. Monogue, Tomasz Z. Jodlowski, James B. Cutrell
JAMA (2020-04-13) https://doi.org/ggr27x
DOI: 10.1001/jama.2020.6019 · PMID: 32282022

42. COVID-19 Research in Brief: December, 2019 to June, 2020
Thiago Carvalho
Nature Medicine (2020-06-26) https://doi.org/gg3kd2
DOI: 10.1038/d41591-020-00026-w · PMID: 32778824

43. Pathophysiology, Transmission, Diagnosis, and Treatment of Coronavirus Disease 2019 (COVID-19)
W. Joost Wiersinga, Andrew Rhodes, Allen C. Cheng, Sharon J. Peacock, Hallie C. Prescott
JAMA (2020-08-25) https://doi.org/gg4ht4
DOI: 10.1001/jama.2020.12839 · PMID: 32648899

44. Coronavirus Disease 2019 (COVID-19) – Symptoms
CDC
Centers for Disease Control and Prevention (2020-12-22) https://www.cdc.gov/coronavirus/2019-ncov/symptoms-testing/symptoms.html

45. WHO Declares COVID-19 a Pandemic
Domenico Cucinotta, Maurizio Vanelli
Acta Bio Medica Atenei Parmensis (2020-03-19) https://doi.org/ggq86h
DOI: 10.23750/abm.v91i1.9397 · PMID: 32191675 · PMCID: PMC7569573

46. Acute lung injury in patients with COVID‐19 infection
Liyang Li, Qihong Huang, Diane C. Wang, David H. Ingbar, Xiangdong Wang
Clinical and Translational Medicine (2020-03-31) https://doi.org/ghqcrz
DOI: 10.1002/ctm2.16 · PMID: 32508022 · PMCID: PMC7240840

47. Structure, Function, and Evolution of Coronavirus Spike Proteins
Fang Li
Annual Review of Virology (2016-09-29) https://doi.org/ggr7gv
DOI: 10.1146/annurev-virology-110615-042301 · PMID: 27578435 · PMCID: PMC5457962

48. A familial cluster of pneumonia associated with the 2019 novel coronavirus indicating person-to-person transmission: a study of a family cluster
Jasper Fuk-Woo Chan, Shuofeng Yuan, Kin-Hang Kok, Kelvin Kai-Wang To, Hin Chu, Jin Yang, Fanfan Xing, Jieling Liu, Cyril Chik-Yan Yip, Rosana Wing-Shan Poon, … Kwok-Yung Yuen
The Lancet (2020-02) https://doi.org/ggjs7j
DOI: 10.1016/s0140-6736(20)30154-9 · PMID: 31986261 · PMCID: PMC7159286

49. Fields virology
Bernard N. Fields, David M. Knipe, Peter M. Howley (editors)
Wolters Kluwer Health/Lippincott Williams & Wilkins (2007)
ISBN: 9780781760607

50. Important Role for the Transmembrane Domain of Severe Acute Respiratory Syndrome Coronavirus Spike Protein during Entry
Rene Broer, Bertrand Boson, Willy Spaan, François-Loïc Cosset, Jeroen Corver
Journal of Virology (2006-02-01) https://doi.org/dvvg2h
DOI: 10.1128/jvi.80.3.1302-1310.2006 · PMID: 16415007 · PMCID: PMC1346921

51. Cryo-EM structure of the 2019-nCoV spike in the prefusion conformation
Daniel Wrapp, Nianshuang Wang, Kizzmekia S. Corbett, Jory A. Goldsmith, Ching-Lin Hsieh, Olubukola Abiona, Barney S. Graham, Jason S. McLellan
Science (2020-03-13) https://doi.org/ggmtk2
DOI: 10.1126/science.abb2507 · PMID: 32075877 · PMCID: PMC7164637

52. Medical microbiology
Samuel Baron (editor)
University of Texas Medical Branch at Galveston (1996)
ISBN: 9780963117212

53. Coronaviruses: An Overview of Their Replication and Pathogenesis
Anthony R. Fehr, Stanley Perlman
Methods in Molecular Biology (2015) https://doi.org/ggpc6n
DOI: 10.1007/978-1-4939-2438-7_1 · PMID: 25720466 · PMCID: PMC4369385

54. Emerging coronaviruses: Genome structure, replication, and pathogenesis
Yu Chen, Qianyun Liu, Deyin Guo
Journal of Medical Virology (2020-02-07) https://doi.org/ggjvwj
DOI: 10.1002/jmv.25681 · PMID: 31967327 · PMCID: PMC7167049

55. SARS coronavirus entry into host cells through a novel clathrin- and caveolae-independent endocytic pathway
Hongliang Wang, Peng Yang, Kangtai Liu, Feng Guo, Yanli Zhang, Gongyi Zhang, Chengyu Jiang
Cell Research (2008-01-29) https://doi.org/bp9275
DOI: 10.1038/cr.2008.15 · PMID: 18227861 · PMCID: PMC7091891

56. Virus Entry by Endocytosis
Jason Mercer, Mario Schelhaas, Ari Helenius
Annual Review of Biochemistry (2010-06-07) https://doi.org/cw4dnb
DOI: 10.1146/annurev-biochem-060208-104626 · PMID: 20196649

57. Mechanisms of Coronavirus Cell Entry Mediated by the Viral Spike Protein
Sandrine Belouzard, Jean K. Millet, Beth N. Licitra, Gary R. Whittaker
Viruses (2012-06-20) https://doi.org/gbbktb
DOI: 10.3390/v4061011 · PMID: 22816037 · PMCID: PMC3397359

58. Phylogenetic Analysis and Structural Modeling of SARS-CoV-2 Spike Protein Reveals an Evolutionary Distinct and Proteolytically Sensitive Activation Loop
Javier A. Jaimes, Nicole M. André, Joshua S. Chappie, Jean K. Millet, Gary R. Whittaker
Journal of Molecular Biology (2020-05) https://doi.org/ggtxhr
DOI: 10.1016/j.jmb.2020.04.009 · PMID: 32320687 · PMCID: PMC7166309

59. Angiotensin-converting enzyme 2 (ACE2) as a SARS-CoV-2 receptor: molecular mechanisms and potential therapeutic target
Haibo Zhang, Josef M. Penninger, Yimin Li, Nanshan Zhong, Arthur S. Slutsky
Intensive Care Medicine (2020-03-03) https://doi.org/ggpx6p
DOI: 10.1007/s00134-020-05985-9 · PMID: 32125455 · PMCID: PMC7079879

60. Infection of Human Airway Epithelia by Sars Coronavirus is Associated with ACE2 Expression and Localization
Hong Peng Jia, Dwight C. Look, Melissa Hickey, Lei Shi, Lecia Pewe, Jason Netland, Michael Farzan, Christine Wohlford-Lenane, Stanley Perlman, Paul B. McCray
Advances in Experimental Medicine and Biology (2006) https://doi.org/dhh5tp
DOI: 10.1007/978-0-387-33012-9_85 · PMID: 17037581 · PMCID: PMC7123641

61. Receptor Recognition Mechanisms of Coronaviruses: a Decade of Structural Studies
Fang Li
Journal of Virology (2015-02-15) https://doi.org/f633jb
DOI: 10.1128/jvi.02615-14 · PMID: 25428871 · PMCID: PMC4338876

62. The spike protein of SARS-CoV — a target for vaccine and therapeutic development
Lanying Du, Yuxian He, Yusen Zhou, Shuwen Liu, Bo-Jian Zheng, Shibo Jiang
Nature Reviews Microbiology (2009-02-09) https://doi.org/d4tq4t
DOI: 10.1038/nrmicro2090 · PMID: 19198616 · PMCID: PMC2750777

63. Structure, Function, and Antigenicity of the SARS-CoV-2 Spike Glycoprotein
Alexandra C. Walls, Young-Jun Park, M. Alejandra Tortorici, Abigail Wall, Andrew T. McGuire, David Veesler
Cell (2020-04) https://doi.org/dpvh
DOI: 10.1016/j.cell.2020.02.058 · PMID: 32155444 · PMCID: PMC7102599

64. Molecular Interactions in the Assembly of Coronaviruses
Cornelis A. M. de Haan, Peter J. M. Rottier
Advances in Virus Research (2005) https://doi.org/cf8chz
DOI: 10.1016/s0065-3527(05)64006-7 · PMID: 16139595 · PMCID: PMC7112327

65. Coronavirus membrane fusion mechanism offers a potential target for antiviral development
Tiffany Tang, Miya Bidon, Javier A. Jaimes, Gary R. Whittaker, Susan Daniel
Antiviral Research (2020-06) https://doi.org/ggr23b
DOI: 10.1016/j.antiviral.2020.104792 · PMID: 32272173 · PMCID: PMC7194977

66. SARS-CoV-2 Cell Entry Depends on ACE2 and TMPRSS2 and Is Blocked by a Clinically Proven Protease Inhibitor
Markus Hoffmann, Hannah Kleine-Weber, Simon Schroeder, Nadine Krüger, Tanja Herrler, Sandra Erichsen, Tobias S. Schiergens, Georg Herrler, Nai-Huei Wu, Andreas Nitsche, … Stefan Pöhlmann
Cell (2020-04) https://doi.org/ggnq74
DOI: 10.1016/j.cell.2020.02.052 · PMID: 32142651 · PMCID: PMC7102627

67. Characterization of spike glycoprotein of SARS-CoV-2 on virus entry and its immune cross-reactivity with SARS-CoV
Xiuyuan Ou, Yan Liu, Xiaobo Lei, Pei Li, Dan Mi, Lili Ren, Li Guo, Ruixuan Guo, Ting Chen, Jiaxin Hu, … Zhaohui Qian
Nature Communications (2020-03-27) https://doi.org/ggqsrf
DOI: 10.1038/s41467-020-15562-9 · PMID: 32221306 · PMCID: PMC7100515

68. Coronaviridae ~ ViralZone page https://viralzone.expasy.org/30?outline

69. Spatiotemporal interplay of severe acute respiratory syndrome coronavirus and respiratory mucosal cells drives viral dissemination in rhesus macaques
L Liu, Q Wei, K Nishiura, J Peng, H Wang, C Midkiff, X Alvarez, C Qin, A Lackner, Z Chen
Mucosal Immunology (2015-12-09) https://doi.org/f8r7dk
DOI: 10.1038/mi.2015.127 · PMID: 26647718 · PMCID: PMC4900951

70. Understanding Viral dsRNA-Mediated Innate Immune Responses at the Cellular Level Using a Rainbow Trout Model
Sarah J. Poynter, Stephanie J. DeWitte-Orr
Frontiers in Immunology (2018-04-23) https://doi.org/gdhpbs
DOI: 10.3389/fimmu.2018.00829 · PMID: 29740439 · PMCID: PMC5924774

71. Ultrastructure and Origin of Membrane Vesicles Associated with the Severe Acute Respiratory Syndrome Coronavirus Replication Complex
Eric J. Snijder, Yvonne van der Meer, Jessika Zevenhoven-Dobbe, Jos J. M. Onderwater, Jannes van der Meulen, Henk K. Koerten, A. Mieke Mommaas
Journal of Virology (2006-06-15) https://doi.org/b2rh4r
DOI: 10.1128/jvi.02501-05 · PMID: 16731931 · PMCID: PMC1472606

72. Structural basis for translational shutdown and immune evasion by the Nsp1 protein of SARS-CoV-2
Matthias Thoms, Robert Buschauer, Michael Ameismeier, Lennart Koepke, Timo Denk, Maximilian Hirschenberger, Hanna Kratzat, Manuel Hayn, Timur Mackens-Kiani, Jingdong Cheng, … Roland Beckmann
Science (2020-09-04) https://doi.org/gg69nq
DOI: 10.1126/science.abc8665 · PMID: 32680882 · PMCID: PMC7402621

73. Immune responses in COVID-19 and potential vaccines: Lessons learned from SARS and MERS epidemic
Asian Pacific Journal of Allergy and Immunology
(2020) https://doi.org/ggpvxw
DOI: 10.12932/ap-200220-0772 · PMID: 32105090

74. Angiotensin-converting enzyme 2 is a functional receptor for the SARS coronavirus
Wenhui Li, Michael J. Moore, Natalya Vasilieva, Jianhua Sui, Swee Kee Wong, Michael A. Berne, Mohan Somasundaran, John L. Sullivan, Katherine Luzuriaga, Thomas C. Greenough, … Michael Farzan
Nature (2003-11) https://doi.org/bqvpjh
DOI: 10.1038/nature02145 · PMID: 14647384 · PMCID: PMC7095016

75. Efficient Activation of the Severe Acute Respiratory Syndrome Coronavirus Spike Protein by the Transmembrane Protease TMPRSS2
Shutoku Matsuyama, Noriyo Nagata, Kazuya Shirato, Miyuki Kawase, Makoto Takeda, Fumihiro Taguchi
Journal of Virology (2010-12-15) https://doi.org/d4hnfr
DOI: 10.1128/jvi.01542-10 · PMID: 20926566 · PMCID: PMC3004351

76. Evidence that TMPRSS2 Activates the Severe Acute Respiratory Syndrome Coronavirus Spike Protein for Membrane Fusion and Reduces Viral Control by the Humoral Immune Response
I. Glowacka, S. Bertram, M. A. Muller, P. Allen, E. Soilleux, S. Pfefferle, I. Steffen, T. S. Tsegaye, Y. He, K. Gnirss, … S. Pohlmann
Journal of Virology (2011-02-16) https://doi.org/bg97wb
DOI: 10.1128/jvi.02232-10 · PMID: 21325420 · PMCID: PMC3126222

77. Increasing Host Cellular Receptor—Angiotensin-Converting Enzyme 2 (ACE2) Expression by Coronavirus may Facilitate 2019-nCoV Infection
Pei-Hui Wang, Yun Cheng
Cold Spring Harbor Laboratory (2020-02-27) https://doi.org/ggscwd
DOI: 10.1101/2020.02.24.963348

78. Detection of SARS-CoV-2 in Different Types of Clinical Specimens
Wenling Wang, Yanli Xu, Ruqin Gao, Roujian Lu, Kai Han, Guizhen Wu, Wenjie Tan
JAMA (2020-03-11) https://doi.org/ggpp6h
DOI: 10.1001/jama.2020.3786 · PMID: 32159775 · PMCID: PMC7066521

79. Epidemiologic Features and Clinical Course of Patients Infected With SARS-CoV-2 in Singapore
Barnaby Edward Young, Sean Wei Xiang Ong, Shirin Kalimuddin, Jenny G. Low, Seow Yen Tan, Jiashen Loh, Oon-Tek Ng, Kalisvar Marimuthu, Li Wei Ang, Tze Minn Mak, … for the Singapore 2019 Novel Coronavirus Outbreak Research Team
JAMA (2020-04-21) https://doi.org/ggnb37
DOI: 10.1001/jama.2020.3204 · PMID: 32125362 · PMCID: PMC7054855

80. Persistence and clearance of viral RNA in 2019 novel coronavirus disease rehabilitation patients
Yun Ling, Shui-Bao Xu, Yi-Xiao Lin, Di Tian, Zhao-Qin Zhu, Fa-Hui Dai, Fan Wu, Zhi-Gang Song, Wei Huang, Jun Chen, … Hong-Zhou Lu
Chinese Medical Journal (2020-05) https://doi.org/ggnnz8
DOI: 10.1097/cm9.0000000000000774 · PMID: 32118639 · PMCID: PMC7147278

81. Postmortem examination of COVID‐19 patients reveals diffuse alveolar damage with severe capillary congestion and variegated findings in lungs and other organs suggesting vascular dysfunction
Thomas Menter, Jasmin D Haslbauer, Ronny Nienhold, Spasenija Savic, Helmut Hopfer, Nikolaus Deigendesch, Stephan Frank, Daniel Turek, Niels Willi, Hans Pargger, … Alexandar Tzankov
Histopathology (2020-07-05) https://doi.org/ggwr32
DOI: 10.1111/his.14134 · PMID: 32364264 · PMCID: PMC7496150

82. Association of Cardiac Injury With Mortality in Hospitalized Patients With COVID-19 in Wuhan, China
Shaobo Shi, Mu Qin, Bo Shen, Yuli Cai, Tao Liu, Fan Yang, Wei Gong, Xu Liu, Jinjun Liang, Qinyan Zhao, … Congxin Huang
JAMA Cardiology (2020-07-01) https://doi.org/ggq8qf
DOI: 10.1001/jamacardio.2020.0950 · PMID: 32211816 · PMCID: PMC7097841

83. The need for urogenital tract monitoring in COVID-19
Shangqian Wang, Xiang Zhou, Tongtong Zhang, Zengjun Wang
Nature Reviews Urology (2020-04-20) https://doi.org/ggv4xb
DOI: 10.1038/s41585-020-0319-7 · PMID: 32313110 · PMCID: PMC7186932

84. Acute kidney injury in SARS-CoV-2 infected patients
Vito Fanelli, Marco Fiorentino, Vincenzo Cantaluppi, Loreto Gesualdo, Giovanni Stallone, Claudio Ronco, Giuseppe Castellano
Critical Care (2020-04-16) https://doi.org/ggv45f
DOI: 10.1186/s13054-020-02872-z · PMID: 32299479 · PMCID: PMC7161433

85. Liver injury in COVID-19: management and challenges
Chao Zhang, Lei Shi, Fu-Sheng Wang
The Lancet Gastroenterology & Hepatology (2020-05) https://doi.org/ggpx6s
DOI: 10.1016/s2468-1253(20)30057-1 · PMID: 32145190 · PMCID: PMC7129165

86. Evidence for Gastrointestinal Infection of SARS-CoV-2
Fei Xiao, Meiwen Tang, Xiaobin Zheng, Ye Liu, Xiaofeng Li, Hong Shan
Gastroenterology (2020-05) https://doi.org/ggpx27
DOI: 10.1053/j.gastro.2020.02.055 · PMID: 32142773 · PMCID: PMC7130181

87. 2019 Novel coronavirus infection and gastrointestinal tract
Qin Yan Gao, Ying Xuan Chen, Jing Yuan Fang
Journal of Digestive Diseases (2020-03) https://doi.org/ggqr86
DOI: 10.1111/1751-2980.12851 · PMID: 32096611 · PMCID: PMC7162053

88. Symptom Profiles of a Convenience Sample of Patients with COVID-19 — United States, January–April 2020
Rachel M. Burke, Marie E. Killerby, Suzanne Newton, Candace E. Ashworth, Abby L. Berns, Skyler Brennan, Jonathan M. Bressler, Erica Bye, Richard Crawford, Laurel Harduar Morano, … Case Investigation Form Working Group
MMWR. Morbidity and Mortality Weekly Report (2020-07-17) https://doi.org/gg8r2m
DOI: 10.15585/mmwr.mm6928a2 · PMID: 32673296 · PMCID: PMC7366851

89. Population-scale Longitudinal Mapping of COVID-19 Symptoms, Behavior, and Testing Identifies Contributors to Continued Disease Spread in the United States
William E. Allen, Han Altae-Tran, James Briggs, Xin Jin, Glen McGee, Andy Shi, Rumya Raghavan, Mireille Kamariza, Nicole Nova, Albert Pereta, … Xihong Lin
Cold Spring Harbor Laboratory (2020-06-11) https://doi.org/gg8skn
DOI: 10.1101/2020.06.09.20126813 · PMID: 32577674 · PMCID: PMC7302230

90. Extrapulmonary manifestations of COVID-19
Aakriti Gupta, Mahesh V. Madhavan, Kartik Sehgal, Nandini Nair, Shiwani Mahajan, Tejasav S. Sehrawat, Behnood Bikdeli, Neha Ahluwalia, John C. Ausiello, Elaine Y. Wan, … Donald W. Landry
Nature Medicine (2020-07-10) https://doi.org/gg4r37
DOI: 10.1038/s41591-020-0968-3 · PMID: 32651579

91. Acute kidney injury in patients hospitalized with COVID-19
Jamie S. Hirsch, Jia H. Ng, Daniel W. Ross, Purva Sharma, Hitesh H. Shah, Richard L. Barnett, Azzour D. Hazzan, Steven Fishbane, Kenar D. Jhaveri, Mersema Abate, … Jia Hwei. Ng
Kidney International (2020-07) https://doi.org/ggx24k
DOI: 10.1016/j.kint.2020.05.006 · PMID: 32416116 · PMCID: PMC7229463

92. Nervous system involvement after infection with COVID-19 and other coronaviruses
Yeshun Wu, Xiaolin Xu, Zijun Chen, Jiahao Duan, Kenji Hashimoto, Ling Yang, Cunming Liu, Chun Yang
Brain, Behavior, and Immunity (2020-07) https://doi.org/ggq7s2
DOI: 10.1016/j.bbi.2020.03.031 · PMID: 32240762 · PMCID: PMC7146689

93. Neurological associations of COVID-19
Mark A Ellul, Laura Benjamin, Bhagteshwar Singh, Suzannah Lant, Benedict Daniel Michael, Ava Easton, Rachel Kneen, Sylviane Defres, Jim Sejvar, Tom Solomon
The Lancet Neurology (2020-09) https://doi.org/d259
DOI: 10.1016/s1474-4422(20)30221-0 · PMID: 32622375 · PMCID: PMC7332267

94. Update on the neurology of COVID‐19
Josef Finsterer, Claudia Stollberger
Journal of Medical Virology (2020-06-02) https://doi.org/gg2qnn
DOI: 10.1002/jmv.26000 · PMID: 32401352 · PMCID: PMC7272942

95. ‐19: A Global Threat to the Nervous System
Igor J. Koralnik, Kenneth L. Tyler
Annals of Neurology (2020-06-23) https://doi.org/gg3hzh
DOI: 10.1002/ana.25807 · PMID: 32506549 · PMCID: PMC7300753

96. Large-Vessel Stroke as a Presenting Feature of Covid-19 in the Young
Thomas J. Oxley, J. Mocco, Shahram Majidi, Christopher P. Kellner, Hazem Shoirah, I. Paul Singh, Reade A. De Leacy, Tomoyoshi Shigematsu, Travis R. Ladner, Kurt A. Yaeger, … Johanna T. Fifi
New England Journal of Medicine (2020-05-14) https://doi.org/ggtsjg
DOI: 10.1056/nejmc2009787 · PMID: 32343504 · PMCID: PMC7207073

97. Incidence of thrombotic complications in critically ill ICU patients with COVID-19
F. A. Klok, M. J. H. A. Kruip, N. J. M. van der Meer, M. S. Arbous, D. A. M. P. J. Gommers, K. M. Kant, F. H. J. Kaptein, J. van Paassen, M. A. M. Stals, M. V. Huisman, H. Endeman
Thrombosis Research (2020-07) https://doi.org/dt2q
DOI: 10.1016/j.thromres.2020.04.013 · PMID: 32291094 · PMCID: PMC7146714

98. Coagulopathy and Antiphospholipid Antibodies in Patients with Covid-19
Yan Zhang, Meng Xiao, Shulan Zhang, Peng Xia, Wei Cao, Wei Jiang, Huan Chen, Xin Ding, Hua Zhao, Hongmin Zhang, … Shuyang Zhang
New England Journal of Medicine (2020-04-23) https://doi.org/ggrgz7
DOI: 10.1056/nejmc2007575 · PMID: 32268022 · PMCID: PMC7161262

99. Abnormal coagulation parameters are associated with poor prognosis in patients with novel coronavirus pneumonia
Ning Tang, Dengju Li, Xiong Wang, Ziyong Sun
Journal of Thrombosis and Haemostasis (2020-04) https://doi.org/ggqxf6
DOI: 10.1111/jth.14768 · PMID: 32073213 · PMCID: PMC7166509

100. Review: Viral infections and mechanisms of thrombosis and bleeding
M. Goeijenbier, M. van Wissen, C. van de Weg, E. Jong, V. E. A. Gerdes, J. C. M. Meijers, D. P. M. Brandjes, E. C. M. van Gorp
Journal of Medical Virology (2012-10) https://doi.org/f37tfr
DOI: 10.1002/jmv.23354 · PMID: 22930518 · PMCID: PMC7166625

101. Immune mechanisms of pulmonary intravascular coagulopathy in COVID-19 pneumonia
Dennis McGonagle, James S O’Donnell, Kassem Sharif, Paul Emery, Charles Bridgewood
The Lancet Rheumatology (2020-07) https://doi.org/ggvd74
DOI: 10.1016/s2665-9913(20)30121-1 · PMID: 32835247 · PMCID: PMC7252093

102. “War to the knife” against thromboinflammation to protect endothelial function of COVID-19 patients
Gabriele Guglielmetti, Marco Quaglia, Pier Paolo Sainaghi, Luigi Mario Castello, Rosanna Vaschetto, Mario Pirisi, Francesco Della Corte, Gian Carlo Avanzi, Piero Stratta, Vincenzo Cantaluppi
Critical Care (2020-06-19) https://doi.org/gg35w7
DOI: 10.1186/s13054-020-03060-9 · PMID: 32560665 · PMCID: PMC7303575

103. COVID-19 update: Covid-19-associated coagulopathy
Richard C. Becker
Journal of Thrombosis and Thrombolysis (2020-05-15) https://doi.org/ggwpp5
DOI: 10.1007/s11239-020-02134-3 · PMID: 32415579 · PMCID: PMC7225095

104. The complement system in COVID-19: friend and foe?
Anuja Java, Anthony J. Apicelli, M. Kathryn Liszewski, Ariella Coler-Reilly, John P. Atkinson, Alfred H. J. Kim, Hrishikesh S. Kulkarni
JCI Insight (2020-08-06) https://doi.org/gg4b5b
DOI: 10.1172/jci.insight.140711 · PMID: 32554923 · PMCID: PMC7455060

105. COVID-19, microangiopathy, hemostatic activation, and complement
Wen-Chao Song, Garret A. FitzGerald
Journal of Clinical Investigation (2020-06-22) https://doi.org/gg4b5c
DOI: 10.1172/jci140183 · PMID: 32459663 · PMCID: PMC7410042

106. Molecular biology of the cell
Bruce Alberts (editor)
Garland Science (2002)
ISBN: 9780815332183

107. Vander’s human physiology: the mechanisms of body function
Eric P Widmaier, Hershel Raff, Kevin T Strang
McGraw-Hill Higher Education (2008)
ISBN: 9780071283663

108. The Innate Immune System: Fighting on the Front Lines or Fanning the Flames of COVID-19?
Julia L. McKechnie, Catherine A. Blish
Cell Host & Microbe (2020-06) https://doi.org/gg28pq
DOI: 10.1016/j.chom.2020.05.009 · PMID: 32464098 · PMCID: PMC7237895

109. Inflammatory responses and inflammation-associated diseases in organs
Linlin Chen, Huidan Deng, Hengmin Cui, Jing Fang, Zhicai Zuo, Junliang Deng, Yinglun Li, Xun Wang, Ling Zhao
Oncotarget (2017-12-14) https://doi.org/ggps2p
DOI: 10.18632/oncotarget.23208 · PMID: 29467962 · PMCID: PMC5805548

110. Into the Eye of the Cytokine Storm
J. R. Tisoncik, M. J. Korth, C. P. Simmons, J. Farrar, T. R. Martin, M. G. Katze
Microbiology and Molecular Biology Reviews (2012-03-05) https://doi.org/f4n9h2
DOI: 10.1128/mmbr.05015-11 · PMID: 22390970 · PMCID: PMC3294426

111. Cytokines, Inflammation, and Pain
Jun-Ming Zhang, Jianxiong An
International Anesthesiology Clinics (2007-Spring) https://doi.org/ft5b7h
DOI: 10.1097/aia.0b013e318034194e · PMID: 17426506 · PMCID: PMC2785020

112. Dynamic balance of pro- and anti-inflammatory signals controls disease and limits pathology
Joseph M. Cicchese, Stephanie Evans, Caitlin Hult, Louis R. Joslyn, Timothy Wessler, Jess A. Millar, Simeone Marino, Nicholas A. Cilfone, Joshua T. Mattila, Jennifer J. Linderman, Denise E. Kirschner
Immunological Reviews (2018-09) https://doi.org/gd4g4p
DOI: 10.1111/imr.12671 · PMID: 30129209 · PMCID: PMC6292442

113. Cytokine Dysregulation, Inflammation and Well-Being
Ilia J. Elenkov, Domenic G. Iezzoni, Adrian Daly, Alan G. Harris, George P. Chrousos
Neuroimmunomodulation (2005) https://doi.org/bsn7kn
DOI: 10.1159/000087104 · PMID: 16166805

114. Chronic inflammation in the etiology of disease across the life span
David Furman, Judith Campisi, Eric Verdin, Pedro Carrera-Bastos, Sasha Targ, Claudio Franceschi, Luigi Ferrucci, Derek W. Gilroy, Alessio Fasano, Gary W. Miller, … George M. Slavich
Nature Medicine (2019-12-05) https://doi.org/gghdcf
DOI: 10.1038/s41591-019-0675-0 · PMID: 31806905 · PMCID: PMC7147972

115. Inpatient care for septicemia or sepsis: a challenge for patients and hospitals.
Margaret Jean Hall, Sonja N Williams, Carol J DeFrances, Aleksandr Golosinskiy
NCHS data brief (2011-06) https://www.ncbi.nlm.nih.gov/pubmed/22142805
PMID: 22142805

116. Cytokine Balance in the Lungs of Patients with Acute Respiratory Distress Syndrome
WILLIAM Y. PARK, RICHARD B. GOODMAN, KENNETH P. STEINBERG, JOHN T. RUZINSKI, FRANK RADELLA, DAVID R. PARK, JEROME PUGIN, SHAWN J. SKERRETT, LEONARD D. HUDSON, THOMAS R. MARTIN
American Journal of Respiratory and Critical Care Medicine (2001-11-15) https://doi.org/ggqfq7
DOI: 10.1164/ajrccm.164.10.2104013 · PMID: 11734443

117. Cytokine release syndrome
Alexander Shimabukuro-Vornhagen, Philipp Gödel, Marion Subklewe, Hans Joachim Stemmler, Hans Anton Schlößer, Max Schlaak, Matthias Kochanek, Boris Böll, Michael S. von Bergwelt-Baildon
Journal for ImmunoTherapy of Cancer (2018-06-15) https://doi.org/ghbncj
DOI: 10.1186/s40425-018-0343-9 · PMID: 29907163 · PMCID: PMC6003181

118. Understanding the Inflammatory Cytokine Response in Pneumonia and Sepsis
John A. Kellum
Archives of Internal Medicine (2007-08-13) https://doi.org/dbxb66
DOI: 10.1001/archinte.167.15.1655 · PMID: 17698689 · PMCID: PMC4495652

119. The Role of Interleukin 6 During Viral Infections
Lauro Velazquez-Salinas, Antonio Verdugo-Rodriguez, Luis L. Rodriguez, Manuel V. Borca
Frontiers in Microbiology (2019-05-10) https://doi.org/ghbnck
DOI: 10.3389/fmicb.2019.01057 · PMID: 31134045 · PMCID: PMC6524401

120. Expression of elevated levels of pro-inflammatory cytokines in SARS-CoV-infected ACE2 + cells in SARS patients: relation to the acute lung injury and pathogenesis of SARS
L He, Y Ding, Q Zhang, X Che, Y He, H Shen, H Wang, Z Li, L Zhao, J Geng, … S Jiang
The Journal of Pathology (2006-11) https://doi.org/bwb8ns
DOI: 10.1002/path.2067 · PMID: 17031779 · PMCID: PMC7167655

121. Up-regulation of IL-6 and TNF-α induced by SARS-coronavirus spike protein in murine macrophages via NF-κB pathway
Wei Wang, Linbai Ye, Li Ye, Baozong Li, Bo Gao, Yingchun Zeng, Lingbao Kong, Xiaonan Fang, Hong Zheng, Zhenghui Wu, Yinglong She
Virus Research (2007-09) https://doi.org/bm7m55
DOI: 10.1016/j.virusres.2007.02.007 · PMID: 17532082 · PMCID: PMC7114322

122. COVID-19: consider cytokine storm syndromes and immunosuppression
Puja Mehta, Daniel F McAuley, Michael Brown, Emilie Sanchez, Rachel S Tattersall, Jessica J Manson
The Lancet (2020-03) https://doi.org/ggnzmc
DOI: 10.1016/s0140-6736(20)30628-0 · PMID: 32192578 · PMCID: PMC7270045

123. Cytokine Storms: Understanding COVID-19
Nilam Mangalmurti, Christopher A. Hunter
Immunity (2020-07) https://doi.org/gg4fd7
DOI: 10.1016/j.immuni.2020.06.017 · PMID: 32610079 · PMCID: PMC7321048

124. Is a “Cytokine Storm” Relevant to COVID-19?
Pratik Sinha, Michael A. Matthay, Carolyn S. Calfee
JAMA Internal Medicine (2020-09-01) https://doi.org/gg3k6r
DOI: 10.1001/jamainternmed.2020.3313 · PMID: 32602883

125. Can we use interleukin-6 (IL-6) blockade for coronavirus disease 2019 (COVID-19)-induced cytokine release syndrome (CRS)?
Bingwen Liu, Min Li, Zhiguang Zhou, Xuan Guan, Yufei Xiang
Journal of Autoimmunity (2020-07) https://doi.org/ggr79c
DOI: 10.1016/j.jaut.2020.102452 · PMID: 32291137 · PMCID: PMC7151347

126. SARS-CoV-2 infection in primary schools in northern France: A retrospective cohort study in an area of high transmission
Arnaud Fontanet, Rebecca Grant, Laura Tondeur, Yoann Madec, Ludivine Grzelak, Isabelle Cailleau, Marie-Noëlle Ungeheuer, Charlotte Renaudat, Sandrine Fernandes Pellerin, Lucie Kuhmel, … Bruno Hoen
Cold Spring Harbor Laboratory (2020-06-29) https://doi.org/gg87nn
DOI: 10.1101/2020.06.25.20140178

127. SARS-CoV-2 Infection in Children
Xiaoxia Lu, Liqiong Zhang, Hui Du, Jingjing Zhang, Yuan Y. Li, Jingyu Qu, Wenxin Zhang, Youjie Wang, Shuangshuang Bao, Ying Li, … Gary W. K. Wong
New England Journal of Medicine (2020-04-23) https://doi.org/ggpt2q
DOI: 10.1056/nejmc2005073 · PMID: 32187458 · PMCID: PMC7121177

128. Systematic review of COVID‐19 in children shows milder cases and a better prognosis than adults
Jonas F. Ludvigsson
Acta Paediatrica (2020-04-14) https://doi.org/ggq8wr
DOI: 10.1111/apa.15270 · PMID: 32202343 · PMCID: PMC7228328

129. Reopening schools during COVID-19
Ronan Lordan, Garret A. FitzGerald, Tilo Grosser
Science (2020-09-03) https://doi.org/ghsv9p
DOI: 10.1126/science.abe5765 · PMID: 32883837

130. Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) Infection in Children and Adolescents
Riccardo Castagnoli, Martina Votto, Amelia Licari, Ilaria Brambilla, Raffaele Bruno, Stefano Perlini, Francesca Rovida, Fausto Baldanti, Gian Luigi Marseglia
JAMA Pediatrics (2020-09-01) https://doi.org/dswz
DOI: 10.1001/jamapediatrics.2020.1467 · PMID: 32320004

131. Neurologic and Radiographic Findings Associated With COVID-19 Infection in Children
Omar Abdel-Mannan, Michael Eyre, Ulrike Löbel, Alasdair Bamford, Christin Eltze, Biju Hameed, Cheryl Hemingway, Yael Hacohen
JAMA Neurology (2020-11-01) https://doi.org/gg339f
DOI: 10.1001/jamaneurol.2020.2687 · PMID: 32609336 · PMCID: PMC7330822

132. Children with Covid-19 in Pediatric Emergency Departments in Italy
Niccolò Parri, Matteo Lenge, Danilo Buonsenso
New England Journal of Medicine (2020-07-09) https://doi.org/ggtp6z
DOI: 10.1056/nejmc2007617 · PMID: 32356945 · PMCID: PMC7206930

133. COVID-19 in 7780 pediatric patients: A systematic review
Ansel Hoang, Kevin Chorath, Axel Moreira, Mary Evans, Finn Burmeister-Morton, Fiona Burmeister, Rija Naqvi, Matthew Petershack, Alvaro Moreira
EClinicalMedicine (2020-07) https://doi.org/gg4hn2
DOI: 10.1016/j.eclinm.2020.100433 · PMID: 32766542 · PMCID: PMC7318942

134. Multisystem Inflammatory Syndrome in Children During the Coronavirus 2019 Pandemic: A Case Series
Kathleen Chiotos, Hamid Bassiri, Edward M Behrens, Allison M Blatz, Joyce Chang, Caroline Diorio, Julie C Fitzgerald, Alexis Topjian, Audrey R Odom John
Journal of the Pediatric Infectious Diseases Society (2020-07) https://doi.org/ggx4pd
DOI: 10.1093/jpids/piaa069 · PMID: 32463092 · PMCID: PMC7313950

135. Clinical Characteristics of 58 Children With a Pediatric Inflammatory Multisystem Syndrome Temporally Associated With SARS-CoV-2
Elizabeth Whittaker, Alasdair Bamford, Julia Kenny, Myrsini Kaforou, Christine E. Jones, Priyen Shah, Padmanabhan Ramnarayan, Alain Fraisse, Owen Miller, Patrick Davies, … for the PIMS-TS Study Group and EUCLIDS and PERFORM Consortia
JAMA (2020-07-21) https://doi.org/gg2v75
DOI: 10.1001/jama.2020.10369 · PMID: 32511692 · PMCID: PMC7281356

136. Toxic shock-like syndrome and COVID-19: Multisystem inflammatory syndrome in children (MIS-C)
Andrea G. Greene, Mona Saleh, Eric Roseman, Richard Sinert
The American Journal of Emergency Medicine (2020-11) https://doi.org/gg2586
DOI: 10.1016/j.ajem.2020.05.117 · PMID: 32532619 · PMCID: PMC7274960

137. Multisystem inflammatory syndrome in children and COVID-19 are distinct presentations of SARS–CoV-2
Caroline Diorio, Sarah E. Henrickson, Laura A. Vella, Kevin O. McNerney, Julie Chase, Chakkapong Burudpakdee, Jessica H. Lee, Cristina Jasen, Fran Balamuth, David M. Barrett, … Hamid Bassiri
Journal of Clinical Investigation (2020-10-05) https://doi.org/gg7mz2
DOI: 10.1172/jci140970 · PMID: 32730233 · PMCID: PMC7598044

138. The Immunology of Multisystem Inflammatory Syndrome in Children with COVID-19
Camila Rosat Consiglio, Nicola Cotugno, Fabian Sardh, Christian Pou, Donato Amodio, Lucie Rodriguez, Ziyang Tan, Sonia Zicari, Alessandra Ruggiero, Giuseppe Rubens Pascucci, … Petter Brodin
Cell (2020-11) https://doi.org/d8fh
DOI: 10.1016/j.cell.2020.09.016 · PMID: 32966765 · PMCID: PMC7474869

139. Acute Heart Failure in Multisystem Inflammatory Syndrome in Children in the Context of Global SARS-CoV-2 Pandemic
Zahra Belhadjer, Mathilde Méot, Fanny Bajolle, Diala Khraiche, Antoine Legendre, Samya Abakka, Johanne Auriau, Marion Grimaud, Mehdi Oualha, Maurice Beghetti, … Damien Bonnet
Circulation (2020-08-04) https://doi.org/ggwkv6
DOI: 10.1161/circulationaha.120.048360 · PMID: 32418446

140. An adult with Kawasaki-like multisystem inflammatory syndrome associated with COVID-19
Sheila Shaigany, Marlis Gnirke, Allison Guttmann, Hong Chong, Shane Meehan, Vanessa Raabe, Eddie Louie, Bruce Solitar, Alisa Femia
The Lancet (2020-07) https://doi.org/gg4sd6
DOI: 10.1016/s0140-6736(20)31526-9 · PMID: 32659211 · PMCID: PMC7351414

141. COVID-19 associated Kawasaki-like multisystem inflammatory disease in an adult
Sabrina Sokolovsky, Parita Soni, Taryn Hoffman, Philip Kahn, Joshua Scheers-Masters
The American Journal of Emergency Medicine (2021-01) https://doi.org/gg5tf4
DOI: 10.1016/j.ajem.2020.06.053 · PMID: 32631771 · PMCID: PMC7315983

142. Systems Approaches to Biology and Disease Enable Translational Systems Medicine
Leroy Hood, Qiang Tian
Genomics, Proteomics & Bioinformatics (2012-08) https://doi.org/f4f599
DOI: 10.1016/j.gpb.2012.08.004 · PMID: 23084773 · PMCID: PMC3844613

143. A systems approach to infectious disease
Manon Eckhardt, Judd F. Hultquist, Robyn M. Kaake, Ruth Hüttenhain, Nevan J. Krogan
Nature Reviews Genetics (2020-02-14) https://doi.org/ggnv63
DOI: 10.1038/s41576-020-0212-5 · PMID: 32060427

144. Differential expression of serum/plasma proteins in various infectious diseases: Specific or nonspecific signatures
Sandipan Ray, Sandip K. Patel, Vipin Kumar, Jagruti Damahe, Sanjeeva Srivastava
PROTEOMICS - Clinical Applications (2014-02) https://doi.org/f2px3h
DOI: 10.1002/prca.201300074 · PMID: 24293340 · PMCID: PMC7168033

145. Imbalanced Host Response to SARS-CoV-2 Drives Development of COVID-19
Daniel Blanco-Melo, Benjamin E. Nilsson-Payant, Wen-Chun Liu, Skyler Uhl, Daisy Hoagland, Rasmus Møller, Tristan X. Jordan, Kohei Oishi, Maryline Panis, David Sachs, … Benjamin R. tenOever
Cell (2020-05) https://doi.org/ggw5tq
DOI: 10.1016/j.cell.2020.04.026 · PMID: 32416070 · PMCID: PMC7227586

146. SARS-CoV-2 ORF3b is a potent interferon antagonist whose activity is further increased by a naturally occurring elongation variant
Yoriyuki Konno, Izumi Kimura, Keiya Uriu, Masaya Fukushi, Takashi Irie, Yoshio Koyanagi, So Nakagawa, Kei Sato
Cold Spring Harbor Laboratory (2020-05-11) https://doi.org/gg2qgs
DOI: 10.1101/2020.05.11.088179

147. Bulk and single-cell gene expression profiling of SARS-CoV-2 infected human cell lines identifies molecular targets for therapeutic intervention
Wyler Emanuel, Mösbauer Kirstin, Franke Vedran, Diag Asija, Gottula Lina Theresa, Arsie Roberto, Klironomos Filippos, Koppstein David, Ayoub Salah, Buccitelli Christopher, … Landthaler Markus
Cold Spring Harbor Laboratory (2020-05-05) https://doi.org/ggxd2g
DOI: 10.1101/2020.05.05.079194

148. Isolation and characterization of SARS-CoV-2 from the first US COVID-19 patient
Jennifer Harcourt, Azaibi Tamin, Xiaoyan Lu, Shifaq Kamili, Senthil Kumar. Sakthivel, Janna Murray, Krista Queen, Ying Tao, Clinton R. Paden, Jing Zhang, … Natalie J. Thornburg
Cold Spring Harbor Laboratory (2020-03-07) https://doi.org/gg2fkm
DOI: 10.1101/2020.03.02.972935 · PMID: 32511316 · PMCID: PMC7239045

149. Proteomics of SARS-CoV-2-infected host cells reveals therapy targets
Denisa Bojkova, Kevin Klann, Benjamin Koch, Marek Widera, David Krause, Sandra Ciesek, Jindrich Cinatl, Christian Münch
Nature (2020-05-14) https://doi.org/dw7s
DOI: 10.1038/s41586-020-2332-7 · PMID: 32408336

150. Potent human neutralizing antibodies elicited by SARS-CoV-2 infection
Bin Ju, Qi Zhang, Xiangyang Ge, Ruoke Wang, Jiazhen Yu, Sisi Shan, Bing Zhou, Shuo Song, Xian Tang, Jinfang Yu, … Linqi Zhang
Cold Spring Harbor Laboratory (2020-03-26) https://doi.org/ggp7t4
DOI: 10.1101/2020.03.21.990770

151. Plasma proteome of severe acute respiratory syndrome analyzed by two-dimensional gel electrophoresis and mass spectrometry
J.-H. Chen, Y.-W. Chang, C.-W. Yao, T.-S. Chiueh, S.-C. Huang, K.-Y. Chien, A. Chen, F.-Y. Chang, C.-H. Wong, Y.-J. Chen
Proceedings of the National Academy of Sciences (2004-11-30) https://doi.org/dtv8sx
DOI: 10.1073/pnas.0407992101 · PMID: 15572443 · PMCID: PMC535397

152. Analysis of multimerization of the SARS coronavirus nucleocapsid protein
Runtao He, Frederick Dobie, Melissa Ballantine, Andrew Leeson, Yan Li, Nathalie Bastien, Todd Cutts, Anton Andonov, Jingxin Cao, Timothy F. Booth, … Xuguang Li
Biochemical and Biophysical Research Communications (2004-04) https://doi.org/dbfwr9
DOI: 10.1016/j.bbrc.2004.02.074 · PMID: 15020242 · PMCID: PMC7111152

153. A SARS-CoV-2-Human Protein-Protein Interaction Map Reveals Drug Targets and Potential Drug-Repurposing
David E. Gordon, Gwendolyn M. Jang, Mehdi Bouhaddou, Jiewei Xu, Kirsten Obernier, Matthew J. O’Meara, Jeffrey Z. Guo, Danielle L. Swaney, Tia A. Tummino, Ruth Hüttenhain, … Nevan J. Krogan
Cold Spring Harbor Laboratory (2020-03-22) https://doi.org/ggpptg
DOI: 10.1101/2020.03.22.002386 · PMID: 32511329 · PMCID: PMC7239059

154. Protein Palmitoylation and Its Role in Bacterial and Viral Infections
Justyna Sobocińska, Paula Roszczenko-Jasińska, Anna Ciesielska, Katarzyna Kwiatkowska
Frontiers in Immunology (2018-01-19) https://doi.org/gcxpp2
DOI: 10.3389/fimmu.2017.02003 · PMID: 29403483 · PMCID: PMC5780409

155. Virus-host interactome and proteomic survey of PMBCs from COVID-19 patients reveal potential virulence factors influencing SARS-CoV-2 pathogenesis
Jingjiao Li, Mingquan Guo, Xiaoxu Tian, Chengrong Liu, Xin Wang, Xing Yang, Ping Wu, Zixuan Xiao, Yafei Qu, Yue Yin, … Qiming Liang
Cold Spring Harbor Laboratory (2020-04-02) https://doi.org/ggrgbv
DOI: 10.1101/2020.03.31.019216

156. The Nuclear Factor NF- B Pathway in Inflammation
T. Lawrence
Cold Spring Harbor Perspectives in Biology (2009-10-07) https://doi.org/fptfvp
DOI: 10.1101/cshperspect.a001651 · PMID: 20457564 · PMCID: PMC2882124

157. The severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) envelope (E) protein harbors a conserved BH3-like sequence
Vincent Navratil, Loïc Lionnard, Sonia Longhi, J. Marie Hardwick, Christophe Combet, Abdel Aouacheria
Cold Spring Harbor Laboratory (2020-06-09) https://doi.org/ggrp43
DOI: 10.1101/2020.04.09.033522

158. Structure of SARS Coronavirus Spike Receptor-Binding Domain Complexed with Receptor
F. Li
Science (2005-09-16) https://doi.org/fww324
DOI: 10.1126/science.1116480 · PMID: 16166518

159. Structural basis for the recognition of SARS-CoV-2 by full-length human ACE2
Renhong Yan, Yuanyuan Zhang, Yaning Li, Lu Xia, Yingying Guo, Qiang Zhou
Science (2020-03-27) https://doi.org/ggpxc8
DOI: 10.1126/science.abb2762 · PMID: 32132184 · PMCID: PMC7164635

160. Structural basis of receptor recognition by SARS-CoV-2
Jian Shang, Gang Ye, Ke Shi, Yushun Wan, Chuming Luo, Hideki Aihara, Qibin Geng, Ashley Auerbach, Fang Li
Nature (2020-03-30) https://doi.org/ggqspv
DOI: 10.1038/s41586-020-2179-y · PMID: 32225175 · PMCID: PMC7328981

161. Crystal structure of the 2019-nCoV spike receptor-binding domain bound with the ACE2 receptor
Jun Lan, Jiwan Ge, Jinfang Yu, Sisi Shan, Huan Zhou, Shilong Fan, Qi Zhang, Xuanling Shi, Qisheng Wang, Linqi Zhang, Xinquan Wang
Cold Spring Harbor Laboratory (2020-02-20) https://doi.org/ggqzp5
DOI: 10.1101/2020.02.19.956235

162. Structural and Functional Basis of SARS-CoV-2 Entry by Using Human ACE2
Qihui Wang, Yanfang Zhang, Lili Wu, Sheng Niu, Chunli Song, Zengyuan Zhang, Guangwen Lu, Chengpeng Qiao, Yu Hu, Kwok-Yung Yuen, … Jianxun Qi
Cell (2020-05) https://doi.org/ggr2cz
DOI: 10.1016/j.cell.2020.03.045 · PMID: 32275855 · PMCID: PMC7144619

163. Receptor Recognition by the Novel Coronavirus from Wuhan: an Analysis Based on Decade-Long Structural Studies of SARS Coronavirus
Yushun Wan, Jian Shang, Rachel Graham, Ralph S. Baric, Fang Li
Journal of Virology (2020-03-17) https://doi.org/ggjvwn
DOI: 10.1128/jvi.00127-20 · PMID: 31996437 · PMCID: PMC7081895

164. Structure of the Hemagglutinin Precursor Cleavage Site, a Determinant of Influenza Pathogenicity and the Origin of the Labile Conformation
Jue Chen, Kon Ho Lee, David A Steinhauer, David J Stevens, John J Skehel, Don C Wiley
Cell (1998-10) https://doi.org/bvgh5b
DOI: 10.1016/s0092-8674(00)81771-7

165. Role of Hemagglutinin Cleavage for the Pathogenicity of Influenza Virus
David A. Steinhauer
Virology (1999-05) https://doi.org/fw3jz4
DOI: 10.1006/viro.1999.9716 · PMID: 10329563

166. Controversy around airborne versus droplet transmission of respiratory viruses
Eunice Y. C. Shiu, Nancy H. L. Leung, Benjamin J. Cowling
Current Opinion in Infectious Diseases (2019-08) https://doi.org/ggbwdb
DOI: 10.1097/qco.0000000000000563

167. Viral infections acquired indoors through airborne, droplet or contact transmission.
Giuseppina La Rosa, Marta Fratini, Simonetta Della Libera, Marcello Iaconelli, Michele Muscillo
Annali dell’Istituto superiore di sanita (2013) https://www.ncbi.nlm.nih.gov/pubmed/23771256
DOI: 10.4415/ann_13_02_03 · PMID: 23771256

168. Transmission of SARS-CoV-2: implications for infection prevention precautions https://www.who.int/news-room/commentaries/detail/transmission-of-sars-cov-2-implications-for-infection-prevention-precautions

169. Questioning Aerosol Transmission of Influenza
Camille Lemieux, Gabrielle Brankston, Leah Gitterman, Zahir Hirji, Michael Gardam
Emerging Infectious Diseases (2007-01) https://doi.org/c2skj8
DOI: 10.3201/eid1301.061202 · PMID: 17370541 · PMCID: PMC2725811

170. Assessing the Dynamics and Control of Droplet- and Aerosol-Transmitted Influenza Using an Indoor Positioning System
Timo Smieszek, Gianrocco Lazzari, Marcel Salathé
Scientific Reports (2019-02-18) https://doi.org/ggnqbc
DOI: 10.1038/s41598-019-38825-y · PMID: 30778136 · PMCID: PMC6379436

171. Influenza A virus transmission via respiratory aerosols or droplets as it relates to pandemic potential
Mathilde Richard, Ron A. M. Fouchier
FEMS Microbiology Reviews (2016-01) https://doi.org/f8cp4h
DOI: 10.1093/femsre/fuv039 · PMID: 26385895 · PMCID: PMC5006288

172. Coronavirus Pathogenesis
Susan R. Weiss, Julian L. Leibowitz
Advances in Virus Research (2011) https://doi.org/ggvvd7
DOI: 10.1016/b978-0-12-385885-6.00009-2 · PMID: 22094080 · PMCID: PMC7149603

173. Role of fomites in SARS transmission during the largest hospital outbreak in Hong Kong
Shenglan Xiao, Yuguo Li, Tze-wai Wong, David S. C. Hui
PLOS ONE (2017-07-20) https://doi.org/gbpgv7
DOI: 10.1371/journal.pone.0181558 · PMID: 28727803 · PMCID: PMC5519164

174. Stability of Middle East respiratory syndrome coronavirus (MERS-CoV) under different environmental conditions
N van Doremalen, T Bushmaker, VJ Munster
Eurosurveillance (2013-09-19) https://doi.org/ggnnjt
DOI: 10.2807/1560-7917.es2013.18.38.20590 · PMID: 24084338

175. MERS coronavirus: diagnostics, epidemiology and transmission
Ian M. Mackay, Katherine E. Arden
Virology Journal (2015-12-22) https://doi.org/f745px
DOI: 10.1186/s12985-015-0439-5 · PMID: 26695637 · PMCID: PMC4687373

176. Transmission routes of 2019-nCoV and controls in dental practice
Xian Peng, Xin Xu, Yuqing Li, Lei Cheng, Xuedong Zhou, Biao Ren
International Journal of Oral Science (2020-03-03) https://doi.org/ggnf47
DOI: 10.1038/s41368-020-0075-9 · PMID: 32127517 · PMCID: PMC7054527

177. Reducing transmission of SARS-CoV-2
Kimberly A. Prather, Chia C. Wang, Robert T. Schooley
Science (2020-06-26) https://doi.org/ggxp9w
DOI: 10.1126/science.abc6197 · PMID: 32461212

178. Aerosol and Surface Stability of SARS-CoV-2 as Compared with SARS-CoV-1
Neeltje van Doremalen, Trenton Bushmaker, Dylan H. Morris, Myndi G. Holbrook, Amandine Gamble, Brandi N. Williamson, Azaibi Tamin, Jennifer L. Harcourt, Natalie J. Thornburg, Susan I. Gerber, … Vincent J. Munster
New England Journal of Medicine (2020-04-16) https://doi.org/ggn88w
DOI: 10.1056/nejmc2004973 · PMID: 32182409 · PMCID: PMC7121658

179. It Is Time to Address Airborne Transmission of Coronavirus Disease 2019 (COVID-19)
Lidia Morawska, Donald K Milton
Clinical Infectious Diseases (2020-07-06) https://doi.org/gg34zn
DOI: 10.1093/cid/ciaa939 · PMID: 32628269 · PMCID: PMC7454469

180. Aerodynamic analysis of SARS-CoV-2 in two Wuhan hospitals
Yuan Liu, Zhi Ning, Yu Chen, Ming Guo, Yingle Liu, Nirmal Kumar Gali, Li Sun, Yusen Duan, Jing Cai, Dane Westerdahl, … Ke Lan
Nature (2020-04-27) https://doi.org/ggtgng
DOI: 10.1038/s41586-020-2271-3 · PMID: 32340022

181. Indirect Virus Transmission in Cluster of COVID-19 Cases, Wenzhou, China, 2020
Jing Cai, Wenjie Sun, Jianping Huang, Michelle Gamber, Jing Wu, Guiqing He
Emerging Infectious Diseases (2020-06) https://doi.org/ggpxkx
DOI: 10.3201/eid2606.200412 · PMID: 32163030 · PMCID: PMC7258486

182. Airborne Transmission of SARS-CoV-2
Michael Klompas, Meghan A. Baker, Chanu Rhee
JAMA (2020-08-04) https://doi.org/gg4ttq
DOI: 10.1001/jama.2020.12458

183. Exaggerated risk of transmission of COVID-19 by fomites
Emanuel Goldman
The Lancet Infectious Diseases (2020-08) https://doi.org/gg6br7
DOI: 10.1016/s1473-3099(20)30561-2 · PMID: 32628907 · PMCID: PMC7333993

184. Clinical characteristics of 24 asymptomatic infections with COVID-19 screened among close contacts in Nanjing, China
Zhiliang Hu, Ci Song, Chuanjun Xu, Guangfu Jin, Yaling Chen, Xin Xu, Hongxia Ma, Wei Chen, Yuan Lin, Yishan Zheng, … Hongbing Shen
Science China Life Sciences (2020-03-04) https://doi.org/dqbn
DOI: 10.1007/s11427-020-1661-4 · PMID: 32146694 · PMCID: PMC7088568

185. Evidence for transmission of COVID-19 prior to symptom onset
Lauren C Tindale, Jessica E Stockdale, Michelle Coombe, Emma S Garlock, Wing Yin Venus Lau, Manu Saraswat, Louxin Zhang, Dongxuan Chen, Jacco Wallinga, Caroline Colijn
eLife (2020-06-22) https://doi.org/gg6dtw
DOI: 10.7554/elife.57149 · PMID: 32568070 · PMCID: PMC7386904

186. Time Kinetics of Viral Clearance and Resolution of Symptoms in Novel Coronavirus Infection
De Chang, Guoxin Mo, Xin Yuan, Yi Tao, Xiaohua Peng, Fu-Sheng Wang, Lixin Xie, Lokesh Sharma, Charles S. Dela Cruz, Enqiang Qin
American Journal of Respiratory and Critical Care Medicine (2020-05-01) https://doi.org/ggq8xs
DOI: 10.1164/rccm.202003-0524le · PMID: 32200654 · PMCID: PMC7193851

187. Temporal dynamics in viral shedding and transmissibility of COVID-19
Xi He, Eric H. Y. Lau, Peng Wu, Xilong Deng, Jian Wang, Xinxin Hao, Yiu Chung Lau, Jessica Y. Wong, Yujuan Guan, Xinghua Tan, … Gabriel M. Leung
Nature Medicine (2020-04-15) https://doi.org/ggr99q
DOI: 10.1038/s41591-020-0869-5 · PMID: 32296168

188. Coronavirus Disease 2019 (COVID-19) - Transmission
CDC
Centers for Disease Control and Prevention (2020-10-28) https://www.cdc.gov/coronavirus/2019-ncov/prevent-getting-sick/how-covid-spreads.html

189. Virological assessment of hospitalized patients with COVID-2019
Roman Wölfel, Victor M. Corman, Wolfgang Guggemos, Michael Seilmaier, Sabine Zange, Marcel A. Müller, Daniela Niemeyer, Terry C. Jones, Patrick Vollmar, Camilla Rothe, … Clemens Wendtner
Nature (2020-04-01) https://doi.org/ggqrv7
DOI: 10.1038/s41586-020-2196-x · PMID: 32235945

190. Clinical predictors and timing of cessation of viral RNA shedding in patients with COVID-19
Cristina Corsini Campioli, Edison Cano Cevallos, Mariam Assi, Robin Patel, Matthew J. Binnicker, John C. O’Horo
Journal of Clinical Virology (2020-09) https://doi.org/gg7m96
DOI: 10.1016/j.jcv.2020.104577 · PMID: 32777762 · PMCID: PMC7405830

191. Presymptomatic SARS-CoV-2 Infections and Transmission in a Skilled Nursing Facility
Melissa M. Arons, Kelly M. Hatfield, Sujan C. Reddy, Anne Kimball, Allison James, Jesica R. Jacobs, Joanne Taylor, Kevin Spicer, Ana C. Bardossy, Lisa P. Oakley, … John A. Jernigan
New England Journal of Medicine (2020-05-28) https://doi.org/ggszfg
DOI: 10.1056/nejmoa2008457 · PMID: 32329971 · PMCID: PMC7200056

192. Prevalence of SARS-CoV-2 Infection in Residents of a Large Homeless Shelter in Boston
Travis P. Baggett, Harrison Keyes, Nora Sporn, Jessie M. Gaeta
JAMA (2020-06-02) https://doi.org/ggtsh3
DOI: 10.1001/jama.2020.6887 · PMID: 32338732 · PMCID: PMC7186911

193. Presumed Asymptomatic Carrier Transmission of COVID-19
Yan Bai, Lingsheng Yao, Tao Wei, Fei Tian, Dong-Yan Jin, Lijuan Chen, Meiyun Wang
JAMA (2020-04-14) https://doi.org/ggmbs8
DOI: 10.1001/jama.2020.2565 · PMID: 32083643 · PMCID: PMC7042844

194. Transmission of COVID-19 in the terminal stages of the incubation period: A familial cluster
Peng Li, Ji-Bo Fu, Ke-Feng Li, Jie-Nan Liu, Hong-Ling Wang, Lei-Jie Liu, Yan Chen, Yong-Li Zhang, She-Lan Liu, An Tang, … Jian-Bo Yan
International Journal of Infectious Diseases (2020-07) https://doi.org/ggq844
DOI: 10.1016/j.ijid.2020.03.027 · PMID: 32194239 · PMCID: PMC7264481

195. A Cohort of SARS-CoV-2 Infected Asymptomatic and Pre-Symptomatic Contacts from COVID-19 Contact Tracing in Hubei Province, China: Short-Term Outcomes
Peng Zhang, Fei Tian, Yuan Wan, Jing Cai, Zhengmin Qian, Ran Wu, Yunquan Zhang, Shiyu Zhang, Huan Li, Mingyan Li, … Hualiang Lin
SSRN Electronic Journal (2020) https://doi.org/ghf3n2
DOI: 10.2139/ssrn.3678556

196. Estimating the asymptomatic proportion of coronavirus disease 2019 (COVID-19) cases on board the Diamond Princess cruise ship, Yokohama, Japan, 2020
Kenji Mizumoto, Katsushi Kagaya, Alexander Zarebski, Gerardo Chowell
Eurosurveillance (2020-03-12) https://doi.org/ggn4bd
DOI: 10.2807/1560-7917.es.2020.25.10.2000180 · PMID: 32183930 · PMCID: PMC7078829

197. Estimated prevalence and viral transmissibility in subjects with asymptomatic SARS-CoV-2 infections in Wuhan, China
Kang Zhang, Weiwei Tong, Xinghuan Wang, Johnson Yiu-Nam Lau
Precision Clinical Medicine (2020-12) https://doi.org/ghjmks
DOI: 10.1093/pcmedi/pbaa032 · PMCID: PMC7499683

198. Investigation of SARS-CoV-2 outbreaks in six care homes in London, April 2020
Shamez N Ladhani, J.Yimmy Chow, Roshni Janarthanan, Jonathan Fok, Emma Crawley-Boevey, Amoolya Vusirikala, Elena Fernandez, Marina Sanchez Perez, Suzanne Tang, Kate Dun-Campbell, … Maria Zambon
EClinicalMedicine (2020-09) https://doi.org/ghbj9v
DOI: 10.1016/j.eclinm.2020.100533 · PMID: 32923993 · PMCID: PMC7480335

199. Clinical and immunological assessment of asymptomatic SARS-CoV-2 infections
Quan-Xin Long, Xiao-Jun Tang, Qiu-Lin Shi, Qin Li, Hai-Jun Deng, Jun Yuan, Jie-Li Hu, Wei Xu, Yong Zhang, Fa-Jin Lv, … Ai-Long Huang
Nature Medicine (2020-06-18) https://doi.org/gg26dx
DOI: 10.1038/s41591-020-0965-6 · PMID: 32555424

200. Suppression of a SARS-CoV-2 outbreak in the Italian municipality of Vo’
Enrico Lavezzo, Elisa Franchin, Constanze Ciavarella, Gina Cuomo-Dannenburg, Luisa Barzon, Claudia Del Vecchio, Lucia Rossi, Riccardo Manganelli, Arianna Loregian, Nicolò Navarin, … Imperial College COVID-19 Response Team
Nature (2020-06-30) https://doi.org/gg3w87
DOI: 10.1038/s41586-020-2488-1 · PMID: 32604404

201. A systematic review and meta-analysis of published research data on COVID-19 infection-fatality rates
Gideon Meyerowitz-Katz, Lea Merone
Cold Spring Harbor Laboratory (2020-05-06) https://doi.org/dv2r
DOI: 10.1101/2020.05.03.20089854

202. Global Covid-19 Case Fatality Rates
The Centre for Evidence-Based Medicine
https://www.cebm.net/covid-19/global-covid-19-case-fatality-rates/

203. Estimating the Global Infection Fatality Rate of COVID-19
Richard Grewelle, Giulio De Leo
Cold Spring Harbor Laboratory (2020-05-18) https://doi.org/ghbvcj
DOI: 10.1101/2020.05.11.20098780

204. Repeated cross-sectional sero-monitoring of SARS-CoV-2 in New York City
Daniel Stadlbauer, Jessica Tan, Kaijun Jiang, Matthew M. Hernandez, Shelcie Fabre, Fatima Amanat, Catherine Teo, Guha Asthagiri Arunkumar, Meagan McMahon, Christina Capuano, … Florian Krammer
Nature (2020-11-03) https://doi.org/ghhtq9
DOI: 10.1038/s41586-020-2912-6 · PMID: 33142304

205. Modeling infectious diseases in humans and animals
Matthew James Keeling, Pejman Rohani
Princeton University Press (2008)
ISBN: 9780691116174

206. The concept of R o in epidemic theory
J. A. P. Heesterbeek, K. Dietz
Statistica Neerlandica (1996-03) https://doi.org/d29ch4
DOI: 10.1111/j.1467-9574.1996.tb01482.x

207. Modeling infectious disease dynamics
Sarah Cobey
Science (2020-05-15) https://doi.org/ggsztw
DOI: 10.1126/science.abb5659 · PMID: 32332062

208. Theoretical ecology: principles and applications
Robert M. May, Angela R. McLean (editors)
Oxford University Press (2007)
ISBN: 9780199209989

209. A contribution to the mathematical theory of epidemics
William Ogilvy Kermack, A. G. McKendrick
Proceedings of the Royal Society of London. Series A, Containing Papers of a Mathematical and Physical Character (1927-08) https://doi.org/10.1098/rspa.1927.0118
DOI: 10.1098/rspa.1927.0118

210. Population biology of infectious diseases: Part I
Roy M. Anderson, Robert M. May
Nature (1979-08-01) https://doi.org/b6z9hc
DOI: 10.1038/280361a0 · PMID: 460412

211. Nowcasting and forecasting the potential domestic and international spread of the 2019-nCoV outbreak originating in Wuhan, China: a modelling study
Joseph T Wu, Kathy Leung, Gabriel M Leung
The Lancet (2020-02) https://doi.org/ggjvr7
DOI: 10.1016/s0140-6736(20)30260-9 · PMID: 32014114 · PMCID: PMC7159271

212. The reproductive number of COVID-19 is higher compared to SARS coronavirus
Ying Liu, Albert A Gayle, Annelies Wilder-Smith, Joacim Rocklöv
Journal of Travel Medicine (2020-03) https://doi.org/ggnntv
DOI: 10.1093/jtm/taaa021 · PMID: 32052846 · PMCID: PMC7074654

213. Substantial undocumented infection facilitates the rapid dissemination of novel coronavirus (SARS-CoV-2)
Ruiyun Li, Sen Pei, Bin Chen, Yimeng Song, Tao Zhang, Wan Yang, Jeffrey Shaman
Science (2020-05-01) https://doi.org/ggn6c2
DOI: 10.1126/science.abb3221 · PMID: 32179701 · PMCID: PMC7164387

214. Epidemiological parameters of coronavirus disease 2019: a pooled analysis of publicly reported individual data of 1155 cases from seven countries
Shujuan Ma, Jiayue Zhang, Minyan Zeng, Qingping Yun, Wei Guo, Yixiang Zheng, Shi Zhao, Maggie H. Wang, Zuyao Yang
Cold Spring Harbor Laboratory (2020-03-24) https://doi.org/ggqhzz
DOI: 10.1101/2020.03.21.20040329

215. Early Transmissibility Assessment of a Novel Coronavirus in Wuhan, China
Maimuna Majumder, Kenneth D. Mandl
SSRN Electronic Journal (2020) https://doi.org/ggqhz3
DOI: 10.2139/ssrn.3524675 · PMID: 32714102 · PMCID: PMC7366781

216. Time-varying transmission dynamics of Novel Coronavirus Pneumonia in China
Tao Liu, Jianxiong Hu, Jianpeng Xiao, Guanhao He, Min Kang, Zuhua Rong, Lifeng Lin, Haojie Zhong, Qiong Huang, Aiping Deng, … Wenjun Ma
Cold Spring Harbor Laboratory (2020-02-13) https://doi.org/dkx9
DOI: 10.1101/2020.01.25.919787

217. Estimation of the reproductive number of novel coronavirus (COVID-19) and the probable outbreak size on the Diamond Princess cruise ship: A data-driven analysis
Sheng Zhang, MengYuan Diao, Wenbo Yu, Lei Pei, Zhaofen Lin, Dechang Chen
International Journal of Infectious Diseases (2020-04) https://doi.org/ggpx56
DOI: 10.1016/j.ijid.2020.02.033 · PMID: 32097725 · PMCID: PMC7110591

218. Estimation of the Transmission Risk of the 2019-nCoV and Its Implication for Public Health Interventions
Biao Tang, Xia Wang, Qian Li, Nicola Luigi Bragazzi, Sanyi Tang, Yanni Xiao, Jianhong Wu
Journal of Clinical Medicine (2020-02-07) https://doi.org/ggmkf4
DOI: 10.3390/jcm9020462 · PMID: 32046137 · PMCID: PMC7074281

219. Estimating the effective reproduction number of the 2019-nCoV in China
Zhidong Cao, Qingpeng Zhang, Xin Lu, Dirk Pfeiffer, Zhongwei Jia, Hongbing Song, Daniel Dajun Zeng
medRxiv (2020-01) https://www.medrxiv.org/content/10.1101/2020.01.27.20018952v1
DOI: 10.1101/2020.01.27.20018952

220. Modelling the epidemic trend of the 2019 novel coronavirus outbreak in China
Mingwang Shen, Zhihang Peng, Yanni Xiao, Lei Zhang
Cold Spring Harbor Laboratory (2020-01-25) https://doi.org/ggqhzw
DOI: 10.1101/2020.01.23.916726

221. Novel coronavirus 2019-nCoV: early estimation of epidemiological parameters and epidemic predictions
Jonathan M. Read, Jessica R. E. Bridgen, Derek A. T. Cummings, Antonia Ho, Chris P. Jewell
Cold Spring Harbor Laboratory (2020-01-28) https://doi.org/dkzb
DOI: 10.1101/2020.01.23.20018549

222. Using early data to estimate the actual infection fatality ratio from COVID-19 in France
Lionel Roques, Etienne Klein, Julien Papaïx, Antoine Sar, Samuel Soubeyrand
Cold Spring Harbor Laboratory (2020-04-16) https://doi.org/ggqhz2
DOI: 10.1101/2020.03.22.20040915

223. Potential roles of social distancing in mitigating the spread of coronavirus disease 2019 (COVID-19) in South Korea
Sang Woo Park, Kaiyuan Sun, Cécile Viboud, Bryan T Grenfell, Jonathan Dushoff
GitHub (2020) https://github.com/parksw3/Korea-analysis/blob/master/v1/korea.pdf

224. Early dynamics of transmission and control of COVID-19: a mathematical modelling study
Adam J Kucharski, Timothy W Russell, Charlie Diamond, Yang Liu, John Edmunds, Sebastian Funk, Rosalind M Eggo, Fiona Sun, Mark Jit, James D Munday, … Stefan Flasche
The Lancet Infectious Diseases (2020-05) https://doi.org/ggptcf
DOI: 10.1016/s1473-3099(20)30144-4 · PMID: 32171059 · PMCID: PMC7158569

225. Estimating the reproduction number of COVID-19 in Iran using epidemic modeling
Ebrahim Sahafizadeh, Samaneh Sartoli
Cold Spring Harbor Laboratory (2020-04-23) https://doi.org/ggqhzx
DOI: 10.1101/2020.03.20.20038422

226. Report 13: Estimating the number of infections and the impact of non-pharmaceutical interventions on COVID-19 in 11 European countries
S Flaxman, S Mishra, A Gandy, H Unwin, H Coupland, T Mellan, H Zhu, T Berah, J Eaton, P Perez Guzman, … S Bhatt
Imperial College London (2020-03-30) https://doi.org/ggrbmf
DOI: 10.25561/77731

227. Projecting hospital utilization during the COVID-19 outbreaks in the United States
Seyed M. Moghadas, Affan Shoukat, Meagan C. Fitzpatrick, Chad R. Wells, Pratha Sah, Abhishek Pandey, Jeffrey D. Sachs, Zheng Wang, Lauren A. Meyers, Burton H. Singer, Alison P. Galvani
Proceedings of the National Academy of Sciences (2020-04-21) https://doi.org/ggq7jc
DOI: 10.1073/pnas.2004064117 · PMID: 32245814 · PMCID: PMC7183199

228. The effect of control strategies to reduce social mixing on outcomes of the COVID-19 epidemic in Wuhan, China: a modelling study
Kiesha Prem, Yang Liu, Timothy W Russell, Adam J Kucharski, Rosalind M Eggo, Nicholas Davies, Mark Jit, Petra Klepac, Stefan Flasche, Samuel Clifford, … Joel Hellewell
The Lancet Public Health (2020-05) https://doi.org/ggp3xq
DOI: 10.1016/s2468-2667(20)30073-6 · PMID: 32220655 · PMCID: PMC7158905

229. Spread and dynamics of the COVID-19 epidemic in Italy: Effects of emergency containment measures
Marino Gatto, Enrico Bertuzzo, Lorenzo Mari, Stefano Miccoli, Luca Carraro, Renato Casagrandi, Andrea Rinaldo
Proceedings of the National Academy of Sciences (2020-05-12) https://doi.org/ggv4j6
DOI: 10.1073/pnas.2004978117 · PMID: 32327608 · PMCID: PMC7229754

230. Covid-19: Temporal variation in transmission during the COVID-19 outbreak
EpiForecasts and the CMMID Covid working group
https://epiforecasts.io/covid/

231. Rt COVID-19
Kevin Systrom, Thomas Vladeck, Mike Krieger
https://rt.live/

232. Emerging SARS-CoV-2 mutation hot spots include a novel RNA-dependent-RNA polymerase variant
Maria Pachetti, Bruna Marini, Francesca Benedetti, Fabiola Giudici, Elisabetta Mauro, Paola Storici, Claudio Masciovecchio, Silvia Angeletti, Massimo Ciccozzi, Robert C. Gallo, … Rudy Ippodrino
Journal of Translational Medicine (2020-04-22) https://doi.org/ggtzrr
DOI: 10.1186/s12967-020-02344-6 · PMID: 32321524 · PMCID: PMC7174922

233. Emergence of genomic diversity and recurrent mutations in SARS-CoV-2
Lucy van Dorp, Mislav Acman, Damien Richard, Liam P. Shaw, Charlotte E. Ford, Louise Ormond, Christopher J. Owen, Juanita Pang, Cedric C. S. Tan, Florencia A. T. Boshier, … François Balloux
Infection, Genetics and Evolution (2020-09) https://doi.org/ggvz4h
DOI: 10.1016/j.meegid.2020.104351 · PMID: 32387564 · PMCID: PMC7199730

234. An integrated national scale SARS-CoV-2 genomic surveillance network
The Lancet Microbe
(2020-07) https://doi.org/d5mg
DOI: 10.1016/s2666-5247(20)30054-9 · PMID: 32835336 · PMCID: PMC7266609

235. Coast-to-Coast Spread of SARS-CoV-2 during the Early Epidemic in the United States
Joseph R. Fauver, Mary E. Petrone, Emma B. Hodcroft, Kayoko Shioda, Hanna Y. Ehrlich, Alexander G. Watts, Chantal B. F. Vogels, Anderson F. Brito, Tara Alpert, Anthony Muyombwe, … Nathan D. Grubaugh
Cell (2020-05) https://doi.org/gg6r9x
DOI: 10.1016/j.cell.2020.04.021 · PMID: 32386545 · PMCID: PMC7204677

236. Introductions and early spread of SARS-CoV-2 in the New York City area
Ana S. Gonzalez-Reiche, Matthew M. Hernandez, Mitchell J. Sullivan, Brianne Ciferri, Hala Alshammary, Ajay Obla, Shelcie Fabre, Giulio Kleiner, Jose Polanco, Zenab Khan, … Harm van Bakel
Science (2020-05-29) https://doi.org/gg5gv7
DOI: 10.1126/science.abc1917 · PMID: 32471856 · PMCID: PMC7259823

237. Spread of SARS-CoV-2 in the Icelandic Population
Daniel F. Gudbjartsson, Agnar Helgason, Hakon Jonsson, Olafur T. Magnusson, Pall Melsted, Gudmundur L. Norddahl, Jona Saemundsdottir, Asgeir Sigurdsson, Patrick Sulem, Arna B. Agustsdottir, … Kari Stefansson
New England Journal of Medicine (2020-06-11) https://doi.org/ggr6wx
DOI: 10.1056/nejmoa2006100 · PMID: 32289214 · PMCID: PMC7175425

238. GISAID - Initiative https://www.gisaid.org/

239. NCBI SARS-CoV-2 Resources https://www.ncbi.nlm.nih.gov/sars-cov-2/

240. COVID-19 Data Portal - accelerating scientific research through data https://www.covid19dataportal.org/

241. Epidemiology, transmission dynamics and control of SARS: the 2002–2003 epidemic
Roy M. Anderson, Christophe Fraser, Azra C. Ghani, Christl A. Donnelly, Steven Riley, Neil M. Ferguson, Gabriel M. Leung, T. H. Lam, Anthony J. Hedley
Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences (2004-07-29) https://doi.org/c2n646
DOI: 10.1098/rstb.2004.1490 · PMID: 15306395 · PMCID: PMC1693389

242. Coronaviruses and gastrointestinal diseases
Xi Luo, Guan-Zhou Zhou, Yan Zhang, Li-Hua Peng, Li-Ping Zou, Yun-Sheng Yang
Military Medical Research (2020-10-14) https://doi.org/ghqfmj
DOI: 10.1186/s40779-020-00279-z · PMID: 33054860 · PMCID: PMC7556584

243. Human (non-severe acute respiratory syndrome) coronavirus infections in hospitalised children in France
Astrid Vabret, Julia Dina, Stéphanie Gouarin, Joëlle Petitjean, Valérie Tripey, Jacques Brouard, François Freymuth
Journal of Paediatrics and Child Health (2008-04) https://doi.org/cxt434
DOI: 10.1111/j.1440-1754.2007.01246.x · PMID: 17999671 · PMCID: PMC7166728

244. Tracking Changes in SARS-CoV-2 Spike: Evidence that D614G Increases Infectivity of the COVID-19 Virus
Bette Korber, Will M. Fischer, Sandrasegaram Gnanakaran, Hyejin Yoon, James Theiler, Werner Abfalterer, Nick Hengartner, Elena E. Giorgi, Tanmoy Bhattacharya, Brian Foley, … Matthew D. Wyles
Cell (2020-08) https://doi.org/gg3wqn
DOI: 10.1016/j.cell.2020.06.043 · PMID: 32697968 · PMCID: PMC7332439

245. Making Sense of Mutation: What D614G Means for the COVID-19 Pandemic Remains Unclear
Nathan D. Grubaugh, William P. Hanage, Angela L. Rasmussen
Cell (2020-08) https://doi.org/gg4gqt
DOI: 10.1016/j.cell.2020.06.040 · PMID: 32697970 · PMCID: PMC7332445

246. Detection of 2019 novel coronavirus (2019-nCoV) by real-time RT-PCR
Victor M Corman, Olfert Landt, Marco Kaiser, Richard Molenkamp, Adam Meijer, Daniel KW Chu, Tobias Bleicker, Sebastian Brünink, Julia Schneider, Marie Luisa Schmidt, … Christian Drosten
Eurosurveillance (2020-01-23) https://doi.org/ggjs7g
DOI: 10.2807/1560-7917.es.2020.25.3.2000045 · PMID: 31992387 · PMCID: PMC6988269

247. Temporal profiles of viral load in posterior oropharyngeal saliva samples and serum antibody responses during infection by SARS-CoV-2: an observational cohort study
Kelvin Kai-Wang To, Owen Tak-Yin Tsang, Wai-Shing Leung, Anthony Raymond Tam, Tak-Chiu Wu, David Christopher Lung, Cyril Chik-Yan Yip, Jian-Piao Cai, Jacky Man-Chun Chan, Thomas Shiu-Hong Chik, … Kwok-Yung Yuen
The Lancet Infectious Diseases (2020-05) https://doi.org/ggp4qx
DOI: 10.1016/s1473-3099(20)30196-1 · PMID: 32213337 · PMCID: PMC7158907

248. Fecal specimen diagnosis 2019 novel coronavirus–infected pneumonia
JingCheng Zhang, SaiBin Wang, YaDong Xue
Journal of Medical Virology (2020-03-12) https://doi.org/ggpx6d
DOI: 10.1002/jmv.25742 · PMID: 32124995 · PMCID: PMC7228355

249. Library preparation for next generation sequencing: A review of automation strategies
J. F. Hess, T. A. Kohl, M. Kotrová, K. Rönsch, T. Paprotka, V. Mohr, T. Hutzenlaub, M. Brüggemann, R. Zengerle, S. Niemann, N. Paust
Biotechnology Advances (2020-07) https://doi.org/ggth2v
DOI: 10.1016/j.biotechadv.2020.107537 · PMID: 32199980

250. Diagnosing COVID-19: The Disease and Tools for Detection
Buddhisha Udugama, Pranav Kadhiresan, Hannah N. Kozlowski, Ayden Malekjahani, Matthew Osborne, Vanessa Y. C. Li, Hongmin Chen, Samira Mubareka, Jonathan B. Gubbay, Warren C. W. Chan
ACS Nano (2020-03-30) https://doi.org/ggq8ds
DOI: 10.1021/acsnano.0c02624 · PMID: 32223179 · PMCID: PMC7144809

251. Molecular Diagnosis of a Novel Coronavirus (2019-nCoV) Causing an Outbreak of Pneumonia
Daniel KW Chu, Yang Pan, Samuel MS Cheng, Kenrie PY Hui, Pavithra Krishnan, Yingzhi Liu, Daisy YM Ng, Carrie KC Wan, Peng Yang, Quanyi Wang, … Leo LM Poon
Clinical Chemistry (2020-04) https://doi.org/ggnbpp
DOI: 10.1093/clinchem/hvaa029 · PMID: 32031583 · PMCID: PMC7108203

252. dPCR: A Technology Review
Phenix-Lan Quan, Martin Sauzade, Eric Brouzes
Sensors (2018-04-20) https://doi.org/ggr39c
DOI: 10.3390/s18041271 · PMID: 29677144 · PMCID: PMC5948698

253. ddPCR: a more accurate tool for SARS-CoV-2 detection in low viral load specimens
Tao Suo, Xinjin Liu, Jiangpeng Feng, Ming Guo, Wenjia Hu, Dong Guo, Hafiz Ullah, Yang Yang, Qiuhan Zhang, Xin Wang, … Yu Chen
Emerging Microbes & Infections (2020-06-07) https://doi.org/ggx2t2
DOI: 10.1080/22221751.2020.1772678 · PMID: 32438868 · PMCID: PMC7448897

254. Highly accurate and sensitive diagnostic detection of SARS-CoV-2 by digital PCR
Lianhua Dong, Junbo Zhou, Chunyan Niu, Quanyi Wang, Yang Pan, Sitong Sheng, Xia Wang, Yongzhuo Zhang, Jiayi Yang, Manqing Liu, … Xiang Fang
Cold Spring Harbor Laboratory (2020-03-30) https://doi.org/ggqnqh
DOI: 10.1101/2020.03.14.20036129

255. Evaluation of COVID-19 RT-qPCR test in multi-sample pools
Idan Yelin, Noga Aharony, Einat Shaer Tamar, Amir Argoetti, Esther Messer, Dina Berenbaum, Einat Shafran, Areen Kuzli, Nagam Gandali, Tamar Hashimshony, … Roy Kishony
Cold Spring Harbor Laboratory (2020-03-27) https://doi.org/ggrn74
DOI: 10.1101/2020.03.26.20039438

256. Analytical Validation of a COVID-19 qRT-PCR Detection Assay Using a 384-well Format and Three Extraction Methods
Andrew C. Nelson, Benjamin Auch, Matthew Schomaker, Daryl M. Gohl, Patrick Grady, Darrell Johnson, Robyn Kincaid, Kylene E. Karnuth, Jerry Daniel, Jessica K. Fiege, … Sophia Yohe
Cold Spring Harbor Laboratory (2020-04-05) https://doi.org/ggs45d
DOI: 10.1101/2020.04.02.022186

257. Nucleic acid detection with CRISPR-Cas13a/C2c2
Jonathan S. Gootenberg, Omar O. Abudayyeh, Jeong Wook Lee, Patrick Essletzbichler, Aaron J. Dy, Julia Joung, Vanessa Verdine, Nina Donghia, Nichole M. Daringer, Catherine A. Freije, … Feng Zhang
Science (2017-04-28) https://doi.org/f93x8p
DOI: 10.1126/science.aam9321 · PMID: 28408723 · PMCID: PMC5526198

258. Development and Evaluation of A CRISPR-based Diagnostic For 2019-novel Coronavirus
Tieying Hou, Weiqi Zeng, Minling Yang, Wenjing Chen, Lili Ren, Jingwen Ai, Ji Wu, Yalong Liao, Xuejing Gou, Yongjun Li, … Teng Xu
Cold Spring Harbor Laboratory (2020-02-25) https://doi.org/gg7km8
DOI: 10.1101/2020.02.22.20025460

259. CRISPR-based surveillance for COVID-19 using genomically-comprehensive machine learning design
Hayden C. Metsky, Catherine A. Freije, Tinna-Solveig F. Kosoko-Thoroddsen, Pardis C. Sabeti, Cameron Myhrvold
Cold Spring Harbor Laboratory (2020-03-02) https://doi.org/ggr3zf
DOI: 10.1101/2020.02.26.967026

260. A Scalable, Easy-to-Deploy, Protocol for Cas13-Based Detection of SARS-CoV-2 Genetic Material
Jennifer N. Rauch, Eric Valois, Sabrina C. Solley, Friederike Braig, Ryan S. Lach, Morgane Audouard, Jose Carlos Ponce-Rojas, Michael S. Costello, Naomi J. Baxter, Kenneth S. Kosik, … Maxwell Z. Wilson
Cold Spring Harbor Laboratory (2020-08-29) https://doi.org/gg7km7
DOI: 10.1101/2020.04.20.052159

261. CRISPR–Cas12-based detection of SARS-CoV-2
James P. Broughton, Xianding Deng, Guixia Yu, Clare L. Fasching, Venice Servellita, Jasmeet Singh, Xin Miao, Jessica A. Streithorst, Andrea Granados, Alicia Sotomayor-Gonzalez, … Charles Y. Chiu
Nature Biotechnology (2020-04-16) https://doi.org/ggv47f
DOI: 10.1038/s41587-020-0513-4 · PMID: 32300245

262. An ultrasensitive, rapid, and portable coronavirus SARS-CoV-2 sequence detection method based on CRISPR-Cas12
Curti Lucia, Pereyra-Bonnet Federico, Gimenez Carla Alejandra
Cold Spring Harbor Laboratory (2020-03-02) https://doi.org/gg7km6
DOI: 10.1101/2020.02.29.971127

263. All-in-One Dual CRISPR-Cas12a (AIOD-CRISPR) Assay: A Case for Rapid, Ultrasensitive and Visual Detection of Novel Coronavirus SARS-CoV-2 and HIV virus
Xiong Ding, Kun Yin, Ziyue Li, Changchun Liu
Cold Spring Harbor Laboratory (2020-03-21) https://doi.org/gg7km5
DOI: 10.1101/2020.03.19.998724 · PMID: 32511323 · PMCID: PMC7239053

264. SARS-CoV-2 detection with CRISPR diagnostics
Lu Guo, Xuehan Sun, Xinge Wang, Chen Liang, Haiping Jiang, Qingqin Gao, Moyu Dai, Bin Qu, Sen Fang, Yihuan Mao, … Wei Li
Cold Spring Harbor Laboratory (2020-04-11) https://doi.org/gg7kns
DOI: 10.1101/2020.04.10.023358

265. Electric-field-driven microfluidics for rapid CRISPR-based diagnostics and its application to detection of SARS-CoV-2
Ashwin Ramachandran, Diego A. Huyke, Eesha Sharma, Malaya K. Sahoo, Niaz Banaei, Benjamin A. Pinsky, Juan G. Santiago
Cold Spring Harbor Laboratory (2020-05-22) https://doi.org/gg7knt
DOI: 10.1101/2020.05.21.109637

266. Rapid, sensitive and specific SARS coronavirus-2 detection: a multi-center comparison between standard qRT-PCR and CRISPR based DETECTR.
Eelke Brandsma, Han JMP Verhagen, Thijs J. W. van de Laar, Eric C. J. Claas, Marion Cornelissen, Emile van den Akker
Cold Spring Harbor Laboratory (2020-07-29) https://doi.org/gg7km4
DOI: 10.1101/2020.07.27.20147249

267. Coronavirus and the race to distribute reliable diagnostics
Cormac Sheridan
Nature Biotechnology (2020-02-19) https://doi.org/ggm4nt
DOI: 10.1038/d41587-020-00002-2 · PMID: 32265548

268. The standard coronavirus test, if available, works well—but can new diagnostics help in this pandemic?
Robert Service
Science (2020-03-22) https://doi.org/ggq9wm
DOI: 10.1126/science.abb8400

269. Laboratory Diagnosis of COVID-19: Current Issues and Challenges
Yi-Wei Tang, Jonathan E. Schmitz, David H. Persing, Charles W. Stratton
Journal of Clinical Microbiology (2020-05-26) https://doi.org/ggq7h8
DOI: 10.1128/jcm.00512-20 · PMID: 32245835 · PMCID: PMC7269383

270. Negative Nasopharyngeal and Oropharyngeal Swabs Do Not Rule Out COVID-19
Poramed Winichakoon, Romanee Chaiwarith, Chalerm Liwsrisakun, Parichat Salee, Aree Goonna, Atikun Limsukon, Quanhathai Kaewpoowat
Journal of Clinical Microbiology (2020-04-23) https://doi.org/ggpw9m
DOI: 10.1128/jcm.00297-20 · PMID: 32102856 · PMCID: PMC7180262

271. A systematic review of antibody mediated immunity to coronaviruses: antibody kinetics, correlates of protection, and association of antibody responses with severity of disease
Angkana T. Huang, Bernardo Garcia-Carreras, Matt D. T. Hitchings, Bingyi Yang, Leah Katzelnick, Susan M Rattigan, Brooke Borgert, Carlos Moreno, Benjamin D. Solomon, Isabel Rodriguez-Barraquer, … Derek A. T. Cummings
Cold Spring Harbor Laboratory (2020-04-17) https://doi.org/ggsfmz
DOI: 10.1101/2020.04.14.20065771 · PMID: 32511434 · PMCID: PMC7217088

272. Longitudinal profile of antibodies against SARS-coronavirus in SARS patients and their clinical significance
Hongying MO, Guangqiao ZENG, Xiaolan REN, Hui LI, Changwen KE, Yaxia TAN, Chaoda CAI, Kefang LAI, Rongchang CHEN, Moira CHAN-YEUNG, Nanshan ZHONG
Respirology (2006-01) https://doi.org/dn23vj
DOI: 10.1111/j.1440-1843.2006.00783.x · PMID: 16423201 · PMCID: PMC7192223

273. Two‐Year Prospective Study of the Humoral Immune Response of Patients with Severe Acute Respiratory Syndrome
Wei Liu, Arnaud Fontanet, Pan‐He Zhang, Lin Zhan, Zhong‐Tao Xin, Laurence Baril, Fang Tang, Hui Lv, Wu‐Chun Cao
The Journal of Infectious Diseases (2006-03-15) https://doi.org/cmzn2k
DOI: 10.1086/500469 · PMID: 16479513 · PMCID: PMC7109932

274. The time course of the immune response to experimental coronavirus infection of man
K. A. Callow, H. F. Parry, M. Sergeant, D. A. J. Tyrrell
Epidemiology and Infection (2009-05-15) https://doi.org/c9pnmg
DOI: 10.1017/s0950268800048019 · PMID: 2170159 · PMCID: PMC2271881

275. Loss of Bcl-6-Expressing T Follicular Helper Cells and Germinal Centers in COVID-19
Naoki Kaneko, Hsiao-Hsuan Kuo, Julie Boucau, Jocelyn R. Farmer, Hugues Allard-Chamard, Vinay S. Mahajan, Alicja Piechocka-Trocha, Kristina Lefteri, Matthew Osborn, Julia Bals, … Shiv Pillai
Cell (2020-10) https://doi.org/gg9rdv
DOI: 10.1016/j.cell.2020.08.025 · PMID: 32877699 · PMCID: PMC7437499

276. A Peptide-Based Magnetic Chemiluminescence Enzyme Immunoassay for Serological Diagnosis of Coronavirus Disease 2019
Xue-fei Cai, Juan Chen, Jie- li Hu, Quan-xin Long, Hai-jun Deng, Ping Liu, Kai Fan, Pu Liao, Bei-zhong Liu, Gui-cheng Wu, … De-qiang Wang
The Journal of Infectious Diseases (2020-07-15) https://doi.org/ggv2fx
DOI: 10.1093/infdis/jiaa243 · PMID: 32382737 · PMCID: PMC7239108

277. Cellex qSARS-CoV-2 IgG/IgM Rapid Test
Cellex
(2020-04-07) https://www.fda.gov/media/136625/download

278. Detection of antibodies against SARS‐CoV‐2 in patients with COVID‐19
Zhe Du, Fengxue Zhu, Fuzheng Guo, Bo Yang, Tianbing Wang
Journal of Medical Virology (2020-04-10) https://doi.org/ggq7m2
DOI: 10.1002/jmv.25820 · PMID: 32243608 · PMCID: PMC7228383

279. A serological assay to detect SARS-CoV-2 seroconversion in humans
Fatima Amanat, Daniel Stadlbauer, Shirin Strohmeier, Thi H. O. Nguyen, Veronika Chromikova, Meagan McMahon, Kaijun Jiang, Guha Asthagiri Arunkumar, Denise Jurczyszak, Jose Polanco, … Florian Krammer
Cold Spring Harbor Laboratory (2020-04-16) https://doi.org/ggpn83
DOI: 10.1101/2020.03.17.20037713 · PMID: 32511441 · PMCID: PMC7239062

280. Coronavirus testing is ramping up. Here are the new tests and how they work.
Stephanie Pappas-Live Science Contributor 31 March 2020
livescience.com https://www.livescience.com/coronavirus-tests-available.html

281. Strong associations and moderate predictive value of early symptoms for SARS-CoV-2 test positivity among healthcare workers, the Netherlands, March 2020
Alma Tostmann, John Bradley, Teun Bousema, Wing-Kee Yiek, Minke Holwerda, Chantal Bleeker-Rovers, Jaap ten Oever, Corianne Meijer, Janette Rahamat-Langendoen, Joost Hopman, … Heiman Wertheim
Eurosurveillance (2020-04-23) https://doi.org/ggthwx
DOI: 10.2807/1560-7917.es.2020.25.16.2000508 · PMID: 32347200 · PMCID: PMC7189649

282. Correlation of Chest CT and RT-PCR Testing for Coronavirus Disease 2019 (COVID-19) in China: A Report of 1014 Cases
Tao Ai, Zhenlu Yang, Hongyan Hou, Chenao Zhan, Chong Chen, Wenzhi Lv, Qian Tao, Ziyong Sun, Liming Xia
Radiology (2020-08) https://doi.org/ggmw6p
DOI: 10.1148/radiol.2020200642 · PMID: 32101510 · PMCID: PMC7233399

283. Performance of Radiologists in Differentiating COVID-19 from Non-COVID-19 Viral Pneumonia at Chest CT
Harrison X. Bai, Ben Hsieh, Zeng Xiong, Kasey Halsey, Ji Whae Choi, Thi My Linh Tran, Ian Pan, Lin-Bo Shi, Dong-Cui Wang, Ji Mei, … Wei-Hua Liao
Radiology (2020-08) https://doi.org/ggnqw4
DOI: 10.1148/radiol.2020200823 · PMID: 32155105 · PMCID: PMC7233414

284. Covid-19: automatic detection from X-ray images utilizing transfer learning with convolutional neural networks
Ioannis D. Apostolopoulos, Tzani A. Mpesiana
Physical and Engineering Sciences in Medicine (2020-04-03) https://doi.org/ggs448
DOI: 10.1007/s13246-020-00865-4 · PMID: 32524445 · PMCID: PMC7118364

285. Healthcare Workers
CDC
Centers for Disease Control and Prevention (2020-02-11) https://www.cdc.gov/coronavirus/2019-ncov/hcp/testing-overview.html

286. NY Forward: a guide to reopening New York & building back better (2020-05-15) https://www.governor.ny.gov/sites/governor.ny.gov/files/atoms/files/NYForwardReopeningGuide.pdf

287. Interpreting Diagnostic Tests for SARS-CoV-2
Nandini Sethuraman, Sundararaj Stanleyraj Jeremiah, Akihide Ryo
JAMA (2020-06-09) https://doi.org/ggt6cw
DOI: 10.1001/jama.2020.8259 · PMID: 32374370

288. Genomic evidence for reinfection with SARS-CoV-2: a case study
Richard L Tillett, Joel R Sevinsky, Paul D Hartley, Heather Kerwin, Natalie Crawford, Andrew Gorzalski, Chris Laverdure, Subhash C Verma, Cyprian C Rossetto, David Jackson, … Mark Pandori
The Lancet Infectious Diseases (2021-01) https://doi.org/ghfgkt
DOI: 10.1016/s1473-3099(20)30764-7 · PMID: 33058797 · PMCID: PMC7550103

289. Neutralizing Monoclonal Antibodies as Promising Therapeutics against Middle East Respiratory Syndrome Coronavirus Infection
Hui-Ju Han, Jian-Wei Liu, Hao Yu, Xue-Jie Yu
Viruses (2018-11-30) https://doi.org/ggp87v
DOI: 10.3390/v10120680 · PMID: 30513619 · PMCID: PMC6315345

290. COVID-19 Data Repository
Center for Systems Science and Engineering at Johns Hopkins University
GitHub https://github.com/CSSEGISandData/COVID-19/tree/master/csse_covid_19_data/csse_covid_19_time_series

291. Isolation of a Novel Coronavirus from a Man with Pneumonia in Saudi Arabia
Ali M. Zaki, Sander van Boheemen, Theo M. Bestebroer, Albert D. M. E. Osterhaus, Ron A. M. Fouchier
New England Journal of Medicine (2012-11-08) https://doi.org/f4czx5
DOI: 10.1056/nejmoa1211721 · PMID: 23075143

292. Causes of Death and Comorbidities in Patients with COVID-19
Sefer Elezkurtaj, Selina Greuel, Jana Ihlow, Edward Michaelis, Philip Bischoff, Catarina Alisa Kunze, Bruno Valentin Sinn, Manuela Gerhold, Kathrin Hauptmann, Barbara Ingold-Heppner, … David Horst
Cold Spring Harbor Laboratory (2020-06-17) https://doi.org/gg926j
DOI: 10.1101/2020.06.15.20131540

293. Evidence-Based Medicine Data Lab COVID-19 TrialsTracker
Nick DeVito, Peter Inglesby
GitHub (2020-03-29) https://github.com/ebmdatalab/covid_trials_tracker-covid
DOI: 10.5281/zenodo.3732709

294. Small Molecules vs Biologics | Drug Development Differences
Nuventra Pharma Sciences2525 Meridian Parkway, Suite 200 Durham
PK / PD and Clinical Pharmacology Consultants (2020-05-13) https://www.nuventra.com/resources/blog/small-molecules-versus-biologics/

295. Drug Discovery: A Historical Perspective
J. Drews
Science (2000-03-17) https://doi.org/d6bvp7
DOI: 10.1126/science.287.5460.1960 · PMID: 10720314

296. Introduction to modern virology
N. J. Dimmock, A. J. Easton, K. N. Leppard
Blackwell Pub (2007)
ISBN: 9781405136457

297. Coronaviruses
Helena Jane Maier, Erica Bickerton, Paul Britton (editors)
Methods in Molecular Biology (2015) https://doi.org/ggqfqx
DOI: 10.1007/978-1-4939-2438-7 · PMID: 25870870 · ISBN: 9781493924370

298. The potential chemical structure of anti‐SARS‐CoV‐2 RNA‐dependent RNA polymerase
Jrhau Lung, Yu‐Shih Lin, Yao‐Hsu Yang, Yu‐Lun Chou, Li‐Hsin Shu, Yu‐Ching Cheng, Hung Te Liu, Ching‐Yuan Wu
Journal of Medical Virology (2020-03-18) https://doi.org/ggp6fm
DOI: 10.1002/jmv.25761 · PMID: 32167173 · PMCID: PMC7228302

299. Broad-spectrum coronavirus antiviral drug discovery
Allison L. Totura, Sina Bavari
Expert Opinion on Drug Discovery (2019-03-08) https://doi.org/gg74z5
DOI: 10.1080/17460441.2019.1581171 · PMID: 30849247 · PMCID: PMC7103675

300. Favipiravir
DrugBank
(2020-06-12) https://www.drugbank.ca/drugs/DB12466

301. In Vitro and In Vivo Activities of Anti-Influenza Virus Compound T-705
Y. Furuta, K. Takahashi, Y. Fukuda, M. Kuno, T. Kamiyama, K. Kozaki, N. Nomura, H. Egawa, S. Minami, Y. Watanabe, … K. Shiraki
Antimicrobial Agents and Chemotherapy (2002-04) https://doi.org/cndw7n
DOI: 10.1128/aac.46.4.977-981.2002 · PMID: 11897578 · PMCID: PMC127093

302. Efficacy of Orally Administered T-705 on Lethal Avian Influenza A (H5N1) Virus Infections in Mice
Robert W. Sidwell, Dale L. Barnard, Craig W. Day, Donald F. Smee, Kevin W. Bailey, Min-Hui Wong, John D. Morrey, Yousuke Furuta
Antimicrobial Agents and Chemotherapy (2007-03) https://doi.org/dm9xr2
DOI: 10.1128/aac.01051-06 · PMID: 17194832 · PMCID: PMC1803113

303. Favipiravir (T-705), a broad spectrum inhibitor of viral RNA polymerase
Yousuke FURUTA, Takashi KOMENO, Takaaki NAKAMURA
Proceedings of the Japan Academy, Series B (2017) https://doi.org/gbxcxw
DOI: 10.2183/pjab.93.027 · PMID: 28769016 · PMCID: PMC5713175

304. Mechanism of Action of T-705 against Influenza Virus
Yousuke Furuta, Kazumi Takahashi, Masako Kuno-Maekawa, Hidehiro Sangawa, Sayuri Uehara, Kyo Kozaki, Nobuhiko Nomura, Hiroyuki Egawa, Kimiyasu Shiraki
Antimicrobial Agents and Chemotherapy (2005-03) https://doi.org/dgbwdh
DOI: 10.1128/aac.49.3.981-986.2005 · PMID: 15728892 · PMCID: PMC549233

305. Activity of T-705 in a Hamster Model of Yellow Fever Virus Infection in Comparison with That of a Chemically Related Compound, T-1106
Justin G. Julander, Kristiina Shafer, Donald F. Smee, John D. Morrey, Yousuke Furuta
Antimicrobial Agents and Chemotherapy (2009-01) https://doi.org/brknds
DOI: 10.1128/aac.01074-08 · PMID: 18955536 · PMCID: PMC2612161

306. In Vitro and In Vivo Activities of T-705 against Arenavirus and Bunyavirus Infections
Brian B. Gowen, Min-Hui Wong, Kie-Hoon Jung, Andrew B. Sanders, Michelle Mendenhall, Kevin W. Bailey, Yousuke Furuta, Robert W. Sidwell
Antimicrobial Agents and Chemotherapy (2007-09) https://doi.org/d98c87
DOI: 10.1128/aac.00356-07 · PMID: 17606691 · PMCID: PMC2043187

307. Favipiravir (T-705) inhibits in vitro norovirus replication
J. Rocha-Pereira, D. Jochmans, K. Dallmeier, P. Leyssen, M. S. J. Nascimento, J. Neyts
Biochemical and Biophysical Research Communications (2012-08) https://doi.org/f369j7
DOI: 10.1016/j.bbrc.2012.07.034 · PMID: 22809499

308. T-705 (Favipiravir) Inhibition of Arenavirus Replication in Cell Culture
Michelle Mendenhall, Andrew Russell, Terry Juelich, Emily L. Messina, Donald F. Smee, Alexander N. Freiberg, Michael R. Holbrook, Yousuke Furuta, Juan-Carlos de la Torre, Jack H. Nunberg, Brian B. Gowen
Antimicrobial Agents and Chemotherapy (2011-02) https://doi.org/cppwsc
DOI: 10.1128/aac.01219-10 · PMID: 21115797 · PMCID: PMC3028760

309. The evolution of nucleoside analogue antivirals: A review for chemists and non-chemists. Part 1: Early structural modifications to the nucleoside scaffold
Katherine L. Seley-Radtke, Mary K. Yates
Antiviral Research (2018-06) https://doi.org/gdpn35
DOI: 10.1016/j.antiviral.2018.04.004 · PMID: 29649496 · PMCID: PMC6396324

310. The Ambiguous Base-Pairing and High Substrate Efficiency of T-705 (Favipiravir) Ribofuranosyl 5′-Triphosphate towards Influenza A Virus Polymerase
Zhinan Jin, Lucas K. Smith, Vivek K. Rajwanshi, Baek Kim, Jerome Deval
PLoS ONE (2013-07-10) https://doi.org/f5br92
DOI: 10.1371/journal.pone.0068347 · PMID: 23874596 · PMCID: PMC3707847

311. Experimental Treatment with Favipiravir for COVID-19: An Open-Label Control Study
Qingxian Cai, Minghui Yang, Dongjing Liu, Jun Chen, Dan Shu, Junxia Xia, Xuejiao Liao, Yuanbo Gu, Qiue Cai, Yang Yang, … Lei Liu
Engineering (2020-10) https://doi.org/ggpprd
DOI: 10.1016/j.eng.2020.03.007 · PMID: 32346491 · PMCID: PMC7185795

312. Lopinavir–ritonavir in patients admitted to hospital with COVID-19 (RECOVERY): a randomised, controlled, open-label, platform trial
Peter W Horby, Marion Mafham, Jennifer L Bell, Louise Linsell, Natalie Staplin, Jonathan Emberson, Adrian Palfreeman, Jason Raw, Einas Elmahi, Benjamin Prudon, … Martin J Landray
The Lancet (2020-10) https://doi.org/fnx2
DOI: 10.1016/s0140-6736(20)32013-4 · PMID: 33031764 · PMCID: PMC7535623

313. A Large, Simple Trial Leading to Complex Questions
David P. Harrington, Lindsey R. Baden, Joseph W. Hogan
New England Journal of Medicine (2020-12-02) https://doi.org/ghnhnx
DOI: 10.1056/nejme2034294 · PMID: 33264557 · PMCID: PMC7727323

314. Remdesivir EUA Letter of Authorization
Denise M Hinton
(2020-05-01) https://www.fda.gov/media/137564/download

315. A Phase 3 Randomized Study to Evaluate the Safety and Antiviral Activity of Remdesivir (GS-5734™) in Participants With Severe COVID-19
Gilead Sciences
clinicaltrials.gov (2020-12-15) https://clinicaltrials.gov/ct2/show/NCT04292899

316. A Multicenter, Adaptive, Randomized Blinded Controlled Trial of the Safety and Efficacy of Investigational Therapeutics for the Treatment of COVID-19 in Hospitalized Adults
National Institute of Allergy and Infectious Diseases (NIAID)
clinicaltrials.gov (2020-12-05) https://clinicaltrials.gov/ct2/show/NCT04280705

317. Remdesivir for the Treatment of Covid-19 — Final Report
John H. Beigel, Kay M. Tomashek, Lori E. Dodd, Aneesh K. Mehta, Barry S. Zingman, Andre C. Kalil, Elizabeth Hohmann, Helen Y. Chu, Annie Luetkemeyer, Susan Kline, … H. Clifford Lane
New England Journal of Medicine (2020-11-05) https://doi.org/dwkd
DOI: 10.1056/nejmoa2007764 · PMID: 32445440 · PMCID: PMC7262788

318. Compassionate Use of Remdesivir for Patients with Severe Covid-19
Jonathan Grein, Norio Ohmagari, Daniel Shin, George Diaz, Erika Asperges, Antonella Castagna, Torsten Feldt, Gary Green, Margaret L. Green, François-Xavier Lescure, … Timothy Flanigan
New England Journal of Medicine (2020-06-11) https://doi.org/ggrm99
DOI: 10.1056/nejmoa2007016 · PMID: 32275812 · PMCID: PMC7169476

319. The antiviral compound remdesivir potently inhibits RNA-dependent RNA polymerase from Middle East respiratory syndrome coronavirus
Calvin J. Gordon, Egor P. Tchesnokov, Joy Y. Feng, Danielle P. Porter, Matthias Götte
Journal of Biological Chemistry (2020-04-10) https://doi.org/ggqm6x
DOI: 10.1074/jbc.ac120.013056 · PMID: 32094225 · PMCID: PMC7152756

320. Coronavirus Susceptibility to the Antiviral Remdesivir (GS-5734) Is Mediated by the Viral Polymerase and the Proofreading Exoribonuclease
Maria L. Agostini, Erica L. Andres, Amy C. Sims, Rachel L. Graham, Timothy P. Sheahan, Xiaotao Lu, Everett Clinton Smith, James Brett Case, Joy Y. Feng, Robert Jordan, … Mark R. Denison
mBio (2018-03-06) https://doi.org/gc45v6
DOI: 10.1128/mbio.00221-18 · PMID: 29511076 · PMCID: PMC5844999

321. Remdesivir and chloroquine effectively inhibit the recently emerged novel coronavirus (2019-nCoV) in vitro
Manli Wang, Ruiyuan Cao, Leike Zhang, Xinglou Yang, Jia Liu, Mingyue Xu, Zhengli Shi, Zhihong Hu, Wu Zhong, Gengfu Xiao
Cell Research (2020-02-04) https://doi.org/ggkbsg
DOI: 10.1038/s41422-020-0282-0 · PMID: 32020029 · PMCID: PMC7054408

322. A Randomized, Controlled Trial of Ebola Virus Disease Therapeutics
Sabue Mulangu, Lori E. Dodd, Richard T. Davey, Olivier Tshiani Mbaya, Michael Proschan, Daniel Mukadi, Mariano Lusakibanza Manzo, Didier Nzolo, Antoine Tshomba Oloma, Augustin Ibanda, … the PALM Writing Group
New England Journal of Medicine (2019-12-12) https://doi.org/ggqmx4
DOI: 10.1056/nejmoa1910993 · PMID: 31774950

323. Broad-spectrum antiviral GS-5734 inhibits both epidemic and zoonotic coronaviruses
Timothy P. Sheahan, Amy C. Sims, Rachel L. Graham, Vineet D. Menachery, Lisa E. Gralinski, James B. Case, Sarah R. Leist, Krzysztof Pyrc, Joy Y. Feng, Iva Trantcheva, … Ralph S. Baric
Science Translational Medicine (2017-06-28) https://doi.org/gc3grb
DOI: 10.1126/scitranslmed.aal3653 · PMID: 28659436 · PMCID: PMC5567817

324. Did an experimental drug help a U.S. coronavirus patient?
Jon Cohen
Science (2020-03-13) https://doi.org/ggqm62
DOI: 10.1126/science.abb7243

325. First 12 patients with coronavirus disease 2019 (COVID-19) in the United States
Stephanie A. Kujawski, Karen K Wong, Jennifer P. Collins, Lauren Epstein, Marie E. Killerby, Claire M. Midgley, Glen R. Abedi, N. Seema Ahmed, Olivia Almendares, Francisco N. Alvarez, … The COVID-19 Investigation Team
Cold Spring Harbor Laboratory (2020-03-12) https://doi.org/ggqm6z
DOI: 10.1101/2020.03.09.20032896

326. First Case of 2019 Novel Coronavirus in the United States
Michelle L. Holshue, Chas DeBolt, Scott Lindquist, Kathy H. Lofy, John Wiesman, Hollianne Bruce, Christopher Spitters, Keith Ericson, Sara Wilkerson, Ahmet Tural, … Satish K. Pillai
New England Journal of Medicine (2020-03-05) https://doi.org/ggjvr6
DOI: 10.1056/nejmoa2001191 · PMID: 32004427 · PMCID: PMC7092802

327. Remdesivir for 5 or 10 Days in Patients with Severe Covid-19
Jason D. Goldman, David C. B. Lye, David S. Hui, Kristen M. Marks, Raffaele Bruno, Rocio Montejano, Christoph D. Spinner, Massimo Galli, Mi-Young Ahn, Ronald G. Nahass, … Aruna Subramanian
New England Journal of Medicine (2020-11-05) https://doi.org/ggz7qv
DOI: 10.1056/nejmoa2015301 · PMID: 32459919 · PMCID: PMC7377062

328. A Phase 3 Randomized Study to Evaluate the Safety and Antiviral Activity of Remdesivir (GS-5734™) in Participants With Moderate COVID-19 Compared to Standard of Care Treatment
Gilead Sciences
clinicaltrials.gov (2020-07-23) https://clinicaltrials.gov/ct2/show/NCT04292730

329. Multi-centre, adaptive, randomized trial of the safety and efficacy of treatments of COVID-19 in hospitalized adults
EU Clinical Trials Register
(2020-03-09) https://www.clinicaltrialsregister.eu/ctr-search/trial/2020-000936-23/FR

330. A Trial of Remdesivir in Adults With Mild and Moderate COVID-19 - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04252664

331. A Phase 3 Randomized, Double-blind, Placebo-controlled, Multicenter Study to Evaluate the Efficacy and Safety of Remdesivir in Hospitalized Adult Patients With Severe COVID-19.
Bin Cao
clinicaltrials.gov (2020-04-13) https://clinicaltrials.gov/ct2/show/NCT04257656

332. FDA Approves First Treatment for COVID-19
Office of the Commissioner
FDA (2020-10-22) https://www.fda.gov/news-events/press-announcements/fda-approves-first-treatment-covid-19

333. Gilead Sciences Statement on the Solidarity Trial https://www.gilead.com/news-and-press/company-statements/gilead-sciences-statement-on-the-solidarity-trial

334. Conflicting results on the efficacy of remdesivir in hospitalized Covid-19 patients: comment on the Adaptive Covid-19 Treatment Trial
Leonarda Galiuto, Carlo Patrono
European Heart Journal (2020-12-07) https://doi.org/ghp4kw
DOI: 10.1093/eurheartj/ehaa934 · PMID: 33306101

335. The “very, very bad look” of remdesivir, the first FDA-approved COVID-19 drug
Jon Cohen, Kai KupferschmidtOct. 28, 2020, 7:05 Pm
Science | AAAS (2020-10-28) https://www.sciencemag.org/news/2020/10/very-very-bad-look-remdesivir-first-fda-approved-covid-19-drug

336. Effect of Remdesivir vs Standard Care on Clinical Status at 11 Days in Patients With Moderate COVID-19
Christoph D. Spinner, Robert L. Gottlieb, Gerard J. Criner, José Ramón Arribas López, Anna Maria Cattelan, Alex Soriano Viladomiu, Onyema Ogbuagu, Prashant Malhotra, Kathleen M. Mullane, Antonella Castagna, … for the GS-US-540-5774 Investigators
JAMA (2020-09-15) https://doi.org/ghhz6g
DOI: 10.1001/jama.2020.16349 · PMID: 32821939 · PMCID: PMC7442954

337. Baricitinib plus Remdesivir for Hospitalized Adults with Covid-19
Andre C. Kalil, Thomas F. Patterson, Aneesh K. Mehta, Kay M. Tomashek, Cameron R. Wolfe, Varduhi Ghazaryan, Vincent C. Marconi, Guillermo M. Ruiz-Palacios, Lanny Hsieh, Susan Kline, … John H. Beigel
New England Journal of Medicine (2020-12-11) https://doi.org/ghpbd2
DOI: 10.1056/nejmoa2031994 · PMID: 33306283 · PMCID: PMC7745180

338. Proteases Essential for Human Influenza Virus Entry into Cells and Their Inhibitors as Potential Therapeutic Agents
Hiroshi Kido, Yuushi Okumura, Hiroshi Yamada, Trong Quang Le, Mihiro Yano
Current Pharmaceutical Design (2007-02-01) https://doi.org/bts3xp
DOI: 10.2174/138161207780162971 · PMID: 17311557

339. Protease inhibitors targeting coronavirus and filovirus entry
Yanchen Zhou, Punitha Vedantham, Kai Lu, Juliet Agudelo, Ricardo Carrion, Jerritt W. Nunneley, Dale Barnard, Stefan Pöhlmann, James H. McKerrow, Adam R. Renslo, Graham Simmons
Antiviral Research (2015-04) https://doi.org/ggr984
DOI: 10.1016/j.antiviral.2015.01.011 · PMID: 25666761 · PMCID: PMC4774534

340. Structure of Mpro from SARS-CoV-2 and discovery of its inhibitors
Zhenming Jin, Xiaoyu Du, Yechun Xu, Yongqiang Deng, Meiqin Liu, Yao Zhao, Bing Zhang, Xiaofeng Li, Leike Zhang, Chao Peng, … Haitao Yang
Nature (2020-04-09) https://doi.org/ggrp42
DOI: 10.1038/s41586-020-2223-y · PMID: 32272481

341. Design of Wide-Spectrum Inhibitors Targeting Coronavirus Main Proteases
Haitao Yang, Weiqing Xie, Xiaoyu Xue, Kailin Yang, Jing Ma, Wenxue Liang, Qi Zhao, Zhe Zhou, Duanqing Pei, John Ziebuhr, … Zihe Rao
PLoS Biology (2005-09-06) https://doi.org/bcm9k7
DOI: 10.1371/journal.pbio.0030324 · PMID: 16128623 · PMCID: PMC1197287

342. The newly emerged SARS-Like coronavirus HCoV-EMC also has an “Achilles’ heel”: current effective inhibitor targeting a 3C-like protease
Zhilin Ren, Liming Yan, Ning Zhang, Yu Guo, Cheng Yang, Zhiyong Lou, Zihe Rao
Protein & Cell (2013-04-03) https://doi.org/ggr7vh
DOI: 10.1007/s13238-013-2841-3 · PMID: 23549610 · PMCID: PMC4875521

343. Crystal structure of SARS-CoV-2 main protease provides a basis for design of improved α-ketoamide inhibitors
Linlin Zhang, Daizong Lin, Xinyuanyuan Sun, Ute Curth, Christian Drosten, Lucie Sauerhering, Stephan Becker, Katharina Rox, Rolf Hilgenfeld
Science (2020-04-24) https://doi.org/ggp9sb
DOI: 10.1126/science.abb3405 · PMID: 32198291 · PMCID: PMC7164518

344. Ebselen, a promising antioxidant drug: mechanisms of action and targets of biological pathways
Gajendra Kumar Azad, Raghuvir S. Tomar
Molecular Biology Reports (2014-05-28) https://doi.org/f6cnq3
DOI: 10.1007/s11033-014-3417-x · PMID: 24867080

345. Safety and efficacy of ebselen for the prevention of noise-induced hearing loss: a randomised, double-blind, placebo-controlled, phase 2 trial
Jonathan Kil, Edward Lobarinas, Christopher Spankovich, Scott K Griffiths, Patrick J Antonelli, Eric D Lynch, Colleen G Le Prell
The Lancet (2017-09) https://doi.org/gbwnbv
DOI: 10.1016/s0140-6736(17)31791-9

346. Potential therapeutic use of ebselen for COVID-19 and other respiratory viral infections
Helmut Sies, Michael J. Parnham
Free Radical Biology and Medicine (2020-08) https://doi.org/ghdx7s
DOI: 10.1016/j.freeradbiomed.2020.06.032 · PMID: 32598985 · PMCID: PMC7319625

347. Selenium Deficiency Is Associated with Mortality Risk from COVID-19
Arash Moghaddam, Raban Arved Heller, Qian Sun, Julian Seelig, Asan Cherkezov, Linda Seibert, Julian Hackler, Petra Seemann, Joachim Diegmann, Maximilian Pilz, … Lutz Schomburg
Nutrients (2020-07-16) https://doi.org/gg5kbc
DOI: 10.3390/nu12072098 · PMID: 32708526 · PMCID: PMC7400921

348. Selenium and viral infection: are there lessons for COVID-19?
Giovanna Bermano, Catherine Méplan, Derry K. Mercer, John E. Hesketh
British Journal of Nutrition (2020-08-06) https://doi.org/ghdx7w
DOI: 10.1017/s0007114520003128 · PMID: 32758306 · PMCID: PMC7503044

349. Selenium and selenoproteins in viral infection with potential relevance to COVID-19
Jinsong Zhang, Ramy Saad, Ethan Will Taylor, Margaret P. Rayman
Redox Biology (2020-10) https://doi.org/ghdx7t
DOI: 10.1016/j.redox.2020.101715 · PMID: 32992282 · PMCID: PMC7481318

350. Target discovery of ebselen with a biotinylated probe
Zhenzhen Chen, Zhongyao Jiang, Nan Chen, Qian Shi, Lili Tong, Fanpeng Kong, Xiufen Cheng, Hao Chen, Chu Wang, Bo Tang
Chemical Communications (2018) https://doi.org/ggrtcm
DOI: 10.1039/c8cc04258f · PMID: 30091742

351. Curing a viral infection by targeting the host: The example of cyclophilin inhibitors
Kai Lin, Philippe Gallay
Antiviral Research (2013-07) https://doi.org/f4237c
DOI: 10.1016/j.antiviral.2013.03.020 · PMID: 23578729 · PMCID: PMC4332838

352. Lisinopril - Drug Usage Statistics
ClinCalc DrugStats Database
https://clincalc.com/DrugStats/Drugs/Lisinopril

353. Hypertension Hot Potato — Anatomy of the Angiotensin-Receptor Blocker Recalls
J. Brian Byrd, Glenn M. Chertow, Vivek Bhalla
New England Journal of Medicine (2019-04-25) https://doi.org/ggvc7g
DOI: 10.1056/nejmp1901657 · PMID: 30865819 · PMCID: PMC7066505

354. Hypertension, the renin–angiotensin system, and the risk of lower respiratory tract infections and lung injury: implications for COVID-19
Reinhold Kreutz, Engi Abd El-Hady Algharably, Michel Azizi, Piotr Dobrowolski, Tomasz Guzik, Andrzej Januszewicz, Alexandre Persu, Aleksander Prejbisz, Thomas Günther Riemer, Ji-Guang Wang, Michel Burnier
Cardiovascular Research (2020-08-01) https://doi.org/ggtwpj
DOI: 10.1093/cvr/cvaa097 · PMID: 32293003 · PMCID: PMC7184480

355. Angiotensin converting enzyme 2 activity and human atrial fibrillation: increased plasma angiotensin converting enzyme 2 activity is associated with atrial fibrillation and more advanced left atrial structural remodelling
Tomos E. Walters, Jonathan M. Kalman, Sheila K. Patel, Megan Mearns, Elena Velkoska, Louise M. Burrell
Europace (2016-10-12) https://doi.org/gbt2jw
DOI: 10.1093/europace/euw246 · PMID: 27738071

356. Cardiovascular Disease, Drug Therapy, and Mortality in Covid-19
Mandeep R. Mehra, Sapan S. Desai, SreyRam Kuy, Timothy D. Henry, Amit N. Patel
New England Journal of Medicine (2020-06-18) https://doi.org/ggtp6v
DOI: 10.1056/nejmoa2007621 · PMID: 32356626 · PMCID: PMC7206931

357. Response by Cohen et al to Letter Regarding Article, “Association of Inpatient Use of Angiotensin-Converting Enzyme Inhibitors and Angiotensin II Receptor Blockers With Mortality Among Patients With Hypertension Hospitalized With COVID-19”
Jordana B. Cohen, Thomas C. Hanff, Andrew M. South, Matthew A. Sparks, Swapnil Hiremath, Adam P. Bress, J. Brian Byrd, Julio A. Chirinos
Circulation Research (2020-06-05) https://doi.org/gg3xsg
DOI: 10.1161/circresaha.120.317205 · PMID: 32496917 · PMCID: PMC7265880

358. Sound Science before Quick Judgement Regarding RAS Blockade in COVID-19
Matthew A. Sparks, Andrew South, Paul Welling, J. Matt Luther, Jordana Cohen, James Brian Byrd, Louise M. Burrell, Daniel Batlle, Laurie Tomlinson, Vivek Bhalla, … Swapnil Hiremath
Clinical Journal of the American Society of Nephrology (2020-05-07) https://doi.org/ggq8gn
DOI: 10.2215/cjn.03530320 · PMID: 32220930 · PMCID: PMC7269218

359. Elimination or Prolongation of ACE Inhibitors and ARB in Coronavirus Disease 2019 - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04338009

360. Stopping ACE-inhibitors in COVID-19 - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04353596

361. Losartan for Patients With COVID-19 Not Requiring Hospitalization - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04311177

362. Losartan for Patients With COVID-19 Requiring Hospitalization - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04312009

363. The CORONAvirus Disease 2019 Angiotensin Converting Enzyme Inhibitor/Angiotensin Receptor Blocker InvestigatiON (CORONACION) Randomized Clinical Trial
Prof John William McEvoy
clinicaltrials.gov (2020-06-26) https://clinicaltrials.gov/ct2/show/NCT04330300

364. Ramipril for the Treatment of COVID-19 - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04366050

365. The Coronavirus Conundrum: ACE2 and Hypertension Edition
Matthew Sparks, Swapnil Hiremath
NephJC http://www.nephjc.com/news/covidace2

366. Lysosomotropic agents as HCV entry inhibitors
Usman A Ashfaq, Tariq Javed, Sidra Rehman, Zafar Nawaz, Sheikh Riazuddin
Virology Journal (2011-04-12) https://doi.org/dr5g4m
DOI: 10.1186/1743-422x-8-163 · PMID: 21481279 · PMCID: PMC3090357

367. New concepts in antimalarial use and mode of action in dermatology
Sunil Kalia, Jan P Dutz
Dermatologic Therapy (2007-07) https://doi.org/fv69cb
DOI: 10.1111/j.1529-8019.2007.00131.x · PMID: 17970883 · PMCID: PMC7163426

368. Chloroquine is a potent inhibitor of SARS coronavirus infection and spread
Martin J Vincent, Eric Bergeron, Suzanne Benjannet, Bobbie R Erickson, Pierre E Rollin, Thomas G Ksiazek, Nabil G Seidah, Stuart T Nichol
Virology Journal (2005) https://doi.org/dvbds4
DOI: 10.1186/1743-422x-2-69 · PMID: 16115318 · PMCID: PMC1232869

369. 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)
Xueting Yao, Fei Ye, Miao Zhang, Cheng Cui, Baoying Huang, Peihua Niu, Xu Liu, Li Zhao, Erdan Dong, Chunli Song, … Dongyang Liu
Clinical Infectious Diseases (2020-08-01) https://doi.org/ggpx7z
DOI: 10.1093/cid/ciaa237 · PMID: 32150618 · PMCID: PMC7108130

370. Mechanism of Action of Hydroxychloroquine in the Antiphospholipid Syndrome
Nadine Müller-Calleja, Davit Manukyan, Wolfram Ruf, Karl Lackner
Blood (2016-12-02) https://doi.org/ggrm82
DOI: 10.1182/blood.v128.22.5023.5023

371. 14th International Congress on Antiphospholipid Antibodies Task Force Report on Antiphospholipid Syndrome Treatment Trends
Doruk Erkan, Cassyanne L. Aguiar, Danieli Andrade, Hannah Cohen, Maria J. Cuadrado, Adriana Danowski, Roger A. Levy, Thomas L. Ortel, Anisur Rahman, Jane E. Salmon, … Michael D. Lockshin
Autoimmunity Reviews (2014-06) https://doi.org/ggp8r8
DOI: 10.1016/j.autrev.2014.01.053 · PMID: 24468415

372. What is the role of hydroxychloroquine in reducing thrombotic risk in patients with antiphospholipid antibodies?
Tzu-Fei Wang, Wendy Lim
Hematology (2016-12-02) https://doi.org/ggrn3k
DOI: 10.1182/asheducation-2016.1.714 · PMID: 27913551 · PMCID: PMC6142483

373. COVID-19: a recommendation to examine the effect of hydroxychloroquine in preventing infection and progression
Dan Zhou, Sheng-Ming Dai, Qiang Tong
Journal of Antimicrobial Chemotherapy (2020-07) https://doi.org/ggq84c
DOI: 10.1093/jac/dkaa114 · PMID: 32196083 · PMCID: PMC7184499

374. Hydroxychloroquine treatment of patients with human immunodeficiency virus type 1
Kirk Sperber, Michael Louie, Thomas Kraus, Jacqueline Proner, Erica Sapira, Su Lin, Vera Stecher, Lloyd Mayer
Clinical Therapeutics (1995-07) https://doi.org/cq2hx9
DOI: 10.1016/0149-2918(95)80039-5

375. Hydroxychloroquine augments early virological response to pegylated interferon plus ribavirin in genotype-4 chronic hepatitis C patients
Gouda Kamel Helal, Magdy Abdelmawgoud Gad, Mohamed Fahmy Abd-Ellah, Mahmoud Saied Eid
Journal of Medical Virology (2016-12) https://doi.org/f889nt
DOI: 10.1002/jmv.24575 · PMID: 27183377 · PMCID: PMC7167065

376. Hydroxychloroquine and azithromycin as a treatment of COVID-19: results of an open-label non-randomized clinical trial
Philippe Gautret, Jean-Christophe Lagier, Philippe Parola, Van Thuan Hoang, Line Meddeb, Morgane Mailhe, Barbara Doudier, Johan Courjon, Valérie Giordanengo, Vera Esteves Vieira, … Didier Raoult
International Journal of Antimicrobial Agents (2020-07) https://doi.org/dp7d
DOI: 10.1016/j.ijantimicag.2020.105949 · PMID: 32205204 · PMCID: PMC7102549

377. Official Statement from International Society of Antimicrobial Chemotherapy
Andreas Voss
(2020-04-03) https://www.isac.world/news-and-publications/official-isac-statement

378. No evidence of rapid antiviral clearance or clinical benefit with the combination of hydroxychloroquine and azithromycin in patients with severe COVID-19 infection
J. M. Molina, C. Delaugerre, J. Le Goff, B. Mela-Lima, D. Ponscarme, L. Goldwirt, N. de Castro
Médecine et Maladies Infectieuses (2020-06) https://doi.org/ggqzrb
DOI: 10.1016/j.medmal.2020.03.006 · PMID: 32240719 · PMCID: PMC7195369

379. Efficacy of hydroxychloroquine in patients with COVID-19: results of a randomized clinical trial
Zhaowei Chen, Jijia Hu, Zongwei Zhang, Shan Jiang, Shoumeng Han, Dandan Yan, Ruhong Zhuang, Ben Hu, Zhan Zhang
Cold Spring Harbor Laboratory (2020-04-10) https://doi.org/ggqm4v
DOI: 10.1101/2020.03.22.20040758

380. The Extent and Consequences of P-Hacking in Science
Megan L. Head, Luke Holman, Rob Lanfear, Andrew T. Kahn, Michael D. Jennions
PLOS Biology (2015-03-13) https://doi.org/4z7
DOI: 10.1371/journal.pbio.1002106 · PMID: 25768323 · PMCID: PMC4359000

381. A pilot study of hydroxychloroquine in treatment of patients with common coronavirus disease-19 (COVID-19)
CHEN Jun, LIU Danping, LIU Li, LIU Ping, XU Qingnian, XIA Lu, LING Yun, HUANG Dan, SONG Shuli, ZHANG Dandan, … LU Hongzhou
Journal of Zhejiang University (Medical Sciences) (2020-03) https://doi.org/10.3785/j.issn.1008-9292.2020.03.03
DOI: 10.3785/j.issn.1008-9292.2020.03.03

382. Breakthrough: Chloroquine phosphate has shown apparent efficacy in treatment of COVID-19 associated pneumonia in clinical studies
Jianjun Gao, Zhenxue Tian, Xu Yang
BioScience Trends (2020-02-29) https://doi.org/ggm3mv
DOI: 10.5582/bst.2020.01047 · PMID: 32074550

383. Targeting the Endocytic Pathway and Autophagy Process as a Novel Therapeutic Strategy in COVID-19
Naidi Yang, Han-Ming Shen
International Journal of Biological Sciences (2020) https://doi.org/ggqspm
DOI: 10.7150/ijbs.45498 · PMID: 32226290 · PMCID: PMC7098027

384. SARS-CoV-2: an Emerging Coronavirus that Causes a Global Threat
Jun Zheng
International Journal of Biological Sciences (2020) https://doi.org/ggqspr
DOI: 10.7150/ijbs.45053 · PMID: 32226285 · PMCID: PMC7098030

385. RETRACTED: Hydroxychloroquine or chloroquine with or without a macrolide for treatment of COVID-19: a multinational registry analysis
Mandeep R Mehra, Sapan S Desai, Frank Ruschitzka, Amit N Patel
The Lancet (2020-05) https://doi.org/ggwzsb
DOI: 10.1016/s0140-6736(20)31180-6 · PMID: 32450107 · PMCID: PMC7255293

386. Retraction—Hydroxychloroquine or chloroquine with or without a macrolide for treatment of COVID-19: a multinational registry analysis
Mandeep R Mehra, Frank Ruschitzka, Amit N Patel
The Lancet (2020-06) https://doi.org/ggzqng
DOI: 10.1016/s0140-6736(20)31324-6 · PMID: 32511943 · PMCID: PMC7274621

387. Life Threatening Severe QTc Prolongation in Patient with Systemic Lupus Erythematosus due to Hydroxychloroquine
John P. O’Laughlin, Parag H. Mehta, Brian C. Wong
Case Reports in Cardiology (2016) https://doi.org/ggqzrc
DOI: 10.1155/2016/4626279 · PMID: 27478650 · PMCID: PMC4960328

388. Keep the QT interval: It is a reliable predictor of ventricular arrhythmias
Dan M. Roden
Heart Rhythm (2008-08) https://doi.org/d5rchx
DOI: 10.1016/j.hrthm.2008.05.008 · PMID: 18675237 · PMCID: PMC3212752

389. Safety of hydroxychloroquine, alone and in combination with azithromycin, in light of rapid wide-spread use for COVID-19: a multinational, network cohort and self-controlled case series study
Jennifer C.E.Lane, James Weaver, Kristin Kostka, Talita Duarte-Salles, Maria Tereza F. Abrahao, Heba Alghoul, Osaid Alser, Thamir M Alshammari, Patricia Biedermann, Edward Burn, … Daniel Prieto-Alhambra
Cold Spring Harbor Laboratory (2020-04-10) https://doi.org/ggrn7s
DOI: 10.1101/2020.04.08.20054551

390. Chloroquine diphosphate in two different dosages as adjunctive therapy of hospitalized patients with severe respiratory syndrome in the context of coronavirus (SARS-CoV-2) infection: Preliminary safety results of a randomized, double-blinded, phase IIb clinical trial ( CloroCovid-19 Study )
Mayla Gabriela Silva Borba, Fernando Fonseca Almeida Val, Vanderson Souza Sampaio, Marcia Almeida Araújo Alexandre, Gisely Cardoso Melo, Marcelo Brito, Maria Paula Gomes Mourão, José Diego Brito-Sousa, Djane Baía-da-Silva, Marcus Vinitius Farias Guerra, … CloroCovid-19 Team
Cold Spring Harbor Laboratory (2020-04-16) https://doi.org/ggr3nj
DOI: 10.1101/2020.04.07.20056424

391. Heart risk concerns mount around use of chloroquine and hydroxychloroquine for Covid-19 treatment
Jacqueline Howard, Elizabeth Cohen, Nadia Kounang, Per Nyberg
CNN (2020-04-14) https://www.cnn.com/2020/04/13/health/chloroquine-risks-coronavirus-treatment-trials-study/index.html

392. WHO Director-General’s opening remarks at the media briefing on COVID-19
World Health Organization
(2020-05-25) https://www.who.int/dg/speeches/detail/who-director-general-s-opening-remarks-at-the-media-briefing-on-covid-19---25-may-2020

393. Hydroxychloroquine in patients mainly with mild to moderate COVID–19: an open–label, randomized, controlled trial
Wei Tang, Zhujun Cao, Mingfeng Han, Zhengyan Wang, Junwen Chen, Wenjin Sun, Yaojie Wu, Wei Xiao, Shengyong Liu, Erzhen Chen, … Qing Xie
Cold Spring Harbor Laboratory (2020-05-07) https://doi.org/ggr68m
DOI: 10.1101/2020.04.10.20060558

394. Outcomes of hydroxychloroquine usage in United States veterans hospitalized with Covid-19
Joseph Magagnoli, Siddharth Narendran, Felipe Pereira, Tammy Cummings, James W. Hardin, S. Scott Sutton, Jayakrishna Ambati
Cold Spring Harbor Laboratory (2020-04-21) https://doi.org/ggspt6
DOI: 10.1101/2020.04.16.20065920 · PMID: 32511622 · PMCID: PMC7276049

395. Effect of Hydroxychloroquine in Hospitalized Patients with Covid-19
The RECOVERY Collaborative Group
New England Journal of Medicine (2020-11-19) https://doi.org/ghd8c7
DOI: 10.1056/nejmoa2022926 · PMID: 33031652 · PMCID: PMC7556338

396. Hydroxychloroquine for Early Treatment of Adults With Mild Coronavirus Disease 2019: A Randomized, Controlled Trial
Oriol Mitjà, Marc Corbacho-Monné, Maria Ubals, Cristian Tebé, Judith Peñafiel, Aurelio Tobias, Ester Ballana, Andrea Alemany, Núria Riera-Martí, Carla A Pérez, … Martí Vall-Mayans
Clinical Infectious Diseases (2020-07-16) https://doi.org/gg5f9x
DOI: 10.1093/cid/ciaa1009 · PMID: 32674126 · PMCID: PMC7454406

397. A Randomized Trial of Hydroxychloroquine as Postexposure Prophylaxis for Covid-19
David R. Boulware, Matthew F. Pullen, Ananta S. Bangdiwala, Katelyn A. Pastick, Sarah M. Lofgren, Elizabeth C. Okafor, Caleb P. Skipper, Alanna A. Nascene, Melanie R. Nicol, Mahsa Abassi, … Kathy H. Hullsiek
New England Journal of Medicine (2020-08-06) https://doi.org/dxkv
DOI: 10.1056/nejmoa2016638 · PMID: 32492293 · PMCID: PMC7289276

398. Efficacy and Safety of Hydroxychloroquine vs Placebo for Pre-exposure SARS-CoV-2 Prophylaxis Among Health Care Workers
Benjamin S. Abella, Eliana L. Jolkovsky, Barbara T. Biney, Julie E. Uspal, Matthew C. Hyman, Ian Frank, Scott E. Hensley, Saar Gill, Dan T. Vogl, Ivan Maillard, … and the Prevention and Treatment of COVID-19 With Hydroxychloroquine (PATCH) Investigators
JAMA Internal Medicine (2020-09-30) https://doi.org/ghd6nj
DOI: 10.1001/jamainternmed.2020.6319 · PMID: 33001138 · PMCID: PMC7527945

399. Mechanisms of action of hydroxychloroquine and chloroquine: implications for rheumatology
Eva Schrezenmeier, Thomas Dörner
Nature Reviews Rheumatology (2020-02-07) https://doi.org/ggzjnh
DOI: 10.1038/s41584-020-0372-x · PMID: 32034323

400. Synthesis and Pharmacology of Anti-Inflammatory Steroidal Antedrugs
M. Omar F. Khan, Henry J. Lee
Chemical Reviews (2008-12-10) https://doi.org/cmkrtc
DOI: 10.1021/cr068203e · PMID: 19035773 · PMCID: PMC2650492

401. Dexamethasone
The Centre for Evidence-Based Medicine
https://www.cebm.net/covid-19/dexamethasone/

402. 16-METHYLATED STEROIDS. I. 16α-METHYLATED ANALOGS OF CORTISONE, A NEW GROUP OF ANTI-INFLAMMATORY STEROIDS
Glen E. Arth, David B. R. Johnston, John Fried, William W. Spooncer, Dale R. Hoff, Lewis H. Sarett
Journal of the American Chemical Society (1958-06) https://doi.org/cj5c82
DOI: 10.1021/ja01545a061

403. Treatment of Rheumatoid Arthritis with Dexamethasone
Abraham Cohen
JAMA (1960-10-15) https://doi.org/csfmhc
DOI: 10.1001/jama.1960.03030070009002 · PMID: 13694317

404. Potential benefits of precise corticosteroids therapy for severe 2019-nCoV pneumonia
Wei Zhou, Yisi Liu, Dongdong Tian, Cheng Wang, Sa Wang, Jing Cheng, Ming Hu, Minghao Fang, Yue Gao
Signal Transduction and Targeted Therapy (2020-02-21) https://doi.org/ggqr84
DOI: 10.1038/s41392-020-0127-9 · PMID: 32296012 · PMCID: PMC7035340

405. Dexamethasone treatment for the acute respiratory distress syndrome: a multicentre, randomised controlled trial
Jesús Villar, Carlos Ferrando, Domingo Martínez, Alfonso Ambrós, Tomás Muñoz, Juan A Soler, Gerardo Aguilar, Francisco Alba, Elena González-Higueras, Luís A Conesa, … Jesús Villar
The Lancet Respiratory Medicine (2020-03) https://doi.org/ggpxzc
DOI: 10.1016/s2213-2600(19)30417-5

406. Clinical evidence does not support corticosteroid treatment for 2019-nCoV lung injury
Clark D Russell, Jonathan E Millar, J Kenneth Baillie
The Lancet (2020-02) https://doi.org/ggks86
DOI: 10.1016/s0140-6736(20)30317-2 · PMID: 32043983 · PMCID: PMC7134694

407. On the use of corticosteroids for 2019-nCoV pneumonia
Lianhan Shang, Jianping Zhao, Yi Hu, Ronghui Du, Bin Cao
The Lancet (2020-02) https://doi.org/ggq356
DOI: 10.1016/s0140-6736(20)30361-5 · PMID: 32122468 · PMCID: PMC7159292

408. Pharmacology of Postoperative Nausea and Vomiting
Eric S. Zabirowicz, Tong J. Gan
Elsevier BV (2019) https://doi.org/ghfkjw
DOI: 10.1016/b978-0-323-48110-6.00034-x

409. Non-traditional cytokines: How catecholamines and adipokines influence macrophages in immunity, metabolism and the central nervous system
Mark A. Barnes, Monica J. Carson, Meera G. Nair
Cytokine (2015-04) https://doi.org/f65c59
DOI: 10.1016/j.cyto.2015.01.008 · PMID: 25703786 · PMCID: PMC4590987

410. Stress Hormones, Proinflammatory and Antiinflammatory Cytokines, and Autoimmunity
ILIA J. ELENKOV, GEORGE P. CHROUSOS
Annals of the New York Academy of Sciences (2002-06) https://doi.org/fmwpx2
DOI: 10.1111/j.1749-6632.2002.tb04229.x · PMID: 12114286

411. Recovery of the Hypothalamic-Pituitary-Adrenal Response to Stress
Arantxa García, Octavi Martí, Astrid Vallès, Silvina Dal-Zotto, Antonio Armario
Neuroendocrinology (2000) https://doi.org/b2cq8n
DOI: 10.1159/000054578 · PMID: 10971146

412. Modulatory effects of glucocorticoids and catecholamines on human interleukin-12 and interleukin-10 production: clinical implications.
IJ Elenkov, DA Papanicolaou, RL Wilder, GP Chrousos
Proceedings of the Association of American Physicians (1996-09) https://www.ncbi.nlm.nih.gov/pubmed/8902882
PMID: 8902882

413. Prevention of infection caused by immunosuppressive drugs in gastroenterology
Katarzyna Orlicka, Eleanor Barnes, Emma L. Culver
Therapeutic Advances in Chronic Disease (2013-04-22) https://doi.org/ggrqd3
DOI: 10.1177/2040622313485275 · PMID: 23819020 · PMCID: PMC3697844

414. Immunosuppression for hyperinflammation in COVID-19: a double-edged sword?
Andrew I Ritchie, Aran Singanayagam
The Lancet (2020-04) https://doi.org/ggq8hs
DOI: 10.1016/s0140-6736(20)30691-7 · PMID: 32220278 · PMCID: PMC7138169

415. Effect of Dexamethasone in Hospitalized Patients with COVID-19 – Preliminary Report
Peter Horby, Wei Shen Lim, Jonathan Emberson, Marion Mafham, Jennifer Bell, Louise Linsell, Natalie Staplin, Christopher Brightling, Andrew Ustianowski, Einas Elmahi, … RECOVERY Collaborative Group
Cold Spring Harbor Laboratory (2020-06-22) https://doi.org/dz5x
DOI: 10.1101/2020.06.22.20137273

416. Dexamethasone in Hospitalized Patients with Covid-19 — Preliminary Report
The RECOVERY Collaborative Group
New England Journal of Medicine (2020-07-17) https://doi.org/gg5c8p
DOI: 10.1056/nejmoa2021436 · PMID: 32678530 · PMCID: PMC7383595

417. Current concepts in the diagnosis and management of cytokine release syndrome
Daniel W. Lee, Rebecca Gardner, David L. Porter, Chrystal U. Louis, Nabil Ahmed, Michael Jensen, Stephan A. Grupp, Crystal L. Mackall
Blood (2014-07-10) https://doi.org/ggsrwk
DOI: 10.1182/blood-2014-05-552729 · PMID: 24876563 · PMCID: PMC4093680

418. Dexamethasone for COVID-19? Not so fast.
T. C. Theoharides
JOURNAL OF BIOLOGICAL REGULATORS AND HOMEOSTATIC AGENTS (2020-08-31) https://doi.org/ghfkjx
DOI: 10.23812/20-editorial_1-5 · PMID: 32551464

419. Formulation and manufacturability of biologics
Steven J Shire
Current Opinion in Biotechnology (2009-12) https://doi.org/cjk8p6
DOI: 10.1016/j.copbio.2009.10.006 · PMID: 19880308

420. Early Development of Therapeutic Biologics - Pharmacokinetics
A. Baumann
Current Drug Metabolism (2006-01-01) https://doi.org/bhcz79
DOI: 10.2174/138920006774832604 · PMID: 16454690

421. Deriving Immune Modulating Drugs from Viruses—A New Class of Biologics
Jordan R. Yaron, Liqiang Zhang, Qiuyun Guo, Michelle Burgin, Lauren N. Schutz, Enkidia Awo, Lyn Wise, Kurt L. Krause, Cristhian J. Ildefonso, Jacek M. Kwiecien, … Alexandra R. Lucas
Journal of Clinical Medicine (2020-03-31) https://doi.org/ghdx73
DOI: 10.3390/jcm9040972 · PMID: 32244484 · PMCID: PMC7230489

422. IL-6 in Inflammation, Immunity, and Disease
T. Tanaka, M. Narazaki, T. Kishimoto
Cold Spring Harbor Perspectives in Biology (2014-09-04) https://doi.org/gftpjs
DOI: 10.1101/cshperspect.a016295 · PMID: 25190079 · PMCID: PMC4176007

423. Hall of Fame among Pro-inflammatory Cytokines: Interleukin-6 Gene and Its Transcriptional Regulation Mechanisms
Yang Luo, Song Guo Zheng
Frontiers in Immunology (2016-12-19) https://doi.org/ggqmgv
DOI: 10.3389/fimmu.2016.00604 · PMID: 28066415 · PMCID: PMC5165036

424. IL-6 Trans-Signaling via the Soluble IL-6 Receptor: Importance for the Pro-Inflammatory Activities of IL-6
Stefan Rose-John
International Journal of Biological Sciences (2012) https://doi.org/f4c4hf
DOI: 10.7150/ijbs.4989 · PMID: 23136552 · PMCID: PMC3491447

425. Interleukin-6 and its receptor: from bench to bedside
Jürgen Scheller, Stefan Rose-John
Medical Microbiology and Immunology (2006-05-31) https://doi.org/ck8xch
DOI: 10.1007/s00430-006-0019-9 · PMID: 16741736

426. Plasticity and cross-talk of Interleukin 6-type cytokines
Christoph Garbers, Heike M. Hermanns, Fred Schaper, Gerhard Müller-Newen, Joachim Grötzinger, Stefan Rose-John, Jürgen Scheller
Cytokine & Growth Factor Reviews (2012-06) https://doi.org/f3z743
DOI: 10.1016/j.cytogfr.2012.04.001 · PMID: 22595692

427. Soluble receptors for cytokines and growth factors: generation and biological function
S Rose-John, PC Heinrich
Biochemical Journal (1994-06-01) https://doi.org/ggqmgd
DOI: 10.1042/bj3000281 · PMID: 8002928 · PMCID: PMC1138158

428. Interleukin-6; pathogenesis and treatment of autoimmune inflammatory diseases
Toshio Tanaka, Masashi Narazaki, Kazuya Masuda, Tadamitsu Kishimoto
Inflammation and Regeneration (2013) https://doi.org/ggqmgt
DOI: 10.2492/inflammregen.33.054

429. Effective treatment of severe COVID-19 patients with tocilizumab
Xiaoling Xu, Mingfeng Han, Tiantian Li, Wei Sun, Dongsheng Wang, Binqing Fu, Yonggang Zhou, Xiaohu Zheng, Yun Yang, Xiuyong Li, … Haiming Wei
Proceedings of the National Academy of Sciences (2020-05-19) https://doi.org/ggv3r3
DOI: 10.1073/pnas.2005615117 · PMID: 32350134 · PMCID: PMC7245089

430. Tocilizumab in patients with severe COVID-19: a retrospective cohort study
Giovanni Guaraldi, Marianna Meschiari, Alessandro Cozzi-Lepri, Jovana Milic, Roberto Tonelli, Marianna Menozzi, Erica Franceschini, Gianluca Cuomo, Gabriella Orlando, Vanni Borghi, … Cristina Mussini
The Lancet Rheumatology (2020-08) https://doi.org/d2pk
DOI: 10.1016/s2665-9913(20)30173-9 · PMID: 32835257 · PMCID: PMC7314456

431. Tocilizumab Treatment for Cytokine Release Syndrome in Hospitalized Patients With Coronavirus Disease 2019
Christina C. Price, Frederick L. Altice, Yu Shyr, Alan Koff, Lauren Pischel, George Goshua, Marwan M. Azar, Dayna Mcmanus, Sheau-Chiann Chen, Shana E. Gleeson, … Maricar Malinis
Chest (2020-10) https://doi.org/gg2789
DOI: 10.1016/j.chest.2020.06.006 · PMID: 32553536

432. Impact of low dose tocilizumab on mortality rate in patients with COVID-19 related pneumonia
Ruggero Capra, Nicola De Rossi, Flavia Mattioli, Giuseppe Romanelli, Cristina Scarpazza, Maria Pia Sormani, Stefania Cossi
European Journal of Internal Medicine (2020-06) https://doi.org/ggx4fm
DOI: 10.1016/j.ejim.2020.05.009 · PMID: 32405160 · PMCID: PMC7219361

433. Tocilizumab therapy reduced intensive care unit admissions and/or mortality in COVID-19 patients
T. Klopfenstein, S. Zayet, A. Lohse, J.-C. Balblanc, J. Badie, P.-Y. Royer, L. Toko, C. Mezher, N. J. Kadiane-Oussou, M. Bossert, … T. Conrozier
Médecine et Maladies Infectieuses (2020-08) https://doi.org/ggvz45
DOI: 10.1016/j.medmal.2020.05.001 · PMID: 32387320 · PMCID: PMC7202806

434. Outcomes in patients with severe COVID-19 disease treated with tocilizumab: a case–controlled study
G Rojas-Marte, M Khalid, O Mukhtar, AT Hashmi, MA Waheed, S Ehrlich, A Aslam, S Siddiqui, C Agarwal, Y Malyshev, … J Shani
QJM: An International Journal of Medicine (2020-08) https://doi.org/gg496t
DOI: 10.1093/qjmed/hcaa206 · PMID: 32569363 · PMCID: PMC7337835

435. Systematic Review and Meta-Analysis of Case-Control Studies from 7,000 COVID-19 Pneumonia Patients Suggests a Beneficial Impact of Tocilizumab with Benefit Most Evident in Non-Corticosteroid Exposed Subjects.
Abdulla Watad, Nicola Luigi Bragazzi, Charlie Bridgewood, Muhammad Mansour, Naim Mahroum, Matteo Riccò, Ahmed Nasr, Amr Hussein, Omer Gendelman, Yehuda Shoenfeld, … Dennis McGonagle
SSRN Electronic Journal (2020) https://doi.org/gg62hz
DOI: 10.2139/ssrn.3642653

436. The Efficacy of IL-6 Inhibitor Tocilizumab in Reducing Severe COVID-19 Mortality: A Systematic Review
Avi Gurion Kaye, Robert Siegel
Cold Spring Harbor Laboratory (2020-09-03) https://doi.org/gg62hv
DOI: 10.1101/2020.07.10.20150938

437. Rationale and evidence on the use of tocilizumab in COVID-19: a systematic review
A. Cortegiani, M. Ippolito, M. Greco, V. Granone, A. Protti, C. Gregoretti, A. Giarratano, S. Einav, M. Cecconi
Pulmonology (2021-01) https://doi.org/gg5xv3
DOI: 10.1016/j.pulmoe.2020.07.003 · PMID: 32713784 · PMCID: PMC7369580

438. New insights and long-term safety of tocilizumab in rheumatoid arthritis
Graeme Jones, Elena Panova
Therapeutic Advances in Musculoskeletal Disease (2018-10-07) https://doi.org/gffsdt
DOI: 10.1177/1759720x18798462 · PMID: 30327685 · PMCID: PMC6178374

439. Tocilizumab during pregnancy and lactation: drug levels in maternal serum, cord blood, breast milk and infant serum
Jumpei Saito, Naho Yakuwa, Kayoko Kaneko, Chinatsu Takai, Mikako Goto, Ken Nakajima, Akimasa Yamatani, Atsuko Murashima
Rheumatology (2019-08) https://doi.org/ggzhks
DOI: 10.1093/rheumatology/kez100 · PMID: 30945743

440. Short-course tocilizumab increases risk of hepatitis B virus reactivation in patients with rheumatoid arthritis: a prospective clinical observation
Le-Feng Chen, Ying-Qian Mo, Jun Jing, Jian-Da Ma, Dong-Hui Zheng, Lie Dai
International Journal of Rheumatic Diseases (2017-07) https://doi.org/f9pbc5
DOI: 10.1111/1756-185x.13010 · PMID: 28160426

441. Why tocilizumab could be an effective treatment for severe COVID-19?
Binqing Fu, Xiaoling Xu, Haiming Wei
Journal of Translational Medicine (2020-04-14) https://doi.org/ggv5c8
DOI: 10.1186/s12967-020-02339-3 · PMID: 32290839 · PMCID: PMC7154566

442. Risk of adverse events including serious infections in rheumatoid arthritis patients treated with tocilizumab: a systematic literature review and meta-analysis of randomized controlled trials
L. Campbell, C. Chen, S. S. Bhagat, R. A. Parker, A. J. K. Ostor
Rheumatology (2010-11-14) https://doi.org/crqn7c
DOI: 10.1093/rheumatology/keq343 · PMID: 21078627

443. Risk of serious infections in tocilizumab versus other biologic drugs in patients with rheumatoid arthritis: a multidatabase cohort study
Ajinkya Pawar, Rishi J Desai, Daniel H Solomon, Adrian J Santiago Ortiz, Sara Gale, Min Bao, Khaled Sarsour, Sebastian Schneeweiss, Seoyoung C Kim
Annals of the Rheumatic Diseases (2019-04) https://doi.org/gg62hx
DOI: 10.1136/annrheumdis-2018-214367 · PMID: 30679153

444. Risk of infections in rheumatoid arthritis patients treated with tocilizumab
Veronika R. Lang, Matthias Englbrecht, Jürgen Rech, Hubert Nüsslein, Karin Manger, Florian Schuch, Hans-Peter Tony, Martin Fleck, Bernhard Manger, Georg Schett, Jochen Zwerina
Rheumatology (2012-05) https://doi.org/d3b3rh
DOI: 10.1093/rheumatology/ker223 · PMID: 21865281

445. Use of Tocilizumab for COVID-19-Induced Cytokine Release Syndrome
Jared Radbel, Navaneeth Narayanan, Pinki J. Bhatt
Chest (2020-07) https://doi.org/ggtxvs
DOI: 10.1016/j.chest.2020.04.024 · PMID: 32343968 · PMCID: PMC7195070

446. Utilizing tocilizumab for the treatment of cytokine release syndrome in COVID-19
Ali Hassoun, Elizabeth Dilip Thottacherry, Justin Muklewicz, Qurrat-ul-ain Aziz, Jonathan Edwards
Journal of Clinical Virology (2020-07) https://doi.org/ggx359
DOI: 10.1016/j.jcv.2020.104443 · PMID: 32425661 · PMCID: PMC7229471

447. Development of therapeutic antibodies for the treatment of diseases
Ruei-Min Lu, Yu-Chyi Hwang, I-Ju Liu, Chi-Chiu Lee, Han-Zen Tsai, Hsin-Jung Li, Han-Chung Wu
Journal of Biomedical Science (2020-01-02) https://doi.org/ggqbpx
DOI: 10.1186/s12929-019-0592-z · PMID: 31894001 · PMCID: PMC6939334

448. Broadly Neutralizing Antiviral Antibodies
Davide Corti, Antonio Lanzavecchia
Annual Review of Immunology (2013-03-21) https://doi.org/gf25g8
DOI: 10.1146/annurev-immunol-032712-095916 · PMID: 23330954

449. Ibalizumab Targeting CD4 Receptors, An Emerging Molecule in HIV Therapy
Simona A. Iacob, Diana G. Iacob
Frontiers in Microbiology (2017-11-27) https://doi.org/gcn3kh
DOI: 10.3389/fmicb.2017.02323 · PMID: 29230203 · PMCID: PMC5711820

450. Product review on the monoclonal antibody palivizumab for prevention of respiratory syncytial virus infection
Bernhard Resch
Human Vaccines & Immunotherapeutics (2017-06-12) https://doi.org/ggqbps
DOI: 10.1080/21645515.2017.1337614 · PMID: 28605249 · PMCID: PMC5612471

451. Chronological evolution of IgM, IgA, IgG and neutralisation antibodies after infection with SARS-associated coronavirus
P.-R. Hsueh, L.-M. Huang, P.-J. Chen, C.-L. Kao, P.-C. Yang
Clinical Microbiology and Infection (2004-12) https://doi.org/cwwg87
DOI: 10.1111/j.1469-0691.2004.01009.x · PMID: 15606632 · PMCID: PMC7129952

452. Neutralizing Antibodies in Patients with Severe Acute Respiratory Syndrome-Associated Coronavirus Infection
Nie Yuchun, Wang Guangwen, Shi Xuanling, Zhang Hong, Qiu Yan, He Zhongping, Wang Wei, Lian Gewei, Yin Xiaolei, Du Liying, … Ding Mingxiao
The Journal of Infectious Diseases (2004-09) https://doi.org/cgqj5b
DOI: 10.1086/423286 · PMID: 15319862 · PMCID: PMC7199490

453. Potent human monoclonal antibodies against SARS CoV, Nipah and Hendra viruses
Ponraj Prabakaran, Zhongyu Zhu, Xiaodong Xiao, Arya Biragyn, Antony S Dimitrov, Christopher C Broder, Dimiter S Dimitrov
Expert Opinion on Biological Therapy (2009-04-08) https://doi.org/b88kw8
DOI: 10.1517/14712590902763755 · PMID: 19216624 · PMCID: PMC2705284

454. Prior Infection and Passive Transfer of Neutralizing Antibody Prevent Replication of Severe Acute Respiratory Syndrome Coronavirus in the Respiratory Tract of Mice
Kanta Subbarao, Josephine McAuliffe, Leatrice Vogel, Gary Fahle, Steven Fischer, Kathleen Tatti, Michelle Packard, Wun-Ju Shieh, Sherif Zaki, Brian Murphy
Journal of Virology (2004-04-01) https://doi.org/b8wr7c
DOI: 10.1128/jvi.78.7.3572-3577.2004 · PMID: 15016880 · PMCID: PMC371090

455. The Effectiveness of Convalescent Plasma and Hyperimmune Immunoglobulin for the Treatment of Severe Acute Respiratory Infections of Viral Etiology: A Systematic Review and Exploratory Meta-analysis
John Mair-Jenkins, Maria Saavedra-Campos, J. Kenneth Baillie, Paul Cleary, Fu-Meng Khaw, Wei Shen Lim, Sophia Makki, Kevin D. Rooney, Jonathan S. Nguyen-Van-Tam, Charles R. Beck, Convalescent Plasma Study Group
Journal of Infectious Diseases (2015-01-01) https://doi.org/f632n7
DOI: 10.1093/infdis/jiu396 · PMID: 25030060 · PMCID: PMC4264590

456. Identification of human neutralizing antibodies against MERS-CoV and their role in virus adaptive evolution
X.-C. Tang, S. S. Agnihothram, Y. Jiao, J. Stanhope, R. L. Graham, E. C. Peterson, Y. Avnir, A. S. C. Tallarico, J. Sheehan, Q. Zhu, … W. A. Marasco
Proceedings of the National Academy of Sciences (2014-04-28) https://doi.org/smr
DOI: 10.1073/pnas.1402074111 · PMID: 24778221 · PMCID: PMC4024880

457. Exceptionally Potent Neutralization of Middle East Respiratory Syndrome Coronavirus by Human Monoclonal Antibodies
Tianlei Ying, Lanying Du, Tina W. Ju, Ponraj Prabakaran, Candy C. Y. Lau, Lu Lu, Qi Liu, Lili Wang, Yang Feng, Yanping Wang, … Dimiter S. Dimitrov
Journal of Virology (2014-07-15) https://doi.org/ggzf5p
DOI: 10.1128/jvi.00912-14 · PMID: 24789777 · PMCID: PMC4097770

458. The Role of ACE2 in Cardiovascular Physiology
Gavin Y. Oudit, Michael A. Crackower, Peter H. Backx, Josef M. Penninger
Trends in Cardiovascular Medicine (2003-04) https://doi.org/bsbp49
DOI: 10.1016/s1050-1738(02)00233-5

459. A human monoclonal antibody blocking SARS-CoV-2 infection
Chunyan Wang, Wentao Li, Dubravka Drabek, Nisreen M. A. Okba, Rien van Haperen, Albert D. M. E. Osterhaus, Frank J. M. van Kuppeveld, Bart L. Haagmans, Frank Grosveld, Berend-Jan Bosch
Cold Spring Harbor Laboratory (2020-03-12) https://doi.org/ggnw4t
DOI: 10.1101/2020.03.11.987958

460. Unexpected Receptor Functional Mimicry Elucidates Activation of Coronavirus Fusion
Alexandra C. Walls, Xiaoli Xiong, Young-Jun Park, M. Alejandra Tortorici, Joost Snijder, Joel Quispe, Elisabetta Cameroni, Robin Gopal, Mian Dai, Antonio Lanzavecchia, … David Veesler
Cell (2019-02) https://doi.org/gft3jg
DOI: 10.1016/j.cell.2018.12.028 · PMID: 30712865 · PMCID: PMC6751136

461. In Vitro Neutralization Is Not Predictive of Prophylactic Efficacy of Broadly Neutralizing Monoclonal Antibodies CR6261 and CR9114 against Lethal H2 Influenza Virus Challenge in Mice
Troy C. Sutton, Elaine W. Lamirande, Kevin W. Bock, Ian N. Moore, Wouter Koudstaal, Muniza Rehman, Gerrit Jan Weverling, Jaap Goudsmit, Kanta Subbarao
Journal of Virology (2017-10-18) https://doi.org/ggqbpt
DOI: 10.1128/jvi.01603-17 · PMID: 29046448 · PMCID: PMC5709608

462. Human neutralizing antibodies against MERS coronavirus: implications for future immunotherapy
Xian-Chun Tang, Wayne A Marasco
Immunotherapy (2015-07) https://doi.org/ggqbpz
DOI: 10.2217/imt.15.33 · PMID: 26098703 · PMCID: PMC5068219

463. A Potent and Broad Neutralizing Antibody Recognizes and Penetrates the HIV Glycan Shield
R. Pejchal, K. J. Doores, L. M. Walker, R. Khayat, P.-S. Huang, S.-K. Wang, R. L. Stanfield, J.-P. Julien, A. Ramos, M. Crispin, … I. A. Wilson
Science (2011-10-13) https://doi.org/bzqv8c
DOI: 10.1126/science.1213256 · PMID: 21998254 · PMCID: PMC3280215

464. Broadly Neutralizing Antibodies against HIV: Back to Blood
Amir Dashti, Anthony L. DeVico, George K. Lewis, Mohammad M. Sajadi
Trends in Molecular Medicine (2019-03) https://doi.org/ggqbpr
DOI: 10.1016/j.molmed.2019.01.007 · PMID: 30792120 · PMCID: PMC6401214

465. Importance of Neutralizing Monoclonal Antibodies Targeting Multiple Antigenic Sites on the Middle East Respiratory Syndrome Coronavirus Spike Glycoprotein To Avoid Neutralization Escape
Lingshu Wang, Wei Shi, James D. Chappell, M. Gordon Joyce, Yi Zhang, Masaru Kanekiyo, Michelle M. Becker, Neeltje van Doremalen, Robert Fischer, Nianshuang Wang, … Barney S. Graham
Journal of Virology (2018-04-27) https://doi.org/ggqbpv
DOI: 10.1128/jvi.02002-17 · PMID: 29514901 · PMCID: PMC5923077

466. Anti–spike IgG causes severe acute lung injury by skewing macrophage responses during acute SARS-CoV infection
Li Liu, Qiang Wei, Qingqing Lin, Jun Fang, Haibo Wang, Hauyee Kwok, Hangying Tang, Kenji Nishiura, Jie Peng, Zhiwu Tan, … Zhiwei Chen
JCI Insight (2019-02-21) https://doi.org/ggqbpw
DOI: 10.1172/jci.insight.123158 · PMID: 30830861 · PMCID: PMC6478436

467. The antiviral effect of interferon-beta against SARS-Coronavirus is not mediated by MxA protein
Martin Spiegel, Andreas Pichlmair, Elke Mühlberger, Otto Haller, Friedemann Weber
Journal of Clinical Virology (2004-07) https://doi.org/cmc3ds
DOI: 10.1016/j.jcv.2003.11.013 · PMID: 15135736 · PMCID: PMC7128634

468. Coronavirus virulence genes with main focus on SARS-CoV envelope gene
Marta L. DeDiego, Jose L. Nieto-Torres, Jose M. Jimenez-Guardeño, Jose A. Regla-Nava, Carlos Castaño-Rodriguez, Raul Fernandez-Delgado, Fernando Usera, Luis Enjuanes
Virus Research (2014-12) https://doi.org/f6wm24
DOI: 10.1016/j.virusres.2014.07.024 · PMID: 25093995 · PMCID: PMC4261026

469. people
Richard marsden
(2020-03) https://www.synairgen.com/wp-content/uploads/2020/03/200318-Synairgen-to-start-trial-of-SNG001-in-COVID-19-imminently-.pdf

470. people
Richard marsden
(2020-07) https://www.synairgen.com/wp-content/uploads/2020/07/200720-Synairgen-announces-positive-results-from-trial-of-SNG001-in-hospitalised-COVID-19-patients.pdf

471. Developing Covid-19 Vaccines at Pandemic Speed
Nicole Lurie, Melanie Saville, Richard Hatchett, Jane Halton
New England Journal of Medicine (2020-05-21) https://doi.org/ggq8bc
DOI: 10.1056/nejmp2005630 · PMID: 32227757

472. WHO | Novel Coronavirus – China
WHO
http://www.who.int/csr/don/12-january-2020-novel-coronavirus-china/en/

473. SARS-CoV-2 vaccines in development
Florian Krammer
Nature (2020-09-23) https://doi.org/ghdprn
DOI: 10.1038/s41586-020-2798-3 · PMID: 32967006

474. Newer Vaccine Technologies Deployed to Develop COVID-19 Shot
Abby Olena
The Scientist Magazine (2020-02-21) https://www.the-scientist.com/news-opinion/newer-vaccine-technologies-deployed-to-develop-covid-19-shot-67152

475. A strategic approach to COVID-19 vaccine R&D
Lawrence Corey, John R. Mascola, Anthony S. Fauci, Francis S. Collins
Science (2020-05-29) https://doi.org/ggwfck
DOI: 10.1126/science.abc5312 · PMID: 32393526

476. An mRNA Vaccine against SARS-CoV-2 — Preliminary Report
Lisa A. Jackson, Evan J. Anderson, Nadine G. Rouphael, Paul C. Roberts, Mamodikoe Makhene, Rhea N. Coler, Michele P. McCullough, James D. Chappell, Mark R. Denison, Laura J. Stevens, … John H. Beigel
New England Journal of Medicine (2020-11-12) https://doi.org/d3tt
DOI: 10.1056/nejmoa2022483 · PMID: 32663912 · PMCID: PMC7377258

477. DNA https://www.who.int/teams/health-product-and-policy-standards/standards-and-specifications/vaccines-quality/dna

478. Safety, Tolerability and Immunogenicity of INO-4800 for COVID-19 in Healthy Volunteers - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04336410

479. Electroporation delivery of DNA vaccines: prospects for success
Niranjan Y Sardesai, David B Weiner
Current Opinion in Immunology (2011-06) https://doi.org/cq8b4p
DOI: 10.1016/j.coi.2011.03.008 · PMID: 21530212 · PMCID: PMC3109217

480. Tolerability of intramuscular and intradermal delivery by CELLECTRA ® adaptive constant current electroporation device in healthy volunteers
Malissa C Diehl, Jessica C Lee, Stephen E Daniels, Pablo Tebas, Amir S Khan, Mary Giffear, Niranjan Y Sardesai, Mark L Bagarazzi
Human Vaccines & Immunotherapeutics (2014-10-27) https://doi.org/ggrj7h
DOI: 10.4161/hv.24702 · PMID: 24051434 · PMCID: PMC3906411

481. Induction of virus-specific cytotoxic T lymphocytesin vivo by liposome-entrapped mRNA
Frédéric Martinon, Sivadasan Krishnan, Gerlinde Lenzen, Rémy Magné, Elisabeth Gomard, Jean-Gérard Guillet, Jean-Paul Lévy, Pierre Meulien
European Journal of Immunology (1993-07) https://doi.org/b6jb3z
DOI: 10.1002/eji.1830230749 · PMID: 8325342

482. Advances in mRNA Vaccines for Infectious Diseases
Cuiling Zhang, Giulietta Maruggi, Hu Shan, Junwei Li
Frontiers in Immunology (2019-03-27) https://doi.org/ggsnm7
DOI: 10.3389/fimmu.2019.00594 · PMID: 30972078 · PMCID: PMC6446947

483. mRNA vaccine delivery using lipid nanoparticles
Andreas M Reichmuth, Matthias A Oberli, Ana Jaklenec, Robert Langer, Daniel Blankschtein
Therapeutic Delivery (2016-05) https://doi.org/f8xfzc
DOI: 10.4155/tde-2016-0006 · PMID: 27075952 · PMCID: PMC5439223

484. Mechanism of action of mRNA-based vaccines
Carlo Iavarone, Derek T. O’hagan, Dong Yu, Nicolas F. Delahaye, Jeffrey B. Ulmer
Expert Review of Vaccines (2017-07-28) https://doi.org/ggsnm6
DOI: 10.1080/14760584.2017.1355245 · PMID: 28701102

485. SARS-CoV-2 Vaccines: Status Report
Fatima Amanat, Florian Krammer
Immunity (2020-04) https://doi.org/ggrdj4
DOI: 10.1016/j.immuni.2020.03.007 · PMID: 32259480 · PMCID: PMC7136867

486. Study in Healthy Adults to Evaluate Gene Activation After Vaccination With GlaxoSmithKline (GSK) Biologicals’ Candidate Tuberculosis (TB) Vaccine GSK 692342 - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT01669096

487. Nucleoside-modified mRNA immunization elicits influenza virus hemagglutinin stalk-specific antibodies
Norbert Pardi, Kaela Parkhouse, Ericka Kirkpatrick, Meagan McMahon, Seth J. Zost, Barbara L. Mui, Ying K. Tam, Katalin Karikó, Christopher J. Barbosa, Thomas D. Madden, … Drew Weissman
Nature Communications (2018-08-22) https://doi.org/gd49qt
DOI: 10.1038/s41467-018-05482-0 · PMID: 30135514 · PMCID: PMC6105651

488. RNA vaccines: an introduction
PHG Foundation
https://www.phgfoundation.org/briefing/rna-vaccines

489. T Follicular Helper Cell Differentiation, Function, and Roles in Disease
Shane Crotty
Immunity (2014-10) https://doi.org/ggsp64
DOI: 10.1016/j.immuni.2014.10.004 · PMID: 25367570 · PMCID: PMC4223692

490. mRNA vaccines — a new era in vaccinology
Norbert Pardi, Michael J. Hogan, Frederick W. Porter, Drew Weissman
Nature Reviews Drug Discovery (2018-01-12) https://doi.org/gcsmgr
DOI: 10.1038/nrd.2017.243 · PMID: 29326426 · PMCID: PMC5906799

491. Phase I, Open-Label, Dose-Ranging Study of the Safety and Immunogenicity of 2019-nCoV Vaccine (mRNA-1273) in Healthy Adults
National Institute of Allergy and Infectious Diseases (NIAID)
clinicaltrials.gov (2020-12-17) https://clinicaltrials.gov/ct2/show/NCT04283461

492. A Study to Evaluate Efficacy, Safety, and Immunogenicity of mRNA-1273 Vaccine in Adults Aged 18 Years and Older to Prevent COVID-19 - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04470427

493. Progress and Prospects on Vaccine Development against SARS-CoV-2
Jinyong Zhang, Hao Zeng, Jiang Gu, Haibo Li, Lixin Zheng, Quanming Zou
Vaccines (2020-03-29) https://doi.org/ggq726
DOI: 10.3390/vaccines8020153 · PMID: 32235387 · PMCID: PMC7349596

494. Towards an understanding of the adjuvant action of aluminium
Philippa Marrack, Amy S. McKee, Michael W. Munks
Nature Reviews Immunology (2009-04) https://doi.org/drcwvf
DOI: 10.1038/nri2510 · PMID: 19247370 · PMCID: PMC3147301

495. DAMP-Inducing Adjuvant and PAMP Adjuvants Parallelly Enhance Protective Type-2 and Type-1 Immune Responses to Influenza Split Vaccination
Tomoya Hayashi, Masatoshi Momota, Etsushi Kuroda, Takato Kusakabe, Shingo Kobari, Kotaro Makisaka, Yoshitaka Ohno, Yusuke Suzuki, Fumika Nakagawa, Michelle S. J. Lee, … Hidetoshi Arima
Frontiers in Immunology (2018-11-20) https://doi.org/gfqq89
DOI: 10.3389/fimmu.2018.02619 · PMID: 30515151 · PMCID: PMC6255964

496. Better Adjuvants for Better Vaccines: Progress in Adjuvant Delivery Systems, Modifications, and Adjuvant–Antigen Codelivery
Zhi-Biao Wang, Jing Xu
Vaccines (2020-03-13) https://doi.org/gg35vj
DOI: 10.3390/vaccines8010128 · PMID: 32183209 · PMCID: PMC7157724

497. Defining trained immunity and its role in health and disease
Mihai G. Netea, Jorge Domínguez-Andrés, Luis B. Barreiro, Triantafyllos Chavakis, Maziar Divangahi, Elaine Fuchs, Leo A. B. Joosten, Jos W. M. van der Meer, Musa M. Mhlanga, Willem J. M. Mulder, … Eicke Latz
Nature Reviews Immunology (2020-03-04) https://doi.org/gg28pr
DOI: 10.1038/s41577-020-0285-6 · PMID: 32132681 · PMCID: PMC7186935

498. Trained Immunity: a Tool for Reducing Susceptibility to and the Severity of SARS-CoV-2 Infection
Mihai G. Netea, Evangelos J. Giamarellos-Bourboulis, Jorge Domínguez-Andrés, Nigel Curtis, Reinout van Crevel, Frank L. van de Veerdonk, Marc Bonten
Cell (2020-05) https://doi.org/gg2584
DOI: 10.1016/j.cell.2020.04.042 · PMID: 32437659 · PMCID: PMC7196902

499. BCG Vaccination to Protect Healthcare Workers Against COVID-19 - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04327206

500. Reducing Health Care Workers Absenteeism in Covid-19 Pandemic Through BCG Vaccine - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04328441

501. BCG Vaccine for Health Care Workers as Defense Against COVID 19 - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04348370

502. Application of BCG Vaccine for Immune-prophylaxis Among Egyptian Healthcare Workers During the Pandemic of COVID-19
Adel Khattab
clinicaltrials.gov (2020-04-17) https://clinicaltrials.gov/ct2/show/NCT04350931

503. Performance Evaluation of BCG Vaccination in Healthcare Personnel to Reduce the Severity of SARS-COV-2 Infection in Medellín, Colombia, 2020
Universidad de Antioquia
clinicaltrials.gov (2020-11-24) https://clinicaltrials.gov/ct2/show/NCT04362124

504. COVID-19: BCG As Therapeutic Vaccine, Transmission Limitation, and Immunoglobulin Enhancement - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04369794

505. Using BCG Vaccine to Protect Health Care Workers in the COVID-19 Pandemic - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04373291

506. Reducing Morbidity and Mortality in Health Care Workers Exposed to SARS-CoV-2 by Enhancing Non-specific Immune Responses Through Bacillus Calmette-Guérin Vaccination, a Randomized Controlled Trial
TASK Applied Science
clinicaltrials.gov (2020-05-06) https://clinicaltrials.gov/ct2/show/NCT04379336

507. Randomized Controlled Trial Evaluating the Efficacy of Vaccination With Bacillus Calmette and Guérin (BCG) in the Prevention of COVID-19 Via the Strengthening of Innate Immunity in Health Care Workers
Assistance Publique - Hôpitaux de Paris
clinicaltrials.gov (2020-08-17) https://clinicaltrials.gov/ct2/show/NCT04384549

508. Study to Assess VPM1002 in Reducing Healthcare Professionals’ Absenteeism in COVID-19 Pandemic - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04387409

509. A Randomized Clinical Trial for Enhanced Trained Immune Responses Through Bacillus Calmette-Guérin Vaccination to Prevent Infections by COVID-19: The ACTIVATE II Trial
Hellenic Institute for the Study of Sepsis
clinicaltrials.gov (2020-07-10) https://clinicaltrials.gov/ct2/show/NCT04414267

510. Reducing Hospital Admission of Elderly in SARS-CoV-2 Pandemic Via the Induction of Trained Immunity by Bacillus Calmette-Guérin Vaccination, a Randomized Controlled Trial
Radboud University
clinicaltrials.gov (2020-06-03) https://clinicaltrials.gov/ct2/show/NCT04417335

511. Study to Assess VPM1002 in Reducing Hospital Admissions and/or Severe Respiratory Infectious Diseases in Elderly in COVID-19 Pandemic - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04435379

512. Efficacy and Safety of VPM1002 in Reducing SARS-CoV-2 (COVID-19) Infection Rate and Severity - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04439045

513. Complex Immune Dysregulation in COVID-19 Patients with Severe Respiratory Failure
Evangelos J. Giamarellos-Bourboulis, Mihai G. Netea, Nikoletta Rovina, Karolina Akinosoglou, Anastasia Antoniadou, Nikolaos Antonakos, Georgia Damoraki, Theologia Gkavogianni, Maria-Evangelia Adami, Paraskevi Katsaounou, … Antonia Koutsoukou
Cell Host & Microbe (2020-06) https://doi.org/ggthxs
DOI: 10.1016/j.chom.2020.04.009 · PMID: 32320677 · PMCID: PMC7172841

514. Type I and Type III Interferons – Induction, Signaling, Evasion, and Application to Combat COVID-19
Annsea Park, Akiko Iwasaki
Cell Host & Microbe (2020-06) https://doi.org/gg2ccp
DOI: 10.1016/j.chom.2020.05.008 · PMID: 32464097 · PMCID: PMC7255347

515. Pfizer and BioNTech to Submit Emergency Use Authorization Request Today to the U.S. FDA for COVID-19 Vaccine | Pfizer https://www.pfizer.com/news/press-release/press-release-detail/pfizer-and-biontech-submit-emergency-use-authorization

516. Moderna Announces First Participants Dosed in Phase 2/3 Study of COVID-19 Vaccine Candidate in Adolescents | Moderna, Inc. https://investors.modernatx.com/news-releases/news-release-details/moderna-announces-first-participants-dosed-phase-23-study-covid/

517. Table 1, Cost-Comparison Table for Biologic Disease-Modifying Drugs for Rheumatoid Arthritis
National Center for Biotechnology Information, U. S. National Library of Medicine 8600 Rockville Pike, Bethesda MD, 20894 Usa
(2015-08) https://www.ncbi.nlm.nih.gov/books/NBK349513/table/T43/

518. A Cost Comparison of Treatments of Moderate to Severe Psoriasis
Cheryl Hankin, Steven Feldman, Andy Szczotka, Randolph Stinger, Leslie Fish, David Hankin
Drug Benefit Trends (2005-05) https://escholarship.umassmed.edu/meyers_pp/385

519. TNF-α inhibition for potential therapeutic modulation of SARS coronavirus infection
Edward Tobinick
Current Medical Research and Opinion (2008-09-22) https://doi.org/bq4cx2
DOI: 10.1185/030079903125002757 · PMID: 14741070

520. Sanofi and Regeneron begin global Kevzara® (sarilumab) clinical trial program in patients with severe COVID-19
Sanofi
(2020-03-16) http://www.news.sanofi.us/2020-03-16-Sanofi-and-Regeneron-begin-global-Kevzara-R-sarilumab-clinical-trial-program-in-patients-with-severe-COVID-19

521. Sarilumab COVID-19 - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04327388

522. COVID-19: combining antiviral and anti-inflammatory treatments
Justin Stebbing, Anne Phelan, Ivan Griffin, Catherine Tucker, Olly Oechsle, Dan Smith, Peter Richardson
The Lancet Infectious Diseases (2020-04) https://doi.org/dph5
DOI: 10.1016/s1473-3099(20)30132-8 · PMID: 32113509 · PMCID: PMC7158903

523. Baricitinib as potential treatment for 2019-nCoV acute respiratory disease
Peter Richardson, Ivan Griffin, Catherine Tucker, Dan Smith, Olly Oechsle, Anne Phelan, Michael Rawling, Edward Savory, Justin Stebbing
The Lancet (2020-02) https://doi.org/ggnrsx
DOI: 10.1016/s0140-6736(20)30304-4 · PMID: 32032529 · PMCID: PMC7137985

524. Lilly Begins Clinical Testing of Therapies for COVID-19 | Eli Lilly and Company https://investor.lilly.com/news-releases/news-release-details/lilly-begins-clinical-testing-therapies-covid-19

525. Baricitinib Combined With Antiviral Therapy in Symptomatic Patients Infected by COVID-19: an Open-label, Pilot Study
Fabrizio Cantini
clinicaltrials.gov (2020-04-19) https://clinicaltrials.gov/ct2/show/NCT04320277

526. Design and Synthesis of Hydroxyferroquine Derivatives with Antimalarial and Antiviral Activities
Christophe Biot, Wassim Daher, Natascha Chavain, Thierry Fandeur, Jamal Khalife, Daniel Dive, Erik De Clercq
Journal of Medicinal Chemistry (2006-05) https://doi.org/db4n83
DOI: 10.1021/jm0601856 · PMID: 16640347

527. An orally bioavailable broad-spectrum antiviral inhibits SARS-CoV-2 in human airway epithelial cell cultures and multiple coronaviruses in mice
Timothy P. Sheahan, Amy C. Sims, Shuntai Zhou, Rachel L. Graham, Andrea J. Pruijssers, Maria L. Agostini, Sarah R. Leist, Alexandra Schäfer, Kenneth H. Dinnon, Laura J. Stevens, … Ralph S. Baric
Science Translational Medicine (2020-04-29) https://doi.org/ggrqd2
DOI: 10.1126/scitranslmed.abb5883 · PMID: 32253226 · PMCID: PMC7164393

528. Antiviral Monoclonal Antibodies: Can They Be More Than Simple Neutralizing Agents?
Mireia Pelegrin, Mar Naranjo-Gomez, Marc Piechaczyk
Trends in Microbiology (2015-10) https://doi.org/f7vzrf
DOI: 10.1016/j.tim.2015.07.005 · PMID: 26433697 · PMCID: PMC7127033

529. Intranasal Treatment with Poly(I{middle dot}C) Protects Aged Mice from Lethal Respiratory Virus Infections
J. Zhao, C. Wohlford-Lenane, J. Zhao, E. Fleming, T. E. Lane, P. B. McCray, S. Perlman
Journal of Virology (2012-08-22) https://doi.org/f4bzfp
DOI: 10.1128/jvi.01410-12 · PMID: 22915814 · PMCID: PMC3486278

530. Main protease structure and XChem fragment screen
Diamond
(2020-05-05) https://www.diamond.ac.uk/covid-19/for-scientists/Main-protease-structure-and-XChem.html

531. COVID-19: The Inflammation Link and the Role of Nutrition in Potential Mitigation
Ioannis Zabetakis, Ronan Lordan, Catherine Norton, Alexandros Tsoupras
Nutrients (2020-05-19) https://doi.org/ggxdq3
DOI: 10.3390/nu12051466 · PMID: 32438620 · PMCID: PMC7284818

532. Could nutrition modulate COVID-19 susceptibility and severity of disease? A systematic review
Philip T. James, Zakari Ali, Andrew E. Armitage, Ana Bonell, Carla Cerami, Hal Drakesmith, Modou Jobe, Kerry S. Jones, Zara Liew, Sophie E. Moore, … Andrew M. Prentice
Cold Spring Harbor Laboratory (2020-10-21) https://doi.org/ghr94g
DOI: 10.1101/2020.10.19.20214395

533. Coronavirus Disease 2019 (COVID-19) and Nutritional Status: The Missing Link?
Renata Silverio, Daniela Caetano Gonçalves, Márcia Fábia Andrade, Marilia Seelaender
Advances in Nutrition (2020-09-25) https://doi.org/ghhqjd
DOI: 10.1093/advances/nmaa125 · PMID: 32975565 · PMCID: PMC7543263

534. Nutritional status of patients with COVID-19
Jae Hyoung Im, Young Soo Je, Jihyeon Baek, Moon-Hyun Chung, Hea Yoon Kwon, Jin-Soo Lee
International Journal of Infectious Diseases (2020-11) https://doi.org/gg7t5t
DOI: 10.1016/j.ijid.2020.08.018 · PMID: 32795605 · PMCID: PMC7418699

535. Optimal Nutritional Status for a Well-Functioning Immune System Is an Important Factor to Protect against Viral Infections
Philip C. Calder, Anitra C. Carr, Adrian F. Gombart, Manfred Eggersdorfer
Nutrients (2020-04-23) https://doi.org/gg29hh
DOI: 10.3390/nu12041181 · PMID: 32340216 · PMCID: PMC7230749

536. Structural Design Principles for Delivery of Bioactive Components in Nutraceuticals and Functional Foods
David Julian McClements, Eric Andrew Decker, Yeonhwa Park, Jochen Weiss
Critical Reviews in Food Science and Nutrition (2009-06-16) https://doi.org/dt68m4
DOI: 10.1080/10408390902841529 · PMID: 19484636

537. Nutraceutical therapies for atherosclerosis
Joe W. E. Moss, Dipak P. Ramji
Nature Reviews Cardiology (2016-07-07) https://doi.org/f9g389
DOI: 10.1038/nrcardio.2016.103 · PMID: 27383080 · PMCID: PMC5228762

538. Nutraceutical-definition and introduction
Ekta K. Kalra
AAPS PharmSci (2015-07-10) https://doi.org/cg5wc9
DOI: 10.1208/ps050325 · PMID: 14621960 · PMCID: PMC2750935

539. Peak dietary supplement sales leveling off during COVID-19 pandemic, but growth still remains strong over last year, market researchers report during webcast
Nutritional Outlook
https://www.nutritionaloutlook.com/view/peak-dietary-supplement-sales-leveling-during-covid-19-pandemic-growth-still-remains-strong

540. Dietary Diversity among Chinese Residents during the COVID-19 Outbreak and Its Associated Factors
Ai Zhao, Zhongyu Li, Yalei Ke, Shanshan Huo, Yidi Ma, Yumei Zhang, Jian Zhang, Zhongxia Ren
Nutrients (2020-06-06) https://doi.org/ghc6d9
DOI: 10.3390/nu12061699 · PMID: 32517210 · PMCID: PMC7352896

541. Lockdown impact: Grocery stores bolstered NZ supplements sales as pharmacies slumped
nutraingredients-asia.com
nutraingredients-asia.com https://www.nutraingredients-asia.com/Article/2020/07/06/Lockdown-impact-Grocery-stores-bolstered-NZ-supplements-sales-as-pharmacies-slumped

542. COVID-19 temporarily bolsters European interest in supplements
.nutritioninsight.com/
https://ni.cnsmedia.com/a/EHHJsDOG2oc=

543. India’s immune health surge: Nation leads APAC in number of new product launches – new data
nutraingredients.com
nutraingredients.com https://www.nutraingredients.com/Article/2020/07/21/India-s-immune-health-surge-Nation-leads-APAC-in-number-of-new-product-launches-new-data

544. Food policy, nutrition and nutraceuticals in the prevention and management of COVID-19: Advice for healthcare professionals
Yasemin Ipek Ayseli, Nazli Aytekin, Derya Buyukkayhan, Ismail Aslan, Mehmet Turan Ayseli
Trends in Food Science & Technology (2020-11) https://doi.org/ghjtcp
DOI: 10.1016/j.tifs.2020.09.001

545. 5 Food and Beverage Trends in Europe During COVID-19 https://kerry.com/insights/kerrydigest/2020/5-food-and-beverage-trends-in-europe-during-covid-19

546. Dietary Supplement Health and Education Act of 1994
National Institutes of Health Office of Dietary Supplements
https://ods.od.nih.gov/About/DSHEA_Wording.aspx

547. Food and Drug Administration Modernization Act (FDAMA) of 1997
Office of the Commissioner
FDA (2018-11-03) https://www.fda.gov/regulatory-information/selected-amendments-fdc-act/food-and-drug-administration-modernization-act-fdama-1997

548. Nutraceuticals - shedding light on the grey area between pharmaceuticals and food
Antonello Santini, Ettore Novellino
Expert Review of Clinical Pharmacology (2018-04-23) https://doi.org/ggwztk
DOI: 10.1080/17512433.2018.1464911 · PMID: 29667442

549. https://eur-lex.europa.eu/legal-content/EN/ALL

550. https://eur-lex.europa.eu/legal-content/EN/TXT

551. EU Register of nutrition and health claims made on foods (v.3.5) https://ec.europa.eu/food/safety/labelling_nutrition/claims/register/public/

552. Nutraceuticals: opening the debate for a regulatory framework
Antonello Santini, Silvia Miriam Cammarata, Giacomo Capone, Angela Ianaro, Gian Carlo Tenore, Luca Pani, Ettore Novellino
British Journal of Clinical Pharmacology (2018-04) https://doi.org/ggwztm
DOI: 10.1111/bcp.13496 · PMID: 29433155 · PMCID: PMC5867125

553. Reviewing the Nutrition and Health Claims Regulation (EC) No. 1924/2006: What do we know about its challenges and potential impact on innovation?
Stefanie Bröring, Sukhada Khedkar, Stefano Ciliberti
International Journal of Food Sciences and Nutrition (2016-08-02) https://doi.org/ghr936
DOI: 10.1080/09637486.2016.1212816 · PMID: 27484163

554. Dietary Supplements: Regulatory Challenges and Research Resources
Johanna Dwyer, Paul Coates, Michael Smith
Nutrients (2018-01-04) https://doi.org/ghr949
DOI: 10.3390/nu10010041 · PMID: 29300341 · PMCID: PMC5793269

555. Noetic Nutraceuticals - 607572 - 05/15/2020
Center for Drug Evaluation and Research
Center for Drug Evaluation and Research (2020-05-18) https://www.fda.gov/inspections-compliance-enforcement-and-criminal-investigations/warning-letters/noetic-nutraceuticals-607572-05152020

556. Regulations.gov Beta https://beta.regulations.gov/document/FDA-2020-S-0023-0068

557. Spartan Enterprises Inc. dba Watershed Wellness Center - 610876 - 10/30/2020
Center for Drug Evaluation and Research
Center for Drug Evaluation and Research (2020-11-02) https://www.fda.gov/inspections-compliance-enforcement-and-criminal-investigations/warning-letters/spartan-enterprises-inc-dba-watershed-wellness-center-610876-10302020

558. FTC Sues California Marketer of $23,000 COVID-19 “Treatment” Plan
Federal Trade Commission
(2020-07-31) https://www.ftc.gov/news-events/press-releases/2020/07/ftc-sues-california-marketer-23000-covid-19-treatment-plan

559. https://cen.acs.org/biological-chemistry/natural-products/oleandrin-compound-touted-possible-COVID/98/web/2020/08

560. Reducing mortality from 2019-nCoV: host-directed therapies should be an option
Alimuddin Zumla, David S Hui, Esam I Azhar, Ziad A Memish, Markus Maeurer
The Lancet (2020-02) https://doi.org/ggkd3b
DOI: 10.1016/s0140-6736(20)30305-6 · PMID: 32035018 · PMCID: PMC7133595

561. Diet Supplementation, Probiotics, and Nutraceuticals in SARS-CoV-2 Infection: A Scoping Review
Fabio Infusino, Massimiliano Marazzato, Massimo Mancone, Francesco Fedele, Claudio Maria Mastroianni, Paolo Severino, Giancarlo Ceccarelli, Letizia Santinelli, Elena Cavarretta, Antonino G. M. Marullo, … Gabriella d’Ettorre
Nutrients (2020-06-08) https://doi.org/gg8k58
DOI: 10.3390/nu12061718 · PMID: 32521760 · PMCID: PMC7352781

562. Potential interventions for novel coronavirus in China: A systematic review
Lei Zhang, Yunhui Liu
Journal of Medical Virology (2020-03-03) https://doi.org/ggpx57
DOI: 10.1002/jmv.25707 · PMID: 32052466 · PMCID: PMC7166986

563. Nutraceuticals have potential for boosting the type 1 interferon response to RNA viruses including influenza and coronavirus
Mark F. McCarty, James J. DiNicolantonio
Progress in Cardiovascular Diseases (2020-05) https://doi.org/ggpwx2
DOI: 10.1016/j.pcad.2020.02.007 · PMID: 32061635 · PMCID: PMC7130854

564. Inflammation and cardiovascular disease: are marine phospholipids the answer?
Ronan Lordan, Shane Redfern, Alexandros Tsoupras, Ioannis Zabetakis
Food & Function (2020) https://doi.org/gg29hg
DOI: 10.1039/c9fo01742a · PMID: 32270798

565. The Potential Beneficial Effect of EPA and DHA Supplementation Managing Cytokine Storm in Coronavirus Disease
Zoltán Szabó, Tamás Marosvölgyi, Éva Szabó, Péter Bai, Mária Figler, Zsófia Verzár
Frontiers in Physiology (2020-06-19) https://doi.org/gg4hz4
DOI: 10.3389/fphys.2020.00752 · PMID: 32636763 · PMCID: PMC7318894

566. Prospective options of algae-derived nutraceuticals as supplements to combat COVID-19 and human coronavirus diseases
Sachitra K. Ratha, Nirmal Renuka, Ismail Rawat, Faizal Bux
Nutrition (2021-03) https://doi.org/ghr93z
DOI: 10.1016/j.nut.2020.111089 · PMID: 33412367 · PMCID: PMC7680017

567. Omega-3 polyunsaturated fatty acids and inflammatory processes: nutrition or pharmacology?
Philip C. Calder
British Journal of Clinical Pharmacology (2013-03) https://doi.org/ggqmgg
DOI: 10.1111/j.1365-2125.2012.04374.x · PMID: 22765297 · PMCID: PMC3575932

568. Proresolving Lipid Mediators and Mechanisms in the Resolution of Acute Inflammation
Christopher D. Buckley, Derek W. Gilroy, Charles N. Serhan
Immunity (2014-03) https://doi.org/f5wntr
DOI: 10.1016/j.immuni.2014.02.009 · PMID: 24656045 · PMCID: PMC4004957

569. COVID-19 and its implications for thrombosis and anticoagulation
Jean M. Connors, Jerrold H. Levy
Blood (2020-06-04) https://doi.org/ggv35b
DOI: 10.1182/blood.2020006000 · PMID: 32339221 · PMCID: PMC7273827

570. COVID-19 and Thrombotic or Thromboembolic Disease: Implications for Prevention, Antithrombotic Therapy, and Follow-Up
Behnood Bikdeli, Mahesh V. Madhavan, David Jimenez, Taylor Chuich, Isaac Dreyfus, Elissa Driggin, Caroline Der Nigoghossian, Walter Ageno, Mohammad Madjid, Yutao Guo, … Gregory Y. H. Lip
Journal of the American College of Cardiology (2020-06) https://doi.org/ggsppk
DOI: 10.1016/j.jacc.2020.04.031 · PMID: 32311448 · PMCID: PMC7164881

571. Thrombosis and COVID-19: The Potential Role of Nutrition
Alexandros Tsoupras, Ronan Lordan, Ioannis Zabetakis
Frontiers in Nutrition (2020-09-25) https://doi.org/ghr945
DOI: 10.3389/fnut.2020.583080 · PMID: 33102511 · PMCID: PMC7545367

572. Specialized pro-resolving mediators: endogenous regulators of infection and inflammation
Maria C. Basil, Bruce D. Levy
Nature Reviews Immunology (2015-12-21) https://doi.org/f9fgtd
DOI: 10.1038/nri.2015.4 · PMID: 26688348 · PMCID: PMC5242505

573. Specialized mediators in infection and lung injury
Shayna Sandhaus, Andrew G. Swick
BioFactors (2020-11-28) https://doi.org/ghr93m
DOI: 10.1002/biof.1691 · PMID: 33249673 · PMCID: PMC7744833

574. The Lipid Mediator Protectin D1 Inhibits Influenza Virus Replication and Improves Severe Influenza
Masayuki Morita, Keiji Kuba, Akihiko Ichikawa, Mizuho Nakayama, Jun Katahira, Ryo Iwamoto, Tokiko Watanebe, Saori Sakabe, Tomo Daidoji, Shota Nakamura, … Yumiko Imai
Cell (2013-03) https://doi.org/f4rbgb
DOI: 10.1016/j.cell.2013.02.027 · PMID: 23477864

575. Pro-resolving lipid mediators are leads for resolution physiology
Charles N. Serhan
Nature (2014-06-04) https://doi.org/ggv53d
DOI: 10.1038/nature13479 · PMID: 24899309 · PMCID: PMC4263681

576. The Specialized Proresolving Mediator 17-HDHA Enhances the Antibody-Mediated Immune Response against Influenza Virus: A New Class of Adjuvant?
Sesquile Ramon, Steven F. Baker, Julie M. Sahler, Nina Kim, Eric A. Feldsott, Charles N. Serhan, Luis Martínez-Sobrido, David J. Topham, Richard P. Phipps
The Journal of Immunology (2014-12-15) https://doi.org/f6spr8
DOI: 10.4049/jimmunol.1302795 · PMID: 25392529 · PMCID: PMC4258475

577. Inflammation resolution: a dual-pronged approach to averting cytokine storms in COVID-19?
Dipak Panigrahy, Molly M. Gilligan, Sui Huang, Allison Gartung, Irene Cortés-Puch, Patricia J. Sime, Richard P. Phipps, Charles N. Serhan, Bruce D. Hammock
Cancer and Metastasis Reviews (2020-05-08) https://doi.org/ggvv7w
DOI: 10.1007/s10555-020-09889-4 · PMID: 32385712 · PMCID: PMC7207990

578. Pro resolving inflammatory effects of the lipid mediators of omega 3 fatty acids and its implication in SARS COVID-19
Pedro-Antonio Regidor, Fernando Gonzalez Santos, Jose Miguel Rizo, Fernando Moreno Egea
Medical Hypotheses (2020-12) https://doi.org/ghr93x
DOI: 10.1016/j.mehy.2020.110340 · PMID: 33069094 · PMCID: PMC7543931

579. Fish Oil-Fed Mice Have Impaired Resistance to Influenza Infection
Nicole M. J. Schwerbrock, Erik A. Karlsson, Qing Shi, Patricia A. Sheridan, Melinda A. Beck
The Journal of Nutrition (2009-08) https://doi.org/dv45f4
DOI: 10.3945/jn.109.108027 · PMID: 19549756 · PMCID: PMC2709305

580. Modulation of host defence against bacterial and viral infections by omega-3 polyunsaturated fatty acids
Marie-Odile Husson, Delphine Ley, Céline Portal, Madeleine Gottrand, Thomas Hueso, Jean-Luc Desseyn, Frédéric Gottrand
Journal of Infection (2016-12) https://doi.org/f9pp2h
DOI: 10.1016/j.jinf.2016.10.001 · PMID: 27746159

581. Bioactive products formed in humans from fish oils
Carsten Skarke, Naji Alamuddin, John A. Lawson, Xuanwen Li, Jane F. Ferguson, Muredach P. Reilly, Garret A. FitzGerald
Journal of Lipid Research (2015-09) https://doi.org/f7pm5g
DOI: 10.1194/jlr.m060392 · PMID: 26180051 · PMCID: PMC4548785

582. Resolving Inflammatory Storm in COVID-19 Patients by Omega-3 Polyunsaturated Fatty Acids - A Single-blind, Randomized, Placebo-controlled Feasibility Study
Magnus Bäck
clinicaltrials.gov (2020-11-27) https://clinicaltrials.gov/ct2/show/NCT04647604

583. Regulation of platelet function and thrombosis by omega-3 and omega-6 polyunsaturated fatty acids
Reheman Adili, Megan Hawley, Michael Holinstat
Prostaglandins & Other Lipid Mediators (2018-11) https://doi.org/ggvv73
DOI: 10.1016/j.prostaglandins.2018.09.005 · PMID: 30266534 · PMCID: PMC6242736

584. Platelet activation and prothrombotic mediators at the nexus of inflammation and atherosclerosis: Potential role of antiplatelet agents
Ronan Lordan, Alexandros Tsoupras, Ioannis Zabetakis
Blood Reviews (2020-04) https://doi.org/ggvv7x
DOI: 10.1016/j.blre.2020.100694 · PMID: 32340775

585. An Investigation on the Effects of Icosapent Ethyl (VascepaTM) on Inflammatory Biomarkers in Individuals With COVID-19 - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04412018

586. Cardiovascular Risk Reduction with Icosapent Ethyl for Hypertriglyceridemia
Deepak L. Bhatt, P. Gabriel Steg, Michael Miller, Eliot A. Brinton, Terry A. Jacobson, Steven B. Ketchum, Ralph T. Doyle, Rebecca A. Juliano, Lixia Jiao, Craig Granowitz, … Christie M. Ballantyne
New England Journal of Medicine (2019-01-03) https://doi.org/gfj3w9
DOI: 10.1056/nejmoa1812792 · PMID: 30415628

587. A Randomised, Double-blind, Placebo Controlled Study of Eicosapentaenoic Acid (EPA-FFA) Gastro-resistant Capsules to Treat Hospitalised Subjects With Confirmed SARS-CoV-2
S.L.A. Pharma AG
clinicaltrials.gov (2020-10-29) https://clinicaltrials.gov/ct2/show/NCT04335032

588. Anti-inflammatory/Antioxidant Oral Nutrition Supplementation on the Cytokine Storm and Progression of COVID-19: A Randomized Controlled Trial
Mahmoud Abulmeaty FACN M. D.
clinicaltrials.gov (2020-09-18) https://clinicaltrials.gov/ct2/show/NCT04323228

589. Functional Role of Dietary Intervention to Improve the Outcome of COVID-19: A Hypothesis of Work
Giovanni Messina, Rita Polito, Vincenzo Monda, Luigi Cipolloni, Nunzio Di Nunno, Giulio Di Mizio, Paolo Murabito, Marco Carotenuto, Antonietta Messina, Daniela Pisanelli, … Francesco Sessa
International Journal of Molecular Sciences (2020-04-28) https://doi.org/ggvb88
DOI: 10.3390/ijms21093104 · PMID: 32354030 · PMCID: PMC7247152

590. Zinc and immunity: An essential interrelation
Maria Maares, Hajo Haase
Archives of Biochemistry and Biophysics (2016-12) https://doi.org/f9c9b5
DOI: 10.1016/j.abb.2016.03.022 · PMID: 27021581

591. Zinc-Dependent Suppression of TNF-α Production Is Mediated by Protein Kinase A-Induced Inhibition of Raf-1, IκB Kinase β, and NF-κB
Verena von Bülow, Svenja Dubben, Gabriela Engelhardt, Silke Hebel, Birgit Plümäkers, Holger Heine, Lothar Rink, Hajo Haase
The Journal of Immunology (2007-09-15) https://doi.org/f3vs45
DOI: 10.4049/jimmunol.179.6.4180 · PMID: 17785857

592. Zinc activates NF-κB in HUT-78 cells
Ananda S. Prasad, Bin Bao, Frances W. J. Beck, Fazlul H. Sarkar
Journal of Laboratory and Clinical Medicine (2001-10) https://doi.org/cnc6fr
DOI: 10.1067/mlc.2001.118108 · PMID: 11574819

593. Innate or Adaptive Immunity? The Example of Natural Killer Cells
E. Vivier, D. H. Raulet, A. Moretta, M. A. Caligiuri, L. Zitvogel, L. L. Lanier, W. M. Yokoyama, S. Ugolini
Science (2011-01-06) https://doi.org/ckzg9g
DOI: 10.1126/science.1198687 · PMID: 21212348 · PMCID: PMC3089969

594. Zinc supplementation decreases incidence of infections in the elderly: effect of zinc on generation of cytokines and oxidative stress
Ananda S Prasad, Frances WJ Beck, Bin Bao, James T Fitzgerald, Diane C Snell, Joel D Steinberg, Lavoisier J Cardozo
The American Journal of Clinical Nutrition (2007-03) https://doi.org/ggqmgs
DOI: 10.1093/ajcn/85.3.837 · PMID: 17344507

595. The Role of Zinc in Antiviral Immunity
Scott A Read, Stephanie Obeid, Chantelle Ahlenstiel, Golo Ahlenstiel
Advances in Nutrition (2019-07) https://doi.org/ggqmgr
DOI: 10.1093/advances/nmz013 · PMID: 31305906 · PMCID: PMC6628855

596. Efficacy of Zinc Against Common Cold Viruses: An Overview
Darrell Hulisz
Journal of the American Pharmacists Association (2004-09) https://doi.org/cf6pmt
DOI: 10.1331/1544-3191.44.5.594.hulisz · PMID: 15496046 · PMCID: PMC7185598

597. Zinc Lozenges May Shorten the Duration of Colds: A Systematic Review
Harri Hemilä
The Open Respiratory Medicine Journal (2011-06-23) https://doi.org/bndmfq
DOI: 10.2174/1874306401105010051 · PMID: 21769305 · PMCID: PMC3136969

598. COVID-19: Poor outcomes in patients with zinc deficiency
Dinesh Jothimani, Ezhilarasan Kailasam, Silas Danielraj, Balaji Nallathambi, Hemalatha Ramachandran, Padmini Sekar, Shruthi Manoharan, Vidyalakshmi Ramani, Gomathy Narasimhan, Ilankumaran Kaliamoorthy, Mohamed Rela
International Journal of Infectious Diseases (2020-11) https://doi.org/ghr93t
DOI: 10.1016/j.ijid.2020.09.014 · PMID: 32920234 · PMCID: PMC7482607

599. Zn2+ Inhibits Coronavirus and Arterivirus RNA Polymerase Activity In Vitro and Zinc Ionophores Block the Replication of These Viruses in Cell Culture
Aartjan J. W. te Velthuis, Sjoerd H. E. van den Worm, Amy C. Sims, Ralph S. Baric, Eric J. Snijder, Martijn J. van Hemert
PLoS Pathogens (2010-11-04) https://doi.org/d95x4g
DOI: 10.1371/journal.ppat.1001176 · PMID: 21079686 · PMCID: PMC2973827

600. The SARS-coronavirus papain-like protease: Structure, function and inhibition by designed antiviral compounds
Yahira M. Báez-Santos, Sarah E. St. John, Andrew D. Mesecar
Antiviral Research (2015-03) https://doi.org/f63hjp
DOI: 10.1016/j.antiviral.2014.12.015 · PMID: 25554382 · PMCID: PMC5896749

601. A Randomized Study Evaluating the Safety and Efficacy of Hydroxychloroquine and Zinc in Combination With Either Azithromycin or Doxycycline for the Treatment of COVID-19 in the Outpatient Setting
Avni Thakore MD
clinicaltrials.gov (2020-12-08) https://clinicaltrials.gov/ct2/show/NCT04370782

602. A Study of Hydroxychloroquine and Zinc in the Prevention of COVID-19 Infection in Military Healthcare Workers - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04377646

603. Therapies to Prevent Progression of COVID-19, Including Hydroxychloroquine, Azithromycin, Zinc, Vitamin D, Vitamin B12 With or Without Vitamin C, a Multi-centre, International, Randomized Trial: The International ALLIANCE Study
National Institute of Integrative Medicine, Australia
clinicaltrials.gov (2020-09-09) https://clinicaltrials.gov/ct2/show/NCT04395768

604. Early Intervention in COVID-19: Favipiravir Verses Standard Care - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04373733

605. Hydroxychloroquine with or without Azithromycin in Mild-to-Moderate Covid-19
Alexandre B. Cavalcanti, Fernando G. Zampieri, Regis G. Rosa, Luciano C. P. Azevedo, Viviane C. Veiga, Alvaro Avezum, Lucas P. Damiani, Aline Marcadenti, Letícia Kawano-Dourado, Thiago Lisboa, … Otavio Berwanger
New England Journal of Medicine (2020-11-19) https://doi.org/gg5343
DOI: 10.1056/nejmoa2019014 · PMID: 32706953 · PMCID: PMC7397242

606. The efficacy and safety of hydroxychloroquine for COVID-19 prophylaxis: A systematic review and meta-analysis of randomized trials
Kimberley Lewis, Dipayan Chaudhuri, Fayez Alshamsi, Laiya Carayannopoulos, Karin Dearness, Zain Chagla, Waleed Alhazzani, for the GUIDE Group
PLOS ONE (2021-01-06) https://doi.org/ghsv36
DOI: 10.1371/journal.pone.0244778 · PMID: 33406138 · PMCID: PMC7787432

607. Effect of hydroxychloroquine with or without azithromycin on the mortality of coronavirus disease 2019 (COVID-19) patients: a systematic review and meta-analysis
Thibault Fiolet, Anthony Guihur, Mathieu Edouard Rebeaud, Matthieu Mulot, Nathan Peiffer-Smadja, Yahya Mahamat-Saleh
Clinical Microbiology and Infection (2021-01) https://doi.org/gg9jk2
DOI: 10.1016/j.cmi.2020.08.022 · PMID: 32860962 · PMCID: PMC7449662

608. Zinc sulfate in combination with a zinc ionophore may improve outcomes in hospitalized COVID-19 patients
Philip M. Carlucci, Tania Ahuja, Christopher Petrilli, Harish Rajagopalan, Simon Jones, Joseph Rahimian
Journal of Medical Microbiology (2020-10-01) https://doi.org/ghnws7
DOI: 10.1099/jmm.0.001250 · PMID: 32930657 · PMCID: PMC7660893

609. The Minimal Effect of Zinc on the Survival of Hospitalized Patients With COVID-19
Jasper Seth Yao, Joseph Alexander Paguio, Edward Christopher Dee, Hanna Clementine Tan, Achintya Moulick, Carmelo Milazzo, Jerry Jurado, Nicolás Della Penna, Leo Anthony Celi
Chest (2021-01) https://doi.org/gg5w36
DOI: 10.1016/j.chest.2020.06.082 · PMID: 32710890 · PMCID: PMC7375307

610. Coronavirus Disease 2019- Using Ascorbic Acid and Zinc Supplementation (COVIDAtoZ) Research Study A Randomized, Open Label Single Center Study
Milind Desai
clinicaltrials.gov (2020-08-12) https://clinicaltrials.gov/ct2/show/NCT04342728

611. Vitamin B12 May Inhibit RNA-Dependent-RNA Polymerase Activity of nsp12 from the COVID-19 Virus
Naveen Narayanan, Deepak T. Nair
Preprints (2020-03-22) https://doi.org/ggqmjc
DOI: 10.20944/preprints202003.0347.v1

612. The Long History of Vitamin C: From Prevention of the Common Cold to Potential Aid in the Treatment of COVID-19
Giuseppe Cerullo, Massimo Negro, Mauro Parimbelli, Michela Pecoraro, Simone Perna, Giorgio Liguori, Mariangela Rondanelli, Hellas Cena, Giuseppe D’Antona
Frontiers in Immunology (2020-10-28) https://doi.org/ghr943
DOI: 10.3389/fimmu.2020.574029 · PMID: 33193359 · PMCID: PMC7655735

613. The Emerging Role of Vitamin C in the Prevention and Treatment of COVID-19
Anitra C. Carr, Sam Rowe
Nutrients (2020-10-27) https://doi.org/ghr95c
DOI: 10.3390/nu12113286 · PMID: 33121019 · PMCID: PMC7693980

614. Vitamin C Mitigates Oxidative Stress and Tumor Necrosis Factor-Alpha in Severe Community-Acquired Pneumonia and LPS-Induced Macrophages
Yuanyuan Chen, Guangyan Luo, Jiao Yuan, Yuanyuan Wang, Xiaoqiong Yang, Xiaoyun Wang, Guoping Li, Zhiguang Liu, Nanshan Zhong
Mediators of Inflammation (2014) https://doi.org/f6nb5f
DOI: 10.1155/2014/426740 · PMID: 25253919 · PMCID: PMC4165740

615. Intravenous infusion of ascorbic acid decreases serum histamine concentrations in patients with allergic and non-allergic diseases
Alexander F. Hagel, Christian M. Layritz, Wolfgang H. Hagel, Hans-Jürgen Hagel, Edith Hagel, Wolfgang Dauth, Jürgen Kressel, Tanja Regnet, Andreas Rosenberg, Markus F. Neurath, … Martin Raithel
Naunyn-Schmiedeberg’s Archives of Pharmacology (2013-05-11) https://doi.org/f48jsb
DOI: 10.1007/s00210-013-0880-1 · PMID: 23666445

616. Vitamin C and Immune Function
Anitra Carr, Silvia Maggini
Nutrients (2017-11-03) https://doi.org/gfzrjs
DOI: 10.3390/nu9111211 · PMID: 29099763 · PMCID: PMC5707683

617. Changes in Leucocyte Ascorbic Acid during the Common Cold
R. Hume, Elspeth Weyers
Scottish Medical Journal (2016-06-25) https://doi.org/ggqrfj
DOI: 10.1177/003693307301800102 · PMID: 4717661

618. ASCORBIC ACID FUNCTION AND METABOLISM DURING COLDS
C. W. M. Wilson
Annals of the New York Academy of Sciences (1975-09) https://doi.org/bjfdtb
DOI: 10.1111/j.1749-6632.1975.tb29312.x · PMID: 1106304

619. Metabolism of ascorbic acid (vitamin C) in subjects infected with common cold viruses
J. E. W. Davies, R. E. Hughes, Eleri Jones, Sylvia E. Reed, J. W. Craig, D. A. J. Tyrrell
Biochemical Medicine (1979-02) https://doi.org/fd22sv
DOI: 10.1016/0006-2944(79)90058-9

620. Vitamin C and Infections
Harri Hemilä
Nutrients (2017-03-29) https://doi.org/gfkb9n
DOI: 10.3390/nu9040339 · PMID: 28353648 · PMCID: PMC5409678

621. Vitamin C and the common cold
Harri Hemilä
British Journal of Nutrition (2007-03-09) https://doi.org/fszhc6
DOI: 10.1079/bjn19920004 · PMID: 1547201

622. Serum Levels of Vitamin C and Vitamin D in a Cohort of Critically Ill COVID-19 Patients of a North American Community Hospital Intensive Care Unit in May 2020: A Pilot Study
Cristian Arvinte, Maharaj Singh, Paul E. Marik
Medicine in Drug Discovery (2020-12) https://doi.org/ghnwqt
DOI: 10.1016/j.medidd.2020.100064 · PMID: 32964205 · PMCID: PMC7499070

623. Vitamin C levels in patients with SARS-CoV-2-associated acute respiratory distress syndrome
Luis Chiscano-Camón, Juan Carlos Ruiz-Rodriguez, Adolf Ruiz-Sanmartin, Oriol Roca, Ricard Ferrer
Critical Care (2020-08-26) https://doi.org/ghbr97
DOI: 10.1186/s13054-020-03249-y · PMID: 32847620 · PMCID: PMC7447967

624. Targeting coagulation activation in severe COVID-19 pneumonia: lessons from bacterial pneumonia and sepsis
Ricardo J. José, Andrew Williams, Ari Manuel, Jeremy S. Brown, Rachel C. Chambers
European Respiratory Review (2020-10-01) https://doi.org/ghr94s
DOI: 10.1183/16000617.0240-2020 · PMID: 33004529 · PMCID: PMC7537941

625. Vitamin C and Microvascular Dysfunction in Systemic Inflammation
Karel Tyml
Antioxidants (2017-06-29) https://doi.org/ghr947
DOI: 10.3390/antiox6030049 · PMID: 28661424 · PMCID: PMC5618077

626. The use of IV vitamin C for patients with COVID-19: a case series
Raul Hiedra, Kevin Bryan Lo, Mohammad Elbashabsheh, Fahad Gul, Robert Matthew Wright, Jeri Albano, Zurab Azmaiparashvili, Gabriel Patarroyo Aponte
Expert Review of Anti-infective Therapy (2020-08-01) https://doi.org/ghr938
DOI: 10.1080/14787210.2020.1794819 · PMID: 32662690 · PMCID: PMC7441798

627. Vitamin C for preventing and treating the common cold
Harri Hemilä, Elizabeth Chalker
Cochrane Database of Systematic Reviews (2013-01-31) https://doi.org/xz5
DOI: 10.1002/14651858.cd000980.pub4 · PMID: 23440782

628. Vitamin C intake and susceptibility to pneumonia
HARRI HEMILÄ
The Pediatric Infectious Disease Journal (1997-09) https://doi.org/fkvs9d
DOI: 10.1097/00006454-199709000-00003 · PMID: 9306475

629. Vitamin C Can Shorten the Length of Stay in the ICU: A Meta-Analysis
Harri Hemilä, Elizabeth Chalker
Nutrients (2019-03-27) https://doi.org/gfzscg
DOI: 10.3390/nu11040708 · PMID: 30934660 · PMCID: PMC6521194

630. Effect of Vitamin C Infusion on Organ Failure and Biomarkers of Inflammation and Vascular Injury in Patients With Sepsis and Severe Acute Respiratory Failure
Alpha A. Fowler, Jonathon D. Truwit, R. Duncan Hite, Peter E. Morris, Christine DeWilde, Anna Priday, Bernard Fisher, Leroy R. Thacker, Ramesh Natarajan, Donald F. Brophy, … Matthew Halquist
JAMA (2019-10-01) https://doi.org/ggqmh8
DOI: 10.1001/jama.2019.11825 · PMID: 31573637 · PMCID: PMC6777268

631. Intravenous high-dose vitamin C for the treatment of severe COVID-19: study protocol for a multicentre randomised controlled trial
Fang Liu, Yuan Zhu, Jing Zhang, Yiming Li, Zhiyong Peng
BMJ Open (2020-07-08) https://doi.org/gg4sgj
DOI: 10.1136/bmjopen-2020-039519 · PMID: 32641343 · PMCID: PMC7348463

632. Pilot Trial of High-dose vitamin C in critically ill COVID-19 patients
Jing Zhang, Xin Rao, Yiming Li, Yuan Zhu, Fang Liu, Guangling Guo, Guoshi Luo, Zhongji Meng, Daniel De Backer, Hui Xiang, Zhi-Yong Peng
Research Square (2020-08-03) https://doi.org/ghr94x
DOI: 10.21203/rs.3.rs-52778/v2

633. High-dose vitamin C infusion for the treatment of critically ill COVID-19
Jing Zhang, Xin Rao, Yiming Li, Yuan Zhu, Fang Liu, Guangling Guo, Guoshi Luo, Zhongji Meng, Daniel De Backer, Hui Xiang, Zhi-Yong Peng
Research Square (2020-08-03) https://doi.org/ghr94v
DOI: 10.21203/rs.3.rs-52778/v1

634. Vitamin D and Infectious Diseases: Simple Bystander or Contributing Factor?
Pedro Gois, Daniela Ferreira, Simon Olenski, Antonio Seguro
Nutrients (2017-06-24) https://doi.org/ggpcwr
DOI: 10.3390/nu9070651 · PMID: 28672783 · PMCID: PMC5537771

635. Vitamin D and Influenza—Prevention or Therapy?
Beata M. Gruber–Bzura
International Journal of Molecular Sciences (2018-08-16) https://doi.org/ggndrj
DOI: 10.3390/ijms19082419 · PMID: 30115864 · PMCID: PMC6121423

636. Immunologic Effects of Vitamin D on Human Health and Disease
Nipith Charoenngam, Michael F. Holick
Nutrients (2020-07-15) https://doi.org/gg45fp
DOI: 10.3390/nu12072097 · PMID: 32679784 · PMCID: PMC7400911

637. Vitamin D and respiratory health
D. A. Hughes, R. Norton
Clinical & Experimental Immunology (2009-10) https://doi.org/b3n6wc
DOI: 10.1111/j.1365-2249.2009.04001.x · PMID: 19737226 · PMCID: PMC2759054

638. Regulation of Immune Function by Vitamin D and Its Use in Diseases of Immunity
An-Sofie Vanherwegen, Conny Gysemans, Chantal Mathieu
Endocrinology and Metabolism Clinics of North America (2017-12) https://doi.org/gcm7h9
DOI: 10.1016/j.ecl.2017.07.010 · PMID: 29080635

639. Vitamin D and the Immune System
Cynthia Aranow
Journal of Investigative Medicine (2015-12-15) https://doi.org/f3wh87
DOI: 10.2310/jim.0b013e31821b8755 · PMID: 21527855

640. Vitamin D in the prevention of acute respiratory infection: Systematic review of clinical studies
David A. Jolliffe, Christopher J. Griffiths, Adrian R. Martineau
The Journal of Steroid Biochemistry and Molecular Biology (2013-07) https://doi.org/ggqmh9
DOI: 10.1016/j.jsbmb.2012.11.017 · PMID: 23220552

641. Vitamin D: modulator of the immune system
Femke Baeke, Tatiana Takiishi, Hannelie Korf, Conny Gysemans, Chantal Mathieu
Current Opinion in Pharmacology (2010-08) https://doi.org/d43qtf
DOI: 10.1016/j.coph.2010.04.001 · PMID: 20427238

642. Evidence that Vitamin D Supplementation Could Reduce Risk of Influenza and COVID-19 Infections and Deaths
William B. Grant, Henry Lahore, Sharon L. McDonnell, Carole A. Baggerly, Christine B. French, Jennifer L. Aliano, Harjit P. Bhattoa
Nutrients (2020-04-02) https://doi.org/ggr2v5
DOI: 10.3390/nu12040988 · PMID: 32252338 · PMCID: PMC7231123

643. Perspective: Vitamin D deficiency and COVID‐19 severity – plausibly linked by latitude, ethnicity, impacts on cytokines, ACE2 and thrombosis
J. M. Rhodes, S. Subramanian, E. Laird, G. Griffin, R. A. Kenny
Journal of Internal Medicine (2020-07-22) https://doi.org/ghc7dh
DOI: 10.1111/joim.13149 · PMID: 32613681 · PMCID: PMC7361294

644. COVID-19 fatalities, latitude, sunlight, and vitamin D
Paul B. Whittemore
American Journal of Infection Control (2020-09) https://doi.org/ghr93r
DOI: 10.1016/j.ajic.2020.06.193 · PMID: 32599103 · PMCID: PMC7319635

645. Editorial: low population mortality from COVID-19 in countries south of latitude 35 degrees North supports vitamin D as a factor determining severity
Jonathan M. Rhodes, Sreedhar Subramanian, Eamon Laird, Rose A. Kenny
Alimentary Pharmacology & Therapeutics (2020-06) https://doi.org/ggtw4b
DOI: 10.1111/apt.15777 · PMID: 32311755 · PMCID: PMC7264531

646. 25-Hydroxyvitamin D Concentrations Are Lower in Patients with Positive PCR for SARS-CoV-2
Antonio D’Avolio, Valeria Avataneo, Alessandra Manca, Jessica Cusato, Amedeo De Nicolò, Renzo Lucchini, Franco Keller, Marco Cantù
Nutrients (2020-05-09) https://doi.org/ggvv76
DOI: 10.3390/nu12051359 · PMID: 32397511 · PMCID: PMC7285131

647. Vitamin D deficiency as risk factor for severe COVID-19: a convergence of two pandemics
D. De Smet, K. De Smet, P. Herroelen, S. Gryspeerdt, G. A. Martens
Cold Spring Harbor Laboratory (2020-05-05) https://doi.org/ggvv75
DOI: 10.1101/2020.05.01.20079376

648. Vitamin D sufficiency, a serum 25-hydroxyvitamin D at least 30 ng/mL reduced risk for adverse clinical outcomes in patients with COVID-19 infection
Zhila Maghbooli, Mohammad Ali Sahraian, Mehdi Ebrahimi, Marzieh Pazoki, Samira Kafan, Hedieh Moradi Tabriz, Azar Hadadi, Mahnaz Montazeri, Mehrad Nasiri, Arash Shirvani, Michael F. Holick
PLOS ONE (2020-09-25) https://doi.org/ghdzx8
DOI: 10.1371/journal.pone.0239799 · PMID: 32976513 · PMCID: PMC7518605

649. Role of vitamin D in preventing of COVID-19 infection, progression and severity
Nurshad Ali
Journal of Infection and Public Health (2020-10) https://doi.org/ghdzw9
DOI: 10.1016/j.jiph.2020.06.021 · PMID: 32605780 · PMCID: PMC7305922

650. Low plasma 25(OH) vitamin D level is associated with increased risk of COVID‐19 infection: an Israeli population‐based study
Eugene Merzon, Dmitry Tworowski, Alessandro Gorohovski, Shlomo Vinker, Avivit Golan Cohen, Ilan Green, Milana Frenkel‐Morgenstern
The FEBS Journal (2020-08-28) https://doi.org/gg7b5c
DOI: 10.1111/febs.15495 · PMID: 32700398 · PMCID: PMC7404739

651. Association of Vitamin D Status and Other Clinical Characteristics With COVID-19 Test Results
David O. Meltzer, Thomas J. Best, Hui Zhang, Tamara Vokes, Vineet Arora, Julian Solway
JAMA Network Open (2020-09-03) https://doi.org/ghdzw6
DOI: 10.1001/jamanetworkopen.2020.19722 · PMID: 32880651 · PMCID: PMC7489852

652. Vitamin D Status in Hospitalized Patients with SARS-CoV-2 Infection
José L Hernández, Daniel Nan, Marta Fernandez-Ayala, Mayte García-Unzueta, Miguel A Hernández-Hernández, Marcos López-Hoyos, Pedro Muñoz-Cacho, José M Olmos, Manuel Gutiérrez-Cuadra, Juan J Ruiz-Cubillán, … Víctor M Martínez-Taboada
The Journal of Clinical Endocrinology & Metabolism (2020-10-27) https://doi.org/ghh737
DOI: 10.1210/clinem/dgaa733 · PMID: 33159440

653. Analysis of vitamin D level among asymptomatic and critically ill COVID-19 patients and its correlation with inflammatory markers
Anshul Jain, Rachna Chaurasia, Narendra Singh Sengar, Mayank Singh, Sachin Mahor, Sumit Narain
Scientific Reports (2020-11-19) https://doi.org/ghm3zn
DOI: 10.1038/s41598-020-77093-z · PMID: 33214648 · PMCID: PMC7677378

654. Low 25-Hydroxyvitamin D Levels on Admission to the Intensive Care Unit May Predispose COVID-19 Pneumonia Patients to a Higher 28-Day Mortality Risk: A Pilot Study on a Greek ICU Cohort
Alice G. Vassiliou, Edison Jahaj, Maria Pratikaki, Stylianos E. Orfanos, Ioanna Dimopoulou, Anastasia Kotanidou
Nutrients (2020-12-09) https://doi.org/ghr95d
DOI: 10.3390/nu12123773 · PMID: 33316914 · PMCID: PMC7764169

655. Vitamin D deficiency as a predictor of poor prognosis in patients with acute respiratory failure due to COVID-19
G. E. Carpagnano, V. Di Lecce, V. N. Quaranta, A. Zito, E. Buonamico, E. Capozza, A. Palumbo, G. Di Gioia, V. N. Valerio, O. Resta
Journal of Endocrinological Investigation (2020-08-09) https://doi.org/gg7kqp
DOI: 10.1007/s40618-020-01370-x · PMID: 32772324 · PMCID: PMC7415009

656. Vitamin D Deficiency and Outcome of COVID-19 Patients
Aleksandar Radujkovic, Theresa Hippchen, Shilpa Tiwari-Heckler, Saida Dreher, Monica Boxberger, Uta Merle
Nutrients (2020-09-10) https://doi.org/ghgfmp
DOI: 10.3390/nu12092757 · PMID: 32927735 · PMCID: PMC7551780

657. Impact of Vitamin D Deficiency on COVID-19—A Prospective Analysis from the CovILD Registry
Alex Pizzini, Magdalena Aichner, Sabina Sahanic, Anna Böhm, Alexander Egger, Gregor Hoermann, Katharina Kurz, Gerlig Widmann, Rosa Bellmann-Weiler, Günter Weiss, … Judith Löffler-Ragg
Nutrients (2020-09-11) https://doi.org/ghr95b
DOI: 10.3390/nu12092775 · PMID: 32932831 · PMCID: PMC7551662

658. Does Serum Vitamin D Level Affect COVID-19 Infection and Its Severity?-A Case-Control Study
Kun Ye, Fen Tang, Xin Liao, Benjamin A. Shaw, Meiqiu Deng, Guangyi Huang, Zhiqiang Qin, Xiaomei Peng, Hewei Xiao, Chunxia Chen, … Jianrong Yang
Journal of the American College of Nutrition (2020-10-13) https://doi.org/ghr935
DOI: 10.1080/07315724.2020.1826005 · PMID: 33048028

659. Lower levels of vitamin D are associated with SARS-CoV-2 infection and mortality in the Indian population: An observational study
Sunali Padhi, Subham Suvankar, Venketesh K. Panda, Abhijit Pati, Aditya K. Panda
International Immunopharmacology (2020-11) https://doi.org/ghr93w
DOI: 10.1016/j.intimp.2020.107001 · PMID: 33182040 · PMCID: PMC7489890

660. Vitamin D Deficiency Is Inversely Associated with COVID-19 Incidence and Disease Severity in Chinese People
Xia Luo, Qing Liao, Ying Shen, Huijun Li, Liming Cheng
The Journal of Nutrition (2021-01) https://doi.org/ghr939
DOI: 10.1093/jn/nxaa332 · PMID: 33188401

661. Vitamin D concentrations and COVID-19 infection in UK Biobank
Claire E. Hastie, Daniel F. Mackay, Frederick Ho, Carlos A. Celis-Morales, Srinivasa Vittal Katikireddi, Claire L. Niedzwiedz, Bhautesh D. Jani, Paul Welsh, Frances S. Mair, Stuart R. Gray, … Jill P. Pell
Diabetes & Metabolic Syndrome: Clinical Research & Reviews (2020-07) https://doi.org/ggvv72
DOI: 10.1016/j.dsx.2020.04.050 · PMID: 32413819 · PMCID: PMC7204679

662. Vitamin D and COVID-19 infection and mortality in UK Biobank
Claire E. Hastie, Jill P. Pell, Naveed Sattar
European Journal of Nutrition (2020-08-26) https://doi.org/ghr93p
DOI: 10.1007/s00394-020-02372-4 · PMID: 32851419 · PMCID: PMC7449523

663. Low serum 25‐hydroxyvitamin D (25[OH]D) levels in patients hospitalized with COVID‐19 are associated with greater disease severity
Grigorios Panagiotou, Su Ann Tee, Yasir Ihsan, Waseem Athar, Gabriella Marchitelli, Donna Kelly, Christopher S. Boot, Nadia Stock, James Macfarlane, Adrian R. Martineau, … Richard Quinton
Clinical Endocrinology (2020-08-06) https://doi.org/gg5gbj
DOI: 10.1111/cen.14276 · PMID: 32621392 · PMCID: PMC7361912

664. Letter in response to the article: Vitamin D concentrations and COVID-19 infection in UK biobank (Hastie et al.)
W. B. Grant, S. L. McDonnell
Diabetes & Metabolic Syndrome: Clinical Research & Reviews (2020-09) https://doi.org/ghc7p4
DOI: 10.1016/j.dsx.2020.05.046 · PMID: 32563941 · PMCID: PMC7293469

665. Vitamin D deficiency in African Americans is associated with a high risk of severe disease and mortality by SARS-CoV-2
Virna Margarita Martín Giménez, Felipe Inserra, León Ferder, Joxel García, Walter Manucha
Journal of Human Hypertension (2020-08-13) https://doi.org/ghr933
DOI: 10.1038/s41371-020-00398-z · PMID: 32792611 · PMCID: PMC7425793

666. Evidence for possible association of vitamin D status with cytokine storm and unregulated inflammation in COVID-19 patients
Ali Daneshkhah, Vasundhara Agrawal, Adam Eshein, Hariharan Subramanian, Hemant Kumar Roy, Vadim Backman
Aging Clinical and Experimental Research (2020-09-02) https://doi.org/ghr93q
DOI: 10.1007/s40520-020-01677-y · PMID: 32876941 · PMCID: PMC7465887

667. Short term, high-dose vitamin D supplementation for COVID-19 disease: a randomised, placebo-controlled, study (SHADE study)
Ashu Rastogi, Anil Bhansali, Niranjan Khare, Vikas Suri, Narayana Yaddanapudi, Naresh Sachdeva, GD Puri, Pankaj Malhotra
Postgraduate Medical Journal (2020-11-12) https://doi.org/ghnhpq
DOI: 10.1136/postgradmedj-2020-139065 · PMID: 33184146

668. “Effect of calcifediol treatment and best available therapy versus best available therapy on intensive care unit admission and mortality among patients hospitalized for COVID-19: A pilot randomized clinical study”
Marta Entrenas Castillo, Luis Manuel Entrenas Costa, José Manuel Vaquero Barrios, Juan Francisco Alcalá Díaz, José López Miranda, Roger Bouillon, José Manuel Quesada Gomez
The Journal of Steroid Biochemistry and Molecular Biology (2020-10) https://doi.org/ghd79r
DOI: 10.1016/j.jsbmb.2020.105751 · PMID: 32871238 · PMCID: PMC7456194

669. COVID-19 rapid evidence summary: vitamin D for COVID-19 | Advice | NICE https://www.nice.org.uk/advice/es28

670. Mathematical analysis of Córdoba calcifediol trial suggests strong role for Vitamin D in reducing ICU admissions of hospitalized COVID-19 patients
Irwin Jungreis, Manolis Kellis
Cold Spring Harbor Laboratory (2020-12-21) https://doi.org/ghr94h
DOI: 10.1101/2020.11.08.20222638

671. High-Dose Cholecalciferol Booster Therapy is Associated with a Reduced Risk of Mortality in Patients with COVID-19: A Cross-Sectional Multi-Centre Observational Study
Stephanie F. Ling, Eleanor Broad, Rebecca Murphy, Joseph M. Pappachan, Satveer Pardesi-Newton, Marie-France Kong, Edward B. Jude
Nutrients (2020-12-11) https://doi.org/ghr95f
DOI: 10.3390/nu12123799 · PMID: 33322317 · PMCID: PMC7763301

672. Effect of Vitamin D 3 Supplementation vs Placebo on Hospital Length of Stay in Patients with Severe COVID-19: A Multicenter, Double-blind, Randomized Controlled Trial
Igor H. Murai, Alan L. Fernandes, Lucas P. Sales, Ana J. Pinto, Karla F. Goessler, Camila S. C. Duran, Carla B. R. Silva, André S. Franco, Marina B. Macedo, Henrique H. H. Dalmolin, … Rosa M. R. Pereira
Cold Spring Harbor Laboratory (2020-11-17) https://doi.org/ghr94j
DOI: 10.1101/2020.11.16.20232397

673. Effect of Vitamin D Administration on Prevention and Treatment of Mild Forms of Suspected Covid-19
Manuel Castillo Garzón
clinicaltrials.gov (2020-04-03) https://clinicaltrials.gov/ct2/show/NCT04334005

674. Improving Vitamin D Status in the Management of COVID-19
Aldo Montano-Loza
clinicaltrials.gov (2020-06-03) https://clinicaltrials.gov/ct2/show/NCT04385940

675. Cholecalciferol to Improve the Outcomes of COVID-19 Patients - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04411446

676. COvid-19 and Vitamin D Supplementation: a Multicenter Randomized Controlled Trial of High Dose Versus Standard Dose Vitamin D3 in High-risk COVID-19 Patients (CoVitTrial) - Full Text View - ClinicalTrials.gov https://clinicaltrials.gov/ct2/show/NCT04344041

677. The LEAD COVID-19 Trial: Low-risk, Early Aspirin and Vitamin D to Reduce COVID-19 Hospitalizations
Louisiana State University Health Sciences Center in New Orleans
clinicaltrials.gov (2020-04-24) https://clinicaltrials.gov/ct2/show/NCT04363840

678. Randomized Double-Blind Placebo-Controlled Proof-of-Concept Trial of a Plant Polyphenol for the Outpatient Treatment of Mild Coronavirus Disease (COVID-19)
Marvin McCreary MD
clinicaltrials.gov (2020-09-22) https://clinicaltrials.gov/ct2/show/NCT04400890

679. Current vitamin D status in European and Middle East countries and strategies to prevent vitamin D deficiency: a position statement of the European Calcified Tissue Society
Paul Lips, Kevin D Cashman, Christel Lamberg-Allardt, Heike Annette Bischoff-Ferrari, Barbara Obermayer-Pietsch, Maria Luisa Bianchi, Jan Stepan, Ghada El-Hajj Fuleihan, Roger Bouillon
European Journal of Endocrinology (2019-04) https://doi.org/ggr42p
DOI: 10.1530/eje-18-0736 · PMID: 30721133

680. Communiqué de l’Académie nationale de Médecine : Vitamine D et Covid-19 – Académie nationale de médecine | Une institution dans son temps https://www.academie-medecine.fr/communique-de-lacademie-nationale-de-medecine-vitamine-d-et-covid-19/

681. COVID-19 and African Americans
Clyde W. Yancy
JAMA (2020-05-19) https://doi.org/ggv494
DOI: 10.1001/jama.2020.6548 · PMID: 32293639

682. Covid-19: Public health agencies review whether vitamin D supplements could reduce risk
Ingrid Torjesen
BMJ (2020-06-19) https://doi.org/ghr94p
DOI: 10.1136/bmj.m2475 · PMID: 32561509

683. Avoidance of vitamin D deficiency to slow the COVID-19 pandemic
Martin Kohlmeier
BMJ Nutrition, Prevention & Health (2020-06) https://doi.org/ghr94q
DOI: 10.1136/bmjnph-2020-000096 · PMID: 33230496 · PMCID: PMC7295862

684. COVID-19 rapid guideline: vitamin D
National Institute for Health and Care Excellence (NICE)
https://www.nice.org.uk/guidance/ng187/resources/covid19-rapid-guideline-vitamin-d-pdf-66142026720709

685. Covid-19: NHS bosses told to assess risk to ethnic minority staff who may be at greater risk
Gareth Iacobucci
BMJ (2020-05-04) https://doi.org/ggv2zq
DOI: 10.1136/bmj.m1820 · PMID: 32366503

686. Prevalence of Vitamin D Deficiency and Associated Risk Factors in the US Population (2011-2012)
Naveen R Parva, Satish Tadepalli, Pratiksha Singh, Andrew Qian, Rajat Joshi, Hyndavi Kandala, Vinod K Nookala, Pramil Cheriyath
Cureus (2018-06-05) https://doi.org/gg7kqq
DOI: 10.7759/cureus.2741 · PMID: 30087817 · PMCID: PMC6075634

687. Vitamin D
COVID-19 Treatment Guidelines
https://www.covid19treatmentguidelines.nih.gov/adjunctive-therapy/vitamin-d/

688. Court Orders Georgia Defendants to Stop Selling Vitamin D Products as Treatments for Covid-19 and Other Diseases (2021-01-08) https://www.justice.gov/opa/pr/court-orders-georgia-defendants-stop-selling-vitamin-d-products-treatments-covid-19-and-other

689. The International Scientific Association for Probiotics and Prebiotics consensus statement on the scope and appropriate use of the term probiotic
Colin Hill, Francisco Guarner, Gregor Reid, Glenn R. Gibson, Daniel J. Merenstein, Bruno Pot, Lorenzo Morelli, Roberto Berni Canani, Harry J. Flint, Seppo Salminen, … Mary Ellen Sanders
Nature Reviews Gastroenterology & Hepatology (2014-06-10) https://doi.org/f6ndv7
DOI: 10.1038/nrgastro.2014.66 · PMID: 24912386

690. The Effect of Probiotics on Prevention of Common Cold: A Meta-Analysis of Randomized Controlled Trial Studies
En-Jin Kang, Soo Young Kim, In-Hong Hwang, Yun-Jeong Ji
Korean Journal of Family Medicine (2013) https://doi.org/gg3knf
DOI: 10.4082/kjfm.2013.34.1.2 · PMID: 23372900 · PMCID: PMC3560336

691. Probiotics and Paraprobiotics in Viral Infection: Clinical Application and Effects on the Innate and Acquired Immune Systems
Osamu Kanauchi, Akira Andoh, Sazaly AbuBakar, Naoki Yamamoto
Current Pharmaceutical Design (2018-05-10) https://doi.org/gdjnpk
DOI: 10.2174/1381612824666180116163411 · PMID: 29345577 · PMCID: PMC6006794

692. Using Probiotics to Flatten the Curve of Coronavirus Disease COVID-2019 Pandemic
David Baud, Varvara Dimopoulou Agri, Glenn R. Gibson, Gregor Reid, Eric Giannoni
Frontiers in Public Health (2020-05-08) https://doi.org/gg3knd
DOI: 10.3389/fpubh.2020.00186 · PMID: 32574290 · PMCID: PMC7227397

693. Next-generation probiotics: the spectrum from probiotics to live biotherapeutics
Paul W. O’Toole, Julian R. Marchesi, Colin Hill
Nature Microbiology (2017-04-25) https://doi.org/ggzggv
DOI: 10.1038/nmicrobiol.2017.57 · PMID: 28440276

694. Mechanisms of Action of Probiotics
Julio Plaza-Diaz, Francisco Javier Ruiz-Ojeda, Mercedes Gil-Campos, Angel Gil
Advances in Nutrition (2019-01) https://doi.org/gft8sh
DOI: 10.1093/advances/nmy063 · PMID: 30721959 · PMCID: PMC6363529

695. Probiotic mechanisms of action
Katrina Halloran, Mark A. Underwood
Early Human Development (2019-08) https://doi.org/gg3jc4
DOI: 10.1016/j.earlhumdev.2019.05.010 · PMID: 31174927

696. Probiotic Mechanisms of Action
Miriam Bermudez-Brito, Julio Plaza-Díaz, Sergio Muñoz-Quezada, Carolina Gómez-Llorente, Angel Gil
Annals of Nutrition and Metabolism (2012) https://doi.org/gg3knb
DOI: 10.1159/000342079 · PMID: 23037511

697. A novel eukaryotic cell culture model to study antiviral activity of potential probiotic bacteria
T BOTIC, T KLINGBERG, H WEINGARTL, A CENCIC
International Journal of Food Microbiology (2007-04-30) https://doi.org/fks5cz
DOI: 10.1016/j.ijfoodmicro.2006.10.044 · PMID: 17261339

698. Oral administration of Lactobacillus brevis KB290 to mice alleviates clinical symptoms following influenza virus infection
N. Waki, N. Yajima, H. Suganuma, B. M. Buddle, D. Luo, A. Heiser, T. Zheng
Letters in Applied Microbiology (2014-01) https://doi.org/f5j37w
DOI: 10.1111/lam.12160 · PMID: 24329975

699. Antiviral activity of Lactobacillus brevis towards herpes simplex virus type 2: Role of cell wall associated components
Paola Mastromarino, Fatima Cacciotti, Alessandra Masci, Luciana Mosca
Anaerobe (2011-12) https://doi.org/bcpvm5
DOI: 10.1016/j.anaerobe.2011.04.022 · PMID: 21621625

700. Critical Adverse Impact of IL-6 in Acute Pneumovirus Infection
Caroline M. Percopo, Michelle Ma, Todd A. Brenner, Julia O. Krumholz, Timothy J. Break, Karen Laky, Helene F. Rosenberg
The Journal of Immunology (2019-02-01) https://doi.org/ghr95h
DOI: 10.4049/jimmunol.1800927 · PMID: 30578308 · PMCID: PMC6365009

701. Antiviral Activity of Exopolysaccharides Produced by Lactic Acid Bacteria of the Genera Pediococcus, Leuconostoc and Lactobacillus against Human Adenovirus Type 5
Biliavska, Pankivska, Povnitsa, Zagorodnya
Medicina (2019-08-22) https://doi.org/ghr948
DOI: 10.3390/medicina55090519 · PMID: 31443536 · PMCID: PMC6780409

702. Prevention of respiratory syncytial virus infection with probiotic lactic acid bacterium Lactobacillus gasseri SBT2055
Kei Eguchi, Naoki Fujitani, Hisako Nakagawa, Tadaaki Miyazaki
Scientific Reports (2019-03-18) https://doi.org/ghr934
DOI: 10.1038/s41598-019-39602-7 · PMID: 30886158 · PMCID: PMC6423325

703. Effect of probiotic on innate inflammatory response and viral shedding in experimental rhinovirus infection – a randomised controlled trial
R. B. Turner, J. A. Woodfolk, L. Borish, J. W. Steinke, J. T. Patrie, L. M. Muehling, S. Lahtinen, M. J. Lehtinen
Beneficial Microbes (2017-04-26) https://doi.org/f955fh
DOI: 10.3920/bm2016.0160 · PMID: 28343401 · PMCID: PMC5797652

704. Immunobiotic lactobacilli reduce viral-associated pulmonary damage through the modulation of inflammation–coagulation interactions
Hortensia Zelaya, Kohichiro Tsukida, Eriko Chiba, Gabriela Marranzino, Susana Alvarez, Haruki Kitazawa, Graciela Agüero, Julio Villena
International Immunopharmacology (2014-03) https://doi.org/f5wd93
DOI: 10.1016/j.intimp.2013.12.020 · PMID: 24394565

705. Nasal priming with immunobiotic lactobacilli improves the adaptive immune response against influenza virus
Fernanda Raya Tonetti, Md. Aminul Islam, Maria Guadalupe Vizoso-Pinto, Hideki Takahashi, Haruki Kitazawa, Julio Villena
International Immunopharmacology (2020-01) https://doi.org/ghr93v
DOI: 10.1016/j.intimp.2019.106115 · PMID: 31841753

706. The potential application of probiotics and prebiotics for the prevention and treatment of COVID-19
Amin N. Olaimat, Iman Aolymat, Murad Al-Holy, Mutamed Ayyash, Mahmoud Abu Ghoush, Anas A. Al-Nabulsi, Tareq Osaili, Vasso Apostolopoulos, Shao-Quan Liu, Nagendra P. Shah
npj Science of Food (2020-10-05) https://doi.org/ghggq4
DOI: 10.1038/s41538-020-00078-9 · PMID: 33083549 · PMCID: PMC7536434

707. Pulmonary-intestinal cross-talk in mucosal inflammatory disease
S Keely, NJ Talley, PM Hansbro
Mucosal Immunology (2011-11-16) https://doi.org/b5knk2
DOI: 10.1038/mi.2011.55 · PMID: 22089028 · PMCID: PMC3243663

708. The role of the lung microbiota and the gut-lung axis in respiratory infectious diseases
Alexia Dumas, Lucie Bernard, Yannick Poquet, Geanncarlo Lugo-Villarino, Olivier Neyrolles
Cellular Microbiology (2018-12) https://doi.org/gfjds9
DOI: 10.1111/cmi.12966 · PMID: 30329198

709. Gut microbiota and Covid-19- possible link and implications
Debojyoti Dhar, Abhishek Mohanty
Virus Research (2020-08) https://doi.org/gg3jc5
DOI: 10.1016/j.virusres.2020.198018 · PMID: 32430279 · PMCID: PMC7217790

710. Oral Microbiome and SARS-CoV-2: Beware of Lung Co-infection
Lirong Bao, Cheng Zhang, Jiajia Dong, Lei Zhao, Yan Li, Jianxun Sun
Frontiers in Microbiology (2020-07-31) https://doi.org/ghr944
DOI: 10.3389/fmicb.2020.01840 · PMID: 32849438 · PMCID: PMC7411080

711. Lung microbiome and coronavirus disease 2019 (COVID-19): Possible link and implications
Saroj Khatiwada, Astha Subedi
Human Microbiome Journal (2020-08) https://doi.org/gg7m83
DOI: 10.1016/j.humic.2020.100073 · PMID: 32835135 · PMCID: PMC7405772

712. Probiotics in respiratory virus infections
L. Lehtoranta, A. Pitkäranta, R. Korpela
European Journal of Clinical Microbiology & Infectious Diseases (2014-03-18) https://doi.org/f583jr
DOI: 10.1007/s10096-014-2086-y · PMID: 24638909 · PMCID: PMC7088122

713. Probiotics for preventing acute upper respiratory tract infections
Qiukui Hao, Bi Rong Dong, Taixiang Wu
Cochrane Database of Systematic Reviews (2015-02-03) https://doi.org/gg3jc3
DOI: 10.1002/14651858.cd006895.pub3 · PMID: 25927096

714. Probiotics for the prevention of respiratory tract infections: a systematic review
Evridiki K. Vouloumanou, Gregory C. Makris, Drosos E. Karageorgopoulos, Matthew E. Falagas
International Journal of Antimicrobial Agents (2009-09) https://doi.org/dn8kw8
DOI: 10.1016/j.ijantimicag.2008.11.005 · PMID: 19179052

715. Effectiveness of probiotics on the duration of illness in healthy children and adults who develop common acute respiratory infectious conditions: a systematic review and meta-analysis
Sarah King, Julie Glanville, Mary Ellen Sanders, Anita Fitzgerald, Danielle Varley
British Journal of Nutrition (2014-04-29) https://doi.org/f57hq5
DOI: 10.1017/s0007114514000075 · PMID: 24780623 · PMCID: PMC4054664

716. Effect of probiotics on the incidence of ventilator-associated pneumonia in critically ill patients: a randomized controlled multicenter trial
Juan Zeng, Chun-Ting Wang, Fu-Shen Zhang, Feng Qi, Shi-Fu Wang, Shuang Ma, Tie-Jun Wu, Hui Tian, Zhao-Tao Tian, Shu-Liu Zhang, … Yu-Ping Wang
Intensive Care Medicine (2016-04-04) https://doi.org/f8jnrt
DOI: 10.1007/s00134-016-4303-x · PMID: 27043237

717. Probiotic Prophylaxis of Ventilator-associated Pneumonia
Lee E. Morrow, Marin H. Kollef, Thomas B. Casale
American Journal of Respiratory and Critical Care Medicine (2010-10-15) https://doi.org/d5hh4t
DOI: 10.1164/rccm.200912-1853oc · PMID: 20522788 · PMCID: PMC2970846

718. Synbiotics modulate gut microbiota and reduce enteritis and ventilator-associated pneumonia in patients with sepsis: a randomized controlled trial
Kentaro Shimizu, Tomoki Yamada, Hiroshi Ogura, Tomoyoshi Mohri, Takeyuki Kiguchi, Satoshi Fujimi, Takashi Asahara, Tomomi Yamada, Masahiro Ojima, Mitsunori Ikeda, Takeshi Shimazu
Critical Care (2018-09-27) https://doi.org/gfdggj
DOI: 10.1186/s13054-018-2167-x · PMID: 30261905 · PMCID: PMC6161427

719. Probiotics for the Prevention of Ventilator-Associated Pneumonia: A Meta-Analysis of Randomized Controlled Trials
Minmin Su, Ying Jia, Yan Li, Dianyou Zhou, Jinsheng Jia
Respiratory Care (2020-05) https://doi.org/gg3kng
DOI: 10.4187/respcare.07097 · PMID: 32127415

720. COVID-19: An Alert to Ventilator-Associated Bacterial Pneumonia
Helvécio Cardoso Corrêa Póvoa, Gabriela Ceccon Chianca, Natalia Lopes Pontes Póvoa Iorio
Adis Journals (2020) https://doi.org/gg3knh
DOI: 10.6084/m9.figshare.12340496

721. The challenge of ventilator-associated pneumonia diagnosis in COVID-19 patients
Bruno François, Pierre-François Laterre, Charles-Edouard Luyt, Jean Chastre
Critical Care (2020-06-05) https://doi.org/gg3knc
DOI: 10.1186/s13054-020-03013-2 · PMID: 32503590 · PMCID: PMC7273812

722. Prophylactic use of probiotics for gastrointestinal disorders in children
Celine Perceval, Hania Szajewska, Flavia Indrio, Zvi Weizman, Yvan Vandenplas
The Lancet Child & Adolescent Health (2019-09) https://doi.org/d2qp
DOI: 10.1016/s2352-4642(19)30182-8

723. Effect of Gastrointestinal Symptoms in Patients With COVID-19
Zili Zhou, Ning Zhao, Yan Shu, Shengbo Han, Bin Chen, Xiaogang Shu
Gastroenterology (2020-06) https://doi.org/ggq8x8
DOI: 10.1053/j.gastro.2020.03.020 · PMID: 32199880 · PMCID: PMC7270807

724. The digestive system is a potential route of 2019-nCov infection: a bioinformatics analysis based on single-cell transcriptomes
Hao Zhang, Zijian Kang, Haiyi Gong, Da Xu, Jing Wang, Zifu Li, Xingang Cui, Jianru Xiao, Tong Meng, Wang Zhou, … Huji Xu
Cold Spring Harbor Laboratory (2020-01-31) https://doi.org/ggjvx2
DOI: 10.1101/2020.01.30.927806

725. Cryo-electron microscopy structures of the SARS-CoV spike glycoprotein reveal a prerequisite conformational state for receptor binding
Miao Gui, Wenfei Song, Haixia Zhou, Jingwei Xu, Silian Chen, Ye Xiang, Xinquan Wang
Cell Research (2016-12-23) https://doi.org/f9m247
DOI: 10.1038/cr.2016.152 · PMID: 28008928 · PMCID: PMC5223232

726. Prolonged presence of SARS-CoV-2 viral RNA in faecal samples
Yongjian Wu, Cheng Guo, Lantian Tang, Zhongsi Hong, Jianhui Zhou, Xin Dong, Huan Yin, Qiang Xiao, Yanping Tang, Xiujuan Qu, … Xi Huang
The Lancet Gastroenterology & Hepatology (2020-05) https://doi.org/ggq8zp
DOI: 10.1016/s2468-1253(20)30083-2 · PMID: 32199469 · PMCID: PMC7158584

727. Enteric involvement of coronaviruses: is faecal–oral transmission of SARS-CoV-2 possible?
Charleen Yeo, Sanghvi Kaushal, Danson Yeo
The Lancet Gastroenterology & Hepatology (2020-04) https://doi.org/ggpx7s
DOI: 10.1016/s2468-1253(20)30048-0 · PMID: 32087098 · PMCID: PMC7130008

728. Characteristics of pediatric SARS-CoV-2 infection and potential evidence for persistent fecal viral shedding
Yi Xu, Xufang Li, Bing Zhu, Huiying Liang, Chunxiao Fang, Yu Gong, Qiaozhi Guo, Xin Sun, Danyang Zhao, Jun Shen, … Sitang Gong
Nature Medicine (2020-03-13) https://doi.org/ggpwx5
DOI: 10.1038/s41591-020-0817-4 · PMID: 32284613 · PMCID: PMC7095102

729. Modulation of rotavirus severe gastroenteritis by the combination of probiotics and prebiotics
Guadalupe Gonzalez-Ochoa, Lilian K. Flores-Mendoza, Ramona Icedo-Garcia, Ricardo Gomez-Flores, Patricia Tamez-Guerra
Archives of Microbiology (2017-06-20) https://doi.org/gbsb4d
DOI: 10.1007/s00203-017-1400-3 · PMID: 28634691 · PMCID: PMC5548957

730. Multicenter Trial of a Combination Probiotic for Children with Gastroenteritis
Stephen B. Freedman, Sarah Williamson-Urquhart, Ken J. Farion, Serge Gouin, Andrew R. Willan, Naveen Poonai, Katrina Hurley, Philip M. Sherman, Yaron Finkelstein, Bonita E. Lee, … Suzanne Schuh
New England Journal of Medicine (2018-11-22) https://doi.org/gfkbsf
DOI: 10.1056/nejmoa1802597 · PMID: 30462939

731. Synbiotic Therapy of Gastrointestinal Symptoms During Covid-19 Infection: A Randomized, Double-blind, Placebo Controlled, Telemedicine Study (SynCov Study)
Medical University of Graz
clinicaltrials.gov (2021-01-14) https://clinicaltrials.gov/ct2/show/NCT04420676

732. Multicentric Study to Assess the Effect of Consumption of Lactobacillus Coryniformis K8 on Healthcare Personnel Exposed to COVID-19
Biosearch S.A.
clinicaltrials.gov (2020-04-28) https://clinicaltrials.gov/ct2/show/NCT04366180

733. The Intestinal Microbiota as a Therapeutic Target in Hospitalized Patients With COVID-19 Infection
Bioithas SL
clinicaltrials.gov (2020-05-14) https://clinicaltrials.gov/ct2/show/NCT04390477

734. Probiotics: definition, scope and mechanisms of action
Gregor Reid
Best Practice & Research Clinical Gastroenterology (2016-02) https://doi.org/f8m79k
DOI: 10.1016/j.bpg.2015.12.001 · PMID: 27048893

735. Health benefits and health claims of probiotics: bridging science and marketing
Ger T. Rijkers, Willem M. de Vos, Robert-Jan Brummer, Lorenzo Morelli, Gerard Corthier, Philippe Marteau
British Journal of Nutrition (2011-08-24) https://doi.org/cb78rx
DOI: 10.1017/s000711451100287x · PMID: 21861940

736. Probiotics and COVID-19: one size does not fit all
Joyce WY Mak, Francis KL Chan, Siew C Ng
The Lancet Gastroenterology & Hepatology (2020-07) https://doi.org/d2qq
DOI: 10.1016/s2468-1253(20)30122-9 · PMID: 32339473 · PMCID: PMC7182525

737. Network-based drug repurposing for novel coronavirus 2019-nCoV/SARS-CoV-2
Yadi Zhou, Yuan Hou, Jiayu Shen, Yin Huang, William Martin, Feixiong Cheng
Cell Discovery (2020-03-16) https://doi.org/ggq84x
DOI: 10.1038/s41421-020-0153-3 · PMID: 32194980 · PMCID: PMC7073332

738. Role of Melatonin on Virus-Induced Neuropathogenesis—A Concomitant Therapeutic Strategy to Understand SARS-CoV-2 Infection
Prapimpun Wongchitrat, Mayuri Shukla, Ramaswamy Sharma, Piyarat Govitrapong, Russel J. Reiter
Antioxidants (2021-01-02) https://doi.org/ghr946
DOI: 10.3390/antiox10010047 · PMID: 33401749

739. Nutraceutical Strategies for Suppressing NLRP3 Inflammasome Activation: Pertinence to the Management of COVID-19 and Beyond
Mark F. McCarty, Simon Bernard Iloki Assanga, Lidianys Lewis Luján, James H. O’Keefe, James J. DiNicolantonio
Nutrients (2020-12-25) https://doi.org/ghr95g
DOI: 10.3390/nu13010047 · PMID: 33375692

740. Update: Here’s what is known about Trump’s COVID-19 treatment
Jon Cohen
Science (2020-10-05) https://doi.org/ghr94n
DOI: 10.1126/science.abf0974

741. Dietary supplements during the COVID-19 pandemic: insights from 1.4M users of the COVID Symptom Study app - a longitudinal app-based community survey
Panayiotis Louca, Benjamin Murray, Kerstin Klaser, Mark S Graham, Mohsen Mazidi, Emily R Leeming, Ellen Thompson, Ruth Bowyer, David A Drew, Long H Nguyen, … Cristina Menni
Cold Spring Harbor Laboratory (2020-11-30) https://doi.org/ghr94k
DOI: 10.1101/2020.11.27.20239087

742. Vitamin D deficiency aggravates COVID-19: systematic review and meta-analysis
Marcos Pereira, Alialdo Dantas Damascena, Laylla Mirella Galvão Azevedo, Tarcio de Almeida Oliveira, Jerusa da Mota Santana
Critical Reviews in Food Science and Nutrition (2020-11-04) https://doi.org/ghr937
DOI: 10.1080/10408398.2020.1841090 · PMID: 33146028

743. ESPEN expert statements and practical guidance for nutritional management of individuals with SARS-CoV-2 infection
Rocco Barazzoni, Stephan C. Bischoff, Joao Breda, Kremlin Wickramasinghe, Zeljko Krznaric, Dorit Nitzan, Matthias Pirlich, Pierre Singer
Clinical Nutrition (2020-06) https://doi.org/ggtzjq
DOI: 10.1016/j.clnu.2020.03.022 · PMID: 32305181 · PMCID: PMC7138149

744. Nutritional status assessment in patients with Covid-19 after discharge from the intensive care unit
Nassim Essabah Haraj, Siham El Aziz, Asma Chadli, Asma Dafir, Amal Mjabber, Ouissal Aissaoui, Lhoucine Barrou, Chafik El Kettani El Hamidi, Afak Nsiri, Rachid AL Harrar, … Moulay Hicham Afif
Clinical Nutrition ESPEN (2020-10) https://doi.org/ghjhdq
DOI: 10.1016/j.clnesp.2020.09.214 · PMCID: PMC7552965

745. Nutrition Status Affects COVID‐19 Patient Outcomes
Mette M Berger
Journal of Parenteral and Enteral Nutrition (2020-07-15) https://doi.org/gg5qv4
DOI: 10.1002/jpen.1954 · PMID: 32613691 · PMCID: PMC7361441

746. Evaluation of Nutrition Risk and Its Association With Mortality Risk in Severely and Critically Ill COVID‐19 Patients
Xiaobo Zhao, Yan Li, Yanyan Ge, Yuxin Shi, Ping Lv, Jianchu Zhang, Gui Fu, Yanfen Zhou, Ke Jiang, Nengxing Lin, … Xin Li
Journal of Parenteral and Enteral Nutrition (2020-07-20) https://doi.org/ghr93n
DOI: 10.1002/jpen.1953 · PMID: 32613660 · PMCID: PMC7361906

747. Multisystem inflammatory syndrome in children: A systematic review
Mubbasheer Ahmed, Shailesh Advani, Axel Moreira, Sarah Zoretic, John Martinez, Kevin Chorath, Sebastian Acosta, Rija Naqvi, Finn Burmeister-Morton, Fiona Burmeister, … Alvaro Moreira
EClinicalMedicine (2020-09) https://doi.org/ghsv27
DOI: 10.1016/j.eclinm.2020.100527 · PMID: 32923992 · PMCID: PMC7473262

748. Nutritional management of COVID-19 patients in a rehabilitation unit
Luigia Brugliera, Alfio Spina, Paola Castellazzi, Paolo Cimino, Pietro Arcuri, Alessandra Negro, Elise Houdayer, Federica Alemanno, Alessandra Giordani, Pietro Mortini, Sandro Iannaccone
European Journal of Clinical Nutrition (2020-05-20) https://doi.org/gg29hf
DOI: 10.1038/s41430-020-0664-x · PMID: 32433599 · PMCID: PMC7237874

749. The frontier between nutrition and pharma: The international regulatory framework of functional foods, food supplements and nutraceuticals
Laura Domínguez Díaz, Virginia Fernández-Ruiz, Montaña Cámara
Critical Reviews in Food Science and Nutrition (2019-03-29) https://doi.org/ggqs3w
DOI: 10.1080/10408398.2019.1592107 · PMID: 30924346

750. Coronavirus Update: FDA and FTC Warn Seven Companies Selling Fraudulent Products that Claim to Treat or Prevent COVID-19
Office of the Commissioner
FDA (2020-03-27) https://www.fda.gov/news-events/press-announcements/coronavirus-update-fda-and-ftc-warn-seven-companies-selling-fraudulent-products-claim-treat-or

751. Coronavirus Disease 2019 (COVID-19) – Prevention & Treatment
CDC
Centers for Disease Control and Prevention (2020-12-31) https://www.cdc.gov/coronavirus/2019-ncov/prevent-getting-sick/prevention.html

752. Potential roles of social distancing in mitigating the spread of coronavirus disease 2019 (COVID-19) in South Korea
Sang Woo Park, Kaiyuan Sun, Cécile Viboud, Bryan T. Grenfell, Jonathan Dushoff
Cold Spring Harbor Laboratory (2020-03-30) https://doi.org/gg3mhg
DOI: 10.1101/2020.03.27.20045815 · PMID: 32511429 · PMCID: PMC7217070

753. Evaluating the Effectiveness of Social Distancing Interventions to Delay or Flatten the Epidemic Curve of Coronavirus Disease
Laura Matrajt, Tiffany Leung
Emerging Infectious Diseases (2020-08) https://doi.org/ggtx3k
DOI: 10.3201/eid2608.201093 · PMID: 32343222 · PMCID: PMC7392458

754. Association of Race With Mortality Among Patients Hospitalized With Coronavirus Disease 2019 (COVID-19) at 92 US Hospitals
Baligh R. Yehia, Angela Winegar, Richard Fogel, Mohamad Fakih, Allison Ottenbacher, Christine Jesser, Angelo Bufalino, Ren-Huai Huang, Joseph Cacchione
JAMA Network Open (2020-08-18) https://doi.org/ghcspt
DOI: 10.1001/jamanetworkopen.2020.18039 · PMID: 32809033 · PMCID: PMC7435340

755. Characteristics of and Important Lessons From the Coronavirus Disease 2019 (COVID-19) Outbreak in China
Zunyou Wu, Jennifer M. McGoogan
JAMA (2020-04-07) https://doi.org/ggmq43
DOI: 10.1001/jama.2020.2648 · PMID: 32091533

756. Critical Care Utilization for the COVID-19 Outbreak in Lombardy, Italy
Giacomo Grasselli, Antonio Pesenti, Maurizio Cecconi
JAMA (2020-04-28) https://doi.org/ggqf6g
DOI: 10.1001/jama.2020.4031 · PMID: 32167538

757. Hospitalization Rates and Characteristics of Patients Hospitalized with Laboratory-Confirmed Coronavirus Disease 2019 — COVID-NET, 14 States, March 1–30, 2020
Shikha Garg, Lindsay Kim, Michael Whitaker, Alissa O’Halloran, Charisse Cummings, Rachel Holstein, Mila Prill, Shua J. Chai, Pam D. Kirley, Nisha B. Alden, … Alicia Fry
MMWR. Morbidity and Mortality Weekly Report (2020-04-17) https://doi.org/ggsppz
DOI: 10.15585/mmwr.mm6915e3 · PMID: 32298251 · PMCID: PMC7755063

758. COVID-19 and Your Health
CDC
Centers for Disease Control and Prevention (2020-02-11) https://www.cdc.gov/coronavirus/2019-ncov/need-extra-precautions/older-adults.html

759. Disparities In Outcomes Among COVID-19 Patients In A Large Health Care System In California
Kristen M. J. Azar, Zijun Shen, Robert J. Romanelli, Stephen H. Lockhart, Kelly Smits, Sarah Robinson, Stephanie Brown, Alice R. Pressman
Health Affairs (2020-07-01) https://doi.org/ggx4mf
DOI: 10.1377/hlthaff.2020.00598 · PMID: 32437224

760. Characteristics Associated with Hospitalization Among Patients with COVID-19 — Metropolitan Atlanta, Georgia, March–April 2020
Marie E. Killerby, Ruth Link-Gelles, Sarah C. Haight, Caroline A. Schrodt, Lucinda England, Danica J. Gomes, Mays Shamout, Kristen Pettrone, Kevin O’Laughlin, Anne Kimball, … CDC COVID-19 Response Clinical Team
MMWR. Morbidity and Mortality Weekly Report (2020-06-26) https://doi.org/gg3k6h
DOI: 10.15585/mmwr.mm6925e1 · PMID: 32584797 · PMCID: PMC7316317

761. Demographic science aids in understanding the spread and fatality rates of COVID-19
Jennifer Beam Dowd, Liliana Andriano, David M. Brazel, Valentina Rotondi, Per Block, Xuejie Ding, Yan Liu, Melinda C. Mills
Proceedings of the National Academy of Sciences (2020-05-05) https://doi.org/ggsd5b
DOI: 10.1073/pnas.2004911117 · PMID: 32300018 · PMCID: PMC7211934

762. ‐19 and Older Adults: What We Know
Zainab Shahid, Ricci Kalayanamitra, Brendan McClafferty, Douglas Kepko, Devyani Ramgobin, Ravi Patel, Chander Shekher Aggarwal, Ramarao Vunnam, Nitasa Sahu, Dhirisha Bhatt, … Rohit Jain
Journal of the American Geriatrics Society (2020-04-20) https://doi.org/ggxgsb
DOI: 10.1111/jgs.16472 · PMID: 32255507 · PMCID: PMC7262251

763. Features of 20 133 UK patients in hospital with covid-19 using the ISARIC WHO Clinical Characterisation Protocol: prospective observational cohort study
Annemarie B Docherty, Ewen M Harrison, Christopher A Green, Hayley E Hardwick, Riinu Pius, Lisa Norman, Karl A Holden, Jonathan M Read, Frank Dondelinger, Gail Carson, … Malcolm G Semple
BMJ (2020-05-22) https://doi.org/ggw4nh
DOI: 10.1136/bmj.m1985 · PMID: 32444460 · PMCID: PMC7243036

764. Impact of sex and gender on COVID-19 outcomes in Europe
Catherine Gebhard, Vera Regitz-Zagrosek, Hannelore K. Neuhauser, Rosemary Morgan, Sabra L. Klein
Biology of Sex Differences (2020-05-25) https://doi.org/ghbvck
DOI: 10.1186/s13293-020-00304-9 · PMID: 32450906 · PMCID: PMC7247289

765. The Sex, Gender and COVID-19 Project | Global Health 50/50 https://globalhealth5050.org/the-sex-gender-and-covid-19-project/

766. Biological sex impacts COVID-19 outcomes
Sabra L. Klein, Santosh Dhakal, Rebecca L. Ursin, Sharvari Deshpande, Kathryn Sandberg, Franck Mauvais-Jarvis
PLOS Pathogens (2020-06-22) https://doi.org/gg3hwv
DOI: 10.1371/journal.ppat.1008570 · PMID: 32569293 · PMCID: PMC7307725

767. Sex-specific clinical characteristics and prognosis of coronavirus disease-19 infection in Wuhan, China: A retrospective study of 168 severe patients
Yifan Meng, Ping Wu, Wanrong Lu, Kui Liu, Ke Ma, Liang Huang, Jiaojiao Cai, Hong Zhang, Yu Qin, Haiying Sun, … Peng Wu
PLOS Pathogens (2020-04-28) https://doi.org/ggv3zn
DOI: 10.1371/journal.ppat.1008520 · PMID: 32343745 · PMCID: PMC7209966

768. Sex differences in renal angiotensin converting enzyme 2 (ACE2) activity are 17β-oestradiol-dependent and sex chromosome-independent
Jun Liu, Hong Ji, Wei Zheng, Xie Wu, Janet J Zhu, Arthur P Arnold, Kathryn Sandberg
Biology of Sex Differences (2010) https://doi.org/bbx3r6
DOI: 10.1186/2042-6410-1-6 · PMID: 21208466 · PMCID: PMC3010099

769. COVID-19 in nursing homes
A Fallon, T Dukelow, SP Kennelly, D O’Neill
QJM: An International Journal of Medicine (2020-06) https://doi.org/ggv4xx
DOI: 10.1093/qjmed/hcaa136 · PMID: 32311049 · PMCID: PMC7188176

770. Vulnerabilities to COVID-19 Among Transgender Adults in the U.S.
Jody L. Herman, Kathryn O’Neill
(2020-04-01) https://escholarship.org/uc/item/55t297mc

771. Association of Blood Glucose Control and Outcomes in Patients with COVID-19 and Pre-existing Type 2 Diabetes
Lihua Zhu, Zhi-Gang She, Xu Cheng, Juan-Juan Qin, Xiao-Jing Zhang, Jingjing Cai, Fang Lei, Haitao Wang, Jing Xie, Wenxin Wang, … Hongliang Li
Cell Metabolism (2020-06) https://doi.org/ggvcc9
DOI: 10.1016/j.cmet.2020.04.021 · PMID: 32369736 · PMCID: PMC7252168

772. Diabetes increases the mortality of patients with COVID-19: a meta-analysis
Zeng-hong Wu, Yun Tang, Qing Cheng
Acta Diabetologica (2020-06-24) https://doi.org/gg3k55
DOI: 10.1007/s00592-020-01546-0 · PMID: 32583078 · PMCID: PMC7311595

773. COVID-19 infection may cause ketosis and ketoacidosis
Juyi Li, Xiufang Wang, Jian Chen, Xiuran Zuo, Hongmei Zhang, Aiping Deng
Diabetes, Obesity and Metabolism (2020-05-18) https://doi.org/ggv4tm
DOI: 10.1111/dom.14057 · PMID: 32314455 · PMCID: PMC7264681

774. COVID-19 pandemic, coronaviruses, and diabetes mellitus
Ranganath Muniyappa, Sriram Gubbi
American Journal of Physiology-Endocrinology and Metabolism (2020-05-01) https://doi.org/ggq79v
DOI: 10.1152/ajpendo.00124.2020 · PMID: 32228322 · PMCID: PMC7191633

775. Severe obesity, increasing age and male sex are independently associated with worse in-hospital outcomes, and higher in-hospital mortality, in a cohort of patients with COVID-19 in the Bronx, New York
Leonidas Palaiodimos, Damianos G. Kokkinidis, Weijia Li, Dimitrios Karamanis, Jennifer Ognibene, Shitij Arora, William N. Southern, Christos S. Mantzoros
Metabolism (2020-07) https://doi.org/ggx229
DOI: 10.1016/j.metabol.2020.154262 · PMID: 32422233 · PMCID: PMC7228874

776. Features of 16,749 hospitalised UK patients with COVID-19 using the ISARIC WHO Clinical Characterisation Protocol
AB Docherty, EM Harrison, CA Green, H Hardwick, R Pius, L Norman, KA Holden, JM Read, F Dondelinger, G Carson, … ISARIC4C Investigators
Cold Spring Harbor Laboratory (2020-04-28) https://doi.org/ggtdtb
DOI: 10.1101/2020.04.23.20076042

777. When Two Pandemics Meet: Why Is Obesity Associated with Increased COVID-19 Mortality?
Sam M. Lockhart, Stephen O’Rahilly
Med (2020-12) https://doi.org/gg3k57
DOI: 10.1016/j.medj.2020.06.005 · PMID: 32838359 · PMCID: PMC7323660

778. Besides population age structure, health and other demographic factors can contribute to understanding the COVID-19 burden
Marília R. Nepomuceno, Enrique Acosta, Diego Alburez-Gutierrez, José Manuel Aburto, Alain Gagnon, Cássio M. Turra
Proceedings of the National Academy of Sciences (2020-06-23) https://doi.org/gg33qx
DOI: 10.1073/pnas.2008760117 · PMID: 32576710 · PMCID: PMC7322063

779. https://www.who.int/publications/i/item/report-of-the-who-china-joint-mission-on-coronavirus-disease-2019

780. The SOFA (Sepsis-related Organ Failure Assessment) score to describe organ dysfunction/failure
J. -L. Vincent, R. Moreno, J. Takala, S. Willatts, A. De Mendonça, H. Bruining, C. K. Reinhart, P. M. Suter, L. G. Thijs
Intensive Care Medicine (1996-07) https://doi.org/bpkxdw
DOI: 10.1007/bf01709751 · PMID: 8844239

781. The Third International Consensus Definitions for Sepsis and Septic Shock (Sepsis-3)
Mervyn Singer, Clifford S. Deutschman, Christopher Warren Seymour, Manu Shankar-Hari, Djillali Annane, Michael Bauer, Rinaldo Bellomo, Gordon R. Bernard, Jean-Daniel Chiche, Craig M. Coopersmith, … Derek C. Angus
JAMA (2016-02-23) https://doi.org/gdrcdh
DOI: 10.1001/jama.2016.0287 · PMID: 26903338 · PMCID: PMC4968574

782. COVID-19 and Racial/Ethnic Disparities
Monica Webb Hooper, Anna María Nápoles, Eliseo J. Pérez-Stable
JAMA (2020-06-23) https://doi.org/ggvzqn
DOI: 10.1001/jama.2020.8598 · PMID: 32391864

783. Covid-19: Black people and other minorities are hardest hit in US
Owen Dyer
BMJ (2020-04-14) https://doi.org/ggv5br
DOI: 10.1136/bmj.m1483 · PMID: 32291262

784. Susceptibility of Southwestern American Indian Tribes to Coronavirus Disease 2019 (COVID‐19)
Monika Kakol, Dona Upson, Akshay Sood
The Journal of Rural Health (2020-06) https://doi.org/ggtzkq
DOI: 10.1111/jrh.12451 · PMID: 32304251 · PMCID: PMC7264672

785. The Fullest Look Yet at the Racial Inequity of Coronavirus
Richard A. Oppel Jr, Robert Gebeloff, K. K. Rebecca Lai, Will Wright, Mitch Smith
The New York Times (2020-07-05) https://www.nytimes.com/interactive/2020/07/05/us/coronavirus-latinos-african-americans-cdc-data.html

786. Addressing inequities in COVID-19 morbidity and mortality: research and policy recommendations
Monica L Wang, Pamela Behrman, Akilah Dulin, Monica L Baskin, Joanna Buscemi, Kassandra I Alcaraz, Carly M Goldstein, Tiffany L Carson, Megan Shen, Marian Fitzgibbon
Translational Behavioral Medicine (2020-06) https://doi.org/gg3389
DOI: 10.1093/tbm/ibaa055 · PMID: 32542349 · PMCID: PMC7337775

787. Historical Environmental Racism, Structural Inequalities, and Dik’os Ntsaaígíí-19 (COVID-19) on Navajo Nation
Nicholet A. Deschine Parkhurst, Kimberly R. Huyser, Aggie J. Yellow Horse
Journal of Indigenous Social Development (2020-11-02) https://journalhosting.ucalgary.ca/index.php/jisd/article/view/70753

788. Protect Indigenous peoples from COVID-19
Lucas Ferrante, Philip M. Fearnside
Science (2020-04-16) https://doi.org/gg3k6f
DOI: 10.1126/science.abc0073 · PMID: 32299940

789. Factors associated with COVID-19-related death using OpenSAFELY
Elizabeth J. Williamson, Alex J. Walker, Krishnan Bhaskaran, Seb Bacon, Chris Bates, Caroline E. Morton, Helen J. Curtis, Amir Mehrkar, David Evans, Peter Inglesby, … Ben Goldacre
Nature (2020-07-08) https://doi.org/gg39n7
DOI: 10.1038/s41586-020-2521-4 · PMID: 32640463

790. Implications of biogeography of human populations for “race” and medicine
Sarah A Tishkoff, Kenneth K Kidd
Nature Genetics (2004-10-26) https://doi.org/d2xq92
DOI: 10.1038/ng1438 · PMID: 15507999

791. African Genetic Diversity: Implications for Human Demographic History, Modern Human Origins, and Complex Disease Mapping
Michael C. Campbell, Sarah A. Tishkoff
Annual Review of Genomics and Human Genetics (2008-09) https://doi.org/cphggp
DOI: 10.1146/annurev.genom.9.081307.164258 · PMID: 18593304 · PMCID: PMC2953791

792. NIH must confront the use of race in science
Michael Yudell, Dorothy Roberts, Rob DeSalle, Sarah Tishkoff, 70 signatories
Science (2020-09-10) https://doi.org/ghcm7s
DOI: 10.1126/science.abd4842 · PMID: 32913094

793. Interferon-Induced Transmembrane Protein 3 Genetic Variant rs12252-C Associated With Disease Severity in Coronavirus Disease 2019
Yonghong Zhang, Ling Qin, Yan Zhao, Ping Zhang, Bin Xu, Kang Li, Lianchun Liang, Chi Zhang, Yanchao Dai, Yingmei Feng, … Ronghua Jin
The Journal of Infectious Diseases (2020-07-01) https://doi.org/ggv3tj
DOI: 10.1093/infdis/jiaa224 · PMID: 32348495 · PMCID: PMC7197559

794. Genomewide Association Study of Severe Covid-19 with Respiratory Failure
The Severe Covid-19 GWAS Group
New England Journal of Medicine (2020-10-15) https://doi.org/gg2pqx
DOI: 10.1056/nejmoa2020283 · PMID: 32558485 · PMCID: PMC7315890

795. APOE e4 Genotype Predicts Severe COVID-19 in the UK Biobank Community Cohort
Chia-Ling Kuo, Luke C Pilling, Janice L Atkins, Jane AH Masoli, João Delgado, George A Kuchel, David Melzer
The Journals of Gerontology: Series A (2020-11) https://doi.org/ggx4ng
DOI: 10.1093/gerona/glaa131 · PMID: 32451547 · PMCID: PMC7314139

796. Genome-wide CRISPR screen reveals host genes that regulate SARS-CoV-2 infection
Jin Wei, Mia Madel Alfajaro, Ruth E. Hanna, Peter C. DeWeirdt, Madison S. Strine, William J. Lu-Culligan, Shang-Min Zhang, Vincent R. Graziano, Cameron O. Schmitz, Jennifer S. Chen, … Craig B. Wilen
Cold Spring Harbor Laboratory (2020-06-17) https://doi.org/dzz3
DOI: 10.1101/2020.06.16.155101 · PMID: 32869025 · PMCID: PMC7457610

797. New insights into genetic susceptibility of COVID-19: an ACE2 and TMPRSS2 polymorphism analysis
Yuan Hou, Junfei Zhao, William Martin, Asha Kallianpur, Mina K. Chung, Lara Jehi, Nima Sharifi, Serpil Erzurum, Charis Eng, Feixiong Cheng
BMC Medicine (2020-07-15) https://doi.org/gg445n
DOI: 10.1186/s12916-020-01673-z · PMID: 32664879 · PMCID: PMC7360473

798. Seroprevalence of anti-SARS-CoV-2 IgG antibodies in Kenyan blood donors
Sophie Uyoga, Ifedayo M. O. Adetifa, Henry K. Karanja, James Nyagwange, James Tuju, Perpetual Wanjiku, Rashid Aman, Mercy Mwangangi, Patrick Amoth, Kadondi Kasera, … George M Warimwe
Cold Spring Harbor Laboratory (2020-07-29) https://doi.org/ghcm7p
DOI: 10.1101/2020.07.27.20162693

799. High SARS-CoV-2 seroprevalence in Health Care Workers but relatively low numbers of deaths in urban Malawi
Marah Grace Chibwana, Khuzwayo Chidiwa Jere, Raphael kamng’ona, Jonathan Mandolo, Vincent Katunga-Phiri, Dumizulu Tembo, Ndaona Mitole, Samantha Musasa, Simon Sichone, Agness Lakudzala, … Kondwani Charles Jambo
Cold Spring Harbor Laboratory (2020-08-05) https://doi.org/ghcm7q
DOI: 10.1101/2020.07.30.20164970 · PMID: 32766597 · PMCID: PMC7402052

800. Africa’s pandemic puzzle: why so few cases and deaths?
Linda Nordling
Science (2020-08-13) https://doi.org/ghcm7r
DOI: 10.1126/science.369.6505.756 · PMID: 32792376

801. Are some ethnic groups more vulnerable to COVID-19 than others? https://www.ifs.org.uk/inequality/chapter/are-some-ethnic-groups-more-vulnerable-to-covid-19-than-others/

802. Quantifying the social distancing privilege gap: a longitudinal study of smartphone movement
Nabarun Dasgupta, Michele Jonsson Funk, Allison Lazard, Benjamin Eugene White, Stephen W. Marshall
Cold Spring Harbor Laboratory (2020-05-08) https://doi.org/gg79qk
DOI: 10.1101/2020.05.03.20084624

803. Uncovering socioeconomic gaps in mobility reduction during the COVID-19 pandemic using location data
Samuel P. Fraiberger, Pablo Astudillo, Lorenzo Candeago, Alex Chunet, Nicholas K. W. Jones, Maham Faisal Khan, Bruno Lepri, Nancy Lozano Gracia, Lorenzo Lucchini, Emanuele Massaro, Aleister Montfort
arXiv (2020-07-28) https://arxiv.org/abs/2006.15195

804. Mobility network models of COVID-19 explain inequities and inform reopening
Serina Chang, Emma Pierson, Pang Wei Koh, Jaline Gerardin, Beth Redbird, David Grusky, Jure Leskovec
Nature (2020-11-10) https://doi.org/ghjmt2
DOI: 10.1038/s41586-020-2923-3 · PMID: 33171481

805. A Basic Demographic Profile of Workers in Frontline Industries
Hye Jin Rho;Shawn Fremstad;Hayley Brown
(2020-06) https://mronline.org/wp-content/uploads/2020/06/2020-04-Frontline-Workers.pdf

806. Differential occupational risk for COVID‐19 and other infection exposure according to race and ethnicity
Devan Hawkins
American Journal of Industrial Medicine (2020-06-15) https://doi.org/gg3rb2
DOI: 10.1002/ajim.23145 · PMID: 32539166 · PMCID: PMC7323065

807. Estimating the burden of United States workers exposed to infection or disease: A key factor in containing risk of COVID-19 infection
Marissa G. Baker, Trevor K. Peckham, Noah S. Seixas
PLOS ONE (2020-04-28) https://doi.org/ggtx7c
DOI: 10.1371/journal.pone.0232452 · PMID: 32343747 · PMCID: PMC7188235

808. Coronavirus (COVID-19) related deaths by occupation, England and Wales: deaths registered up to and including 20 April 2020
Ben Windsor-Shellard, Jasveer Kaur
(2020-05-11) https://www.ons.gov.uk/peoplepopulationandcommunity/healthandsocialcare/causesofdeath/bulletins/coronaviruscovid19relateddeathsbyoccupationenglandandwales/deathsregistereduptoandincluding20april2020

809. Which occupations have the highest potential exposure to the coronavirus (COVID-19)?
Office for National Statistics
(2020-05-11) https://www.ons.gov.uk/employmentandlabourmarket/peopleinwork/employmentandemployeetypes/articles/whichoccupationshavethehighestpotentialexposuretothecoronaviruscovid19/2020-05-11

810. Disparities in the risk and outcomes from COVID-19
Public Health England
(2020-06-12) https://assets.publishing.service.gov.uk/government/uploads/system/uploads/attachment_data/file/892085/disparities_review.pdf

811. Exclusive: deaths of NHS staff from covid-19 analysed
Tim Cook, Emira Kursumovic, Simon Lennane2020-04-22T12:42:00+01:00
Health Service Journal https://www.hsj.co.uk/exclusive-deaths-of-nhs-staff-from-covid-19-analysed/7027471.article

812. https://act.nationalnursesunited.org/page

813. Racial Disparity in COVID-19 Deaths: Seeking Economic Roots with Census data.
John McLaren
National Bureau of Economic Research (2020-06-22) https://www.nber.org/papers/w27407

814. Mortality, Admissions, and Patient Census at SNFs in 3 US Cities During the COVID-19 Pandemic
Michael L. Barnett, Lissy Hu, Thomas Martin, David C. Grabowski
JAMA (2020-08-04) https://doi.org/gg3387
DOI: 10.1001/jama.2020.11642 · PMID: 32579161 · PMCID: PMC7315390

815. COVID-19 in Prisons and Jails in the United States
Laura Hawks, Steffie Woolhandler, Danny McCormick
JAMA Internal Medicine (2020-08-01) https://doi.org/ggtxw6
DOI: 10.1001/jamainternmed.2020.1856 · PMID: 32343355

816. COVID-19 Cases and Deaths in Federal and State Prisons
Brendan Saloner, Kalind Parish, Julie A. Ward, Grace DiLaura, Sharon Dolovich
JAMA (2020-08-11) https://doi.org/gg4dcv
DOI: 10.1001/jama.2020.12528 · PMID: 32639537 · PMCID: PMC7344796

817. State Rates of Incarceration Race & Ethnicity_updated2
nlizanna
(2018-03-24) https://www.issuelab.org/resources/695/695.pdf

818. Under One Roof: A Review of Research on Intergenerational Coresidence and Multigenerational Households in the United States
Jennifer Reid Keene, Christie D. Batson
Sociology Compass (2010-08) https://doi.org/fsrcr7
DOI: 10.1111/j.1751-9020.2010.00306.x

819. Chaos and the macrosetting: The role of poverty and socioeconomic status.
Gary W. Evans, John Eckenrode, Lyscha A. Marcynyszyn
American Psychological Association (APA) (2010-01-12) https://doi.org/c76n3g
DOI: 10.1037/12057-014

820. Housing insecurity among urban fathers
Marah A. Curtis, Amanda B. Geller
Columbia University (2010) https://doi.org/ghdjn2
DOI: 10.7916/d8wh2w9t

821. Housing and Employment Insecurity among the Working Poor
Matthew Desmond, Carl Gershenson
Social Problems (2016-02) https://doi.org/f8crm2
DOI: 10.1093/socpro/spv025

822. Obesity and its comorbid conditions
Lalita Khaodhiar, Karen C. McCowen, George L. Blackburn
Clinical Cornerstone (1999-01) https://doi.org/bpp37d
DOI: 10.1016/s1098-3597(99)90002-9

823. Aging, Male Sex, Obesity, and Metabolic Inflammation Create the Perfect Storm for COVID-19
Franck Mauvais-Jarvis
Diabetes (2020-09) https://doi.org/gg47zk
DOI: 10.2337/dbi19-0023 · PMID: 32669390 · PMCID: PMC7458034

824. Non-communicable disease syndemics: poverty, depression, and diabetes among low-income populations
Emily Mendenhall, Brandon A Kohrt, Shane A Norris, David Ndetei, Dorairaj Prabhakaran
The Lancet (2017-03) https://doi.org/gddg84
DOI: 10.1016/s0140-6736(17)30402-6 · PMID: 28271846 · PMCID: PMC5491333

825. Obesity and poverty paradox in developed countries
Wioletta Żukiewicz-Sobczak, Paula Wróblewska, Jacek Zwoliński, Jolanta Chmielewska-Badora, Piotr Adamczuk, Ewelina Krasowska, Jerzy Zagórski, Anna Oniszczuk, Jacek Piątek, Wojciech Silny
Annals of Agricultural and Environmental Medicine (2014-09-04) https://doi.org/f6jhzc
DOI: 10.5604/12321966.1120608 · PMID: 25292135

826. Impact of the COVID‐19 Pandemic on Unhealthy Eating in Populations with Obesity
Nathaniel J. S. Ashby
Obesity (2020-08-20) https://doi.org/ghd6qc
DOI: 10.1002/oby.22940 · PMID: 32589788 · PMCID: PMC7361200

827. Fast Food Patronage and Obesity Prevalence During the COVID‐19 Pandemic: An Alternative Explanation
Candice A. Myers, Stephanie T. Broyles
Obesity (2020-09-03) https://doi.org/gg6v84
DOI: 10.1002/oby.22993 · PMID: 32741130 · PMCID: PMC7435526

828. The global food syndemic: The impact of food insecurity, Malnutrition and obesity on the healthspan amid the COVID-19 pandemic
Martha I. Huizar, Ross Arena, Deepika R. Laddu
Progress in Cardiovascular Diseases (2020-07) https://doi.org/gg4r3h
DOI: 10.1016/j.pcad.2020.07.002 · PMID: 32653438 · PMCID: PMC7347484

829. Stress, chronic inflammation, and emotional and physical well-being: Concurrent effects and chronic sequelae
George P. Chrousos
Journal of Allergy and Clinical Immunology (2000-11) https://doi.org/bgx7hn
DOI: 10.1067/mai.2000.110163 · PMID: 11080744

830. Chronic psychological stress and the regulation of pro-inflammatory cytokines: A glucocorticoid-resistance model.
Gregory E. Miller, Sheldon Cohen, A. Kim Ritchey
Health Psychology (2002) https://doi.org/dj5r8b
DOI: 10.1037/0278-6133.21.6.531

831. Chronic stress, daily stressors, and circulating inflammatory markers.
Jean-Philippe Gouin, Ronald Glaser, William B. Malarkey, David Beversdorf, Janice Kiecolt-Glaser
Health Psychology (2012-03) https://doi.org/dkz9tr
DOI: 10.1037/a0025536 · PMID: 21928900 · PMCID: PMC3253267

832. Turning Up the Heat
Gregory E. Miller, Ekin Blackwell
Current Directions in Psychological Science (2016-06-24) https://doi.org/bft9mv
DOI: 10.1111/j.1467-8721.2006.00450.x

833. Sick of Poverty
Robert Sapolsky
Scientific American (2005-12) https://doi.org/fxf5kp
DOI: 10.1038/scientificamerican1205-92 · PMID: 16323696

834. Exposure to air pollution and COVID-19 mortality in the United States: A nationwide cross-sectional study
Xiao Wu, Rachel C Nethery, M Benjamin Sabath, Danielle Braun, Francesca Dominici
Cold Spring Harbor Laboratory (2020-04-27) https://doi.org/ggrpcj
DOI: 10.1101/2020.04.05.20054502 · PMID: 32511651 · PMCID: PMC7277007

835. Global estimates of mortality associated with long-term exposure to outdoor fine particulate matter
Richard Burnett, Hong Chen, Mieczysław Szyszkowicz, Neal Fann, Bryan Hubbell, C. Arden Pope, Joshua S. Apte, Michael Brauer, Aaron Cohen, Scott Weichenthal, … Joseph V. Spadaro
Proceedings of the National Academy of Sciences (2018-09-18) https://doi.org/gfgbcx
DOI: 10.1073/pnas.1803222115 · PMID: 30181279 · PMCID: PMC6156628

836. Early-Life Air Pollution Exposure, Neighborhood Poverty, and Childhood Asthma in the United States, 1990–2014
Nicole Kravitz-Wirtz, Samantha Teixeira, Anjum Hajat, Bongki Woo, Kyle Crowder, David Takeuchi
International Journal of Environmental Research and Public Health (2018-05-30) https://doi.org/gdvwp9
DOI: 10.3390/ijerph15061114 · PMID: 29848979 · PMCID: PMC6025399

837. Early life stress, air pollution, inflammation, and disease: An integrative review and immunologic model of social-environmental adversity and lifespan health
Hector A. Olvera Alvarez, Laura D. Kubzansky, Matthew J. Campen, George M. Slavich
Neuroscience & Biobehavioral Reviews (2018-09) https://doi.org/gd46bm
DOI: 10.1016/j.neubiorev.2018.06.002 · PMID: 29874545 · PMCID: PMC6082389

838. Covid-19 and Disparities in Nutrition and Obesity
Matthew J. Belanger, Michael A. Hill, Angeliki M. Angelidi, Maria Dalamaga, James R. Sowers, Christos S. Mantzoros
New England Journal of Medicine (2020-09-10) https://doi.org/gg475x
DOI: 10.1056/nejmp2021264 · PMID: 32668105

839. Systemic racism, chronic health inequities, and ‐19: A syndemic in the making?
Clarence C. Gravlee
American Journal of Human Biology (2020-08-04) https://doi.org/ghcxwk
DOI: 10.1002/ajhb.23482 · PMID: 32754945 · PMCID: PMC7441277

840. Coronavirus Disease (COVID-19): A primer for emergency physicians
Summer Chavez, Brit Long, Alex Koyfman, Stephen Y. Liang
The American Journal of Emergency Medicine (2020-03) https://doi.org/ggr22z
DOI: 10.1016/j.ajem.2020.03.036 · PMID: 32265065 · PMCID: PMC7102516

841. Insurers May Only Pay For Coronavirus Tests When They’re “Medically Necessary”
NPR.org
https://www.npr.org/sections/health-shots/2020/06/19/880543755/insurers-may-only-pay-for-coronavirus-tests-when-theyre-medically-necessary

842. Private Health Insurance Coverage in the COVID-19 Public Health Emergency | Commonwealth Fund https://www.commonwealthfund.org/blog/2020/private-health-insurance-coverage-covid-19-public-health-emergency

843. COVID-19 and racial disparities
Monica Shah, Muskaan Sachdeva, Roni P. Dodiuk-Gad
Journal of the American Academy of Dermatology (2020-07) https://doi.org/ggtwm7
DOI: 10.1016/j.jaad.2020.04.046 · PMID: 32305444 · PMCID: PMC7162783

844. FAQs for COVID-19 Claims Reimbursement to Health Care Providers and Facilities for Testing, Treatment and Vaccine Administration
Official web site of the U.S. Health Resources & Services Administration
https://www.hrsa.gov/coviduninsuredclaim/frequently-asked-questions

845. Potential association between COVID-19 mortality and health-care resource availability
Yunpeng Ji, Zhongren Ma, Maikel P Peppelenbosch, Qiuwei Pan
The Lancet Global Health (2020-04) https://doi.org/ggqscd
DOI: 10.1016/s2214-109x(20)30068-1 · PMID: 32109372 · PMCID: PMC7128131

846. Combating COVID-19: health equity matters
Zhicheng Wang, Kun Tang
Nature Medicine (2020-03-26) https://doi.org/ggs4p6
DOI: 10.1038/s41591-020-0823-6 · PMID: 32284617

847. lockup black
chen0307
(2020-10-14) https://www.census.gov/content/dam/Census/library/publications/2019/demo/p60-267.pdf

848. An Ethical Dilemma in SARS-Cov-2 Pandemic : Who Gets the Ventilator?
Dumache Raluca, Ciocan Veronica, Muresan Camelia Oana, Enache Alexandra
European Scientific Journal ESJ (2020-07-31) https://doi.org/ghfprk
DOI: 10.19044/esj.2020.v16n21p24

849. Planning Hospital Needs for Ventilators and Respiratory Therapists in the COVID-19 Crisis
John Raffensperger, Marygail Brauner, R. Briggs
Rand Corporation (2020) https://doi.org/ghfprp
DOI: 10.7249/pea228-1

850. Fair Allocation of Vaccines, Ventilators and Antiviral Treatments: Leaving No Ethical Value Behind in Health Care Rationing
Parag A. Pathak, Tayfun Sönmez, M. Utku Ünver, M. Bumin Yenmez
arXiv (2020-08-04) https://arxiv.org/abs/2008.00374

851. Reallocating ventilators during the coronavirus disease 2019 pandemic: Is it ethical?
Quyen Chu, Ricardo Correa, Tracey L. Henry, Kyle A. McGregor, Hanni Stoklosa, Loren Robinson, Sachin Jha, Alagappan Annamalai, Benson S. Hsu, Rohit Gupta, … SreyRam Kuy
Surgery (2020-09) https://doi.org/ghfprb
DOI: 10.1016/j.surg.2020.04.044 · PMID: 32616345 · PMCID: PMC7205622

852. Ethics Lessons From Seattle’s Early Experience With COVID-19
Denise M. Dudzinski, Benjamin Y. Hoisington, Crystal E. Brown
The American Journal of Bioethics (2020-06-18) https://doi.org/ghfprc
DOI: 10.1080/15265161.2020.1764137 · PMID: 32552455

853. Rationing Limited Healthcare Resources in the COVID‐19 Era and Beyond: Ethical Considerations Regarding Older Adults
Timothy W. Farrell, Leslie Francis, Teneille Brown, Lauren E. Ferrante, Eric Widera, Ramona Rhodes, Tony Rosen, Ula Hwang, Leah J. Witt, Niranjan Thothala, … Debra Saliba
Journal of the American Geriatrics Society (2020-06-14) https://doi.org/ggvt7z
DOI: 10.1111/jgs.16539 · PMID: 32374466 · PMCID: PMC7267288

854. Paediatric ethical issues during the COVID‐19 pandemic are not just about ventilator triage
Marlyse F. Haward, Gregory P. Moore, John Lantos, Annie Janvier
Acta Paediatrica (2020-05-20) https://doi.org/ggv24n
DOI: 10.1111/apa.15334 · PMID: 32364256 · PMCID: PMC7267437

855. Ethical Challenges Arising in the COVID-19 Pandemic: An Overview from the Association of Bioethics Program Directors (ABPD) Task Force
Amy L. McGuire, Mark P. Aulisio, F. Daniel Davis, Cheryl Erwin, Thomas D. Harter, Reshma Jagsi, Robert Klitzman, Robert Macauley, Eric Racine, Susan M. Wolf, … The COVID-19 Task Force of the Association of Bioethics Program Directors (ABPD)
The American Journal of Bioethics (2020-06-08) https://doi.org/gg6c5k
DOI: 10.1080/15265161.2020.1764138 · PMID: 32511078

856. Disability, Ethics, and Health Care in the COVID-19 Pandemic
Maya Sabatello, Teresa Blankmeyer Burke, Katherine E. McDonald, Paul S. Appelbaum
American Journal of Public Health (2020-10) https://doi.org/ghfprm
DOI: 10.2105/ajph.2020.305837 · PMID: 32816541 · PMCID: PMC7483109

857. Allocating Ventilators During the COVID-19 Pandemic and Conscientious Objection
Mark Wicclair
The American Journal of Bioethics (2020-07-27) https://doi.org/gg6nk4
DOI: 10.1080/15265161.2020.1777347 · PMID: 32716798

858. Colorblind Algorithms: Racism in the Era of COVID-19
J. Corey Williams, Nientara Anderson, Myra Mathis, Ezelle Sanford, Jeffrey Eugene, Jessica Isom
Journal of the National Medical Association (2020-10) https://doi.org/ghfpq8
DOI: 10.1016/j.jnma.2020.05.010 · PMID: 32563687

859. Structural Racism, Social Risk Factors, and Covid-19 — A Dangerous Convergence for Black Americans
Leonard E. Egede, Rebekah J. Walker
New England Journal of Medicine (2020-09-17) https://doi.org/gg56nc
DOI: 10.1056/nejmp2023616 · PMID: 32706952 · PMCID: PMC7747672

860. Allocating Remdesivir Under Scarcity: Social Justice or More Systemic Racism
Eli Weber, Mark J. Bliton
The American Journal of Bioethics (2020-08-25) https://doi.org/ghfprf
DOI: 10.1080/15265161.2020.1795538 · PMID: 32840451

861. Revisiting the equity debate in COVID-19: ICU is no panacea
Angela Ballantyne, Wendy A Rogers, Vikki Entwistle, Cindy Towns
Journal of Medical Ethics (2020-10) https://doi.org/gg33nq
DOI: 10.1136/medethics-2020-106460 · PMID: 32571847 · PMCID: PMC7335695

862. Ethical Dilemmas in Covid-19 Medical Care: Is a Problematic Triage Protocol Better or Worse than No Protocol at All?
Sheri Fink
The American Journal of Bioethics (2020-07-27) https://doi.org/gg6nqn
DOI: 10.1080/15265161.2020.1788663 · PMID: 32716771

863. Developing an Ethics Framework for Allocating Remdesivir in the COVID-19 Pandemic
Sarah Lim, Debra A. DeBruin, Jonathon P. Leider, Nneka Sederstrom, Ruth Lynfield, Jason V. Baker, Susan Kline, Sarah Kesler, Stacey Rizza, Joel Wu, … Susan M. Wolf
Mayo Clinic Proceedings (2020-09) https://doi.org/ghfpq9
DOI: 10.1016/j.mayocp.2020.06.016 · PMID: 32861338 · PMCID: PMC7305893

864. Ethically Allocating COVID-19 Drugs Via Pre-approval Access and Emergency Use Authorization
Jamie Webb, Lesha D. Shah, Holly Fernandez Lynch
The American Journal of Bioethics (2020-08-25) https://doi.org/ghfprd
DOI: 10.1080/15265161.2020.1795529

865. (2011-07-27) https://www.cdc.gov/about/advisory/pdf/VentDocument_Release.pdf

866. Adopting an Anti-Racist Model of COVID-19 Drug Allocation and Prioritization
Akilah A. Jefferson
The American Journal of Bioethics (2020-08-25) https://doi.org/ghggz5
DOI: 10.1080/15265161.2020.1795541

867. Equitably Sharing the Benefits and Burdens of Research: Covid‐19 Raises the Stakes
Carl H. Coleman
Ethics & Human Research (2020-05-14) https://doi.org/ghgg2q
DOI: 10.1002/eahr.500055 · PMID: 32410347 · PMCID: PMC7272984

868. Ensuring global access to COVID-19 vaccines
Gavin Yamey, Marco Schäferhoff, Richard Hatchett, Muhammad Pate, Feng Zhao, Kaci Kennedy McDade
The Lancet (2020-05) https://doi.org/ggq7mf
DOI: 10.1016/s0140-6736(20)30763-7 · PMID: 32243778 · PMCID: PMC7271264

869. The Equitable Distribution of COVID-19 Therapeutics and Vaccines
Thomas J. Bollyky, Lawrence O. Gostin, Margaret A. Hamburg
JAMA (2020-06-23) https://doi.org/ggvcvt
DOI: 10.1001/jama.2020.6641 · PMID: 32379268

870. Recruitment and participation in clinical trials: Socio-demographic, rural/urban, and health care access predictors
Claudia R. Baquet, Patricia Commiskey, C. Daniel Mullins, Shiraz I. Mishra
Cancer Detection and Prevention (2006-01) https://doi.org/bk2k4g
DOI: 10.1016/j.cdp.2005.12.001 · PMID: 16495020 · PMCID: PMC3276312

871. COVID-19 vaccine trials in Africa
Munyaradzi Makoni
The Lancet Respiratory Medicine (2020-11) https://doi.org/fgzk
DOI: 10.1016/s2213-2600(20)30401-x

872. Racial/ethnic differences in clinical trial enrollment, refusal rates, ineligibility, and reasons for decline among patients at sites in the National Cancer Institute’s Community Cancer Centers Program
Aisha T. Langford, Ken Resnicow, Eileen P. Dimond, Andrea M. Denicoff, Diane St. Germain, Worta McCaskill-Stevens, Rebecca A. Enos, Angela Carrigan, Kathy Wilkinson, Ronald S. Go
Cancer (2014-03-15) https://doi.org/ghgg2p
DOI: 10.1002/cncr.28483 · PMID: 24327389 · PMCID: PMC3947654

873. Participation in Cancer Clinical Trials
Vivek H. Murthy, Harlan M. Krumholz, Cary P. Gross
JAMA (2004-06-09) https://doi.org/bbh8h7
DOI: 10.1001/jama.291.22.2720 · PMID: 15187053

874. Participation in Surgical Oncology Clinical Trials: Gender-, Race/Ethnicity-, and Age-based Disparities
John H. Stewart, Alain G. Bertoni, Jennifer L. Staten, Edward A. Levine, Cary P. Gross
Annals of Surgical Oncology (2007-08-08) https://doi.org/cq7wqs
DOI: 10.1245/s10434-007-9500-y · PMID: 17682824

875. Inclusion, Analysis, and Reporting of Sex and Race/Ethnicity in Clinical Trials: Have We Made Progress?
Stacie E. Geller, Abby Koch, Beth Pellettieri, Molly Carnes
Journal of Women’s Health (2011-03) https://doi.org/dhzxk7
DOI: 10.1089/jwh.2010.2469 · PMID: 21351877 · PMCID: PMC3058895

876. The Representation of Gender and Race/Ethnic Groups in Randomized Clinical Trials of Individuals with Systemic Lupus Erythematosus
Titilola Falasinnu, Yashaar Chaichian, Michelle B. Bass, Julia F. Simard
Current Rheumatology Reports (2018-03-17) https://doi.org/ghjmpz
DOI: 10.1007/s11926-018-0728-2 · PMID: 29550947 · PMCID: PMC5857270

877. Racial Disproportionality in Covid Clinical Trials
Daniel B. Chastain, Sharmon P. Osae, Andrés F. Henao-Martínez, Carlos Franco-Paredes, Joeanna S. Chastain, Henry N. Young
New England Journal of Medicine (2020-08-27) https://doi.org/gg7vcf
DOI: 10.1056/nejmp2021971 · PMID: 32780573

878. Othering and Being Othered in the Context of Health Care Services
Joy L. Johnson, Joan L. Bottorff, Annette J. Browne, Sukhdev Grewal, B. Ann Hilton, Heather Clarke
Health Communication (2004-04) https://doi.org/cvxqm4
DOI: 10.1207/s15327027hc1602_7 · PMID: 15090288

879. A decade of studying implicit racial/ethnic bias in healthcare providers using the implicit association test
Ivy W. Maina, Tanisha D. Belton, Sara Ginzberg, Ajit Singh, Tiffani J. Johnson
Social Science & Medicine (2018-02) https://doi.org/gdfwd9
DOI: 10.1016/j.socscimed.2017.05.009 · PMID: 28532892

880. A Systematic Review of the Impact of Physician Implicit Racial Bias on Clinical Decision Making
Erin Dehon, Nicole Weiss, Jonathan Jones, Whitney Faulconer, Elizabeth Hinton, Sarah Sterling
Academic Emergency Medicine (2017-08) https://doi.org/gbw5pk
DOI: 10.1111/acem.13214 · PMID: 28472533

881. Aversive racism and medical interactions with Black patients: A field study
Louis A. Penner, John F. Dovidio, Tessa V. West, Samuel L. Gaertner, Terrance L. Albrecht, Rhonda K. Dailey, Tsveti Markova
Journal of Experimental Social Psychology (2010-03) https://doi.org/dc5342
DOI: 10.1016/j.jesp.2009.11.004 · PMID: 20228874 · PMCID: PMC2835170

882. Intersection of Bias, Structural Racism, and Social Determinants With Health Care Inequities
Tiffani J. Johnson
Pediatrics (2020-08) https://doi.org/ghjmqz
DOI: 10.1542/peds.2020-003657 · PMID: 32690807

883. A Systematic Review Of The Food And Drug Administration’s “Exception From Informed Consent” Pathway
William B. Feldman, Spencer Phillips Hey, Aaron S. Kesselheim
Health Affairs (2018-10) https://doi.org/ghjmqw
DOI: 10.1377/hlthaff.2018.0501 · PMID: 30273035

884. The legacy of the tuskegee syphilis experiments for emergency exception from informed consent
Terri A. Schmidt
Annals of Emergency Medicine (2003-01) https://doi.org/fw3kvs
DOI: 10.1067/mem.2003.17 · PMID: 12514686

885. CDC officials are considering a plan to distribute COVID-19 vaccines to the most vulnerable first — including people of color
Sarah Al-Arshani
Business Insider https://www.businessinsider.com/cdc-official-considering-giving-covid-19-vaccine-most-vulnerable-first-2020-10

886. Racial Differences in T-Lymphocyte Response to Glucocorticoids
Monica J. Federico, Ronina A. Covar, Eleanor E. Brown, Donald Y. M. Leung, Joseph D. Spahn
Chest (2005-02) https://doi.org/bjfcf6
DOI: 10.1378/chest.127.2.571 · PMID: 15705998

887. ENDOCRINOLOGY OF THE STRESS RESPONSE
Evangelia Charmandari, Constantine Tsigos, George Chrousos
Annual Review of Physiology (2005-03-17) https://doi.org/brcm9n
DOI: 10.1146/annurev.physiol.67.040403.120816 · PMID: 15709959

888. Chronic stress, glucocorticoid receptor resistance, inflammation, and disease risk
S. Cohen, D. Janicki-Deverts, W. J. Doyle, G. E. Miller, E. Frank, B. S. Rabin, R. B. Turner
Proceedings of the National Academy of Sciences (2012-04-02) https://doi.org/f4n6r8
DOI: 10.1073/pnas.1118355109 · PMID: 22474371 · PMCID: PMC3341031

889. Opportunities and obstacles for deep learning in biology and medicine
Travers Ching, Daniel S. Himmelstein, Brett K. Beaulieu-Jones, Alexandr A. Kalinin, Brian T. Do, Gregory P. Way, Enrico Ferrero, Paul-Michael Agapow, Michael Zietz, Michael M. Hoffman, … Casey S. Greene
Journal of The Royal Society Interface (2018-04-04) https://doi.org/gddkhn
DOI: 10.1098/rsif.2017.0387 · PMID: 29618526 · PMCID: PMC5938574

890. Opportunities and obstacles for deep learning in biology and medicine [update in progress]
Casey S. Greene, Daniel S. Himmelstein, Daniel C. Elton, Brock C. Christensen, Anthony Gitter, Alexander J. Titus, Joshua J. Levy
Manubot (2020-11-10) https://greenelab.github.io/deep-review/

891. ismms-himc/covid-19_sinai_reviews
Human Immune Monitoring Center at Mount Sinai
(2020-07-09) https://github.com/ismms-himc/covid-19_sinai_reviews

892. Advancing scientific knowledge in times of pandemics
Nicolas Vabret, Robert Samstein, Nicolas Fernandez, Miriam Merad, The Sinai Immunology Review Project, Trainees, Faculty
Nature Reviews Immunology (2020-04-23) https://doi.org/ghjs79
DOI: 10.1038/s41577-020-0319-0 · PMID: 32327718 · PMCID: PMC7187143

893. Undergraduate Mentoring - American Physician Scientists Association https://www.physicianscientists.org/page/summer-research-pilot-program

894. greenelab/covid19-review
Greene Laboratory
(2021-01-21) https://github.com/greenelab/covid19-review

895. SARS-CoV-2 and COVID-19: An Evolving Review of Diagnostics and Therapeutics
Halie M. Rando, Casey S. Greene, Michael P. Robson, Simina M. Boca, Nils Wellhausen, Ronan Lordan, Christian Brueffer, Sandipan Ray, Lucy D\’Agostino McGowan, Anthony Gitter, … Rishi Raj Goel
Manubot (2021-01-21) https://greenelab.github.io/covid19-review/

896. Matplotlib: A 2D Graphics Environment
John D. Hunter
Computing in Science & Engineering (2007) https://doi.org/drbjhg
DOI: 10.1109/mcse.2007.55

897. Using the MAARIE Framework To Read the Research Literature
M. Corcoran
American Journal of Occupational Therapy (2006-07-01) https://doi.org/bqh97x
DOI: 10.5014/ajot.60.4.367 · PMID: 16915865

898. Potent binding of 2019 novel coronavirus spike protein by a SARS coronavirus-specific human monoclonal antibody
Xiaolong Tian, Cheng Li, Ailing Huang, Shuai Xia, Sicong Lu, Zhengli Shi, Lu Lu, Shibo Jiang, Zhenlin Yang, Yanling Wu, Tianlei Ying
Cold Spring Harbor Laboratory (2020-01-28) https://doi.org/ggjqfd
DOI: 10.1101/2020.01.28.923011

899. Integrative Bioinformatics Analysis Provides Insight into the Molecular Mechanisms of 2019-nCoV
Xiang He, Lei Zhang, Qin Ran, Junyi Wang, Anying Xiong, Dehong Wu, Feng Chen, Guoping Li
Cold Spring Harbor Laboratory (2020-02-05) https://doi.org/ggrbd8
DOI: 10.1101/2020.02.03.20020206

900. Diarrhea may be underestimated: a missing link in 2019 novel coronavirus
Weicheng Liang, Zhijie Feng, Shitao Rao, Cuicui Xiao, Ze-Xiao Lin, Qi Zhang, Qi Wei
Cold Spring Harbor Laboratory (2020-02-17) https://doi.org/ggrbdw
DOI: 10.1101/2020.02.03.20020289

901. Specific ACE2 Expression in Cholangiocytes May Cause Liver Damage After 2019-nCoV Infection
Xiaoqiang Chai, Longfei Hu, Yan Zhang, Weiyu Han, Zhou Lu, Aiwu Ke, Jian Zhou, Guoming Shi, Nan Fang, Jia Fan, … Fei Lan
Cold Spring Harbor Laboratory (2020-02-04) https://doi.org/ggq626
DOI: 10.1101/2020.02.03.931766

902. Recapitulation of SARS-CoV-2 Infection and Cholangiocyte Damage with Human Liver Organoids
Bing Zhao, Chao Ni, Ran Gao, Yuyan Wang, Li Yang, Jinsong Wei, Ting Lv, Jianqing Liang, Qisheng Zhang, Wei Xu, … Xinhua Lin
Cold Spring Harbor Laboratory (2020-03-17) https://doi.org/ggq648
DOI: 10.1101/2020.03.16.990317

903. ACE2 expression by colonic epithelial cells is associated with viral infection, immunity and energy metabolism
Jun Wang, Shanmeizi Zhao, Ming Liu, Zhiyao Zhao, Yiping Xu, Ping Wang, Meng Lin, Yanhui Xu, Bing Huang, Xiaoyu Zuo, … Yuxia Zhang
Cold Spring Harbor Laboratory (2020-02-07) https://doi.org/ggrfbx
DOI: 10.1101/2020.02.05.20020545

904. The Pathogenicity of SARS-CoV-2 in hACE2 Transgenic Mice
Linlin Bao, Wei Deng, Baoying Huang, Hong Gao, Jiangning Liu, Lili Ren, Qiang Wei, Pin Yu, Yanfeng Xu, Feifei Qi, … Chuan Qin
Cold Spring Harbor Laboratory (2020-02-28) https://doi.org/dph2
DOI: 10.1101/2020.02.07.939389

905. Caution on Kidney Dysfunctions of COVID-19 Patients
Zhen Li, Ming Wu, Jiwei Yao, Jie Guo, Xiang Liao, Siji Song, Jiali Li, Guangjie Duan, Yuanxiu Zhou, Xiaojun Wu, … Junan Yan
Cold Spring Harbor Laboratory (2020-03-27) https://doi.org/ggq627
DOI: 10.1101/2020.02.08.20021212

906. Acute renal impairment in coronavirus-associated severe acute respiratory syndrome
Kwok Hong Chu, Wai Kay Tsang, Colin S. Tang, Man Fai Lam, Fernand M. Lai, Ka Fai To, Ka Shun Fung, Hon Lok Tang, Wing Wa Yan, Hilda W. H. Chan, … Kar Neng Lai
Kidney International (2005-02) https://doi.org/b7tgtx
DOI: 10.1111/j.1523-1755.2005.67130.x · PMID: 15673319 · PMCID: PMC7112337

907. Single-cell Analysis of ACE2 Expression in Human Kidneys and Bladders Reveals a Potential Route of 2019-nCoV Infection
Wei Lin, Longfei Hu, Yan Zhang, Joshua D. Ooi, Ting Meng, Peng Jin, Xiang Ding, Longkai Peng, Lei Song, Zhou Xiao, … Yong Zhong
Cold Spring Harbor Laboratory (2020-02-18) https://doi.org/ggq629
DOI: 10.1101/2020.02.08.939892

908. The immune vulnerability landscape of the 2019 Novel Coronavirus, SARS-CoV-2
James Zhu, Jiwoong Kim, Xue Xiao, Yunguan Wang, Danni Luo, Shuang Jiang, Ran Chen, Lin Xu, He Zhang, Lenny Moise, … Yang Xie
Cold Spring Harbor Laboratory (2020-09-04) https://doi.org/ggq628
DOI: 10.1101/2020.02.08.939553 · PMID: 32908981 · PMCID: PMC7480032

909. Tissue distribution of ACE2 protein, the functional receptor for SARS coronavirus. A first step in understanding SARS pathogenesis
I Hamming, W Timens, MLC Bulthuis, AT Lely, GJ Navis, H van Goor
The Journal of Pathology (2004-06) https://doi.org/bhpzc3
DOI: 10.1002/path.1570 · PMID: 15141377 · PMCID: PMC7167720

910. Clinical Course and Outcomes of Critically Ill Patients With Middle East Respiratory Syndrome Coronavirus Infection
Yaseen M. Arabi, Ahmed A. Arifi, Hanan H. Balkhy, Hani Najm, Abdulaziz S. Aldawood, Alaa Ghabashi, Hassan Hawa, Adel Alothman, Abdulaziz Khaldi, Basel Al Raiy
Annals of Internal Medicine (2014-03-18) https://doi.org/ggptxw
DOI: 10.7326/m13-2486 · PMID: 24474051

911. Neutrophil-to-Lymphocyte Ratio Predicts Severe Illness Patients with 2019 Novel Coronavirus in the Early Stage
Jingyuan Liu, Yao Liu, Pan Xiang, Lin Pu, Haofeng Xiong, Chuansheng Li, Ming Zhang, Jianbo Tan, Yanli Xu, Rui Song, … Xianbo Wang
Cold Spring Harbor Laboratory (2020-02-12) https://doi.org/ggrbdx
DOI: 10.1101/2020.02.10.20021584

912. Dysregulation of Immune Response in Patients With Coronavirus 2019 (COVID-19) in Wuhan, China
Chuan Qin, Luoqi Zhou, Ziwei Hu, Shuoqi Zhang, Sheng Yang, Yu Tao, Cuihong Xie, Ke Ma, Ke Shang, Wei Wang, Dai-Shi Tian
Clinical Infectious Diseases (2020-08-01) https://doi.org/ggpxcf
DOI: 10.1093/cid/ciaa248 · PMID: 32161940 · PMCID: PMC7108125

913. Characteristics of lymphocyte subsets and cytokines in peripheral blood of 123 hospitalized patients with 2019 novel coronavirus pneumonia (NCP)
Suxin Wan, Qingjie Yi, Shibing Fan, Jinglong Lv, Xianxiang Zhang, Lian Guo, Chunhui Lang, Qing Xiao, Kaihu Xiao, Zhengjun Yi, … Yongping Chen
Cold Spring Harbor Laboratory (2020-02-12) https://doi.org/ggq63b
DOI: 10.1101/2020.02.10.20021832

914. Longitudinal Characteristics of Lymphocyte Responses and Cytokine Profiles in the Peripheral Blood of SARS-CoV-2 Infected Patients
Jing Liu, Sumeng Li, Jia Liu, Boyun Liang, Xiaobei Wang, Wei Li, Hua Wang, Qiaoxia Tong, Jianhua Yi, Lei Zhao, … Xin Zheng
SSRN Electronic Journal (2020) https://doi.org/ggq655
DOI: 10.2139/ssrn.3539682

915. Epidemiological and Clinical Characteristics of 17 Hospitalized Patients with 2019 Novel Coronavirus Infections Outside Wuhan, China
Jie Li, Shilin Li, Yurui Cai, Qin Liu, Xue Li, Zhaoping Zeng, Yanpeng Chu, Fangcheng Zhu, Fanxin Zeng
Cold Spring Harbor Laboratory (2020-02-12) https://doi.org/ggq63c
DOI: 10.1101/2020.02.11.20022053

916. ACE2 Expression in Kidney and Testis May Cause Kidney and Testis Damage After 2019-nCoV Infection
Caibin Fan, Kai Li, Yanhong Ding, Wei Lu, Jianqing Wang
Cold Spring Harbor Laboratory (2020-02-13) https://doi.org/ggq63d
DOI: 10.1101/2020.02.12.20022418

917. Aberrant pathogenic GM-CSF + T cells and inflammatory CD14 + CD16 + monocytes in severe pulmonary syndrome patients of a new coronavirus
Yonggang Zhou, Binqing Fu, Xiaohu Zheng, Dongsheng Wang, Changcheng Zhao, Yingjie qi, Rui Sun, Zhigang Tian, Xiaoling Xu, Haiming Wei
Cold Spring Harbor Laboratory (2020-02-20) https://doi.org/ggq63f
DOI: 10.1101/2020.02.12.945576

918. Clinical Characteristics of 2019 Novel Infected Coronavirus Pneumonia: A Systemic Review and Meta-analysis
Kai Qian, Yi Deng, Yong-Hang Tai, Jun Peng, Hao Peng, Li-Hong Jiang
Cold Spring Harbor Laboratory (2020-02-17) https://doi.org/ggrgbq
DOI: 10.1101/2020.02.14.20021535

919. Longitudinal characteristics of lymphocyte responses and cytokine profiles in the peripheral blood of SARS-CoV-2 infected patients
Jing Liu, Sumeng Li, Jia Liu, Boyun Liang, Xiaobei Wang, Hua Wang, Wei Li, Qiaoxia Tong, Jianhua Yi, Lei Zhao, … Xin Zheng
Cold Spring Harbor Laboratory (2020-02-22) https://doi.org/ggq63g
DOI: 10.1101/2020.02.16.20023671

920. Clinical and immunologic features in severe and moderate forms of Coronavirus Disease 2019
Guang Chen, Di Wu, Wei Guo, Yong Cao, Da Huang, Hongwu Wang, Tao Wang, Xiaoyun Zhang, Huilong Chen, Haijing Yu, … Qin Ning
Cold Spring Harbor Laboratory (2020-02-19) https://doi.org/ggq63h
DOI: 10.1101/2020.02.16.20023903

921. SARS-CoV-2 and SARS-CoV Spike-RBD Structure and Receptor Binding Comparison and Potential Implications on Neutralizing Antibody and Vaccine Development
Chunyun Sun, Long Chen, Ji Yang, Chunxia Luo, Yanjing Zhang, Jing Li, Jiahui Yang, Jie Zhang, Liangzhi Xie
Cold Spring Harbor Laboratory (2020-02-20) https://doi.org/ggq63j
DOI: 10.1101/2020.02.16.951723

922. Protection of Rhesus Macaque from SARS-Coronavirus challenge by recombinant adenovirus vaccine
Yiyou Chen, Qiang Wei, Ruobing Li, Hong Gao, Hua Zhu, Wei Deng, Linlin Bao, Wei Tong, Zhe Cong, Hong Jiang, Chuan Qin
Cold Spring Harbor Laboratory (2020-02-21) https://doi.org/ggq63k
DOI: 10.1101/2020.02.17.951939

923. Reduction and Functional Exhaustion of T Cells in Patients with Coronavirus Disease 2019 (COVID-19)
Bo Diao, Chenhui Wang, Yingjun Tan, Xiewan Chen, Ying Liu, Lifen Ning, Li Chen, Min Li, Yueping Liu, Gang Wang, … Yongwen Chen
Cold Spring Harbor Laboratory (2020-02-20) https://doi.org/ggq63m
DOI: 10.1101/2020.02.18.20024364

924. Clinical characteristics of 25 death cases with COVID-19: a retrospective review of medical records in a single medical center, Wuhan, China
Xun Li, Luwen Wang, Shaonan Yan, Fan Yang, Longkui Xiang, Jiling Zhu, Bo Shen, Zuojiong Gong
Cold Spring Harbor Laboratory (2020-02-25) https://doi.org/ggq63n
DOI: 10.1101/2020.02.19.20025239

925. SARS-CoV-2 infection does not significantly cause acute renal injury: an analysis of 116 hospitalized patients with COVID-19 in a single hospital, Wuhan, China
Luwen Wang, Xun Li, Hui Chen, Shaonan Yan, Yan Li, Dong Li, Zuojiong Gong
Cold Spring Harbor Laboratory (2020-02-23) https://doi.org/ggq63p
DOI: 10.1101/2020.02.19.20025288

926. Clinical Characteristics of 138 Hospitalized Patients With 2019 Novel Coronavirus–Infected Pneumonia in Wuhan, China
Dawei Wang, Bo Hu, Chang Hu, Fangfang Zhu, Xing Liu, Jing Zhang, Binbin Wang, Hui Xiang, Zhenshun Cheng, Yong Xiong, … Zhiyong Peng
JAMA (2020-03-17) https://doi.org/ggkh48
DOI: 10.1001/jama.2020.1585 · PMID: 32031570 · PMCID: PMC7042881

927. Clinical characteristics of 2019 novel coronavirus infection in China
Wei-jie Guan, Zheng-yi Ni, Yu Hu, Wen-hua Liang, Chun-quan Ou, Jian-xing He, Lei Liu, Hong Shan, Chun-liang Lei, David S. C. Hui, … Nan-shan Zhong
Cold Spring Harbor Laboratory (2020-02-09) https://doi.org/ggkj9s
DOI: 10.1101/2020.02.06.20020974

928. Potential T-cell and B-cell Epitopes of 2019-nCoV
Ethan Fast, Russ B. Altman, Binbin Chen
Cold Spring Harbor Laboratory (2020-03-18) https://doi.org/ggq63q
DOI: 10.1101/2020.02.19.955484

929. Structure, function and antigenicity of the SARS-CoV-2 spike glycoprotein
Alexandra C. Walls, Young-Jun Park, M. Alexandra Tortorici, Abigail Wall, Andrew T. McGuire, David Veesler
Cold Spring Harbor Laboratory (2020-02-20) https://doi.org/ggrgbr
DOI: 10.1101/2020.02.19.956581

930. Breadth of concomitant immune responses underpinning viral clearance and patient recovery in a non-severe case of COVID-19
Irani Thevarajan, Thi HO Nguyen, Marios Koutsakos, Julian Druce, Leon Caly, Carolien E van de Sandt, Xiaoxiao Jia, Suellen Nicholson, Mike Catton, Benjamin Cowie, … Katherine Kedzierska
Cold Spring Harbor Laboratory (2020-02-23) https://doi.org/ggq63r
DOI: 10.1101/2020.02.20.20025841

931. The landscape of lung bronchoalveolar immune cells in COVID-19 revealed by single-cell RNA sequencing
Minfeng Liao, Yang Liu, Jin Yuan, Yanling Wen, Gang Xu, Juanjuan Zhao, Lin Chen, Jinxiu Li, Xin Wang, Fuxiang Wang, … Zheng Zhang
Cold Spring Harbor Laboratory (2020-02-26) https://doi.org/ggq63s
DOI: 10.1101/2020.02.23.20026690

932. Influenza A Virus Infection Induces Hyperresponsiveness in Human Lung Tissue-Resident and Peripheral Blood NK Cells
Marlena Scharenberg, Sindhu Vangeti, Eliisa Kekäläinen, Per Bergman, Mamdoh Al-Ameri, Niclas Johansson, Klara Sondén, Sara Falck-Jones, Anna Färnert, Hans-Gustaf Ljunggren, … Nicole Marquardt
Frontiers in Immunology (2019-05-17) https://doi.org/ggq656
DOI: 10.3389/fimmu.2019.01116 · PMID: 31156653 · PMCID: PMC6534051

933. Clinical features of patients infected with 2019 novel coronavirus in Wuhan, China
Chaolin Huang, Yeming Wang, Xingwang Li, Lili Ren, Jianping Zhao, Yi Hu, Li Zhang, Guohui Fan, Jiuyang Xu, Xiaoying Gu, … Bin Cao
The Lancet (2020-02) https://doi.org/ggjfnn
DOI: 10.1016/s0140-6736(20)30183-5 · PMID: 31986264 · PMCID: PMC7159299

934. Alveolar Macrophages in the Resolution of Inflammation, Tissue Repair, and Tolerance to Infection
Benoit Allard, Alice Panariti, James G. Martin
Frontiers in Immunology (2018-07-31) https://doi.org/gd3bnz
DOI: 10.3389/fimmu.2018.01777 · PMID: 30108592 · PMCID: PMC6079255

935. PPAR-γ in Macrophages Limits Pulmonary Inflammation and Promotes Host Recovery following Respiratory Viral Infection
Su Huang, Bibo Zhu, In Su Cheon, Nick P. Goplen, Li Jiang, Ruixuan Zhang, R. Stokes Peebles, Matthias Mack, Mark H. Kaplan, Andrew H. Limper, Jie Sun
Journal of Virology (2019-04-17) https://doi.org/ggq652
DOI: 10.1128/jvi.00030-19 · PMID: 30787149 · PMCID: PMC6475778

936. Can routine laboratory tests discriminate 2019 novel coronavirus infected pneumonia from other community-acquired pneumonia?
Yunbao Pan, Guangming Ye, Xiantao Zeng, Guohong Liu, Xiaojiao Zeng, Xianghu Jiang, Jin Zhao, Liangjun Chen, Shuang Guo, Qiaoling Deng, … Xinghuan Wang
Cold Spring Harbor Laboratory (2020-02-25) https://doi.org/ggq63t
DOI: 10.1101/2020.02.25.20024711

937. Correlation Analysis Between Disease Severity and Inflammation-related Parameters in Patients with COVID-19 Pneumonia
Jing Gong, Hui Dong, Qingsong Xia, Zhaoyi Huang, Dingkun Wang, Yan Zhao, Wenhua Liu, Shenghao Tu, Mingmin Zhang, Qi Wang, Fuer Lu
Cold Spring Harbor Laboratory (2020-02-26) https://doi.org/ggq63v
DOI: 10.1101/2020.02.25.20025643

938. An Effective CTL Peptide Vaccine for Ebola Zaire Based on Survivors’ CD8+ Targeting of a Particular Nucleocapsid Protein Epitope with Potential Implications for COVID-19 Vaccine Design
CV Herst, S Burkholz, J Sidney, A Sette, PE Harris, S Massey, T Brasel, E Cunha-Neto, DS Rosa, WCH Chao, … R Rubsamen
Cold Spring Harbor Laboratory (2020-04-06) https://doi.org/ggq63x
DOI: 10.1101/2020.02.25.963546

939. Epitope-based peptide vaccine design and target site characterization against novel coronavirus disease caused by SARS-CoV-2
Lin Li, Ting Sun, Yufei He, Wendong Li, Yubo Fan, Jing Zhang
Cold Spring Harbor Laboratory (2020-02-27) https://doi.org/ggnqwt
DOI: 10.1101/2020.02.25.965434

940. The definition and risks of Cytokine Release Syndrome-Like in 11 COVID-19-Infected Pneumonia critically ill patients: Disease Characteristics and Retrospective Analysis
Wang Wenjun, Liu Xiaoqing, Wu Sipei, Lie Puyi, Huang Liyan, Li Yimin, Cheng Linling, Chen Sibei, Nong Lingbo, Lin Yongping, He Jianxing
Cold Spring Harbor Laboratory (2020-02-27) https://doi.org/ggrgbs
DOI: 10.1101/2020.02.26.20026989

941. Clinical characteristics of 36 non-survivors with COVID-19 in Wuhan, China
Ying Huang, Rui Yang, Ying Xu, Ping Gong
Cold Spring Harbor Laboratory (2020-03-05) https://doi.org/ggq63z
DOI: 10.1101/2020.02.27.20029009

942. Risk factors related to hepatic injury in patients with corona virus disease 2019
Lu Li, Shuang Li, Manman Xu, Pengfei Yu, Sujun Zheng, Zhongping Duan, Jing Liu, Yu Chen, Junfeng Li
Cold Spring Harbor Laboratory (2020-03-10) https://doi.org/ggq632
DOI: 10.1101/2020.02.28.20028514

943. Detectable serum SARS-CoV-2 viral load (RNAaemia) is closely associated with drastically elevated interleukin 6 (IL-6) level in critically ill COVID-19 patients
Xiaohua Chen, Binghong Zhao, Yueming Qu, Yurou Chen, Jie Xiong, Yong Feng, Dong Men, Qianchuan Huang, Ying Liu, Bo Yang, … Feng Li
Cold Spring Harbor Laboratory (2020-03-03) https://doi.org/ggq633
DOI: 10.1101/2020.02.29.20029520

944. Prognostic factors in the acute respiratory distress syndrome
Wei Chen, Lorraine B Ware
Clinical and Translational Medicine (2015-07-02) https://doi.org/ggq653
DOI: 10.1186/s40169-015-0065-2 · PMID: 26162279 · PMCID: PMC4534483

945. Lymphopenia predicts disease severity of COVID-19: a descriptive and predictive study
Li Tan, Qi Wang, Duanyang Zhang, Jinya Ding, Qianchuan Huang, Yi-Quan Tang, Qiongshu Wang, Hongming Miao
Cold Spring Harbor Laboratory (2020-03-03) https://doi.org/ggq634
DOI: 10.1101/2020.03.01.20029074

946. The potential role of IL-6 in monitoring severe case of coronavirus disease 2019
Tao Liu, Jieying Zhang, Yuhui Yang, Hong Ma, Zhengyu Li, Jiaoyu Zhang, Ji Cheng, Xiaoyun Zhang, Yanxia Zhao, Zihan Xia, … Jianhua Yi
Cold Spring Harbor Laboratory (2020-03-10) https://doi.org/ggq635
DOI: 10.1101/2020.03.01.20029769

947. Clinical and Laboratory Profiles of 75 Hospitalized Patients with Novel Coronavirus Disease 2019 in Hefei, China
Zonghao Zhao, Jiajia Xie, Ming Yin, Yun Yang, Hongliang He, Tengchuan Jin, Wenting Li, Xiaowu Zhu, Jing Xu, Changcheng Zhao, … Xiaoling Ma
Cold Spring Harbor Laboratory (2020-03-06) https://doi.org/ggq636
DOI: 10.1101/2020.03.01.20029785

948. Exuberant elevation of IP-10, MCP-3 and IL-1ra during SARS-CoV-2 infection is associated with disease severity and fatal outcome
Yang Yang, Chenguang Shen, Jinxiu Li, Jing Yuan, Minghui Yang, Fuxiang Wang, Guobao Li, Yanjie Li, Li Xing, Ling Peng, … Yingxia Liu
Cold Spring Harbor Laboratory (2020-03-06) https://doi.org/ggq637
DOI: 10.1101/2020.03.02.20029975

949. Antibody responses to SARS-CoV-2 in patients of novel coronavirus disease 2019
Juanjuan Zhao, Quan Yuan, Haiyan Wang, Wei Liu, Xuejiao Liao, Yingying Su, Xin Wang, Jing Yuan, Tingdong Li, Jinxiu Li, … Zheng Zhang
Cold Spring Harbor Laboratory (2020-03-02) https://doi.org/ggrbj6
DOI: 10.1101/2020.03.02.20030189

950. Restoration of leukomonocyte counts is associated with viral clearance in COVID-19 hospitalized patients
Xiaoping Chen, Jiaxin Ling, Pingzheng Mo, Yongxi Zhang, Qunqun Jiang, Zhiyong Ma, Qian Cao, Wenjia Hu, Shi Zou, Liangjun Chen, … Yong Xiong
Cold Spring Harbor Laboratory (2020-03-06) https://doi.org/ggq639
DOI: 10.1101/2020.03.03.20030437

951. Effects of Systemically Administered Hydrocortisone on the Human Immunome
Matthew J. Olnes, Yuri Kotliarov, Angélique Biancotto, Foo Cheung, Jinguo Chen, Rongye Shi, Huizhi Zhou, Ena Wang, John S. Tsang, Robert Nussenblatt, The CHI Consortium
Scientific Reports (2016-03-14) https://doi.org/f8dmvw
DOI: 10.1038/srep23002 · PMID: 26972611 · PMCID: PMC4789739

952. Procalcitonin in patients with severe coronavirus disease 2019 (COVID-19): A meta-analysis
Giuseppe Lippi, Mario Plebani
Clinica Chimica Acta (2020-06) https://doi.org/ggpxp7
DOI: 10.1016/j.cca.2020.03.004 · PMID: 32145275 · PMCID: PMC7094472

953. Clinical findings in critically ill patients infected with SARS-CoV-2 in Guangdong Province, China: a multi-center, retrospective, observational study
Yonghao Xu, Zhiheng Xu, Xuesong Liu, Lihua Cai, Haichong Zheng, Yongbo Huang, Lixin Zhou, Linxi Huang, Yun Lin, Liehua Deng, … Yimin Li
Cold Spring Harbor Laboratory (2020-03-06) https://doi.org/ggq64b
DOI: 10.1101/2020.03.03.20030668

954. Multi-epitope vaccine design using an immunoinformatics approach for 2019 novel coronavirus (SARS-CoV-2)
Ye Feng, Min Qiu, Liang Liu, Shengmei Zou, Yun Li, Kai Luo, Qianpeng Guo, Ning Han, Yingqiang Sun, Kui Wang, … Fan Mo
Cold Spring Harbor Laboratory (2020-06-30) https://doi.org/ggq64c
DOI: 10.1101/2020.03.03.962332

955. Clinical Features of Patients Infected with the 2019 Novel Coronavirus (COVID-19) in Shanghai, China
Min Cao, Dandan Zhang, Youhua Wang, Yunfei Lu, Xiangdong Zhu, Ying Li, Honghao Xue, Yunxiao Lin, Min Zhang, Yiguo Sun, … Hongzhou Lu
Cold Spring Harbor Laboratory (2020-03-06) https://doi.org/ggq64d
DOI: 10.1101/2020.03.04.20030395 · PMID: 32511465 · PMCID: PMC7255784

956. Serological detection of 2019-nCoV respond to the epidemic: A useful complement to nucleic acid testing
Jin Zhang, Jianhua Liu, Na Li, Yong Liu, Rui Ye, Xiaosong Qin, Rui Zheng
Cold Spring Harbor Laboratory (2020-03-06) https://doi.org/ggq64f
DOI: 10.1101/2020.03.04.20030916

957. Human Kidney is a Target for Novel Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) Infection
Bo Diao, Chenhui Wang, Rongshuai Wang, Zeqing Feng, Yingjun Tan, Huiming Wang, Changsong Wang, Liang Liu, Ying Liu, Yueping Liu, … Yongwen Chen
Cold Spring Harbor Laboratory (2020-04-10) https://doi.org/ggq64g
DOI: 10.1101/2020.03.04.20031120

958. COVID-19 early warning score: a multi-parameter screening tool to identify highly suspected patients
Cong-Ying Song, Jia Xu, Jian-Qin He, Yuan-Qiang Lu
Cold Spring Harbor Laboratory (2020-03-08) https://doi.org/ggq64h
DOI: 10.1101/2020.03.05.20031906

959. LY6E impairs coronavirus fusion and confers immune control of viral disease
Stephanie Pfaender, Katrina B. Mar, Eleftherios Michailidis, Annika Kratzel, Dagny Hirt, Philip V’kovski, Wenchun Fan, Nadine Ebert, Hanspeter Stalder, Hannah Kleine-Weber, … Volker Thiel
Cold Spring Harbor Laboratory (2020-03-07) https://doi.org/dpvn
DOI: 10.1101/2020.03.05.979260 · PMID: 32511345 · PMCID: PMC7255780

960. A preliminary study on serological assay for severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) in 238 admitted hospital patients
Lei Liu, Wanbing Liu, Yaqiong Zheng, Xiaojing Jiang, Guomei Kou, Jinya Ding, Qiongshu Wang, Qianchuan Huang, Yinjuan Ding, Wenxu Ni, … Shangen Zheng
Cold Spring Harbor Laboratory (2020-03-08) https://doi.org/ggq64j
DOI: 10.1101/2020.03.06.20031856

961. Monoclonal antibodies for the S2 subunit of spike of SARS-CoV cross-react with the newly-emerged SARS-CoV-2
Zhiqiang Zheng, Vanessa M. Monteil, Sebastian Maurer-Stroh, Chow Wenn Yew, Carol Leong, Nur Khairiah Mohd-Ismail, Suganya Cheyyatraivendran Arularasu, Vincent Tak Kwong Chow, Raymond Lin Tzer Pin, Ali Mirazimi, … Yee-Joo Tan
Cold Spring Harbor Laboratory (2020-03-07) https://doi.org/ggrbj7
DOI: 10.1101/2020.03.06.980037

962. Mortality of COVID-19 is Associated with Cellular Immune Function Compared to Immune Function in Chinese Han Population
Qiang Zeng, Yong-zhe Li, Gang Huang, Wei Wu, Sheng-yong Dong, Yang Xu
Cold Spring Harbor Laboratory (2020-03-13) https://doi.org/ggq64k
DOI: 10.1101/2020.03.08.20031229

963. Retrospective Analysis of Clinical Features in 101 Death Cases with COVID-19
Hua Fan, Lin Zhang, Bin Huang, Muxin Zhu, Yong Zhou, Huan Zhang, Xiaogen Tao, Shaohui Cheng, Wenhu Yu, Liping Zhu, Jian Chen
Cold Spring Harbor Laboratory (2020-03-17) https://doi.org/ggq64n
DOI: 10.1101/2020.03.09.20033068

964. Relationship between the ABO Blood Group and the COVID-19 Susceptibility
Jiao Zhao, Yan Yang, Hanping Huang, Dong Li, Dongfeng Gu, Xiangfeng Lu, Zheng Zhang, Lei Liu, Ting Liu, Yukun Liu, … Peng George Wang
Cold Spring Harbor Laboratory (2020-03-27) https://doi.org/ggpn3d
DOI: 10.1101/2020.03.11.20031096

965. The inhaled corticosteroid ciclesonide blocks coronavirus RNA replication by targeting viral NSP15
Shutoku Matsuyama, Miyuki Kawase, Naganori Nao, Kazuya Shirato, Makoto Ujike, Wataru Kamitani, Masayuki Shimojima, Shuetsu Fukushi
Cold Spring Harbor Laboratory (2020-03-12) https://doi.org/ggq64p
DOI: 10.1101/2020.03.11.987016

966. Immune phenotyping based on neutrophil-to-lymphocyte ratio and IgG predicts disease severity and outcome for patients with COVID-19
Bicheng Zhang, Xiaoyang Zhou, Chengliang Zhu, Fan Feng, Yanru Qiu, Jia Feng, Qingzhu Jia, Qibin Song, Bo Zhu, Jun Wang
Cold Spring Harbor Laboratory (2020-03-16) https://doi.org/ggq64q
DOI: 10.1101/2020.03.12.20035048

967. Lack of Reinfection in Rhesus Macaques Infected with SARS-CoV-2
Linlin Bao, Wei Deng, Hong Gao, Chong Xiao, Jiayi Liu, Jing Xue, Qi Lv, Jiangning Liu, Pin Yu, Yanfeng Xu, … Chuan Qin
Cold Spring Harbor Laboratory (2020-05-01) https://doi.org/ggn8r8
DOI: 10.1101/2020.03.13.990226

968. A highly conserved cryptic epitope in the receptor-binding domains of SARS-CoV-2 and SARS-CoV
Meng Yuan, Nicholas C. Wu, Xueyong Zhu, Chang-Chun D. Lee, Ray T. Y. So, Huibin Lv, Chris K. P. Mok, Ian A. Wilson
Cold Spring Harbor Laboratory (2020-03-14) https://doi.org/ggq64s
DOI: 10.1101/2020.03.13.991570

969. SARS-CoV-2 invades host cells via a novel route: CD147-spike protein
Ke Wang, Wei Chen, Yu-Sen Zhou, Jian-Qi Lian, Zheng Zhang, Peng Du, Li Gong, Yang Zhang, Hong-Yong Cui, Jie-Jie Geng, … Zhi-Nan Chen
Cold Spring Harbor Laboratory (2020-03-14) https://doi.org/ggq64t
DOI: 10.1101/2020.03.14.988345

970. CD147 (EMMPRIN/Basigin) in kidney diseases: from an inflammation and immune system viewpoint
Tomoki Kosugi, Kayaho Maeda, Waichi Sato, Shoichi Maruyama, Kenji Kadomatsu
Nephrology Dialysis Transplantation (2015-07) https://doi.org/ggq624
DOI: 10.1093/ndt/gfu302 · PMID: 25248362

971. The roles of CyPA and CD147 in cardiac remodelling
Hongyan Su, Yi Yang
Experimental and Molecular Pathology (2018-06) https://doi.org/ggq622
DOI: 10.1016/j.yexmp.2018.05.001 · PMID: 29772453

972. Cancer-related issues of CD147.
Ulrich H Weidle, Werner Scheuer, Daniela Eggle, Stefan Klostermann, Hannes Stockinger
Cancer genomics & proteomics https://www.ncbi.nlm.nih.gov/pubmed/20551248
PMID: 20551248

973. Blood single cell immune profiling reveals the interferon-MAPK pathway mediated adaptive immune response for COVID-19
Lulin Huang, Yi Shi, Bo Gong, Li Jiang, Xiaoqi Liu, Jialiang Yang, Juan Tang, Chunfang You, Qi Jiang, Bo Long, … Zhenglin Yang
Cold Spring Harbor Laboratory (2020-03-17) https://doi.org/ggq64v
DOI: 10.1101/2020.03.15.20033472

974. Cross-reactive antibody response between SARS-CoV-2 and SARS-CoV infections
Huibin Lv, Nicholas C. Wu, Owen Tak-Yin Tsang, Meng Yuan, Ranawaka A. P. M. Perera, Wai Shing Leung, Ray T. Y. So, Jacky Man Chun Chan, Garrick K. Yip, Thomas Shiu Hong Chik, … Chris K. P. Mok
Cold Spring Harbor Laboratory (2020-03-17) https://doi.org/ggq64w
DOI: 10.1101/2020.03.15.993097 · PMID: 32511317 · PMCID: PMC7239046

975. The feasibility of convalescent plasma therapy in severe COVID- 19 patients: a pilot study
Kai Duan, Bende Liu, Cesheng Li, Huajun Zhang, Ting Yu, Jieming Qu, Min Zhou, Li Chen, Shengli Meng, Yong Hu, … Xiaoming Yang
Cold Spring Harbor Laboratory (2020-03-23) https://doi.org/dqrs
DOI: 10.1101/2020.03.16.20036145

976. Hydroxychloroquine and azithromycin as a treatment of COVID-19: results of an open-label non-randomized clinical trial
Philippe Gautret, Jean-Christophe Lagier, Philippe Parola, Van Thuan Hoang, Line Meddeb, Morgane Mailhe, Barbara Doudier, Johan Courjon, Valérie Giordanengo, Vera Esteves Vieira, … Didier Raoult
Cold Spring Harbor Laboratory (2020-03-20) https://doi.org/dqbv
DOI: 10.1101/2020.03.16.20037135

977. Chloroquine: Modes of action of an undervalued drug
Rodolfo Thomé, Stefanie Costa Pinto Lopes, Fabio Trindade Maranhão Costa, Liana Verinaud
Immunology Letters (2013-06) https://doi.org/f5b5cr
DOI: 10.1016/j.imlet.2013.07.004 · PMID: 23891850

978. Patients with LRBA deficiency show CTLA4 loss and immune dysregulation responsive to abatacept therapy
B. Lo, K. Zhang, W. Lu, L. Zheng, Q. Zhang, C. Kanellopoulou, Y. Zhang, Z. Liu, J. M. Fritz, R. Marsh, … M. B. Jordan
Science (2015-07-23) https://doi.org/f7kc8d
DOI: 10.1126/science.aaa1663 · PMID: 26206937

979. The sequence of human ACE2 is suboptimal for binding the S spike protein of SARS coronavirus 2
Erik Procko
Cold Spring Harbor Laboratory (2020-05-11) https://doi.org/ggrbj8
DOI: 10.1101/2020.03.16.994236 · PMID: 32511321 · PMCID: PMC7239051

980. Comparative Pathogenesis Of COVID-19, MERS And SARS In A Non-Human Primate Model
Barry Rockx, Thijs Kuiken, Sander Herfst, Theo Bestebroer, Mart M. Lamers, Dennis de Meulder, Geert van Amerongen, Judith van den Brand, Nisreen M. A. Okba, Debby Schipper, … Bart L. Haagmans
Cold Spring Harbor Laboratory (2020-03-17) https://doi.org/ggq649
DOI: 10.1101/2020.03.17.995639

981. Lethal Infection of K18-hACE2 Mice Infected with Severe Acute Respiratory Syndrome Coronavirus
Paul B. McCray, Lecia Pewe, Christine Wohlford-Lenane, Melissa Hickey, Lori Manzel, Lei Shi, Jason Netland, Hong Peng Jia, Carmen Halabi, Curt D. Sigmund, … Stanley Perlman
Journal of Virology (2007-01-15) https://doi.org/b2dr3s
DOI: 10.1128/jvi.02012-06 · PMID: 17079315 · PMCID: PMC1797474

982. Investigating the Impact of Asymptomatic Carriers on COVID-19 Transmission
Jacob B. Aguilar, Jeremy Samuel Faust, Lauren M. Westafer, Juan B. Gutierrez
Cold Spring Harbor Laboratory (2020-03-31) https://doi.org/ggqnvp
DOI: 10.1101/2020.03.18.20037994

983. Antibody responses to SARS-CoV-2 in COVID-19 patients: the perspective application of serological tests in clinical practice
Quan-xin Long, Hai-jun Deng, Juan Chen, Jie-li Hu, Bei-zhong Liu, Pu Liao, Yong Lin, Li-hua Yu, Zhan Mo, Yin-yin Xu, … Ai-long Huang
Cold Spring Harbor Laboratory (2020-03-20) https://doi.org/ggpvz3
DOI: 10.1101/2020.03.18.20038018

984. Heat inactivation of serum interferes with the immunoanalysis of antibodies to SARS-CoV-2
Xiumei Hu, Taixue An, Bo Situ, Yuhai Hu, Zihao Ou, Qiang Li, Xiaojing He, Ye Zhang, Peifu Tian, Dehua Sun, … Lei Zheng
Cold Spring Harbor Laboratory (2020-03-16) https://doi.org/ggq646
DOI: 10.1101/2020.03.12.20034231

985. SARS-CoV-2 specific antibody responses in COVID-19 patients
Nisreen M. A. Okba, Marcel A. Müller, Wentao Li, Chunyan Wang, Corine H. GeurtsvanKessel, Victor M. Corman, Mart M. Lamers, Reina S. Sikkema, Erwin de Bruin, Felicity D. Chandler, … Bart L. Haagmans
Cold Spring Harbor Laboratory (2020-03-20) https://doi.org/ggpvz2
DOI: 10.1101/2020.03.18.20038059

986. A brief review of antiviral drugs evaluated in registered clinical trials for COVID-19
Drifa Belhadi, Nathan Peiffer-Smadja, François-Xavier Lescure, Yazdan Yazdanpanah, France Mentré, Cédric Laouénan
Cold Spring Harbor Laboratory (2020-03-27) https://doi.org/ggq65b
DOI: 10.1101/2020.03.18.20038190

987. ACE-2 Expression in the Small Airway Epithelia of Smokers and COPD Patients: Implications for COVID-19
Janice M. Leung, Chen X. Yang, Anthony Tam, Tawimas Shaipanich, Tillie-Louise Hackett, Gurpreet K. Singhera, Delbert R. Dorscheid, Don D. Sin
Cold Spring Harbor Laboratory (2020-03-23) https://doi.org/dqx2
DOI: 10.1101/2020.03.18.20038455

988. Dynamic profile of severe or critical COVID-19 cases
Yang Xu
Cold Spring Harbor Laboratory (2020-03-20) https://doi.org/ggrbj9
DOI: 10.1101/2020.03.18.20038513

989. Association between Clinical, Laboratory and CT Characteristics and RT-PCR Results in the Follow-up of COVID-19 patients
Hang Fu, Huayan Xu, Na Zhang, Hong Xu, Zhenlin Li, Huizhu Chen, Rong Xu, Ran Sun, Lingyi Wen, Linjun Xie, … Yingkun Guo
Cold Spring Harbor Laboratory (2020-03-23) https://doi.org/ggq65c
DOI: 10.1101/2020.03.19.20038315

990. An orally bioavailable broad-spectrum antiviral inhibits SARS-CoV-2 and multiple endemic, epidemic and bat coronavirus
Timothy P. Sheahan, Amy C. Sims, Shuntai Zhou, Rachel L. Graham, Collin S. Hill, Sarah R. Leist, Alexandra Schäfer, Kenneth H. Dinnon, Stephanie A. Montgomery, Maria L. Agostini, … Ralph S. Baric
Cold Spring Harbor Laboratory (2020-03-20) https://doi.org/ggrbkb
DOI: 10.1101/2020.03.19.997890

991. Identification of antiviral drug candidates against SARS-CoV-2 from FDA-approved drugs
Sangeun Jeon, Meehyun Ko, Jihye Lee, Inhee Choi, Soo Young Byun, Soonju Park, David Shum, Seungtaek Kim
Cold Spring Harbor Laboratory (2020-03-28) https://doi.org/ggq65h
DOI: 10.1101/2020.03.20.999730

992. Respiratory disease and virus shedding in rhesus macaques inoculated with SARS-CoV-2
Vincent J. Munster, Friederike Feldmann, Brandi N. Williamson, Neeltje van Doremalen, Lizzette Pérez-Pérez, Jonathan Schulz, Kimberly Meade-White, Atsushi Okumura, Julie Callison, Beniah Brumbaugh, … Emmie de Wit
Cold Spring Harbor Laboratory (2020-03-21) https://doi.org/ggq65j
DOI: 10.1101/2020.03.21.001628 · PMID: 32511299 · PMCID: PMC7217148

993. Ocular conjunctival inoculation of SARS-CoV-2 can cause mild COVID-19 in Rhesus macaques
Wei Deng, Linlin Bao, Hong Gao, Zhiguang Xiang, Yajin Qu, Zhiqi Song, Shunran Gong, Jiayi Liu, Jiangning Liu, Pin Yu, … Chuan Qin
Cold Spring Harbor Laboratory (2020-03-30) https://doi.org/ggq64r
DOI: 10.1101/2020.03.13.990036

994. ACE2 Expression is Increased in the Lungs of Patients with Comorbidities Associated with Severe COVID-19
Bruna G. G. Pinto, Antonio E. R. Oliveira, Youvika Singh, Leandro Jimenez, Andre N A. Gonçalves, Rodrigo L. T. Ogava, Rachel Creighton, Jean Pierre Schatzmann Peron, Helder I. Nakaya
Cold Spring Harbor Laboratory (2020-03-27) https://doi.org/ggq65k
DOI: 10.1101/2020.03.21.20040261 · PMID: 32511627 · PMCID: PMC7276054

995. Meplazumab treats COVID-19 pneumonia: an open-labelled, concurrent controlled add-on clinical trial
Huijie Bian, Zhao-Hui Zheng, Ding Wei, Zheng Zhang, Wen-Zhen Kang, Chun-Qiu Hao, Ke Dong, Wen Kang, Jie-Lai Xia, Jin-Lin Miao, … Ping Zhu
Cold Spring Harbor Laboratory (2020-07-15) https://doi.org/ggq65m
DOI: 10.1101/2020.03.21.20040691

996. CD147 facilitates HIV-1 infection by interacting with virus-associated cyclophilin A
T. Pushkarsky, G. Zybarth, L. Dubrovsky, V. Yurchenko, H. Tang, H. Guo, B. Toole, B. Sherry, M. Bukrinsky
Proceedings of the National Academy of Sciences (2001-05-15) https://doi.org/cc4c7p
DOI: 10.1073/pnas.111583198 · PMID: 11353871 · PMCID: PMC33473

997. CD147/EMMPRIN Acts as a Functional Entry Receptor for Measles Virus on Epithelial Cells
Akira Watanabe, Misako Yoneda, Fusako Ikeda, Yuri Terao-Muto, Hiroki Sato, Chieko Kai
Journal of Virology (2010-05-01) https://doi.org/dpcsqg
DOI: 10.1128/jvi.02168-09 · PMID: 20147391 · PMCID: PMC2863760

998. Basigin is a receptor essential for erythrocyte invasion by Plasmodium falciparum
Cécile Crosnier, Leyla Y. Bustamante, S. Josefin Bartholdson, Amy K. Bei, Michel Theron, Makoto Uchikawa, Souleymane Mboup, Omar Ndir, Dominic P. Kwiatkowski, Manoj T. Duraisingh, … Gavin J. Wright
Nature (2011-11-09) https://doi.org/dm59hf
DOI: 10.1038/nature10606 · PMID: 22080952 · PMCID: PMC3245779

999. Function of HAb18G/CD147 in Invasion of Host Cells by Severe Acute Respiratory Syndrome Coronavirus
Zhinan Chen, Li Mi, Jing Xu, Jiyun Yu, Xianhui Wang, Jianli Jiang, Jinliang Xing, Peng Shang, Airong Qian, Yu Li, … Ping Zhu
The Journal of Infectious Diseases (2005-03) https://doi.org/cd8snd
DOI: 10.1086/427811 · PMID: 15688292 · PMCID: PMC7110046

1000. CD147 mediates intrahepatic leukocyte aggregation and determines the extent of liver injury
Christine Yee, Nathan M. Main, Alexandra Terry, Igor Stevanovski, Annette Maczurek, Alison J. Morgan, Sarah Calabro, Alison J. Potter, Tina L. Iemma, David G. Bowen, … Nicholas A. Shackel
PLOS ONE (2019-07-10) https://doi.org/ggq654
DOI: 10.1371/journal.pone.0215557 · PMID: 31291257 · PMCID: PMC6619953

1001. Characterisation of the transcriptome and proteome of SARS-CoV-2 using direct RNA sequencing and tandem mass spectrometry reveals evidence for a cell passage induced in-frame deletion in the spike glycoprotein that removes the furin-like cleavage site
Andrew D. Davidson, Maia Kavanagh Williamson, Sebastian Lewis, Deborah Shoemark, Miles W. Carroll, Kate Heesom, Maria Zambon, Joanna Ellis, Phillip A. Lewis, Julian A. Hiscox, David A. Matthews
Cold Spring Harbor Laboratory (2020-03-24) https://doi.org/ggq65n
DOI: 10.1101/2020.03.22.002204

1002. Modifications to the Hemagglutinin Cleavage Site Control the Virulence of a Neurotropic H1N1 Influenza Virus
Xiangjie Sun, Longping V. Tse, A Damon Ferguson, Gary R. Whittaker
Journal of Virology (2010-09-01) https://doi.org/drs2zt
DOI: 10.1128/jvi.00797-10 · PMID: 20554779 · PMCID: PMC2919019

1003. The architecture of SARS-CoV-2 transcriptome
Dongwan Kim, Joo-Yeon Lee, Jeong-Sun Yang, Jun Won Kim, V. Narry Kim, Hyeshik Chang
Cold Spring Harbor Laboratory (2020-03-14) https://doi.org/ggpx9q
DOI: 10.1101/2020.03.12.988865

1004. First Clinical Study Using HCV Protease Inhibitor Danoprevir to Treat Naïve and Experienced COVID-19 Patients
Hongyi Chen, Zhicheng Zhang, Li Wang, Zhihua Huang, Fanghua Gong, Xiaodong Li, Yahong Chen, Jinzi J. Wu
Cold Spring Harbor Laboratory (2020-03-24) https://doi.org/ggrgbt
DOI: 10.1101/2020.03.22.20034041

1005. Preclinical Characteristics of the Hepatitis C Virus NS3/4A Protease Inhibitor ITMN-191 (R7227)
Scott D. Seiwert, Steven W. Andrews, Yutong Jiang, Vladimir Serebryany, Hua Tan, Karl Kossen, P. T. Ravi Rajagopalan, Shawn Misialek, Sarah K. Stevens, Antitsa Stoycheva, … Lawrence M. Blatt
Antimicrobial Agents and Chemotherapy (2008-12) https://doi.org/btpg52
DOI: 10.1128/aac.00699-08 · PMID: 18824605 · PMCID: PMC2592891

1006. Efficacy and Safety of All-oral, 12-week Ravidasvir Plus Ritonavir-boosted Danoprevir and Ribavirin in Treatment-naïve Noncirrhotic HCV Genotype 1 Patients: Results from a Phase 2/3 Clinical Trial in China
Xiaoyuan Xu, Bo Feng, Yujuan Guan, Sujun Zheng, Jifang Sheng, Xingxiang Yang, Yuanji Ma, Yan Huang, Yi Kang, Xiaofeng Wen, … Lai Wei
Journal of Clinical and Translational Hepatology (2019-09-30) https://doi.org/ggrbkd
DOI: 10.14218/jcth.2019.00033 · PMID: 31608212 · PMCID: PMC6783683

1007. Potentially highly potent drugs for 2019-nCoV
Duc Duy Nguyen, Kaifu Gao, Jiahui Chen, Rui Wang, Guo-Wei Wei
Cold Spring Harbor Laboratory (2020-02-13) https://doi.org/ggrbj5
DOI: 10.1101/2020.02.05.936013 · PMID: 32511344 · PMCID: PMC7255774

1008. Serology characteristics of SARS-CoV-2 infection since the exposure and post symptoms onset
Bin Lou, Ting-Dong Li, Shu-Fa Zheng, Ying-Ying Su, Zhi-Yong Li, Wei Liu, Fei Yu, Sheng-Xiang Ge, Qian-Da Zou, Quan Yuan, … Yu Chen
Cold Spring Harbor Laboratory (2020-03-27) https://doi.org/ggrbkc
DOI: 10.1101/2020.03.23.20041707

1009. SARS-CoV-2 launches a unique transcriptional signature from in vitro, ex vivo, and in vivo systems
Daniel Blanco-Melo, Benjamin E. Nilsson-Payant, Wen-Chun Liu, Rasmus Møller, Maryline Panis, David Sachs, Randy A. Albrecht, Benjamin R. tenOever
Cold Spring Harbor Laboratory (2020-03-24) https://doi.org/ggq65q
DOI: 10.1101/2020.03.24.004655

1010. A New Predictor of Disease Severity in Patients with COVID-19 in Wuhan, China
Ying Zhou, Zhen Yang, Yanan Guo, Shuang Geng, Shan Gao, Shenglan Ye, Yi Hu, Yafei Wang
Cold Spring Harbor Laboratory (2020-03-27) https://doi.org/ggq65r
DOI: 10.1101/2020.03.24.20042119

1011. Metabolic disturbances and inflammatory dysfunction predict severity of coronavirus disease 2019 (COVID-19): a retrospective study
Shuke Nie, Xueqing Zhao, Kang Zhao, Zhaohui Zhang, Zhentao Zhang, Zhan Zhang
Cold Spring Harbor Laboratory (2020-03-26) https://doi.org/ggq65s
DOI: 10.1101/2020.03.24.20042283

1012. Viral Kinetics and Antibody Responses in Patients with COVID-19
Wenting Tan, Yanqiu Lu, Juan Zhang, Jing Wang, Yunjie Dan, Zhaoxia Tan, Xiaoqing He, Chunfang Qian, Qiangzhong Sun, Qingli Hu, … Guohong Deng
Cold Spring Harbor Laboratory (2020-03-26) https://doi.org/ggq65t
DOI: 10.1101/2020.03.24.20042382

1013. Global profiling of SARS-CoV-2 specific IgG/ IgM responses of convalescents using a proteome microarray
He-wei Jiang, Yang Li, Hai-nan Zhang, Wei Wang, Dong Men, Xiao Yang, Huan Qi, Jie Zhou, Sheng-ce Tao
Cold Spring Harbor Laboratory (2020-03-27) https://doi.org/ggq65g
DOI: 10.1101/2020.03.20.20039495

1014. COVID-19 infection induces readily detectable morphological and inflammation-related phenotypic changes in peripheral blood monocytes, the severity of which correlate with patient outcome
Dan Zhang, Rui Guo, Lei Lei, Hongjuan Liu, Yawen Wang, Yili Wang, Hongbo Qian, Tongxin Dai, Tianxiao Zhang, Yanjun Lai, … Jinsong Hu
Cold Spring Harbor Laboratory (2020-03-26) https://doi.org/ggq65v
DOI: 10.1101/2020.03.24.20042655

1015. Correlation between universal BCG vaccination policy and reduced morbidity and mortality for COVID-19: an epidemiological study
Aaron Miller, Mac Josh Reandelar, Kimberly Fasciglione, Violeta Roumenova, Yan Li, Gonzalo H. Otazu
Cold Spring Harbor Laboratory (2020-03-28) https://doi.org/ggq65w
DOI: 10.1101/2020.03.24.20042937

1016. Non-specific effects of BCG vaccine on viral infections
S. J. C. F. M. Moorlag, R. J. W. Arts, R. van Crevel, M. G. Netea
Clinical Microbiology and Infection (2019-12) https://doi.org/ggq62z
DOI: 10.1016/j.cmi.2019.04.020 · PMID: 31055165

1017. BCG vaccination to reduce the impact of COVID-19 in healthcare workers (The BRACE Trial)
Murdoch Children’s Research Institute
https://www.mcri.edu.au/BRACE

1018. Non-neuronal expression of SARS-CoV-2 entry genes in the olfactory system suggests mechanisms underlying COVID-19-associated anosmia
David H. Brann, Tatsuya Tsukahara, Caleb Weinreb, Marcela Lipovsek, Koen Van den Berge, Boying Gong, Rebecca Chance, Iain C. Macaulay, Hsin-jung Chou, Russell Fletcher, … Sandeep Robert Datta
Cold Spring Harbor Laboratory (2020-05-18) https://doi.org/ggqr4m
DOI: 10.1101/2020.03.25.009084

1019. Cigarette smoke exposure and inflammatory signaling increase the expression of the SARS-CoV-2 receptor ACE2 in the respiratory tract
Joan C. Smith, Erin L. Sausville, Vishruth Girish, Monet Lou Yuan, Kristen M. John, Jason M. Sheltzer
Cold Spring Harbor Laboratory (2020-04-26) https://doi.org/ggq65x
DOI: 10.1101/2020.03.28.013672

1020. The comparative superiority of IgM-IgG antibody test to real-time reverse transcriptase PCR detection for SARS-CoV-2 infection diagnosis
Rui Liu, Xinghui Liu, Huan Han, Muhammad Adnan Shereen, Zhili Niu, Dong Li, Fang Liu, Kailang Wu, Zhen Luo, Chengliang Zhu
Cold Spring Harbor Laboratory (2020-03-30) https://doi.org/ggqtp5
DOI: 10.1101/2020.03.28.20045765

12 Appendix A

This appendix contains reviews produced by the Immunology Institute of the Icahn School of Medicine

12.1 Potent binding of 2019 novel coronavirus spike protein by a SARS coronavirus-specific human monoclonal antibody

Tian et al. Emerg Microbes Infect 2020 [898]

12.1.1 Keywords

12.1.2 Summary

Considering the relatively high identity of the receptor binding domain (RBD) of the spike proteins from 2019-nCoV and SARS-CoV (73%), this study aims to assess the cross-reactivity of several anti-SARS-CoV monoclonal antibodies with 2019-nCoV. The results showed that the SARS-CoV-specific antibody CR3022 can potently bind 2019-nCoV RBD.

12.1.3 Main Findings

The structure of the 2019-nCoV spike RBD and its conformation in complex with the receptor angiotensin-converting enzyme (ACE2) was modeled in silico and compared with the SARS-CoV RBD structure. The models predicted very similar RBD-ACE2 interactions for both viruses. The binding capacity of representative SARS-CoV-RBD specific monoclonal antibodies (m396, CR3014, and CR3022) to recombinant 2019-nCoV RBD was then investigated by ELISA and their binding kinetics studied using biolayer interferometry. The analysis showed that only CR3022 was able to bind 2019-nCoV RBD with high affinity (KD of 6.3 nM), however it did not interfere with ACE2 binding. Antibodies m396 and CR3014, which target the ACE2 binding site of SARS-CoV failed to bind 2019-nCoV spike protein.

12.1.4 Limitations

The 2019-nCoV RBD largely differ from the SARS-CoV at the C-terminus residues, which drastically impact the cross-reactivity of antibodies described for other B beta-coronaviruses, including SARS-CoV. This study claims that CR3022 antibody could be a potential candidate for therapy. However, none of the antibodies assayed in this work showed cross-reactivity with the ACE2 binding site of 2019-nCoV, essential for the replication of this virus. Furthermore, neutralization assays with 2019-nCoV virus or pseudovirus were not performed. Although the use of neutralizing antibodies is an interesting approach, these results suggest that it is critical the development of novel monoclonal antibodies able to specifically bind 2019-nCoV spike protein.

12.1.5 Credit

Review by D.L.O as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.2 Integrative Bioinformatics Analysis Provides Insight into the Molecular Mechanisms of 2019-nCoV

He et al. medRxiv [899]

12.2.1 Keywords

12.2.2 Main Findings

The authors used bioinformatics tools to identify features of ACE2 expression in the lungs of different patent groups: healthy, smokers, patients with chronic airway disease (i.e., COPD) or asthma. They used gene expression data publicly available from GEO that included lung tissues, bronchoalveolar lavage, bronchial epithelial cells, small airway epithelial cells, or SARS-Cov infected cells.

The authors describe no significant differences in ACE2 expression in lung tissues of Healthy, COPD, and Asthma groups (p=0.85); or in BAL of Healthy and COPD (p=0.48); or in epithelial brushings of Healthy and Mild/Moderate/Severe Asthma (p=0.99). ACE2 was higher in the small airway epithelium of long-term smokers vs non-smokers (p<0.001). Consistently, there was a trend of higher ACE2 expression in the bronchial airway epithelial cells 24h post-acute smoking exposure (p=0.073). Increasing ACE2 expression at 24h and 48h compared to 12h post SARS-Cov infection (p=0.026; n=3 at each time point) was also detected.

15 lung samples’ data from healthy participants were separated into high and low ACE2 expression groups. “High” ACE2 expression was associated with the following GO pathways: innate and adaptive immune responses, B cell mediated immunity, cytokine secretion, and IL-1, IL-10, IL-6, IL-8 cytokines. The authors speculate that a high basal ACE2 expression will increase susceptibility to SARS-CoV infection.

In 3 samples SARS-Cov infection was associated with IL-1, IL-10 and IL-6 cytokine production (GO pathways) at 24h. And later, at 48h, with T-cell activation and T-cell cytokine production. It is unclear whether those changes were statistically significant.

The authors describe a time course quantification of immune infiltrates in epithelial cells infected with SARS-Cov infection. They state that in healthy donors ACE2 expression did not correlate with the immune cell infiltration. However, in SARS-Cov samples, at 48h they found that ACE2 correlated with neutrophils, NK-, Th17-, Th2-, Th1- cells, and DCs. Again, while authors claim significance, the corresponding correlation coefficients and p-values are not presented in the text or figures. In addition, the source of the data for this analysis is not clear.

Using network analysis, proteins SRC, FN1, MAPK3, LYN, MBP, NLRC4, NLRP1 and PRKCD were found to be central (Hub proteins) in the regulating network of cytokine secretion after coronavirus infection. Authors conclude this indicates that these molecules were critically important in ACE2-induced inflammatory response. Additionally, authors speculate that the increased expression of ACE2 affected RPS3 and SRC, which were the two hub genes involved in viral replication and inflammatory response.

12.2.3 Limitations

The methods section is very limited and does not describe any of the statistical analyses; and description of the construction of the regulatory protein networks is also limited. For the findings in Figures 2 authors claim significance, which is not supported by p-values or coefficients. For the sample selection, would be useful if sample sizes and some of the patients’ demographics (e.g. age) were described.

For the analysis of high vs low ACE2 expression in healthy subjects, it is not clear what was the cut off for ‘high’ expression and how it was determined. Additionally, further laboratory studies are warranted to confirm that high ACE2 gene expression would have high correlation with the amount of ACE2 protein on cell surface. For the GO pathway analysis significance was set at p<0.05, but not adjusted for multiple comparisons.

There were no samples with SARS-CoV-2 infection. While SARS-Cov and SARS-CoV-2 both use ACE2 to enter the host cells, the analysis only included data on SARS-Cov and any conclusions about SARS-CoV2 are limited.

Upon checking GSE accession numbers of the datasets references, two might not be cited correctly: GSE37758 (“A spergillus niger: Control (fructose) vs. steam-exploded sugarcane induction (SEB)”" was used in this paper as “lung tissue” data) and GSE14700 (“Steroid Pretreatment of Organ Donors to Prevent Postischemic Renal Allograft Failure: A Randomized, Controlled Trial” – was used as SARS-Cov infection data).

12.2.4 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

Liang et al. medRxiv [900]

12.3.1 Keywords

12.3.2 Main Findings

This study examined the incidence of diarrhea in patients infected with SARS-CoV-2 across three recently published cohorts and found that there are statistically significant differences by Fisher’s exact test. They report that this could be due to subjective diagnosis criterion for diarrhea or from patients first seeking medical care from gastroenterologist. In order to minimize nosocomial infections arising from unsuspected patients with diarrhea and gain comprehensive understanding of transmission routes for this viral pathogen, they compared the transcriptional levels of ACE2 of various human tissues from NCBI public database as well as in small intestine tissue from CD57BL/6 mice using single cell sequencing. They show that ACE2 expression is not only increased in the human small intestine, but demonstrate a particular increase in mice enterocytes positioned on the surface of the intestinal lining exposed to viral pathogens. Given that ACE2 is the viral receptor for SARS-CoV-2 and also reported to regulate diarrhea, their data suggests the small intestine as a potential transmission route and diarrhea as a potentially underestimated symptom in COVID19 patients that must be carefully monitored. Interestingly, however, they show that ACE2 expression level is not elevated in human lung tissue.

12.3.3 Limitations

Although this study demonstrates a statistical difference in the incidence of diarrhea across three separate COVID19 patient cohorts, their conclusions are limited by a small sample size. Specifically, the p-value computed by Fisher’s exact test is based on a single patient cohort of only six cases of which 33% are reported to have diarrhea, while the remaining two larger cohorts with 41 and 99 cases report 3% and 2% diarrhea incidence, respectively. Despite showing significance, they would need to acquire larger sample sizes and cohorts to minimize random variability and draw meaningful conclusions. Furthermore, they do not address why ACE2 expression level is not elevated in human lung tissue despite it being a major established route of transmission for SARS-CoV-2. It could be helpful to validate this result by looking at ACE2 expression in mouse lung tissue. Finally, although this study is descriptive and shows elevated ACE2 expression in small intestinal epithelial cells, it does not establish a mechanistic link to SARS-CoV-2 infection of the host. Overall, their claim that infected patients exhibiting diarrhea pose an increased risk to hospital staff needs to be further substantiated.

12.3.4 Significance

This study provides a possible transmission route and a potentially underappreciated clinical symptom for SARS-CoV-2 for better clinical management and control of COVID19.

12.3.5 Credit

Summary generated as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.4 Specific ACE2 Expression in Cholangiocytes May Cause Liver Damage After 2019-nCoV Infection

Chai et al. bioRxiv [901]

12.4.1 Keywords

12.4.2 Summary

Using both publicly available scRNA-seq dataset of liver samples from colorectal patients and scRNA-sequencing of four liver samples from healthy volunteers, the authors show that ACE2 is significantly enriched in the majority of cholangiocytes (59.7 %) but not in hepatocytes (2.6%).

12.4.3 Main Findings

Using bioinformatics approaches of RNASeq analysis, this study reveals that ACE2 dominates in cholangiocytes and is present at very low levels in hepatocytes.

12.4.4 Limitations

The study does not provide mechanistic insights into how SARS-CoV-2 can infect and replicate in cholangiocytes and the types of intrinsic anti-viral responses induced by cholangiocytes when infected. In addition, because the study relies on the assumption that SARS-CoV-2 infects cells only through ACE2, it cannot discount the possibility that the virus can infect hepatocytes through mechanisms other than ACE2-mediated entry. Furthermore, because the scRNA-seq analysis were performed on healthy liver samples, one cannot draw any definitive conclusions about gene expression states (including ACE2 expression in liver cell types) in system-wide inflammatory contexts.

12.4.5 Significance

This article with other studies on liver damage in COVID patients suggests that liver damage observed in COVID patients is more due to inflammatory cytokines than direct infection of the liver. Even if cholangiocytes are infectable by SARS-CoV-2 (which was demonstrated by human liver ductal organoid study ([902]), published clinical data show no significant increase in bile duct injury related indexes (i.e. alkaline phosphatase, gamma-glutamyl transpeptidase and total bilirubin). In sum, it underscores the importance of future studies characterizing cellular responses of extra-pulmonary organs in the context of COVID or at least in viral lung infections..

12.4.6 Credit

Summary generated by Chang Moon as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.5 ACE2 expression by colonic epithelial cells is associated with viral infection, immunity and energy metabolism

Wang et al. medRxiv. [903]

12.5.1 Keywords

12.5.2 Main Findings

Colonic enterocytes primarily express ACE2. Cellular pathways associated with ACE2 expression include innate immune signaling, HLA up regulation, energy metabolism and apoptotic signaling.

12.5.3 Limitations

This is a study of colonic biopsies taken from 17 children with and without IBD and analyzed using scRNAseq to look at ACE2 expression and identify gene families correlated with ACE2 expression. The authors find ACE2 expression to be primarily in colonocytes. It is not clear why both healthy and IBD patients were combined for the analysis. Biopsies were all of children so extrapolation to adults is limited. The majority of genes found to be negatively correlated with ACE2 expression include immunoglobulin genes (IGs). IG expression will almost certainly be low in colonocytes irrespective of ACE2 expression.

12.5.4 Significance

This study performs a retrospective analysis of ACE2 expression using an RNAseq dataset from intestinal biopsies of children with and without IBD. The implications for the CoV-19 epidemic are modest, but do provide support that ACE2 expression is specific to colonocytes in the intestines. The ontological pathway analysis provides some limited insights into gene expression associated with ACE2.

12.5.5 Credit

Summary generated as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.6 The Pathogenicity of 2019 Novel Coronavirus in hACE2 Transgenic Mice

Bao et al. bioRxiv [904]

12.6.1 Keywords

12.6.2 Main Findings

Using a transgenic human Angiotensin-converting enzyme 2 (hACE2) mouse that has previously been shown susceptible to infection by SARS-CoV, Bao et al. create a model of pandemic 2019-nCoV strain coronavirus. The model includes interstitial hyperplasia in lung tissue, moderate inflammation in bronchioles and blood vessels, and histology consistent with viral pneumonia at 3 days post infection. Wildtype did not experience these symptoms. In addition, viral antigen and hACE2 receptor were found to co-localize the lung by immunofluorescence 3-10 days post infection only in the hACE2 infected mice.

12.6.3 Limitations

The characterization of the infection remains incomplete, as well as lacking characterization of the immune response other than the presence of a single antiviral antibody. Though they claim to fulfill Koch’s postulates, they only isolate the virus and re-infect Vero cells, rather than naive mice.

12.6.4 Significance

This paper establishes a murine model for 2019-nCoV infection with symptoms consistent with viral pneumonia. Though not fully characterized, this model allows in vivo analysis of viral entry and pathology that is important for the development of vaccines and antiviral therapeutics.

12.6.5 Credit

Review by Dan Fu Ruan, Evan Cody and Venu Pothula as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.7 Caution on Kidney Dysfunctions of 2019-nCoV Patients

Li et al. medRxiv. [905]

12.7.1 Keywords

CoVID-19, 2019-nCoV, SARS-CoV-2, kidney, clinical, creatinine, proteinuria, albuminuria, CT

12.7.2 Main Findings

12.7.3 Limitations

12.7.4 Significance

12.7.5 Credit

Review by Samarth Hegde as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.8 Profiling the immune vulnerability landscape of the 2019 Novel Coronavirus

Zhu et al. bioRxiv [908]

12.8.1 Keywords

12.8.2 Main Findings

This study harnesses bioinformatic profiling to predict the potential of COV2 viral proteins to be presented on MHC I and II and to form linear B-cell epitopes. These estimates suggest a T-cell antigenic profile distinct from SARS-CoV or MERS-CoV, identify focused regions of the virus with a high density of predicted epitopes, and provide preliminary evidence for adaptive immune pressure in the genetic evolution of the virus.

12.8.3 Limitations

While the study performs a comprehensive analysis of potential epitopes within the virus genome, the analysis relies solely on bioinformatic prediction to examine MHC binding affinity and B-cell epitope potential and does not capture the immunogenicity or recognition of these epitopes. Future experimental validation in data from patients infected with SARS-CoV-2 will be important to validate and refine these findings. Thus some of the potential conclusions stated, including viral evolution toward lower immunogenicity or a dominant role for CD4+ T-cells rather than CD8+ T-cells in viral clearance, require further valiadtion.

12.8.4 Significance

These findings may help direct peptide vaccine design toward relevant epitopes and provide intriguing evidence of viral evolution in response to immune pressure.

12.8.5 Credit

Summary generated as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.9 Single-cell Analysis of ACE2 Expression in Human Kidneys and Bladders Reveals a Potential Route of 2019-nCoV Infection

Lin et al. bioRxiv [907]

12.9.1 Keywords

12.9.2 Main Findings

12.9.3 Limitations

12.9.4 Significance

12.9.5 Credit

Review by Samarth Hegde as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.10 Neutrophil-to-Lymphocyte Ratio Predicts Severe Illness Patients with 2019 Novel Coronavirus in the Early Stage

Liu et al. medRxiv [911]

12.10.1 Keywords

12.10.2 Main Findings

This study aimed to find prognostic biomarkers of COVID-19 pneumonia severity. Sixty-one (61) patients with COVID-19 treated in January at a hospital in Beijing, China were included. On average, patients were seen within 5 days from illness onset. Samples were collected on admission; and then patients were monitored for the development of severe illness with a median follow-up of 10 days].

Patients were grouped as “mild” (N=44) or “moderate/severe” (N=17) according to symptoms on admission and compared for different clinical/laboratory features. “Moderate/severe” patients were significantly older (median of 56 years old, compared to 41 years old). Whereas comorbidies rates were largely similar between the groups, except for hypertension, which was more frequent in the severe group (p= 0.056). ‘Severe’ patients had higher counts of neutrophils, and serum glucose levels; but lower lymphocyte counts, sodium and serum chlorine levels. The ratio of neutrophils to lymphocytes (NLR) was also higher for the ‘severe’ group. ‘Severe’ patients had a higher rate of bacterial infections (and antibiotic treatment) and received more intensive respiratory support and treatment.

26 clinical/laboratory variables were used to select NLR and age as the best predictors of the severe disease. Predictive cutoffs for a severe illness as NLR ≥ 3.13 or age ≥ 50 years.

12.10.3 Limitations

Identification of early biomarkers is important for making clinical decisions, but large sample size and validation cohorts are necessary to confirm findings. It is worth noting that patients classified as “mild” showed pneumonia by imaging and fever, and in accordance with current classifications this would be consistent with “moderate” cases. Hence it would be more appropriate to refer to the groups as “moderate” vs “severe/critical”. Furthermore, there are several limitations that could impact the interpretation of the results: e.g. classification of patients was based on symptoms presented on admission and not based on disease progression, small sample size, especially the number of ‘severe’ cases (with no deaths among these patients). Given the small sample size, the proposed NLR and age cut offs might not hold for a slightly different set of patients. For example, in a study of >400 patients, ‘non-severe’ and ‘severe’ NLR were 3.2 and 5.5, respectively [912].

12.10.4 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.11 Characteristics of lymphocyte subsets and cytokines in peripheral blood of 123 hospitalized patients with 2019 novel coronavirus pneumonia (NCP)

Wan et. al. medRxiv [913]

12.11.1 Keywords

12.11.2 Main Findings

The authors analyzed lymphocyte subsets and cytokines of 102 patients with mild disease and 21 with severe disease. CD8+T cells and CD4+T cells were significantly reduced in both cohort. particularly in severe patients. The cytokines IL6 and IL10 were significantly elevated in severe patients as compared to mild. No significant differences were observed in frequency of B cells and NK cells.

The authors argue that the measurement of T cell frequencies and cytokine levels of IL6 and IL10 can be used to predict progression of disease from Mild to severe Cov-2 infection.

12.11.3 Limitations

The study demonstrates in a limited cohort similar associations to several other reported studies. The authors didn’t compare the changes in lymphocyte and cytokine with healthy individual (Covid-19 Negative) rather used an internal standard value. The recently preprint in LANCET shows The degree of lymphopenia and a pro-inflammatory cytokine storm is higher in severe COVID-19 patients than in mild cases, and is associated with the disease severity [914].

12.11.4 Significance

This translational data identifies key cytokines and lymphopenia associated with disease severity although mechanism and key cellular players are still unknown. Higher level IL-6 production in severe patient suggests potential role of Tocilizumab (anti-IL6R) biologic although clinical trial will be necessary.

12.11.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.12 Epidemiological and Clinical Characteristics of 17 Hospitalized Patients with 2019 Novel Coronavirus Infections Outside Wuhan, China

Li et al. medRxiv [915]

12.12.1 Keywords

12.12.2 Major Findings

These authors looked at 17 hospitalized patients with COVID-19 confirmed by RT-PCR in Dazhou, Sichuan. Patients were admitted between January 22 and February 10 and the final data were collected on February 11. Of the 17 patients, 12 remained hospitalized while 5 were discharged after meeting national standards. The authors observed no differences based on the sex of the patients but found that the discharged patients were younger in age (p = 0.026) and had higher lymphocyte counts (p = 0.005) and monocyte counts (p = 0.019) upon admission.

12.12.3 Limitations

This study is limited in the sample size of the study and the last data collection point was only one day after some of the patients were admitted.

12.12.4 Significance

These findings have been somewhat supported by subsequent studies that show that older age and an immunocompromised state are more likely to result in a more severe clinical course with COVID-19. However, other studies have been published that report on larger numbers of cases.

12.12.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.13 ACE2 Expression in Kidney and Testis May Cause Kidney and Testis Damage After 2019-nCoV Infection

[916]

12.13.1 Keywords

12.13.2 Main Findings

12.13.3 Limitations

12.13.4 Significance

12.13.5 Credit

Review by Samarth Hegde as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.14 Aberrant pathogenic GM-CSF+ T cells and inflammatory CD14+CD16+ monocytes in severe pulmonary syndrome patients of a new coronavirus

[917]

12.14.1 Keywords

12.14.2 Main Findings

The authors of this study sought to characterize the immune mechanism causing severe pulmonary disease and mortality in 2019-nCoV (COVID-19) patients. Peripheral blood was collected from hospitalized ICU (n=12) and non-ICU (n=21) patients with confirmed 2019-nCoV and from healthy controls (n=10) in The First Affiliated Hospital of University of Science and Technology China (Hefei, Anhui). Immune analysis was conducted by flow cytometry. 2019-nCoV patients had decreased lymphocyte, monocyte, and CD4 T cell counts compared to healthy controls. ICU patients had fewer lymphocytes than non-ICU patients. CD4 T cells of 2019-nCoV patients expressed higher levels of activation markers (OX40, CD69, CD38, CD44) and exhaustion markers (PD-1 and Tim3) than those of healthy controls. CD4 cells of ICU patients expressed significantly higher levels of OX40, PD-1, and Tim3 than those of non-ICU patients. 2019-nCoV patients had higher percentages of CD4 T cells co-expressing GM-CSF and IL-6 compared to healthy controls, while ICU patients had a markedly higher percentage of GM-CSF+ IFN-γ+ CD4 T cells than non-ICU patients. The CD4 T cells of nCoV patients and healthy controls showed no differences in TNF-α secretion.

The CD8 T cells of 2019-nCoV patients also showed higher expression of activation markers CD69, CD38, and CD44, as well as exhaustion markers PD-1 and Tim3, compared to healthy controls. CD8 T cells of ICU patients expressed higher levels of GM-CSF than those of non-ICU patients and healthy controls. No IL-6 or TNF-α was found in the CD8 T cells of any group. There were no differences in numbers of NK cells or B cells in 2019-nCoV patients and healthy controls, nor was there any GM-CSF or IL-6 secretion from these cells in either group.

Percentages of CD14+ CD16+ GM-CSF+ and CD14+ CD16+ IL-6+ inflammatory monocytes were significantly increased in nCoV patients compared to healthy controls; in particular, patients in the ICU had greater percentages of CD14+ CD16+ IL-6+ monocytes than non-ICU patients. The authors suggest that in 2019-nCoV patients, pathogenic Th1 cells produce GM-CSF, recruiting CD14+ CD16+ inflammatory monocytes that secrete high levels of IL-6. These may enter pulmonary circulation and damage lung tissue while initiating the cytokine storm that causes mortality in severe cases. This is consistent with the cytokine storm seen in similar coronaviruses, as IL-6, IFN-γ, and GM-CSF are key inflammatory mediators seen in patients with SARS-CoV-1 and MERS-CoV.

12.14.3 Limitations

Though the results of this study open questions for further investigation, this is an early study on a small cohort of patients, and as such there are a number of limitations. The study included only 12 ICU patients and 21 non-ICU patients, and ideally would be repeated with a much larger patient cohort. Though the authors make claims about differences in lymphocyte and monocyte counts between patients and healthy controls, they did not report baseline laboratory findings for the control group. Additionally, severity of disease was classified based on whether or not patients were in the ICU. It would be interesting to contextualize the authors’ immunological findings with more specific metrics of disease severity or time course. Noting mortality, time from disease onset, pre-existing conditions, or severity of lung pathology in post-mortem tissue samples would paint a fuller picture of how to assess risk level and the relationship between severity of disease and immunopathology. Another limitation is the selection of cytokines and immune markers for analysis, as the selection criteria were based on the cell subsets and cytokine storm typically seen in SARS-CoV-1 and MERS-CoV patients. Unbiased cytokine screens and immune profiling may reveal novel therapeutic targets that were not included in this study.

12.14.4 Significance

This study identifies potential therapeutic targets that could prevent acute respiratory disease syndrome (ARDS) and mortality in patients most severely affected by COVID-19. The authors propose testing monoclonal antibodies against IL6-R or GM-CSF to block recruitment of inflammatory monocytes and the subsequent cytokine storm in these patients.

12.14.5 Credit

Review by Gabrielle Lubitz as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.15 Clinical Characteristics of 2019 Novel Infected Coronavirus Pneumonia:A Systemic Review and Meta-analysis

Qian et al. medRxiv. [918]

12.15.1 Keywords

12.15.2 Main Findings

The authors performed a meta analysis of literature on clinical, laboratory and radiologic characteristics of patients presenting with pneumonia related to SARSCoV2 infection, published up to Feb 6 2020. They found that symptoms that were mostly consistent among studies were sore throat, headache, diarrhea and rhinorrhea. Fever, cough, malaise and muscle pain were highly variable across studies. Leukopenia (mostly lymphocytopenia) and increased white blood cells were highly variable across studies. They identified three most common patterns seen on CT scan, but there was high variability across studies. Consistently across the studies examined, the authors found that about 75% of patients need supplemental oxygen therapy, about 23% mechanical ventilation and about 5% extracorporeal membrane oxygenation (ECMO). The authors calculated a staggering pooled mortality incidence of 78% for these patients.

12.15.3 Limitations

The authors mention that the total number of studies included in this meta analysis is nine, however they also mentioned that only three studies reported individual patient data. It is overall unclear how many patients in total were included in their analysis. This is mostly relevant as they reported an incredibly high mortality (78%) and mention an absolute number of deaths of 26 cases overall. It is not clear from their report how the mortality rate was calculated.

The data is based on reports from China and mostly from the Wuhan area, which somewhat limits the overall generalizability and applicability of these results.

12.15.4 Significance

This meta analysis offers some important data for clinicians to refer to when dealing with patients with COVID-19 and specifically with pneumonia. It is very helpful to set expectations about the course of the disease.

12.15.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.16 Longitudinal characteristics of lymphocyte responses and cytokine profiles in the peripheral blood of SARS-CoV-2 infected patients

Liu et al. medRxiv [919]

12.16.1 Keywords

12.16.2 Main Findings

Liu et al. enrolled a cohort of 40 patients from Wuhan including 27 mild cases and 13 severe cases of COVID-19. They performed a 16-day kinetic analysis of peripheral blood from time of disease onset. Patients in the severe group were older (medium age of 59.7, compared to 48.7 in mild group) and more likely to have hypertension as a co-morbidity. Lymphopenia was observed in 44.4% of the mild patients and 84.6% of the severe patients. Lymphopenia was due to low T cell count, specially CD8 T cells. Severe patients showed higher neutrophil counts and an increase of cytokines in the serum (IL2, IL6, IL10 and IFNγ). The authors measured several other clinical laboratory parameters were also higher in severe cases compared to mild, but concluded that neutrophil to CD8 T cell ratio (N8R) as the best prognostic factor to identify the severe cases compared to other receiver operating characteristic (ROC).

12.16.3 Limitations

This was a small cohort (N=40), and two of the patients initially included in the severe group (N=13) passed away and were excluded from the analysis due to lack of longitudinal data. However, it would be most important to be able to identify patients with severe disease with higher odds of dying. It seems that the different time points analyzed relate to hospital admission, which the authors describe as disease onset. The time between first symptoms and first data points is not described. It would have been important to analyze how the different measured parameters change according to health condition, and not just time (but that would require a larger cohort). The predictive value of N8R compared to the more commonly used NLR needs to be assessed in other independent and larger cohorts. Lastly, it is important to note that pneumonia was detected in patients included in the “mild” group, but according to the Chinese Clinical Guidance for COVID-19 Pneumonia Diagnosis and Treatment (7th edition) this group should be considered “moderate”.

12.16.4 Significance

Lymphopenia and cytokine storm have been described to be detrimental in many other infections including SARS-CoV1 and MERS-CoV. However, it was necessary to confirm that this dramatic immune response was also observed in the SARS-CoV2 infected patients. These results and further validation of the N8R ratio as a predictor of disease severity will contribute for the management of COVID19 patients and potential development of therapies.

12.16.5 Credit

Review by Pauline Hamon as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.17 Clinical and immunologic features in severe and moderate forms of Coronavirus Disease 2019

Chen et al. medRxiv [920]

12.17.1 Keywords

12.17.2 Main Findings

This study retrospectively evaluated clinical, laboratory, hematological, biochemical and immunologic data from 21 subjects admitted to the hospital in Wuhan, China (late December/January) with confirmed SARS-CoV-2 infection. The aim of the study was to compare ‘severe’ (n=11, ~64 years old) and ‘moderate’ (n=10, ~51 years old) COVID-19 cases. Disease severity was defined by patients’ blood oxygen level and respiratory output. They were classified as ‘severe’ if SpO2 93% or respiratory rates 30 per min.

In terms of the clinical laboratory measures, ‘severe’ patients had higher CRP and ferritin, alanine and aspartate aminotransferases, and lactate dehydrogenase but lower albumin concentrations.

The authors then compared plasma cytokine levels (ELISA) and immune cell populations (PBMCs, Flow Cytometry). ‘Severe’ cases had higher levels of IL-2R, IL-10, TNFa, and IL-6 (marginally significant). For the immune cell counts, ‘severe’ group had higher neutrophils, HLA-DR+ CD8 T cells and total B cells; and lower total lymphocytes, CD4 and CD8 T cells (except for HLA-DR+), CD45RA Tregs, and IFNy-expressing CD4 T cells. No significant differences were observed for IL-8, counts of NK cells, CD45+RO Tregs, IFNy-expressing CD8 T and NK cells.

12.17.3 Limitations

Several potential limitations should be noted: 1) Blood samples were collected 2 days post hospital admission and no data on viral loads were available; 2) Most patients were administered medications (e.g. corticosteroids), which could have affected lymphocyte counts. Medications are briefly mentioned in the text of the manuscript; authors should include medications as part of Table 1. 3) ‘Severe’ cases were significantly older and 4/11 ‘severe’ patients died within 20 days. Authors should consider a sensitivity analysis of biomarkers with the adjustment for patients’ age.

12.17.4 Significance

Although the sample size was small, this paper presented a broad range of clinical, biochemical, and immunologic data on patients with COVID-19. One of the main findings is that SARS-CoV-2 may affect T lymphocytes, primarily CD4+ T cells, resulting in decreased IFNy production. Potentially, diminished T lymphocytes and elevated cytokines can serve as biomarkers of severity of COVID-19.

12.17.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.18 SARS-CoV-2 and SARS-CoV Spike-RBD Structure and Receptor Binding Comparison and Potential Implications on Neutralizing Antibody and Vaccine Development

Sun et al. bioRxiv [921]

12.18.1 Keywords

12.18.2 Main Findings

This study compared the structure of SARS-CoV and SARS-CoV-2 Spike (S) protein receptor binding domain (RBD) and interactions with ACE2 using computational modeling, and interrogated cross-reactivity and cross-neutralization of SARS-CoV-2 by antibodies against SARS-CoV. While SARS-CoV and SARS-CoV-2 have over 70 % sequence homology and share the same human receptor ACE2, the receptor binding motif (RBM) is only 50% homologous.

Computational prediction of the SARS-CoV-2 and ACE2 interactions based on the previous crystal structure data of SARS-CoV, and measurement of binding affinities against human ACE2 using recombinant SARS-CoV and SARS-CoV-2 S1 peptides, demonstrated similar binding of the two S1 peptides to ACE2, explaining the similar transmissibility of SARS-CoV and SARS-CoV-2 and consistent with previous data (Wall et al Cell 2020).

The neutralization activity of SARS-CoV-specific rabbit polyclonal antibodies were about two-order of magnitude less efficient to neutralize SARS-CoV-2 than SARS-CoV, and four potently neutralizing monoclonal antibodies against SARS-CoV had poor binding and neutralizing activity against SARS-CoV-2. In contrast, 3 poor SARS-CoV-binding monoclonal antibodies show some efficiency to bind and neutralize SARS-CoV-2. The results suggest that that antibodies to more conserved regions outside the RBM motif might possess better cross-protective neutralizing activities between two strains.

12.18.3 Limitations

It would have been helpful to show the epitopes recognized by the monoclonal antibodies tested on both SARS-CoV, SARS-CoV-2 to be able to make predictions for induction of broadly neutralizing antibodies. The data on monoclonal antibody competition with ACE2 for binding to SARS-CoV RBD should have also included binding on SARS-CoV2, especially for the three monoclonal antibodies that showed neutralization activity for SARS-CoV2. Because of the less homology in RBM sequences between viruses, it still may be possible that these antibodies would recognize the ACE2 RBD in SARS-CoV-2.

12.18.4 Significance

It is noteworthy that immunization to mice and rabbit with SARS-CoV S1 or RBD protein could induce monoclonal antibodies to cross-bind and cross-neutralize SARS-CoV-2 even if they are not ACE2-blocking. If these types of antibodies could be found in human survivors or in the asymptomatic populations as well, it might suggest that exposure to previous Coronavirus strains could have induced cross-neutralizing antibodies and resulted in the protection from severe symptoms in some cases of SARS-CoV2.

12.18.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.19 Protection of Rhesus Macaque from SARS-Coronavirus challenge by recombinant adenovirus vaccine

Chen et al. bioRxiv [922]

12.19.1 Keywords

12.19.2 Main Findings

Rhesus macaques were immunized intramuscularly twice (week 0 and week 4) with SV8000 carrying the information to express a S1-orf8 fusion protein and the N protein from the BJ01 strain of SARS-CoV-1. By week 8, immunized animals had signs of immunological protection (IgG and neutralization titers) against SARS-CoV-1 and were protected against challenge with the PUMC-1 strain, with fewer detectable symptoms of respiratory distress, lower viral load, shorter periods of viral persistence, and less pathology in the lungs compared to non-immunized animals.

12.19.3 Limitations

The authors should write clearer descriptions of the methods used in this article. They do not describe how the IgG titers or neutralization titers were determined. There are some issues with the presentation of data, for example, in Figure 1a, y-axis should not be Vmax; forming cells and 1d would benefit from showing error bars. Furthermore, although I inferred that the animals were challenged at week 8, the authors did not explicitly detail when the animals were challenged. The authors should explain the design of their vaccine, including the choice of antigens and vector. The authors also do not include a description of the ethical use of animals in their study.

12.19.4 Significance

The authors describe a vaccine for SARS-CoV-1 with no discussion of possible implications for the current SARS-CoV-2 pandemic. Could a similar vaccine be designed to protect against SARS-CoV-2 and would the concerns regarding emerging viral mutations that the authors describe as a limitation for SARS-CoV-1 also be true in the context of SARS-CoV-2?

12.19.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.20 Reduction and Functional Exhaustion of T cells in Patients with Coronavirus Disease 2019 (COVID-19)

[923]

12.20.1 Keywords

12.20.2 Main Findings

Based on a retrospective study of 522 COVID patients and 40 healthy controls from two hospitals in Wuhan, China, authors show both age-dependent and clinical severity-dependent decrease in T cell numbers with elderly patients and patients who are in ICU-care showing the most dramatic decrease in T cell counts. Cytokine profiling of COVID patients reveal that TNF-α, IL-6 and IL-10 are increased in infected patients with patients in the ICU showing the highest levels. Interestingly, these three cytokine levels were inversely correlated with T cell counts and such inverse relationship was preserved throughout the disease progression. Surface staining of exhaustion markers (PD-1 and Tim-3) and flow cytometry of stained peripheral blood of 14 patients and 3 healthy volunteers demonstrate that T cells of COVID patients have increased expression of PD-1 with patients in ICU having the highest number of CD8+PD-1+ cells than their counterparts in non-ICU groups.

12.20.3 Limitations

Compared to the number of patients, number of control (n= 40) is small and is not controlled for age. Additional data linking inflammatory cytokines and the quality of the adaptive response including humoral and antigen specific T cell response is much needed. T cell exhaustion study relies on marker-dependent labeling of T cell functionality of a very limited sample size (n=17)—a functional/mechanistic study of these T cells from PBMCs would have bolstered their claims.

12.20.4 Significance

Limited but contains interesting implications. It is already known in literature that in the context of acute respiratory viral infections CD8 T cells exhibit exhaustion-like phenotypes which further underscores the importance of mechanistic studies that can elucidate how COVID infection leads to lymphopenia and T cell exhaustion-like phenotype.

However, as authors have noted, the data does point to an interesting question: How these inflammatory cytokines (TNF-α, IL-6 and IL-10) correlate with or affect effective viral immunity and what types of cells produce these cytokines? Answering that question will help us refine our targets for immune-modulatory therapies especially in patients suffering from cytokine storms.

12.20.5 Credit

This review by Chang Moon was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.21 Clinical Characteristics of 25 death cases infected with COVID-19 pneumonia: a retrospective review of medical records in a single medical center, Wuhan, China

[924]

12.21.1 Keywords

12.21.2 Main Findings

Most common chronic conditions among 25 patients that died from COVID-19 related respiratory failure were hypertension (64%) and diabetes (40%). Disease progression was marked by progressive organ failure, starting first with lung dysfunction, then heart (e.g. increased cTnI and pro-BNP), followed by kidney (e.g. increased BUN, Cr), and liver (e.g. ALT, AST). 72% of patients had neutrophilia and 88% also had lymphopenia. General markers of inflammation were also increased (e.g. PCT, D-Dimer, CRP, LDH, and SAA).

12.21.3 Limitations

The limitations of this study include small sample size and lack of measurements for some tests for several patients. This study would also have been stronger with comparison of the same measurements to patients suffering from less severe disease to further validate and correlate proposed biomarkers with disease severity.

12.21.4 Significance

This study identifies chronic conditions (i.e. hypertension and diabetes) that strongly correlates with disease severity. In addition to general markers of inflammation, the authors also identify concomitant neutrophilia and lymphopenia among their cohort of patients. This is a potentially interesting immunological finding because we would typically expect increased lymphocytes during a viral infection. Neutrophilia may also be contributing to cytokine storm. In addition, PCT was elevated in 90.5% of patients, suggesting a role for sepsis or secondary bacterial infection in COVID-19 related respiratory failure.

12.21.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.22 SARS-CoV-2 infection does not significantly cause acute renal injury: an analysis of 116 hospitalized patients with COVID-19 in a single hospital, Wuhan, China

[925]

12.22.1 Keywords

12.22.2 Main Findings

12.22.3 Limitations

12.22.4 Significance

12.22.5 Credit

Review by Samarth Hegde as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.23 Potential T-cell and B-Cell Epitopes of 2019-nCoV

[928]

12.23.1 Keywords

12.23.2 Main Findings

The authors use 2 neural network algorithms, NetMHCpan4 and MARIA, to identify regions within the COVID-19 genome that are presentable by HLA. They identify 405 viral epitopes that are presentable on MHC-I and MHC-II and validate using known epitopes from SARS-CoV. To determine whether immune surveillance drives viral mutations to evade MHC presentation, the authors analyzed 68 viral genomes from 4 continents. They identified 93 point mutations that occurred preferentially in regions predicted to be presented by MHC-I (p=0.02) suggesting viral evolution to evade CD8 T-cell mediated killing. 2 nonsense mutations were also identified that resulted in loss of presentation of an associated antigen (FGDSVEEVL) predicted to be good antigen for presentation across multiple HLA alleles.

To identify potential sites of neutralizing antibody binding, the authors used homology modeling to the SARS-CoV’s spike protein (S protein) to determine the putative structure of the CoV2 spike protein. They used Discotope2 to identify antibody binding sites on the protein surface in both the down and up conformations of the S protein. The authors validate this approach by first identifying antibody binding site in SARS-CoV S protein. In both the down and up conformation of the CoV2 S protein, the authors identified a potential antibody binding site on the S protein receptor binding domain (RBD) of the ACE2 receptor (residues 440-460, 494-506). While RBDs in both SARS-CoV and CoV2 spike proteins may be important for antibody binding, the authors note that SARS-CoV has larger attack surfaces than CoV2. These results were later validated on published crystal structures of the CoV2 S protein RBD and human ACE2. Furthermore, analysis of 68 viral genomes did not identify any mutations in this potential antibody binding site in CoV2.

Finally, the authors compile a list of potential peptide vaccine candidates across the viral genome that can be presented by multiple HLA alleles. Several of the peptides showed homology to SARS-CoV T-cell and B-cell epitopes.

12.23.3 Limitations

While the authors used computational methods of validation, primarily through multiple comparisons to published SARS-CoV structures and epitopes, future work should include experimental validation of putative T-cell and B-cell epitopes.

12.23.4 Significance

The authors identified potential T-cell and B-cell epitopes that may be good candidates for peptide based vaccines against CoV2. They also made interesting observations in comparing SARS-CoV and CoV2 potential antibody binding sites, noting that SARS-CoV had larger attack surfaces for potential neutralizing antibody binding. One of the highlights of this paper was the authors’ mutation analysis of 68 viral genomes from 4 continents. This analysis not only validated their computational method for identifying T-cell epitopes, but showed that immune surveillance likely drives viral mutation in MHC-I binding peptides. The smaller attack surface may point to potential mechanisms of immune evasion by CoV2. However, absence of mutations in the RBD of CoV2 and the small number of mutations in peptides presentable to T cells suggests that vaccines against multiple epitopes could still elicit robust immunity against CoV2.

12.23.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.24 Structure, Function, and Antigenicity of the SARSCoV-2 Spike Glycoprotein

Walls et al. bioRxiv. [929] now [63]

12.24.1 Keywords

12.24.2 Main Findings

The authors highlight a human angiotensin-converting enzyme 2 (hACE2), as a potential receptor used by the current Severe Acute respiratory syndrome coronavirus-2 (SARS-CoV-2) as a host factor that allows the virus target human cells. This virus-host interaction facilitates the infection of human cells with a high affinity comparable with SARS-CoV. The authors propose this mechanism as a probable explanation of the efficient transmission of SARS-CoV-2 between humans. Besides, Walls and colleagues described SARS-CoV-2 S glycoprotein S by Cryo-EM along with neutralizing polyclonal response against SAR-CoV-2 S from mice immunized with SAR-CoV and blocking SAR-CoV-2 S-mediated entry into VeroE6 infected cells.**

12.24.3 Limitations

The SARS-CoV-2 depends on the cell factors ACE2 and TMPRSS2, this last, according to a recent manuscript by Markus Hoffman et al., Cell, 2020. The authors used green monkey (VeroE6) and hamster (BHK) cell lines in the experiments to drive its conclusions to humans; however, it is well known the caucasian colon adenocarcinoma human cell line (CaCo-2), highly express the hACE2 receptor as the TMPRSS2 protease as well. In humans, ACE2 protein is highly expressed in the gastrointestinal tract, which again, makes the CaCo-2 cell line suitable for the following SARS-CoV-2 studies.

12.24.4 Significance

The results propose a functional receptor used by SARS-CoV-2 to infect humans worldwide and defining two distinct conformations of spike (S) glycoprotein by cryogenic electron microscopy (Cryo-EM). This study might help establish a precedent for initial drug design and treatment of the current global human coronavirus epidemic.

12.24.5 Credit

Review by postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.25 Breadth of concomitant immune responses underpinning viral clearance and patient recovery in a non-severe case of COVID-19

Thevarajan et al. medRxiv [930]

12.25.1 Keywords

12.25.2 Main Findings

The authors characterized the immune response in peripheral blood of a 47-year old COVID-19 patient.

SARS-CoV2 was detected in nasopharyngeal swab, sputum and faeces samples, but not in urine, rectal swab, whole blood or throat swab. 7 days after symptom onset, the nasopharyngeal swab test turned negative, at day 10 the radiography infiltrates were cleared and at day 13 the patient became asymptomatic.

Immunofluorescence staining shows from day 7 the presence of COVID-19-binding IgG and IgM antibodies in plasma, that increase until day 20.

Flow cytometry on whole blood reveals a plasmablast peak at day 8, a gradual increase in T follicular helper cells, stable HLA-DR+ NK frequencies and decreased monocyte frequencies compared to healthy counterparts. The expression of CD38 and HLA-DR peaked on T cells at D9 and was associated with higher production of cytotoxic mediators by CD8+ T cells.

IL-6 and IL-8 were undetectable in plasma.

The authors further highlight the presence of the IFITM3 SNP-rs12252-C/C variant in this patient, which is associated with higher susceptibility to influenza virus.

12.25.3 Limitations

These results need to be confirmed in additional patients.

COVID-19 patients have increased infiltration of macrophages in their lungs [931]. Monitoring monocyte proportions in blood earlier in the disease might help to evaluate their eventual migration to the lungs.

The stable concentration of HLA-DR+ NK cells in blood from day 7 is not sufficient to rule out NK cell activation upon SARS-CoV2 infection. In response to influenza A virus, NK cells express higher levels of activation markers CD69 and CD38, proliferate better and display higher cytotoxicity [932]. Assessing these parameters in COVID-19 patients is required to better understand NK cell role in clearing this infection.

Neutralization potential of the COVID-19-binding IgG and IgM antibodies should be assessed in future studies.

This patient was able to clear the virus, while presenting a SNP associated with severe outcome following influenza infection. The association between this SNP and outcome upon SARS-CoV2 infection should be further investigated.

12.25.4 Significance

This study is among the first to describe the appearance of COVID-19-binding IgG and IgM antibodies upon infection. The emergence of new serological assays might contribute to monitor more precisely the seroconversion kinetics of COVID-19 patients [279]. Further association studies between IFITM3 SNP-rs12252-C/C variant and clinical data might help to refine the COVID-19 outcome prediction tools.

12.25.5 Credit

Review by Bérengère Salomé as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.26 The landscape of lung bronchoalveolar immune cells in COVID-19 revealed by single-cell RNA sequencing

Liao et al. medRxiv [931]

12.26.1 Keywords

12.26.2 Main Findings

The authors performed single-cell RNA sequencing (scRNAseq) on bronchoalveolar lavage fluid (BAL) from 6 COVID-19 patients (n=3 mild cases, n=3 severe cases). Data was compared to previously generated scRNAseq data from healthy donor lung tissue (n=8).

Clustering analysis of the 6 patients revealed distinct immune cell organization between mild and severe disease. Specifically, they found that transcriptional clusters annotated as tissue resident alveolar macrophages were strongly reduced while monocytes-derived FCN1+SPP1+ inflammatory macrophages dominated the BAL of patients with severe COVID19 diseases. They show that inflammatory macrophages upregulated interferon-signaling genes, monocytes recruiting chemokines including CCL2, CCL3, CCL4 as well as IL-6, TNF, IL-8 and profibrotic cytokine TGF-β, while alveolar macrophages expressed lipid metabolism genes, such as PPARG.

The lymphoid compartment was overall enriched in lungs from patients. Clonally expanded CD8 T cells were enriched in mild cases suggesting that CD8 T cells contribute to viral clearance as in Flu infection, whereas proliferating T cells were enriched in severe cases.

SARS-CoV-2 viral transcripts were detected in severe patients, but considered here as ambient contaminations.

12.26.3 Limitations

These results are based on samples from 6 patients and should therefore be confirmed in the future in additional patients. Longitudinal monitoring of BAL during disease progression or resolution would have been most useful.

The mechanisms underlying the skewing of the macrophage compartment in patients towards inflammatory macrophages should be investigated in future studies.

Deeper characterization of the lymphoid subsets is required. The composition of the “proliferating” cluster and how these cells differ from conventional T cell clusters should be assessed. NK and CD8 T cell transcriptomic profile, in particular the expression of cytotoxic mediator and immune checkpoint transcripts, should be compared between healthy and diseased lesions.

12.26.4 Significance

COVID-19 induces a robust inflammatory cytokine storm in patients that contributes to severe lung tissue damage and ARDS [933]. Accumulation of monocyte-derived inflammatory macrophages at the expense of Alveolar macrophages known to play an anti-inflammatory role following respiratory viral infection, in part through the PPARγ pathway [934,935] are likely contributing to lung tissue injuries. These data suggest that reduction of monocyte accumulation in the lung tissues could help modulate COVID-19-induced inflammation. Further analysis of lymphoid subsets is required to understand the contribution of adaptive immunity to disease outcome.

12.26.5 Credit

Review by Bérengère Salomé and Assaf Magen as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.27 Can routine laboratory tests discriminate 2019 novel coronavirus infected pneumonia from other community-acquired pneumonia?

Pan et al. medRxiv [936]

12.27.1 Keywords

12.27.2 Main Findings

In an attempt to use standard laboratory testing for the discrimination between “Novel Coronavirus Infected Pneumonia” (NCIP) and a usual community acquired pneumonia (CAP), the authors compared laboratory testing results of 84 NCIP patients with those of a historical group of 316 CAP patients from 2018 naturally COVID-19 negative. The authors describe significantly lower white blood- as well as red blood- and platelet counts in NCIP patients. When analyzing differential blood counts, lower absolute counts were measured in all subsets of NCIP patients. With regard to clinical chemistry parameters, they found increased AST and bilirubin in NCIP patients as compared to CAP patients.

12.27.3 Limitations

The authors claim to describe a simple method to rapidly assess a pre-test probability for NCIP. However, the study has substantial weakpoints. The deviation in clinical laboratory values in NCIP patients described here can usually be observed in severely ill patients. The authors do not comment on how severely ill the patients tested here were in comparison to the historical control. Thus, the conclusion that the tests discriminate between CAP and NCIP lacks justification.

12.27.4 Significance

The article strives to compare initial laboratory testing results in patients with COVID-19 pneumonia as compared to patients with a usual community acquired pneumonia. The implications of this study for the current clinical situation seem restricted due to a lack in clinical information and the use of a control group that might not be appropriate.

12.27.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

[937]

12.28.1 Keywords

12.28.2 Main Findings

This study is a cross-sectional analysis of 100 patients with COVID-19 pneumonia, divided into mild (n = 34), severe (n = 34), and critical (n = 32) disease status based on clinical definitions.

The criteria used to define disease severity are as follows:

  1. Severe – any of the following: respiratory distress or respiratory rate ≥ 30 respirations/minute; oxygen saturation ≤ 93% at rest; oxygen partial pressure (PaO2)/oxygen concentration (FiO2) in arterial blood ≤ 300mmHg, progression of disease on imaging to >50% lung involvement in the short term.

  2. Critical – any of the following: respiratory failure that requires mechanical ventilation; shock; other organ failure that requires treatment in the ICU.

  3. Patients with pneumonia who test positive for COVID-19 who do not have the symptoms delineated above are considered mild.

Peripheral blood inflammatory markers were correlated to disease status. Disease severity was significantly associated with levels of IL-2R, IL-6, IL-8, IL-10, TNF-α, CRP, ferroprotein, and procalcitonin. Total WBC count, lymphocyte count, neutrophil count, and eosinophil count were also significantly correlated with disease status. Since this is a retrospective, cross-sectional study of clinical laboratory values, these data may be extrapolated for clinical decision making, but without studies of underlying cellular causes of these changes this study does not contribute to a deeper understanding of SARS-CoV-2 interactions with the immune system.

It is also notable that the mean age of patients in the mild group was significantly different from the mean ages of patients designated as severe or critical (p < 0.001). The mean patient age was not significantly different between the severe and critical groups. However, IL-6, IL-8, procalcitonin (Table 2), CRP, ferroprotein (Figure 3A, 3B), WBC count, and neutrophil count (Figure 4A, 4B) were all significantly elevated in the critical group compared to severe. These data suggest underlying differences in COVID-19 progression that is unrelated to age.

12.28.3 Significance

Given the inflammatory profile outlined in this study, patients who have mild or severe COVID-19 pneumonia, who also have any elevations in the inflammatory biomarkers listed above, should be closely monitored for potential progression to critical status.

12.28.4 Credit

This review by JJF was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.29 An Effective CTL Peptide Vaccine for Ebola Zaire Based on Survivors’ CD8+ Targeting of a Particular Nucleocapsid Protein Epitope with Potential Implications for COVID-19 Vaccine Design

Herst et al. bioRxiv [938]

12.29.1 Keywords

12.29.2 Main Findings

Vaccination of mice with a single dose of a 9-amino-acid peptide NP44-52 located in a conserved region of ebolavirus (EBOV) nucleocapsid protein (NP) confers CD8+ T-cell-mediated immunity against mouse adapted EBOV (maEBOV). Bioinformatic analyses predict multiple conserved CD8+ T cell epitopes in the SARS-CoV-2 NP, suggesting that a similar approach may be feasible for vaccine design against SARS-CoV-2.

The authors focus on a site within a 20-peptide region of EBOV NP which was commonly targeted by CD8+ T cells in a group of EBOV survivors carrying the HLA-A*30:01:01 allele. To justify the testing of specific vaccine epitopes in a mouse challenge setting, the authors cite known examples of human pathogen-derived peptide antigens that are also recognized by C57BL/6 mice, as well as existing data surrounding known mouse immunogenicity of peptides related to this EBOV NP region. Testing 3 distinct 9mer peptides over an 11 amino-acid window and comparing to vaccination with the 11mer with a T-cell reactivity readout demonstrated that optimizing peptide length and position for immunogenicity may be crucial, likely due to suboptimal peptide processing and MHC-class-I loading.

Vaccines for maEBOV challenge studies were constructed by packaging NP44-52 in d,l poly(lactic-co-glycolic) acid microspheres. CpG was also packaged within the microspheres, while Monophosphoryl Lipid A (a TLR4 ligand) was added to the injectate solution. A second peptide consisting of a predicted MHC-II epitope from the EBOV VG19 protein was added using a separate population of microspheres, and the formulation was injected by intraperitoneal administration. The vaccine was protective against a range of maEBOV doses up to at least 10,000 PFU. Survival was anticorrelated with levels of IL6, MCP-1 (CCL2), IL9, and GM-CSF, which recapitulated trends seen in human EBOV infection.

While HLA-A*30:01:01 is only present in a minority of humans, the authors state that MHC binding algorithms predict NP44-52 to be a strong binder of a set of more common HLA-A*02 alleles. The authors predict that a peptide vaccine based on the proposed formulation could elicit responses in up to 50% of people in Sudan or 30% of people in North America.

SARS-CoV-2 NP, meanwhile, has conserved regions which may provide peptide-vaccine candidates. Scanning the SARS-CoV-2 NP sequence for HLA-binding 9mers identified 53 peptides with predicted binding affinity < 500nM, including peptides that are predicted to bind to HLA-class-I alleles of 97% of humans, 7 of which have previously been tested in-vitro.

The results support previously appreciated correlations between certain cytokines and disease severity, specifically IL6 which relates to multiple trial therapies. Prediction of HLA-class-I binding of SARS-CoV-2 NP peptides suggests the plausibility of a peptide vaccine targeting conserved regions of SARS-CoV-2 NP although further validation in previously infected patient samples will be essential.

12.29.3 Credit

Review by Andrew M. Leader as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.30 Epitope-based peptide vaccines predicted against novel coronavirus disease caused by SARS-CoV-2

Li et al. bioRxiv. [939]

12.30.1 Keywords

12.30.2 Main Findings

This study employs a series of bioinformatic pipelines to identify T and B cell epitopes on spike (S) protein of SARS-CoV-2 and assess their properties for vaccine potential. To identify B cell epitopes, they assessed structural accessibility, hydrophilicity, and beta-turn and flexibility which are all factors that promote their targeting by antibodies. To identify T cell epitopes, they filtered for peptides with high antigenicity score and capacity to bind 3 or more MHC alleles. Using the protein digest server, they also demonstrated that their identified T and B cell epitopes are stable, having multiple non-digesting enzymes per epitope. Epitopes were also determined to be non-allergenic and non-toxin as assessed by Allergen FP 1.0 and ToxinPred, respectively. For T cell epitopes, they assessed the strength of epitope-HLA interaction via PepSite. Overall, they predict four B cell and eleven T cell epitopes (two MHC I and nine MHC II binding) to pass stringent computational thresholds as candidates for vaccine development. Furthermore, they performed sequence alignment between all identified SARS-CoV-2 S protein mutations and predicted epitopes, and showed that the epitopes are conserved across 134 isolates from 38 locations worldwide. However, they report that these conserved epitopes may soon become obsolete given the known mutation rate of related SARS-CoV is estimated to be 4x10-4/site/year, underscoring the urgency of anti-viral vaccine development.

12.30.3 Limitations

While spike (S) protein may have a critical role in viral entry into host cells and their epitope prediction criterion were comprehensive, this study did not examine other candidate SARS-CoV-2 proteins. This point is particularly important given that a single epitope may not be sufficient to induce robust immune memory, and recent approaches involve multi-epitope vaccine design. Furthermore, their study only included a direct implementation of various published methods, but did not validate individual bioinformatic tools with controls to demonstrate robustness. Finally, it is critical that these predicted epitopes are experimentally validated before any conclusions can be drawn about their potential as vaccine candidates or their clinical efficacy.

12.30.4 Significance

This study provides a computational framework to rapidly identify epitopes that may serve as potential vaccine candidates for treating SARS-CoV-2.

12.30.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.31 The definition and risks of Cytokine Release Syndrome-Like in 11 COVID-19-Infected Pneumonia critically ill patients: Disease Characteristics and Retrospective Analysis

Wang Jr. et al. medRxiv. [940]

12.31.1 Keywords

12.31.2 Main Findings

This study describes the occurrence of a cytokine release syndrome-like (CRSL) toxicity in ICU patients with COVID-19 pneumonia. The median time from first symptom to acute respiratory distress syndrome (ARDS) was 10 days. All patients had decreased CD3, CD4 and CD8 cells, and a significant increase of serum IL-6. Furthermore, 91% had decreased NK cells. The changes in IL-6 levels preceded those in CD4 and CD8 cell counts. All of these parameters correlated with the area of pulmonary inflammation in CT scan images. Mechanical ventilation increased the numbers of CD4 and CD8 cells, while decreasing the levels of IL-6, and improving the immunological parameters.

12.31.3 Limitations

The number of patients included in this retrospective single center study is small (n=11), and the follow-up period very short (25 days). Eight of the eleven patients were described as having CRSL, and were treated by intubation (7) or ECMO (2). Nine patients were still in the intensive care unit at the time of publication of this article, so their disease outcome is unknown.

12.31.4 Significance

The authors define a cytokine release syndrome-like toxicity in patients with COVID-19 with clinical radiological and immunological criteria: 1) decrease of circulating CD4, CD8 and NK cells; 2) substantial increase of IL-6 in peripheral blood; 3) continuous fever; 4) organ and tissue damage. This event seems to occur very often in critically ill patients with COVID-19 pneumonia. Interestingly, the increase of IL-6 in the peripheral blood preceded other laboratory alterations, thus, IL-6 might be an early biomarker for the severity of COVID-19 pneumonia. The manuscript will require considerable editing for organization and clarity.

12.31.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.32 Clinical characteristics of 36 non-survivors with COVID-19 in Wuhan, China

Huang et al. medRxiv. [941]

12.32.1 Keywords

12.32.2 Main Findings

This is a simple study reporting clinical characteristics of patients who did not survive COVID-19. All patients (mean age=69.22 years) had acute respiratory distress syndrome (ARDS) and their median time from onset to ARDS was 11 days. The median time from onset to death was 17 days. Most patients were older male (70% male) with co-morbidities and only 11 % were smokers. 75% patients showed bilateral pneumonia. Many patients had chronic diseases, including hypertension (58.33%). cardiovascular disease (22.22%) and diabetes (19.44%). Typical clinical feature measured in these patients includes lymphopenia and elevated markers of inflammation.

12.32.3 Limitations

As noted by the authors, the conclusions of this study are very limited because this is single-centered study focusing on a small cohort of patients who did not survive. Many clinical parameters observed by the authors (such* as increase levels of serum CRP, PCT, IL-6) have also been described in other COVID19 patients who survived the infection

12.32.4 Significance

This study is essentially descriptive and may be useful for clinical teams monitoring COVID19 patients.

12.32.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

[942]

12.33.1 Keywords

12.33.2 Main Findings

Based on a retrospective study of 85 hospitalized COVID patients in a Beijing hospital, authors showed that patients with elevated ALT levels (n = 33) were characterized by significantly higher levels of lactic acid and CRP as well as lymphopenia and hypoalbuminemia compared to their counterparts with normal ALT levels. Proportion of severe and critical patients in the ALT elevation group was significantly higher than that of normal ALT group. Multivariate logistic regression performed on clinical factors related to ALT elevation showed that CRP \(\geq\) 20mg/L and low lymphocyte count (<1.1*10^9 cells/L) were independently related to ALT elevation—a finding that led the authors to suggest cytokine storm as a major mechanism of liver damage.

12.33.3 Limitations

The article’s most attractive claim that liver damage seen in COVID patients is caused by cytokine storm (rather than direct infection of the liver) hinges solely on their multivariate regression analysis. Without further mechanistic studies a) demonstrating how high levels of inflammatory cytokines can induce liver damage and b) contrasting types of liver damage incurred by direct infection of the liver vs. system-wide elevation of inflammatory cytokines, their claim remains thin. It is also worth noting that six of their elevated ALT group (n=33) had a history of liver disease (i.e. HBV infection, alcoholic liver disease, fatty liver) which can confound their effort to pin down the cause of hepatic injury to COVID.

12.33.4 Significance

Limited. This article confirms a rich body of literature describing liver damage and lymphopenia in COVID patients.

12.33.5 Credit

Review by Chang Moon as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.34 Detectable serum SARS-CoV-2 viral load (RNAaemia) is closely associated with drastically elevated interleukin 6 (IL-6) level in critically ill COVID-19 patients

[943]

12.34.1 Keywords

12.34.2 Main Findings

48 adult patients diagnosed with Covid19 according to Chinese guidelines for Covid19 diagnosis and treatment version 6 were included in this study. Patients were further sub-divided into three groups based on clinical symptoms and disease severity: (1) mild, positive Covid19 qPCR with no or mild clinical symptoms (fever; respiratory; radiological abnormalities); (2) severe, at least one of the following: shortness of breath/respiratory rate >30/min, oxygen saturation SaO2<93%, Horowitz index paO2/FiO2 < 300 mmHg (indicating moderate pulmonary damage); and (3) critically ill, at least one additional complicating factor: respiratory failure with need for mechanical ventilation; systemic shock; multi-organ failure and transfer to ICU. Serum samples and throat-swaps were collected from all 48 patients enrolled. SARS-CoV-2 RNA was assessed by qPCR with positive results being defined as Ct values < 40, and serum interleukin-6 (IL-6) was quantified using a commercially available detection kit. Briefly, patient characteristics in this study confirm previous reports suggesting that higher age and comorbidities are significant risk factors of clinical severity. Of note, 5 out of 48 of patients (10.41%), all in the critically ill category, were found to have detectable serum SARS-CoV-2 RNA levels, so-called RNAaemia. Moreover, serum IL-6 levels in these patients were found to be substantially higher and this correlated with the presence of detectable SARS-CoV-2 RNA levels. The authors hypothesize that viral RNA might be released from acutely damages tissues in moribund patients during the course of Covid19 and that RNaemia along with IL-6 could potentially be used as a prognostic marker.

12.34.3 Limitations

While this group’s report generally confirms some of the major findings of a more extensive study, published in early February 2020, [933], there are limitations that should be taken into account. First, the number of patients enrolled is relatively small; second, interpretation of these data would benefit from inclusion of information about study specifics as well as providing relevant data on the clinical course of these patients other than the fact that some were admitted to ICU (i.e. demographics on how many patients needed respiratory support, dialysis, APACHE Ii/III or other standard ICU scores as robust prognostic markers for mortality etc). It also remains unclear at which time point the serum samples were taken, i.e. whether at admission, when the diagnosis was made or during the course of the hospital stay (and potentially after onset of therapy, which could have affected both IL-6 and RNA levels). The methods section lacks important information on the qPCR protocol employed, including primers and cycling conditions used. From a technical point of view, Ct values >35 seem somewhat non-specific (although Ct <40 was defined as the CDC cutoff as well) indicating that serum RNA levels are probably very low, therefore stressing the need for highly specific primers and high qPCR efficiency. In addition, the statistical tests used (t-tests, according to the methods section) do not seem appropriate as the organ-specific data such as BUN and troponin T values seem to be not normally distributed across groups (n= 5 RNAaemia+ vs. n= 43 RNAaemia-). Given the range of standard deviations and the differences in patient sample size, it is difficult to believe that these data are statistically significantly different.

12.34.4 Significance

This study is very rudimentary and lacks a lot of relevant clinical details. However, it corroborates some previously published observations regarding RNAemia and IL-6 by another group. Generally, regarding future studies, it would be important to address the question of IL-6 and other inflammatory cytokine dynamics in relation to Covid19 disease kinetics (high levels of IL-6, IL-8 and plasma leukotriene were shown to have prognostic value at the onset of ARDS ; serum IL-2 and IL-15 have been associated with mortality; reviewed by Chen W & Ware L, Clin Transl Med. 2015 [944]).

12.34.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.35 Lymphopenia predicts disease severity of COVID-19: a descriptive and predictive study

[945]

12.35.1 Keywords

12.35.2 Main Findings

Based on a retrospective study of 162 COVID patients from a local hospital in Wuhan, China, the authors show an inverse correlation between lymphocyte % (LYM%) of patients and their disease severity. The authors have also tracked LYM% of 70 cases (15 deaths; 15 severe; 40 moderate) throughout the disease progression with fatal cases showing no recovery of lymphocytes ( <5%) even after 17-19 days post-onset. The temporal data of LYM % in COVID patients was used to construct a Time-Lymphocyte% model which is used to categorize and predict patients’ disease severity and progression. The model was validated using 92 hospitalized cases and kappa statistic test was used to assess agreement between predicted disease severity and the assigned clinical severity (k = 0.49).

12.35.3 Limitations

Time-Lymphocyte % Model (TLM) that authors have proposed as a predictive model for clinical severity is very simple in its construction and derives from correlative data of 162 patients. In order for the model to be of use, it needs validation using a far more robust data set and possibly a mechanistic study on how COVID leads to lymphopenia in the first place. In addition, it should be noted that no statistical test assessing significance of LYM % values between disease severities was performed.

12.35.4 Significance

This article is of limited significance as it simply reports similar descriptions of COVID patients made in previous literature that severe cases are characterized by lymphopenia.

12.35.5 Credit

Review by Chang Moon as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.36 The potential role of IL-6 in monitoring severe case of coronavirus disease 2019

Liu et al. medRxiv. [946]

12.36.1 Keywords

12.36.2 Main Findings

Study on blood biomarkers on 80 COVID19 patients (69 severe and 11 non-severe). Patients with severe symptoms at admission (baseline) showed obvious lymphocytopenia and significantly increased interleukin-6 (IL-6) and CRP, which was positively correlated with symptoms severity. IL-6 at baseline positively correlates with CRP, LDH, ferritin and D-Dimer abundance in blood.

Longitudinal analysis of 30 patients (before and after treatment) showed significant reduction of IL-6 in remission cases.

12.36.3 Limitations

Limited sample size at baseline, especially for the non-severe leads to question on representativeness. The longitudinal study method is not described in detail and suffers from non-standardized treatment. Limited panel of pro-inflammatory cytokine was analyzed. Patients with severe disease show a wide range of altered blood composition and biomarkers of inflammation, as well as differences in disease course (53.6% were cured, about 10% developed acute respiratory distress syndrome). The authors comment on associations between IL-6 levels and outcomes, but these were not statistically significant (maybe due to the number of patients, non-standardized treatments, etc.) and data is not shown. Prognostic biomarkers could have been better explored. Study lacks multivariate analysis.

12.36.4 Significance

IL-6 could be used as a pharmacodynamic marker of disease severity. Cytokine Release Syndrome (CRS) is a well-known side effect for CAR-T cancer therapy and there are several effective drugs to manage CRS. Drugs used to manage CRS could be tested to treat the most severe cases of COVID19.

12.36.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.37 Clinical and Laboratory Profiles of 75 Hospitalized Patients with Novel Coronavirus Disease 2019 in Hefei, China

Zhao et al. medRxiv. [947]

12.37.1 Keywords

12.37.2 Main Findings

The authors of this study provide a comprehensive analysis of clinical laboratory assessments in 75 patients (median age 47 year old) hospitalized for Corona virus infection in China measuring differential blood counts including T-cell subsets (CD4, CD8), coagulation function, basic blood chemistry, of infection-related biomarkers including CRP, Procalcitonin (PCT) (Precursor of calcitonin that increases during bacterial infection or tissue injury), IL-6 and erythrocyte sedimentation rate as well as clinical parameters. Among the most common hematological changes they found increased neutrophils, reduced CD4 and CD8 lymphocytes, increased LDH, CRP and PCT

When looking at patients with elevated IL-6, the authors describe significantly reduced CD4 and CD8 lymphocyte counts and elevated CRP and PCT levels were significantly increased in infected patients suggesting that increased IL-6 may correlate well with disease severity in COVID-19 infections

12.37.3 Limitations

The authors performed an early assessment of clinical standard parameters in patients infected with COVID-19. Overall, the number of cases (75) is rather low and the snapshot approach does not inform about dynamics and thus potential relevance in the assessment of treatment options in this group of patients.

12.37.4 Significance

The article summarizes provides a good summary of some of the common changes in immune cells inflammatory cytokines in patients with a COVID-19 infection and. Understanding how these changes can help predict severity of disease and guide therapy including IL-6 cytokine receptor blockade using Tocilizumab or Sarilumab will be important to explore.

12.37.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.38 Exuberant elevation of IP-10, MCP-3 and IL-1ra during SARS-CoV-2 infection is associated with disease severity and fatal outcome

Yang et al. medRxiv [948]

12.38.1 Keywords

12.38.2 Summary

Plasma cytokine analysis (48 cytokines) was performed on COVID-19 patient plasma samples, who were sub-stratified as severe (N=34), moderate (N=19), and compared to healthy controls (N=8). Patients were monitored for up to 24 days after illness onset: viral load (qRT-PCR), cytokine (multiplex on subset of patients), lab tests, and epidemiological/clinical characteristics of patients were reported.

12.38.3 Main Findings

12.38.4 Limitations

Collection time of clinical data and lab results not reported directly (likely 4 days (2,6) after illness onset), making it very difficult to determine if cytokines were predictive of patient outcome or reflective of patient compensatory immune response (likely the latter). Small N for cytokine analysis (N=2 fatal and N=5 severe/critical, and N=7 moderate or discharged). Viral treatment strategy not clearly outlined.

12.38.5 Expanded Results

NOTE: Moderate COVID-19 was classified by fever, respiratory manifestations, and radiological findings consistent with pneumonia while severe patients had one or more of the following: 1) respiratory distraction, resting O2 saturation, or 3) arterial PaO2/FiO2 < 300 mmHg.

Cytokine Results (Human Cytokine Screening Panel, Bio-Rad):

Lab results:

Clinical features (between moderate vs. severe patient groups):

12.38.6 Significance

Outline of pathological time course (implicating innate immunity esp.) and identification key cytokines associated with disease severity and prognosis (+ comorbidities). Anti-IP-10 as a possible therapeutic intervention (ex: Eldelumab).

12.38.7 Credit

Review by Natalie Vaninov as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.39 Antibody responses to SARS-CoV-2 in patients of novel coronavirus disease 2019

Zhao Jr. et al. medRxiv. [949]

12.39.1 Keywords

12.39.2 Main Findings

This study examined antibody responses in the blood of COVID-19 patients during the early SARS CoV2 outbreak in China. Total 535 plasma samples were collected from 173 patients (51.4% female) and were tested for seroconversion rate using ELISA. Authors also compared the sensitivity of RNA and antibody tests over the course of the disease . The key findings are:

12.39.3 Limitations

Because different types of ELISA assays were used for determining antibody concentrations at different time points after disease onset, sequential seroconversion of total Ab, IgM and IgG may not represent actual temporal differences but rather differences in the affinities of the assays used. Also, due to the lack of blood samples collected from patients in the later stage of illness, how long the antibodies could last remain unknown. For investigative dynamics of antibodies, more samples were required.

12.39.4 Significance

Total and IgG antibody titers could be used to understand the epidemiology of SARS CoV-2 infection and to assist in determining the level of humoral immune response in patients.

The findings provide strong clinical evidence for routine serological and RNA testing in the diagnosis and clinical management of COVID-19 patients. The understanding of antibody responses and their half-life during and after SARS CoV2 infection is important and warrants further investigations.

12.39.5 Credit

This review was undertaken by Zafar Mahmood and edited by K Alexandropoulos as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.40 Restoration of leukomonocyte counts is associated with viral clearance in COVID-19 hospitalized patients

Chen et al. medRxiv [950]

12.40.1 Keywords

12.40.2 Main Findings

The authors collected data on 25 COVID-19 patients (n=11 men, n=14 women) using standard laboratory tests and flow cytometry. All patients were treated with antibiotics. Twenty-four of the 25 patients were also treated with anti-viral Umefinovir and 14 of the patients were treated with corticosteroids. 14 patients became negative for the virus after 8-14 days of treatment. The same treatment course was extended to 15-23 days for patients who were still positive for the virus at day 14.

The authors found a negative association between age and resolution of infection. Patients with hypertension, diabetes, malignancy or chronic liver disease were all unable to clear the virus at day 14, though not statistically significant.

Elevated procalcitonin and a trend for increased IL-6 were also found in peripheral blood prior to the treatment.

A trend for lower NK cell, T cell and B cell counts in patients was also reported. B cell, CD4 and CD8 T cell counts were only increased upon treatment in patients who cleared the virus. NK cell frequencies remained unchanged after treatment in all the patients.

12.40.3 Limitations

73% of the patients who remained positive for SARS-CoV2 after the 1st treatment, and 43% of all patients who cleared the virus were treated with corticosteroids. Corticosteroids have strong effects on the immune compartment in blood [951]. The authors should have accounted for corticosteroid treatment when considering changes in T, NK and B cell frequencies.

Assessing if IL-6 concentrations were back to baseline levels following treatment would have provided insights into the COVID-19 cytokine storm biology. Patients with higher baseline levels of IL-6 have been reported to have lower CD8 and CD4 T cell frequencies [947]. Correlating IL-6 with cell counts before and after treatment would thus have also been of interest. The report of the laboratory measures in table 2 is incomplete and should include the frequencies of patients with increased/decreased levels for each parameter.

Correction is needed for the 1st paragraph of the discussion as data does not support NK cell restoration upon treatment in patients who cleared the virus. NK cells remain unchanged after the 1st treatment course and only seem to increase in 2 out of 6 donors after the 2nd treatment course in those patients.

12.40.4 Significance

Previous reports suggest an association between disease severity and elevated IL-6 or pro-calcitonin concentrations in COVID-19 patients [943,952]. IL-6 receptor blockade is also being administered to patients enrolled in clinical trials (NCT04317092). This report thus contributes to highlight elevated concentrations of these analytes in COVID-19 patients. Mechanisms underlying the association between viral clearance and restoration of the T cell and B cell frequencies suggests viral-driven immune dysregulation, which needs to be investigated in further studies.

12.40.5 Credit

Review by Bérengère Salomé as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.41 Clinical findings in critically ill patients infected with SARS-CoV-2 in Guangdong Province, China: a multi-center, retrospective, observational study

Xu et al. medRxiv. [953]

12.41.1 Keywords

12.41.2 Main Findings:

This work analyses laboratory and clinical data from 45 patients treated in the in ICU in a single province in China. Overall, 44% of the patients were intubated within 3 days of ICU admission with only 1 death.

Lymphopenia was noted in 91% of patient with an inverse correlation with LDH.

Lymphocyte levels are negatively correlated with Sequential Organ Failure Assessment (SOFA) score (clinical score, the higher the more critical state), LDH levels are positively correlated to SOFA score. Overall, older patients (>60yo), with high SOFA score, high LDH levels and low lymphocytes levels at ICU admission are at higher risk of intubation.

Of note, convalescent plasma was administered to 6 patients but due to limited sample size no conclusion can be made.

12.41.3 Limitations

While the study offers important insights into disease course and clinical lab correlates of outcome, the cohort is relatively small and is likely skewed towards a less-severe population compared to other ICU reports given the outcomes observed. Analysis of laboratory values and predictors of outcomes in larger cohorts will be important to make triage and treatment decisions. As with many retrospective analyses, pre-infection data is limited and thus it is not possible to understand whether lymphopenia was secondary to underlying comorbidities or infection.

Well-designed studies are necessary to evaluate the effect of convalescent plasma administration.

12.41.4 Significance

This clinical data enables the identification of at-risk patients and gives guidance for research for treatment options. Indeed, further work is needed to better understand the causes of the lymphopenia and its correlation with outcome.

12.41.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.42 Immune Multi-epitope vaccine design using an immunoinformatics approach for 2019 novel coronavirus in China (SARS-CoV-2)

[954]

12.42.1 Keywords

12.42.2 Main Findings

Using in silico bioinformatic tools, this study identified putative antigenic B-cell epitopes and HLA restricted T-cell epitopes from the spike, envelope and membrane proteins of SARS-CoV-2, based on the genome sequence available on the NCBI database. T cell epitopes were selected based on predicted affinity for the more common HLA-I alleles in the Chinese population. Subsequently, the authors designed vaccine peptides by bridging selected B-cell epitopes and adjacent T-cell epitopes. Vaccine peptides containing only T-cell epitopes were also generated.

From 61 predicted B-cell epitopes, only 19 were exposed on the surface of the virion and had a high antigenicity score. A total of 499 T-cell epitopes were predicted. Based on the 19 B-cell epitopes and their 121 adjacent T-cell epitopes, 17 candidate vaccine peptides were designed. Additionally, another 102 vaccine peptides containing T-cell epitopes only were generated. Based on the epitope counts and HLA score, 13 of those were selected. Thus, a total of 30 peptide vaccine candidates were designed.

12.42.3 Limitations

While this study provides candidates for the development of vaccines against SARS-CoV-2, in vitro and in vivo trials are required to validate the immunogenicity of the selected B and T cell epitopes. This could be done using serum and cells from CoV-2-exposed individuals, and in preclinical studies. The implication of this study for the current epidemic are thus limited. Nevertheless, further research on this field is greatly needed.

12.42.4 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.43 Clinical Features of Patients Infected with the 2019 Novel Coronavirus (COVID-19) in Shanghai, China

Cao et al. medRxiv [955]

12.43.1 Keywords

12.43.2 Main Findings

This single-center cohort study analyzes the clinical and laboratory features of 198 patients with confirmed COVID-19 infection in Shanghai, China and correlated these parameters with clinical disease severity, including subsequent intensive care unit (ICU) admission. 19 cases (9.5%) required ICU admission after developing respiratory failure or organ dysfunction. Age, male sex, underlying cardiovascular disease, and high symptom severity (high fever, dyspnea) were all significantly correlated with ICU admission. Additionally, ICU admission was more common in patients who presented with lymphopenia and elevated neutrophil counts, among other laboratory abnormalities. Flow cytometric analysis revealed that patients admitted to the ICU had significantly reduced circulating CD3+ T cell, CD4+ T cell, CD8+ T cell, and CD45+ leukocyte populations compared to the cohort of patients not requiring ICU admission.

12.43.3 Limitations

The limitations of this study include the relatively small sample size and lack of longitudinal testing. The authors also did not assess whether respiratory comorbidity – such as asthma or chronic obstructive lung disease – in addition to immunosuppression affected ICU admission likelihood.

12.43.4 Significance

COVID-19 has already sickened thousands across the globe, though the severity of these infections is markedly diverse, ranging from mild symptoms to respiratory failure requiring maximal intervention. Understanding what clinical, laboratory, and immunologic factors predict the clinical course of COVID-19 infection permits frontline providers to distribute limited medical resources more effectively.

12.43.5 Credit

Review by Andrew Charap as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine at Mount Sinai.

12.44 Serological detection of 2019-nCoV respond to the epidemic: A useful complement to nucleic acid testing

Zhang et al. medRxiv. [956]

12.44.1 Keywords

12.44.2 Main Finding

This study showed that both anti-2019-nCov IgM and IgG were detected by automated chemiluminescent immunoassay in the patients who had been already confirmed as positive by nucleic acid detection, while single positivity of IgM or IgG were detected in a very few cases in the other population including 225 non-COVID-19 cases. In addition to the increase of anti-2019-nCov IgM 7-12 days after morbidity, the increase of IgG was detected in three patients with COVID-19 within a very short of time (0-1 day).

12.44.3 Limitations

The limitation of this study is only 3 confirmed COVID-19 cases were included, so that the relationship between anti-2019-nCov antibodies and disease progression might not be clearly defined. Another limitation is that they did not show the course of 2019-nCov specific antibodies in the cases with positive for COVID-19 but without clinical symptoms.

12.44.4 Significance

The detection of anti-2019-nCov antibodies can be an alternative method to diagnose and treat COVID-19 more comprehensively by distinguish non COVID-19 patients. It may be helpful to understand the course of individual cases with COVID-19 to predict the prognosis if more cases will be evaluated.

12.44.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.45 Human Kidney is a Target for Novel Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) Infection

[957]

12.45.1 Keywords

12.45.2 Main Finding

Analyzing the eGFR (effective glomerular flow rate) of 85 Covid-19 patients and characterizing tissue damage and viral presence in post-mortem kidney samples from 6 Covid-19 patients, the authors conclude that significant damage occurs to the kidney, following Covid-19 infection. This is in contrast to the SARS infection from the 2003 outbreak. They determine this damage to be more prevalent in patients older than 60 years old, as determined by analysis of eGFR. H&E and IHC analysis in 6 Covid-19 patients revealed that damage was in the tubules, not the glomeruli of the kidneys and suggested that macrophage accumulation and C5b-9 deposition are key to this process. 

12.45.3 Limitations

Severe limitations include that the H&E and IHC samples were performed on post-mortem samples of unknown age, thus we cannot assess how/if age correlates with kidney damage, upon Covid-19 infection. Additionally, eGFR was the only in-vivo measurement. Blood urea nitrogen and proteinuria are amongst other measurements that could have been obtained from patient records. An immune panel of the blood was not performed to assess immune system activation. Additionally, patients are only from one hospital. 

12.45.4 Significance

This report makes clear that kidney damage is prevalent in Covid-19 patients and should be accounted for. 

12.45.5 Credit

Review by Dan Fu Ruan, Evan Cody and Venu Pothula as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine at Mount Sinai.

12.46 COVID-19 early warning score: a multi-parameter screening tool to identify highly suspected patients

Song et al. medRxiv. [958]

12.46.1 Keywords

12.46.2 Main Findings

The aim of this study was to identify diagnostic or prognostic criteria which could identify patients with COVID-19 and predict patients who would go on to develop severe respiratory disease. The authors use EMR data from individuals taking a COVID-19 test at Zhejiang hospital, China in late January/Early February. A large number of clinical parameters were different between individuals with COVID-19 and also between ‘severe’ and ‘non-severe’ infections and the authors combine these into a multivariate linear model to derive a weighted score, presumably intended for clinical use.

12.46.3 Limitations

Unfortunately, the paper is lacking a lot of crucial information, making it impossible to determine the importance or relevance of the findings. Most importantly, the timings of the clinical measurements are not described relative to the disease course, so it is unclear if the differences between ‘severe’ and ‘non-severe’ infections are occurring before progression to severe disease (which would make them useful prognostic markers), or after (which would not).

12.46.4 Significance

This paper is one of many retrospective studies coming from hospitals in China studying individuals with COVID-19. Because of the sparse description of the study design, this paper offers little new information. However, studies like this could be very valuable and we would strongly encourage the authors to revise this manuscript to include more information about the timeline of clinical measurements in relation to disease onset and more details of patient outcomes.

12.46.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.47 LY6E impairs coronavirus fusion and confers immune control of viral disease

Pfaender et al. bioRxiv. [959]

12.47.1 Keywords

12.47.2 Main Findings

Screening a cDNA library of >350 human interferon-stimulated genes for antiviral activity against endemic human coronavirus HCoV-229E (associated with the common cold), Pfaender S & Mar K et al. identify lymphocyte antigen 6 complex, locus E (Ly6E) as an inhibitor of cellular infection of Huh7 cells, a human hepatoma cell line susceptible to HCoV-229E and other coronaviruses. In a series of consecutive in vitro experiments including both stable Ly6E overexpression and CRISPR-Cas9-mediated knockout the authors further demonstrate that Ly6E reduces cellular infection by various other coronaviruses including human SARS-CoV and SARS-CoV-2 as well as murine CoV mouse hepatitis virus (MHV). Their experiments suggest that this effect is dependent on Ly6E inhibition of CoV strain-specific spike protein-mediated membrane fusion required for viral cell entry.

To address the function of Ly6E in vivo, hematopoietic stem cell-specific Ly6E knock-out mice were generated by breeding Ly6Efl/fl mice (referred to as functional wild-type mice) with transgenic Vav-iCre mice (offspring referred to as Ly6E HSC ko mice); wild-type and Ly6E HSC ko mice of both sexes were infected intraperitoneally with varying doses of the natural murine coronavirus MHV, generally causing a wide range of diseases in mice including hepatitis, enteritis and encephalomyelitis. Briefly, compared to wild-type controls, mice lacking hematopoietic cell-expressed Ly6E were found to present with a more severe disease phenotype as based on serum ALT levels (prognostic of liver damage), liver histopathology, and viral titers in the spleen. Moreover, bulk RNAseq analysis of infected liver and spleen tissues indicated changes in gene expression pathways related to tissue damage and antiviral immune responses as well as a reduction of genes associated with type I IFN response and inflammation. Finally, the authors report substantial differences in the numbers of hepatic and splenic APC subsets between wild-type and knockout mice following MHV infection and show that Ly6E-deficient B cells and to a lesser extent also DCs are particularly susceptible to MHV infection in vitro.

12.47.3 Limitations

Experiments and data in this study are presented in an overall logical and coherent fashion; however, some observations and the conclusions drawn are problematic and should be further addressed & discussed by the authors. Methodological & formal limitations include relatively low replicate numbers as well as missing technical replicates for some in vitro experiments (cf. Fig. legend 1; Fig. legend 2e); the omission of “outliers” in Fig. legend 2 without an apparent rationale as to why this approach was chosen; the lack of detection of actual Ly6E protein levels in Ly6E HSC ko or wild-type mice; and most importantly, missing information on RNAseq data collection & analysis in the method section and throughout the paper. A more relevant concern though is that the interpretation of the experimental data presented and the language used tend to overrate and at times overgeneralize findings: for example, while the authors demonstrate statistically significant, Ly6E-mediated reduction of coronavirus titers in stable cells lines in vitro, it remains unclear whether a viral titer reduction by one log decade would be of actual biological relevance in face of high viral titers in vivo. After high-dose intraperitoneal MHV infection in vivo, early viral titers in Ly6E HSC knockout vs. wt mice only showed an elevation in the spleen (~1.5 log decades) but not liver of the ko mice (other tissue not evaluated), and while ko mice presented with only modestly increased liver pathology, both male and female ko mice exhibited significantly higher mortality. Thus, the manuscript tile statement that “Ly6E … confers immune control of viral disease” is supported by only limited in vivo data, and gain-of-function experiments (eg. Ly6E overexpression) were not performed. Of additional note here, tissue tropism and virulence differ greatly among various MHV strains and isolates whereas dose, route of infection, age, genetic background and sex of the mice used may additionally affect disease outcome and phenotype (cf. Taguchi F & Hirai-Yuki A, https://doi.org/10.3389/fmicb.2012.00068; Kanolkhar A et al, https://jvi.asm.org/content/ 83/18/9258). Observations attributed to hematopoietic stem cell-specific Ly6E deletion could therefore be influenced by the different genetic backgrounds of floxed and cre mice used, and although it appears that littermates wt and ko littermates were used in the experiments, the potentially decisive impact of strain differences should at least have been discussed. Along these lines, it should also be taken into account that the majority of human coronaviruses cause respiratory symptoms, which follow a different clinical course engaging other primary cellular mediators than the hepatotropic murine MHV disease studied here. It therefore remains highly speculative how the findings reported in this study will translate to human disease and it would therefore be important to test other routes of MHV infection and doses that have been described to produce a more comparable phenotype to human coronavirus disease (cf. Kanolkhar A et al, https://jvi.asm.org/content/ 83/18/9258). Another important shortcoming of this study is the lack of any information on functional deficits or changes in Ly6E-deficient immune cells and how this might relate to the phenotype observed. Overall, the in vitro experiments are more convincing than the in vivo studies which appear somewhat limited.

12.47.4 Significance

Despite some shortcomings, the experiments performed in this study suggest a novel and somewhat unexpected role of Ly6E in the protection against coronaviruses across species. These findings are of relevance and should be further explored in ongoing research on potential coronavirus therapies. Yet an important caveat pertains to the authors’ suggestion that “therapeutic mimicking of Ly6E action” may constitute a first line of defense against novel coronaviruses since their own prior work demonstrated that Ly6E can enhance rather than curtail infection with influenza A and other viruses.

12.47.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.48 A preliminary study on serological assay for severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) in 238 admitted hospital patients

Liu et al. medRxiv. [960]

12.48.1 Keywords

12.48.2 Main Findings

While RT-PCR is being used currently to routinely diagnose infection with SARS-CoV-2, there are significant limitations to the use of a nucleic acid test that lead to a high false-negative rate. This article describes ELISAs that can measure IgM and IgG antibodies against the N protein of SARS-CoV-2 to test samples from 238 patients (153 positive by RT-PCR and 85 negative by RT-PCR) at different times after symptom onset. The positivity rate of the IgM and/or IgG ELISAs was greater than that of the RT-PCR (81.5% compared to 64.3%) with similar positive rates in the confirmed and suspected cases (83% and 78.8%, respectively), suggesting that many of the suspected but RT-PCR-negative cases were also infected. The authors also found that the ELISAs have higher positive rates later after symptom onset while RT-PCR is more effective as a diagnostic test early during the infection.

12.48.3 Limitations

I cannot identify any limitations to this study.

12.48.4 Significance

The authors make a strong case for using a combination of ELISA and RT-PCR for diagnosis of infection with SARS-CoV-2, especially considering the dynamics of positivity rates of RT-PCR and ELISA. Fewer false-negative diagnoses would improve infection control and patient management.

12.48.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.49 Monoclonal antibodies for the S2 subunit of spike of SARS-CoV cross-react with the newly-emerged SARS-CoV-2

[961]

12.49.1 Keywords

12.49.2 Main Findings

Whole genome sequencing-based comparisons of the 2003 Severe Acute Respiratory Syndrome Coronavirus (SARS-CoV) and the 2019 SARS-CoV-2 revealed conserved receptor binding domain (RBD) and host cell receptor, angiotensin-converting enzyme 2 (ACE2). In line with this, the authors tested cross-reactivity of murine monoclonal antibodies (mAbs) previously generated against the SARS-CoV spike (S) glycoprotein involved in viral entry. One of the screened mAb, 1A9, was able to bind and cross-neutralize multiple strains of SARS-CoV, as well as, detect the S protein in SARS-CoV-2-infected cells. mAb 1A9 was generated using an immunogenic fragment in the S2 subunit of SARS-CoV and binds through a novel epitope within the S2 subunit at amino acids 1111-1130. It is important to note that CD8+ T lymphocyte epitopes overlap with these residues, suggesting that S2 subunit could be involved in inducing both, humoral and cell-mediated immunity.

12.49.3 Limitations

The authors used previously generated mouse mAbs against the S protein in SARS-CoV expressed in mammalian cell line. Future experimental validation using COVID-19 patient samples is needed to validate these findings. In addition, the results of these studies are predominantly based on in vitro experiments and so, evaluating the effects of the mAb 1A9 in an animal model infected with this virus will help us better understand the host immune responses in COVID-19 and potential therapeutic vaccines.

12.49.4 Significance

This study identified mAbs that recognize the new coronavirus, SARS-Cov-2. These cross-reactive mAbs will help in developing diagnostic assays for COVID-19.

12.49.5 Credit

This review was undertaken by Tamar Plitt and Katherine Lindblad as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.50 Mortality of COVID-19 is Associated with Cellular Immune Function Compared to Immune Function in Chinese Han Population

Zeng et al. medRxiv. [962]

12.50.1 Keywords

12.50.2 Main Findings

Retrospective study of the clinical characteristics of 752 patients infected with COVID-19 at Chinese PLA General Hospital, Peking Union Medical College Hospital, and affiliated hospitals at Shanghai University of medicine & Health Sciences. This study is the first one that compares PB from healthy controls from the same regions in Shanghai and Beijing, and infected COVID-19 patients to standardize a reference range of WBCs of people at high risk.

12.50.3 Limitations

Lower levels of leukocyte counts -B cells, CD4 and CD8 T cells- correlated with mortality (WBCs are significantly lower in severe or critical UCI patients vs mild ones). Based on 14,117 normal controls in Chinese Han population (ranging in age from 18-86) it is recommended that reference ranges of people at high risk of COVID-19 infection are CD3+ lymphocytes below 900 cells/mm3, CD4+ lymphocytes below 500 cells/mm3, and CD8+ lymphocytes below 300 cells/mm3. Importantly, this study also reported that the levels of D-dimer, C-reactive protein and IL-6 were elevated in COVID-19 pts., indicating clot formation, severe inflammation and cytokine storm.

12.50.4 Significance

This study sets a threshold to identify patients at risk by analyzing their levels of leukocytes, which is an easy and fast approach to stratify individuals that require hospitalization. Although the study is limited (only counts of WBC are analyzed and not its profile) the data is solid and statistically robust to correlate levels of lymphopenia with mortality.

12.50.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.51 Retrospective Analysis of Clinical Features in 101 Death Cases with COVID-19

Chen et al. medRxiv. [963]

12.51.1 Keywords

12.51.2 Main Findings

This is a retrospective study involving 101 death cases with COVID-19 in Wuhan Jinyintan Hospital. The aim was to describe clinical, epidemiological and laboratory features of fatal cases in order to identify the possible primary mortality causes related to COVID-19.

Among 101 death cases, 56.44% were confirmed by RT-PCR and 43.6% by clinical diagnostics. Males dominated the number of deaths and the average age was 65.46 years. All patients died of respiratory failure and multiple organs failure, except one (acute coronary syndrome). The predominant comorbidities were hypertension (42.57%) and diabetes (22.77%). 25.74% of the patients presented more than two underlying diseases. 82% of patients presented myocardial enzymes abnormalities at admission and further increase in myocardial damage indicators with disease progression: patients with elevated Troponin I progressed faster to death. Alterations in coagulation were also detected. Indicators of liver and kidney damage increased 48 hours before death. The authors studied the deceased patients’ blood type and presented the following results: type A (44.44%), type B (29.29%), type AB (8.08%) and type O (18.19%), which is inconsistent with the distribution in Han population in Wuhan.

Clinical analysis showed that the most common symptom was fever (91.9%), followed by cough and dyspnea. The medium time from onset of symptoms to acute respiratory distress syndrome (ARDS) development was 12 days. Unlike SARS, only 2 patients with COVID-19 had diarrhea. 98% presented abnormal lung imaging at admission and most had double-lung abnormalities. Related to the laboratorial findings some inflammatory indicators gradually increased during the disease progression, such as IL-6 secretion in the circulation, procalcitonin (PCT) and C-reactive protein (CRP), while platelets numbers decreased. The authors also reported an initial lymphopenia that was followed by an increase in the lymphocytes numbers. Neutrophil count increased with disease progression.

The patients received different treatments such as antiviral drugs (60.40%), glucocorticoids, thymosin and immunoglobulins. All patients received antibiotic treatment and some received antifungal drugs. All patients received oxygen therapy (invasive or non-invasive ones).

12.51.3 Limitations

This study involves just fatal patients, lacking comparisons with other groups of patients e.g. patients that recovered from COVID-19. The authors didn’t discuss the different approaches used for treatments and how these may affect the several parameters measured. The possible relationship between the increase of inflammatory indicators and morbidities of COVID-19 are not discussed.

12.51.4 Significance

This study has the largest cohort of fatal cases reported so far. The authors show that COVID-19 causes fatal respiratory distress syndrome and multiple organ failure. This study highlights prevalent myocardial damage and indicates that cardiac function of COVID-19 patients should be carefully monitored. The data suggest that Troponin I should be further investigated as an early indicator of patients with high risk of accelerated health deterioration. Secondary bacterial and fungal infections were frequent in critically ill patients and these need to be carefully monitored in severe COVID-19 patients. Differences in blood type distribution were observed, suggesting that type A is detrimental while type O is protective – but further studies are needed to confirm these findings and elucidate if blood type influences infection or disease severity. Several inflammatory indicators (neutrophils, PCT, CRP and IL-6, D-dimer) increased according to disease severity and should be assessed as biomarkers and to better understand the biology of progression to severe disease.

12.51.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.52 Relationship between the ABO Blood Group and the COVID-19 Susceptibility

Zhao et al. medRxiv. [964]

12.52.1 Keywords

12.52.2 Main Findings

These authors compared the ABO blood group of 2,173 patients with RT-PCR-confirmed COVID-19 from hospitals in Wuhan and Shenzhen with the ABO blood group distribution in unaffected people in the same cities from previous studies (2015 and 2010 for Wuhan and Shenzhen, respectively). They found that people with blood group A are statistically over-represented in the number of those infected and who succumb to death while those with blood group O are statistically underrepresented with no influence of age or sex.

12.52.3 Limitations

This study compares patients with COVID-19 to the general population but relies on data published 5 and 10 years ago for the control. The mechanisms that the authors propose may underlie the differences they observed require further study.

12.52.4 Significance

Risk stratification based on blood group may be beneficial for patients and also healthcare workers in infection control. Additionally, investigating the mechanism behind these findings could lead to better developing prophylactic and therapeutic targets for COVID-19.

12.52.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.53 The inhaled corticosteroid ciclesonide blocks coronavirus RNA replication by targeting viral NSP15

Matsuyama et al. bioRxiv [965]

12.53.1 Keywords

12.53.2 Main Findings

This study reconsiders the use of inhaled corticosteroids in the treatment of pneumonia by coronavirus. Corticosteroids were associated with increased mortality for SARS in 2003 and for MERS in 2013, probably due to that fact that systemic corticosteroids suppress the innate immune system, resulting in increased viral replication. However, some steroid compounds might block coronavirus replication. The authors screened steroids from a chemical library and assessed the viral growth suppression and drug cytotoxicity. Ciclesonide demonstrated low cytotoxicity and potent suppression of MERS-CoV viral growth. The commonly used systemic steroids cortisone, prednisolone and dexamethasone did not suppress viral growth, nor did the commonly used inhaled steroid fluticasone. To identify the drug target of virus replication, the authors conducted 11 consecutive MERS-CoV passages in the presence of ciclesonide or mometasone, and they could generate a mutant virus that developed resistance to ciclesonide, but not to mometasone. Afterwards, they performed next-generation sequencing and identified an amino acid substitution in nonstructural protein 15 (NSP15) as the predicted mechanism for viral resistance to ciclesonide. The authors were able to successfully generate a recombinant virus carrying that amino acid substitution, which overcome the antiviral effect of ciclesonide, suggesting that ciclosenide interacts with NSP15. The mutant virus was inhibited by mometasone, suggesting that the antiviral target of mometasone is different from that of ciclesonide. Lastly, the effects of ciclesonide and mometason on suppressing the replication of SARS-CoV-2 were evaluated. Both compounds were found to suppress viral replication with a similar efficacy to lopinavir.

12.53.3 Limitations

Most of the experiments, including the identification of the mutation in NSP15 were conducted with MERS-CoV. This is not the closest related virus to SARS-CoV-2, as that would be SARS-CoV. Thus, to repeat the initial experiments with SARS-CoV, or preferably SARS-CoV-2, is essential. The manuscript should address this and, therefore, it will require considerable editing for organization and clarity. Also, in terms of cell immunogenic epitopes, while SARS-CoV-2 spike protein contains several predicted B and T cell immunogenic epitopes that are shared with other coronaviruses, some studies have shown critical differences between MERS-CoV, SARS-CoV and SARS-CoV-2. A main criticism is that the authors only used VeroE6/TMPRSS2 cells to gauge the direct cytotoxic effects of viral replication. To evaluate this in other cell lines, including human airway epithelial cells, is crucial, as the infectivity of coronavirus strains greatly varies in different cell lines,

12.53.4 Significance

Nevertheless, these findings encourage evaluating ciclesonide and mometasone as better options for patients with COVID-19 in need of inhaled steroids, especially as an alternative to other corticosteroids that have been shown to increase viral replication in vitro. This should be evaluated in future clinical studies.

12.53.5 Credit

This review was undertaken by Alvaro Moreira, MD as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.54 A human monoclonal antibody blocking SARS-CoV-2 infection **

Wang et al. bioRxiv. [459]

12.54.1 Keywords

12.54.2 Main Findings

The authors reported a human monoclonal antibody that neutralizes SARS-CoV-2 and SARS-Cov which belong to same family of corona viruses. For identifying mAbs, supernatants of a collection of 51 hybridomas raised against the spike protein of SARS-CoV (SARS-S) were screened by ELISA for cross-reactivity against the spike protein of SARS-CoN2 (SARS2-S). Hybridomas were derived from immunized transgenic H2L2 mice (chimeric for fully human VH-VL and rat constant region). Four SARS-S hybridomas displayed cross-reactivity with SARS2-S, one of which (47D11) exhibited cross-neutralizing activity for SARS-S and SARS2-S pseudotyped VSV infection. A recombinant, fully human IgG1 isotype antibody was generated and used for further characterization.

The humanized 47D11 antibody inhibited infection of VeroE6 cells with SARS-CoV and SARS-CoV-2 with IC50 values of 0.19 and 0.57 μg/ml respectively. 47D11 mAb bound a conserved epitope on the spike receptor binding domain (RBD) explaining its ability to cross-neutralize SARS-CoV and SARS-CoV-2. 47D11 was shown to target the S1B RBD of SARS-S and SARS2-S with similar affinities. Interestingly, binding of 47D11 to SARS-S1B and SARS2-S1B did not interfere with S1B binding to ACE2 receptor-expressing cells assayed by flow cytometry.

12.54.3 Limitations

These results show that the human 47D11 antibody neutralizes SARS-CoV and SARS-Cov2 infectivity via an as yet unknown mechanism that is different from receptor binding interference. Alternative mechanisms were proposed but these as yet remain to be tested in the context of SARS-CoV2. From a therapeutic standpoint and in the absence of in vivo data, it is unclear whether the 47D11 ab can alter the course of infection in an infected host through virus clearance or protect an uninfected host that is exposed to the virus. There is a precedent for the latter possibility as it relates to SARS-CoV that was cited by the authors and could turn out to be true for SARS-CoV2.

12.54.4 Significance

This study enabled the identification of novel neutralizing antibody against COV-that could potentially be used as first line of treatment in the near future to reduce the viral load and adverse effects in infected patients. In addition, neutralizing antibodies such as 47D11 represent promising reagents for developing antigen-antibody-based detection test kits and assays.

12.54.5 Credit

This review was edited by K. Alexandropoulos as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

Heat inactivation of serum interferes with the immunoanalysis of antibodies to SARS-CoV-2

Heat inactivation, immunochromatography, diagnosis, serum antibodies, IgM, IgG

Summary

The use of heat inactivation to neutralize pathogens in serum samples collected from suspected COVID-19 patients reduces the sensitivity of a fluorescent immunochromatographic assay to detect anti-SARS-CoV-2 IgM and IgG.

Major findings

Coronaviruses can be killed by heat inactivation, and this is an important safety precaution in laboratory manipulation of clinical samples. However, the effect of this step on downstream SARS-CoV-2-specific serum antibody assays has not been examined. The authors tested the effect of heat inactivation (56 deg C for 30 minutes) versus no heat inactivation on a fluorescence immunochromatography assay. Heat inactivation reduced all IgM measurements by an average of 54% and most IgG measurements (22/36 samples, average reduction of 50%), consistent with the lower thermal stability of IgM than that of IgG. Heat inactivation caused a subset of IgM but not IgG readings to fall below a specified positivity threshold.

Limitations

Limitations included the use of only one type of assay for testing heat inactivated vs non-inactivated sera, and the use of the same baseline for heat inactivated and non-inactivated sera. The results indicate that heat inactivation affects the quantification of SARS-CoV-2-antibody response, specially IgM, but still allows to distinguish positive specific IgG. Therefore, the effect of heat inactivation should be studied when designing assays that quantitatively associate immunoglobulin levels (especially IgM) to immune state.

Review by Andrew M. Leader as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn school of medicine, Mount Sinai.

12.55 Immune phenotyping based on neutrophil-to-lymphocyte ratio and IgG predicts disease severity and outcome for patients with COVID-19

Zhang et al. medRxiv [966]

12.55.1 Keywords

12.55.2 Main Findings

In a cohort of 222 patients, anti-SARS-CoV-2 IgM and IgG levels were analyzed during acute and convalescent phases (up to day 35) and correlated to the diseases’ severity. The same was done with neutrophil-to-lymphocyte ratio. High IgG levels and high neutrophil-to-lymphocyte ratio in convalescence were both independently associated to the severity of the disease. The simultaneous occurrence of both of these laboratory findings correlated even stronger to the diseases’ severity.

Severe cases with high neutrophil-to-lymphocyte ratios had clearly higher levels of IL-6. The authors propose that a robust IgG response leads to immune-mediated tissue damage, thus explaining the worse outcome in patients with overexuberant antibody response.

12.55.3 Limitations

A main criticism is that the criteria for stratifying patients in severe vs. non-severe are not described. The only reference related to this is the difference between the percentage of patients who needed mechanical ventilation, which was greater in patients with both high IgG levels and high neutrophil-to-lymphocyte ratio. No patient with both low IgG levels and low neutrophil-to-lymphocyte ratio was treated with mechanical ventilation.

The proposed correlation of severity with IL-2 and IL-10 levels is not very strong.

Furthermore, although mostly ignored in the paper’s discussion, one of the most interesting findings is that an early increase in anti-SARS-CoV-2 IgM levels also seems to correlate with severe disease. However, as only median values are shown for antibody kinetics curves, the extent of variation in acute phase cannot be assessed.

12.55.4 Significance

Anti-SARS-CoV-2 IgG levels and with neutrophil-to-lymphocyte ratio predict severity of COVID-19 independently of each other. An additive predictive value of both variables is noticeable. Importantly, an early-on increase in anti-SARS-CoV-2 IgM levels also seem to predict outcome.

12.55.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.56 Reinfection could not occur in SARS-2 CoV-2 infected rhesus macaques

Bao et al. bioRxiv [967]

12.56.1 Keywords

12.56.2 Main Findings

This study addresses the issue or acquired immunity after a primary COVID-19 infection in rhesus monkeys. Four Chinese rhesus macaques were intratracheally infected with SARS-CoV-2 and two out of the four were re-infected at 28 days post initial infection (dpi) with the same viral dose after confirming the recovery by the absence of clinical symptoms, radiological abnormalities and viral detection (2 negative RT-PCR tests). While the initial infection led the viral loads in nasal and pharyngeal swabs that reach approximately 6.5 log10 RNA copies/ml at 3 dpi in all four monkeys, viral loads in the swabs tested negative after reinfection in the two reinfected monkeys. In addition, the necropsies from a monkey (M1) at 7 days after primary infection, and another monkey (M3) at 5 days post reinfection, revealed the histopathological damages and viral replication in the examined tissues from M1, while no viral replication as well as no histological damages were detected in the tissues from M3. Furthermore, sera from three monkeys at 21 and 28 dpi exhibited neutralizing activity against SARS-CoV-2 in vitro, suggesting the production of protective neutralizing antibodies in these monkeys. Overall, this study indicates that primary infection with SARS-CoV-2 may protect from subsequent exposure to the same virus.

12.56.3 Limitations

In human, virus has been detected by nasopharyngeal swabs until 9 to 15 days after the onset of symptoms. In the infected monkeys in this study, virus were detected from day 1 after the infection, declining to undetectable level by day 15 post infection. It may suggest that there is a faster viral clearance mechanism in monkeys, therefore the conclusions of reinfection protection for humans need to be carefully considered. In addition, only two monkeys were re-infected in this study and the clinical signs of these monkeys were not similar: M3 did not show weight loss and M4 showed relatively higher fever on the day of infection and the day of re-challenge.

12.56.4 Significance

This study showed clear viral clearance and no indications of relapse or viremia after a secondary infection with SARS-CoV-2 in a Chinese rhesus macaque model. These results support the idea that patients with full recovery (two negative RT-PCR results) may also be protected from secondary SARS-CoV-2 infection. Recovered patients may be able to reintegrate to normal public life and provide protective serum perhaps even if having had a mild infection. The results are also encouraging for successful vaccine development against SARS-CoV-2.

12.56.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.57 A highly conserved cryptic epitope in the receptor-binding domains of SARS-CoV-2 and SARS-CoV

[968]

12.57.1 Keywords

12.57.2 Main Findings

Given the sequence similarity of the surface spike glycoprotein (S) of SARS-CoV-2 and SAR-CoV, Yuan et al. (2020) propose that neutralizing antibodies isolated from convalescent SARS-CoV patients may offer insight into cross-reactive antibodies targeting SARS-CoV-2. In particular, they find that the receptor-binding domain (RBD) of SARS-CoV-2 S protein shares 86% sequence similarity with the RBD of SARS-CoV S protein that binds to the CR3022 neutralizing antibody. CR3022 also displays increased affinity for the “up” conformation of the SARS-CoV-2 S protein compared to the “down” conformation as it does for the SARS-CoV S protein. Therefore, the authors propose that this cross-reactive antibody may confer some degree of protection in vivo even if it fails to neutralize in vitro.

12.57.3 Limitations

Although the authors offer a logical rationale for identifying cross-reactive neutralizing antibodies derived from SARS-CoV, their study using only CR3022 failed to demonstrate whether this approach will be successful. After all, CR3022 failed to neutralize in vitro despite the binding affinity to a similar epitope on SARS-CoV-2. They would benefit from testing more candidates and using an in vivo model to demonstrate their claim that protection may be possible in the absence neutralization if combinations are used in vivo.

12.57.4 Significance

The ability to make use of previously characterized neutralizing antibodies for conserved epitopes can expedite drug design and treatment options.

12.57.5 Credit

This review was undertaken by Dan Fu Ruan, Evan Cody and Venu Pothula as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.58 Highly accurate and sensitive diagnostic detection of SARS-CoV-2 by digital PCR

Dong et al. medRxiv [254]

12.58.1 Keywords

12.58.2 Main Findings

The authors present a digital PCR (dPCR) diagnostic test for SARS-CoV-2 infection. In 103 individuals that were confirmed in a follow-up to be infected, the standard qPCR test had a positivity rate of 28.2% while the dPCR test detected 87.4% of the infections by detecting an additional 61 positive cases. The authors also tested samples from close contacts (early in infection stage) and convalescing individuals (late in infection stage) and were able to detect SARS-CoV-2 nucleic acid in many more samples using dPCR compared to qPCR.

12.58.3 Limitations

I did not detect limitations.

12.58.4 Significance

The authors make a strong case for the need for a highly sensitive and accurate confirmatory method for diagnosing COVID-19 during this outbreak and present a potential addition to the diagnostic arsenal. They propose a dPCR test that they present has a dramatically lower false negative rate than the standard RT-qPCR tests and can be especially beneficial in people with low viral load, whether they are in the earlier or later stages of infection.

12.58.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.59 SARS-CoV-2 invades host cells via a novel route: CD147-spike protein

Wang et al. bioRxiv [969]

12.59.1 Keywords

12.59.2 Main Findings

The authors propose a novel mechanism of SARS-CoV-2 viral entry through the interaction of the viral spike protein (SP) and the immunoglobulin superfamily protein CD147 (also known as Basigin). Using an in-house developed humanized antibody against CD147 (maplazumab), they show that blocking CD147 decreases viral replication in Vero E6 cells. Using surface plasmon resonance (SPR), ELISA, and Co-IP assays, they show that the spike protein of SARS-CoV-2 directly interacts with CD147. Lastly, they utilize immune-election microscopy to show spike protein and CD147 localize to viral inclusion bodies of Vero E6 cells.

12.59.3 Limitations

The authors claim that an anti-CD147 antibody (Meplazumab) inhibits SARS-CoV-2 replication by testing cell growth and viral load in cells infected with SARS-CoV-2, however there are key pieces of this experiment that are missing. First, the authors fail to use a non-specific antibody control. Second, the authors claim that viral replication is inhibited, and that they test this by qPCR, however this data is not shown. To further prove specificity, the authors should introduce CD147 to non-susceptible cells and show that they become permissive.

The authors claim that there is a direct interaction between CD147 and SP through SPR, ELISA, and Co-IP, and this data seems generally convincing. The electron microscopy provides further correlative evidence that SARS-CoV-2 may interact with CD147 as they are both found in the same viral inclusion body. A quantification of this data would make the findings more robust.

Finally, the data in this paper lacks replicates, error bars, and statistics to show that the data are reproducible and statistically significant.

12.59.4 Significance

It has been shown in various studies that SARS-CoV-2 binds to the cell surface protein ACE2 for cell entry, yet ACE2 is highly expressed in heart, kidney, and intestinal cells, raising the concern that blocking ACE2 would result in harmful side effects [970] CD147 on the other hand is highly expressed in various tumor types, inflamed tissues, and pathogen infected cells, suggesting that the inhibition of CD147 would not result in major side effects [971,972] The research in this paper has resulted in an ongoing clinical trail in China to test the safety and efficacy of anti-CD147 Meplazumab to treat COVID-19. (ClinicalTrails.gov identifier NCT04275245).

12.59.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.60 Blood single cell immune profiling reveals that interferon-MAPK pathway mediated adaptive immune response for COVID-19

Huang et al. medRxiv [973]

12.60.1 Keywords

12.60.2 Main Findings

The authors performed single-cell RNA sequencing (scRNAseq) of peripheral blood mononuclear cells isolated from whole blood samples of COVID-19 patients (n=10). Data was compared to scRNAseq of samples collected from patients with influenza A (n=1), acute pharyngitis (n=1), and cerebral infarction (n=1), as well as, three healthy controls. COVID-19 patients were categorized into those with moderate (n=6), severe (n=1), critical (n=1), and cured (n=2) disease. Analysis across all COVID-19 disease levels revealed 56 different cellular subtypes, among 17 immune cell types; comparisons between each category to the normal controls revealed increased proportions of CD1c+ dendritic cells, CD8+ CTLs, and plasmacytoid dendritic cells and a decrease in proportions of B cells and CD4+ T cells.

TCR sequencing revealed that greater clonality is associated with milder COVID-19 disease; BCR sequencing revealed that COVID-19 patients have circulating antibodies against known viral antigens, including EBV, HIV, influenza A, and other RNA viruses. This may suggest that the immune response to SARS-CoV-2 infection elicits production of antibodies against known RNA viruses.

Excluding enriched pathways shared by COVID-19 patients and patients with other conditions (influenza A, acute pharyngitis, and cerebral infarction), the authors identified the interferon-MAPK signaling pathway as a major response to SARS-CoV-2 infection. The authors performed quantitative real-time reverse transcriptase polymerase chain reaction (RT-PCR) for interferon-MAPK signaling genes: IRF27, BST2, and FOS. These samples were collected from a separate cohort of COVID-19 patients (critical, n=3; severe, n=3; moderate, n=19; mild, n=3; and cured, n=10; and healthy controls, n=5). Notably, consistent with the original scRNAseq data, FOS showed up-regulation in COVID-19 patients and down-regulation in cured patients. The authors propose that FOS may be a candidate marker gene for curative COVID-19 disease.

12.60.3 Limitations

The sample size of this study is limited. To further delineate differences in the immune profile of peripheral blood of COVID-19 patients, a greater sample size is needed, and longitudinal samples are needed, as well. A better understanding of the immunological interactions in cured patients, for example, would require a profile before and after improvement.

Moreover, the conclusions drawn from this scRNAseq study point to potential autoimmunity and immune deficiency to distinguish different severities of COVID-19 disease. However, this requires an expanded number of samples and a more robust organization of specific immune cell subtypes that can be compared across different patients. Importantly, this criterion is likely needed to ensure greater specificity in identifying markers for COVID-19 infection and subsequent immune response.

12.60.4 Significance

At the single-cell level, COVID-19 disease has been characterized in the lung, but a greater understanding of systemic immunological responses is furthered in this study. Type I interferon is an important signaling molecule for the anti-viral response. The identification of the interferon-MAPK signaling pathway and the differential expression of MAPK regulators between patients of differing COVID-19 severity and compared to cured patients may underscore the importance of either immune deficiency or autoimmunity in COVID-19 disease.

12.60.5 Credit

This review was undertaken by Matthew D. Park as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.61 Cross-reactive antibody response between SARS-CoV-2 and SARS-CoV infection.

Lv et al. bioRxiv [974]

12.61.1 Keywords

SARS-CoV-2, SARS-CoV, spike protein, RBD, cross-reactivity, cross-neutralization, antibody, human patients, mouse

12.61.2 Main Findings

The authors explore the antigenic differences between SARS-CoV-2 and SARS-CoV by analyzing plasma samples from SARS-CoV-2 (n = 15) and SARS-CoV (n = 7) patients. Cross-reactivity in antibody binding to the spike protein between SARS-CoV-2 and SARS-CoV was found to be common, mostly targeting non-RBD regions in plasma from SARS-CoV-2 patients. Only one SARS-CoV-2 plasma sample was able to cross-neutralize SARS-CoV, with low neutralization activity. No cross-neutralization response was detected in plasma from SARS-CoV patients.

To further investigate the cross-reactivity of antibody responses to SARS-CoV-2 and SARS-CoV, the authors analyzed the antibody response of plasma collected from mice infected or immunized with SARS-CoV-2 or SARS-CoV (n = 5 or 6 per group). Plasma from mice immunized with SARS-CoV-2 displayed cross-reactive responses to SARS-CoV S ectodomain and, to a lesser extent, SARS-CoV RBD. Similarly, plasma from mice immunized with SARS-CoV displayed cross-reactive responses to SARS-CoV-2 S ectodomain. Cross-neutralization activity was not detected in any of the mouse plasma samples.

12.61.3 Limitations

The size of each patient cohort is insufficient to accurately determine the frequency of cross-reactivity and cross-neutralization in the current SARS-CoV-2 pandemic. Recruitment of additional patients from a larger range of geographical regions and time points would also enable exploration into the effect of the genetic diversity and evolution of the SARS-CoV-2 virus on cross-reactivity. This work would also benefit from the mapping of specific epitopes for each sample. Future studies may determine whether the non-neutralizing antibody responses can confer in vitro protection or lead to antibody-dependent disease enhancement.

12.61.4 Significance

The cross-reactive antibody responses to S protein in the majority of SARS-CoV-2 patients is an important consideration for development of serological assays and vaccine development during the current outbreak. The limited extent of cross-neutralization demonstrated in this study indicates that vaccinating to cross-reactive conserved epitopes may have limited efficacy, presenting a key concern for the development of a more universal coronavirus vaccine to address the global health risk of novel coronavirus outbreaks.

12.61.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.62 The feasibility of convalescent plasma therapy in severe COVID-19 patients: a pilot study

Duan et al. medRxiv [975]

12.62.1 Keywords

12.62.2 Main Findings

This is the first report to date of convalescent plasma therapy as a therapeutic against COVID-19 disease. This is a feasibility pilot study. The authors report the administration and clinical benefit of 200 mL of convalescent plasma (CP) (1:640 titer) derived from recently cured donors (CP selected among 40 donors based on high neutralizing titer and ABO compatibility) to 10 severe COVID-19 patients with confirmed viremia. The primary endpoint was the safety of CP transfusion. The secondary endpoint were clinical signs of improvement based on symptoms and laboratory parameters.

The authors reported use of methylene blue photochemistry to inactivate any potential residual virus in the plasma samples, without compromising neutralizing antibodies, and no virus was detected before transfusion.

The authors report the following:

12.62.3 Limitations

It is important to note that most recipients had high neutralization titers of antibodies before plasma transfusion and even without transfusion it would be expected to see an increase in neutralizing antibodies over time. In addition to the small sample set number (n=10), there are additional limitations to this pilot study:

  1. All patients received concurrent therapy, in addition to the CP transfusion. Therefore, it is unclear whether a combinatorial or synergistic effect between these standards of care and CP transfusion contributed to the clearance of viremia and improvement of symptoms in these COVID-19 patients.

  2. The kinetics of viral clearance was not investigated, with respect to the administration of CP transfusion. So, the definitive impact of CP transfusion on immune dynamics and subsequent viral load is not well defined.

  3. Comparison with a small historical control group is not ideal.

12.62.4 Significance

For the first time, a pilot study provides promising results involving the use of convalescent plasma from cured COVID-19 patients to treat others with more severe disease. The authors report that the administration of a single, high-dose of neutralizing antibodies is safe. In addition, there were encouraging results with regards to the reduction of viral load and improvement of clinical outcomes. It is, therefore, necessary to expand this type of study with more participants, in order to determine optimal dose and treatment kinetics. It is important to note that CP has been studied to treat H1N1 influenza, SARS-CoV-1, and MERS-CoV, although it has not been proven to be effective in treating these infections.

12.62.5 Credit

Review by Matthew D. Park and revised by Alice O. Kamphorst and Maria A. Curotto de Lafaille as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.63 Hydroxychloroquine and azithromycin as a treatment of COVID-19: results of an open label non-randomized clinical trial

[976]

12.63.1 Keywords

12.63.2 Main Findings

This study was a single-arm, open label clinical trial with 600 mg hydroxychloroquine (HCQ) in the treatment arm (n = 20). Patients who refused participation or patients from another center not treated with HCQ were included as negative controls (n = 16). Among the patients in the treatment arm, 6 received concomitant azithromycin to prevent superimposed bacterial infection. The primary endpoint was respiratory viral loads on day 6 post enrollment, measured by nasopharyngeal swab followed by real-time reverse transcription-PCR.

HCQ alone was able to significantly reduce viral loads by day 6 (n = 8/14, 57.1% complete clearance, p < 0.001); azithromycin appears to be synergistic with HCQ, as 6/6 patients receiving combined treatment had complete viral clearance (p < 0.001).

12.63.3 Limitations

Despite what is outlined above, this study has a number of limitations that must be considered. First, there were originally n = 26 patients in the treatment arm, with 6 lost to follow up for the following reasons: 3 transferred to ICU, 1 discharge, 1 self-discontinued treatment d/t side effects, and 1 patient expired. Total length of clinical follow up was 14 days, but the data beyond day 6 post-inclusion are not shown.

Strikingly, in supplementary table 1, results of the real-time RT-PCR are listed for the control and treatment arms from D0 – D6. However, the data are not reported in a standard way, with a mix of broadly positive or negative result delineation with Ct (cycle threshold) values, the standard output of real time PCR. It is impossible to compare what is defined as a positive value between the patients in the control and treatment arms without a standardized threshold for a positive test. Further, the starting viral loads reported at D0 in the groups receiving HCQ or HCQ + azithromycin were significantly different (ct of 25.3 vs 26.8 respectively), which could explain in part the differences observed in the response to treatment between 2 groups. Finally, patients in the control arm from outside the primary medical center in this study (Marseille) did not actually have samples tested by PCR daily. Instead, positive test results from every other day were extrapolated to mean positive results on the day before and after testing as well (Table 2, footnote a).

Taken together, the results of this study suggest that HCQ represents a promising treatment avenue for COVID-19 patients. However, the limited size of the trial, and the way in which the results were reported does not allow for other medical centers to extrapolate a positive or negative result in the treatment of their own patients with HCQ +/- azithromycin. Further larger randomized clinical trials will be required to ascertain the efficacy of HCQ +/- azithromycin in the treatment of COVID-19.

12.63.4 Significance

Chloroquine is thought to inhibit viral infection, including SARS-Cov-2, by increasing pH within endosomes and lysosomes, altering the biochemical conditions required for viral fusion [321,977]. However, chloroquine also has immuno-modulatory effects that I think may play a role. Chloroquine has been shown to increase CTLA-4 expression at the cell surface by decreasing its degradation in the endo-lysosome pathway; AP-1 traffics the cytoplasmic tail of CTLA-4 to lysosomes, but in conditions of increased pH, the protein machinery required for degradation is less functional [978]. As such, more CTLA-4 remains in endosomes and is trafficked back to the cell surface. It is possible that this may also contribute to patient recovery via reduction of cytokine storm, in addition to the direct anti-viral effects of HCQ.

12.63.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.64 Recapitulation of SARS-CoV-2 Infection and Cholangiocyte Damage with Human Liver Organoids

[902]

12.64.1 Keywords

12.64.2 Main Findings

12.64.3 Limitations:

12.64.4 Significance

12.64.5 Credit

Review by Samarth Hegde as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.65 The sequence of human ACE2 is suboptimal for binding the S spike protein of SARS coronavirus 2

[979]

12.65.1 Keywords

12.65.2 Main Findings

Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) infects cells through S spike glycoprotein binding angiotensin-converting enzyme (ACE2) on host cells. S protein can bind both membrane-bound ACE2 and soluble ACE2 (sACE2), which can serve as a decoy that neutralizes infection. Recombinant sACE2 is now being tested in clinical trials for COVID-19. To determine if a therapeutic sACE2 with higher affinity for S protein could be designed, authors generated a library containing every amino acid substitution possible at the 117 sites spanning the binding interface with S protein. The ACE2 library was expressed in human Expi293F cells and cells were incubated with medium containing the receptor binding domain (RBD) of SARS-CoV-2 fused to GFP. Cells with high or low affinity mutant ACE2 receptor compared to affinity of wild type ACE2 for the RBD were FACS sorted and transcripts from these sorted populations were deep sequenced. Deep mutagenesis identified numerous mutations in ACE2 that enhance RBD binding. This work serves to identify putative high affinity ACE2 therapeutics for the treatment of CoV-2.

12.65.3 Limitations

The authors generated a large library of mutated ACE2, expressed them in human Expi293F cells, and performed deep mutagenesis to identify enhanced binders for the RBD of SARS-CoV-2 S protein. While these data serve as a useful resource, the ability of the high affinity ACE2 mutants identified to serve as therapeutics needs further validation in terms of conformational stability when purified as well as efficacy/safety both in vitro and in vivo. Additionally, authors mentioned fusing the therapeutic ACE2 to Fc receptors to elicit beneficial host immune responses, which would need further design and validation.

12.65.4 Significance

This study identified structural ACE2 mutants that have potential to serve as therapeutics in the treatment of SARS-CoV-2 upon further testing and validation.

12.65.5 Credit

This review was undertaken by Katherine Lindblad and Tamar Plitt as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

Title:  A serological assay to detect SARS-Cov-2 seroconversion in humans

Immunology keywords:  specific serological assay - ELISA - seroconversion - antibody titers

Note: the authors of this review work in the same institution as the authors of the study

Main findings:  

Production of recombinant whole Spike (S) protein and the smaller Receptor Binding Domain (RBD) based on the sequence of Wuhan-Hu-1 SARS-CoV-2 isolate. The S protein was modified to allow trimerization and increase stability. The authors compared the antibody reactivity of 59 banked human serum samples (non-exposed) and 3 serum samples from confirmed SARS-CoV-2 infected patients. All Covid-19 patient sera reacted to the S protein and RBD domain compared to the control sera.

The authors also characterized the antibody isotypes from the Covid-19 patients, and observed stronger IgG3 response than IgG1. IgM and IgA responses were also prevalent.

Limitations of the study:  

The authors analyzed a total of 59 control human serum samples, and samples from only three different patients to test for reactivity against the RBD domain and full-length spike protein. It will be important to follow up with a larger number of patient samples to confirm the data obtained. Furthermore, it would be interesting to assess people at different age groups and determine whether unexposed control kids have a higher “background”.

Applications of the assay described in this study in diagnosis are limited, since antibody response should start to be detectable only one to two weeks after infection. Future studies will be required to assess how long after infection this assay allow to detect anti-CoV2 antibodies. Finally, while likely, the association of seroconversion with protective immunity against SARS-Cov-2 infection still needs to be fully established.

Relevance:

This study has strong implications in the research against SARS-Cov-2. First, it is now possible to perform serosurveys and determine who has been infected, allowing a more accurate estimate of infection prevalence and death rate. Second, if it is confirmed that re-infection does not happen (or is rare), this assay can be used as a tool to screen healthcare workers and prioritize immune ones to work with infected patients. Third, potential convalescent plasma donors can now be screened to help treating currently infected patients. Of note, this assay does not involve live virus handling. experimentally, this is an advantage as the assay does not require the precautions required by manipulation of live virus. Finally, the recombinant proteins described in this study represent new tools that can be used for further applications, including vaccine development.

12.66 COMPARATIVE PATHOGENESIS OF COVID-19, MERS AND SARS IN A NON-HUMAN PRIMATE MODEL

[980]

12.66.1 Keywords

12.66.2 Main Findings

This work assesses SARS-CoV-2 infection in young adult and aged cynomolgus macaques. 4 macaques per age group were infected with low-passage clinical sample of SARS-CoV-2 by intranasal and intratracheal administration. Viral presence was assessed in nose, throat and rectum through RT-PCR and viral culture. SARS-CoV-2 replication was confirmed in the respiratory track (including nasal samples), and it was also detected in ileum. Viral nucleocapsid detection by IHC showed infection of type I and II pneumocytes and epithelia. Virus was found to peak between 2 and 4 days after administration and reached higher levels in aged vs. young animals. The early peak is consistent with data in patients and contrasts to SARS-CoV replication. SARS-CoV-2 reached levels below detection between 8 and 21 days after inoculation and macaques established antibody immunity against the virus by day 14. There were histopathological alteration in lung, but no overt clinical signs. At day 4 post inoculation of SARS-CoV-2, two of four animals presented foci of pulmonary consolidation, with limited areas of alveolar edema and pneumonia, as well as immune cell infiltration. In sum, cynomolgus macaques are permissive to SARS- CoV-2 and develop lung pathology (less severe than SARC-CoV, but more severe than MERS-CoV).

12.66.3 Limitations

Even though cynomolgus macaques were permissive to SARS-CoV-2 replication, it is unclear if the viral load reaches levels comparable to humans and there wasn’t overt clinical pathology.

12.66.4 Significance

The development of platforms in which to carry out relevant experimentation on SARS-CoV-2 pathophysiology is of great urgency. Cynomolgus macaques offer an environment in which viral replication can happen, albeit in a limited way and without translating into clinically relevant symptoms. Other groups are contributing to SARS-CoV2 literature using this animal model [967], potentially showing protection against reinfection in cured macaques. Therefore, this platform could be used to examine SARS-CoV2 pathophysiology while studies in other animal models are also underway [904,981].

12.66.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.67 Investigating the Impact of Asymptomatic Carriers on COVID-19 Transmission

[982]

12.67.1 Keywords

12.67.2 Main Findings

Multiple studies reported the same level of infectiousness between symptomatic and asymptomatic carriers of SARS-CoV-2. Given that asymptomatic and undocumented carriers escape public health surveillance systems, a better mathematical model of transmission is needed to determine a more accurate estimate of the basic reproductive number (R0) of the virus to assess the contagiousness of virus. The authors developed a SEYAR dynamical model for transmission of the new coronavirus that takes into account asymptomatic and undocumented carriers. The model was validated using data reported from thirteen countries during the first three weeks of community transmission. While current studies estimate R0 to be around 3, this model indicates that the value could range between 5.5 to 25.4.

12.67.3 Limitations

The SEYAR model realistically depicts transmission of the virus only during the initial stages of the disease. More data is necessary to better fit the model with current trends. In addition, multiple factors (e.g. behavioral patterns, surveillance capabilities, environmental and socioeconomic factors) affect transmission of the virus and so, these factors must be taken into consideration when estimating the R0.

12.67.4 Significance

Public health authorities use the basic reproductive number to determine the severity of disease. An accurate estimate of R0 will inform intervention strategies. This model can be applied to different locations to assess the potential impact of COVID-19.

12.67.5 Credit

This review was undertaken by Tamar Plitt and Katherine Lindblad as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.68 Antibody responses to SARS-CoV-2 in COVID-19 patients: the perspective application of serological tests in clinical practice

Long et al. medRxiv [983]

12.68.1 Keywords

12.68.2 Main Findings

This study investigated the profile of the acute antibody response against SARS-CoV-2 and provided proposals for serologic tests in clinical practice. Magnetic Chemiluminescence Enzyme Immunoassay was used to evaluate IgM and IgG seroconversion in 285 hospital admitted patients who tested positive for SARS-CoV-2 by RT-PCR and in 52 COVID-19 suspected patients that tested negative by RT-PCR. A follow up study with 63 patients was performed to investigate longitudinal effects. In addition, IgG and IgM titers were evaluated in a cohort of close contacts (164 persons) of an infected couple.

The median day of seroconversion for both IgG and IgM was 13 days after symptom onset. Patients varied in the order of IgM/ IgG seroconversion and there was no apparent correlation of order with age, severity, or hospitalization time. This led the authors to conclude that for diagnosis IgM and IgG should be detected simultaneously at the early phase of infection.

IgG titers, but not IgM titers were higher in severe patients compared to non-severe patients after controlling for days post-symptom onset. Importantly, 12% of COVID-19 patients (RT-PCR confirmed) did not meet the WHO serological diagnosis criterion of either seroconversion or > 4-fold increase in IgG titer in sequential samples. This suggests the current serological criteria may be too stringent for COVID-19 diagnosis.

Of note, 4 patients from a group of 52 suspects (negative RT-PCR test) had anti-SARS-Cov-2 IgM and IgG. Similarly, 4.3% (7/162) of “close contacts” who had negative RT-PCR tests were positive for IgG and/or IgM. This highlights the usefulness of a serological assay to identify asymptomatic infections and/or infections that are missed by RT-PCR.

12.68.3 Limitations

This group’s report generally confirms the findings of others that have evaluated the acute antibody response to SARS-Cov-2. However, these data would benefit from inclusion of data on whether the participants had a documented history of viral infection. Moreover, serum samples that were collected prior to SARS-Cov-2 outbreak from patients with other viral infections would serve as a useful negative control for their assay. Methodological limitations include that only one serum sample per case was tested as well as the heat inactivation of serum samples prior to testing. It has previously been reported that heat inactivation interferes with the level of antibodies to SARS-Cov-2 and their protocol may have resulted in diminished quantification of IgM, specifically [984].

12.68.4 Significance

Understanding the features of the antibody responses against SARS-CoV is useful in the development of a serological test for the diagnosis of COVID-19. This paper addresses the need for additional screening methods that can detect the presence of infection despite lower viral titers. Detecting the production of antibodies, especially IgM, which are produced rapidly after infection can be combined with PCR to enhance detection sensitivity and accuracy and map the full spread of infection in communities, Moreover, serologic assays would be useful to screen health care workers in order to identify those with immunity to care for patients with COVID19.

12.68.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.69 SARS-CoV-2 specific antibody responses in COVID-19 patients

[985]

12.69.1 Keywords

12.69.2 Main findings

Antibodies specific to SARS-CoV-2 S protein, the S1 subunit and the RBD (receptor-binding domain) were detected in all SARS-CoV-2 patient sera by 13 to 21 days post onset of disease. Antibodies specific to SARS-CoV N protein (90% similarity to SARS-CoV-2) were able to neutralize SARS-CoV-2 by PRNT (plaque reduction neutralizing test). SARS-CoV-2 serum cross-reacted with SARS-CoV S and S1 proteins, and to a lower extent with MERS-CoV S protein, but not with the MERS-CoV S1 protein, consistent with an analysis of genetic similarity. No reactivity to SARS-CoV-2 antigens was observed in serum from patients with ubiquitous human CoV infections (common cold) or to non-CoV viral respiratory infections.

12.69.3 Limitations

Authors describe development of a serological ELISA based assay for the detection of neutralizing antibodies towards regions of the spike and nucleocapsid domains of the SARS-CoV-2 virus. Serum samples were obtained from PCR-confirmed COVID-19 patients. Negative control samples include a cohort of patients with confirmed recent exposure to non-CoV infections (i.e. adenovirus, bocavirus, enterovirus, influenza, RSV, CMV, EBV) as well as a cohort of patients with confirmed infections with ubiquitous human CoV infections known to cause the common cold. The study also included serum from patients with previous MERS-CoV and SARS-CoV zoonotic infections. This impressive patient cohort allowed the authors to determine the sensitivity and specificity of the development of their in-house ELISA assay. Of note, seroconversion was observed as early as 13 days following COVID-19 onset but the authors were not clear how disease onset was determined.

12.69.4 Significance

Validated serological tests are urgently needed to map the full spread of SARS-CoV-2 in the population and to determine the kinetics of the antibody response to SARS-CoV-2. Furthermore, clinical trials are ongoing using plasma from patients who have recovered from SARS-CoV-2 as a therapeutic option. An assay such as the one described in this study could be used to screen for strong antibody responses in recovered patients. Furthermore, the assay could be used to screen health care workers for antibody responses to SARS-CoV-2 as personal protective equipment continues to dwindle. The challenge going forward will be to standardize and scale-up the various in-house ELISA’s being developed in independent laboratories across the world.

12.70 A brief review of antiviral drugs evaluated in registered clinical trials for COVID-19

Belhadi et al. [986]

12.70.1 Keywords

12.70.2 Main Findings

Summary of clinical trials registered as of March7, 2020 from U.S, Chinese, Korean, Iranian and European registries. Out of the 353 studies identified, 115 were selected for data extraction. 80% of the trials were randomized with parallel assignment and the median number of planned inclusions was 63 (IRQ, 36-120). Most frequent therapies in the trials included; 1) antiviral drugs [lopinavir/ritonavir (n-15); umifenovir (n=9); favipiravir (n=7); redmesivir (n=5)]; 2) anti-malaria drugs [chloroquine (n-11); hydroxychloroquine (n=7)}; immunosuppressant drugs [methylprednisolone (n=5)]; and stem cell therapies (n=23). Medians of the total number of planned inclusions per trial for these therapies were also included. Stem cells and lopunavir/ritonavir were the most frequently evaluated candidate therapies (23 and 15 trials respectively), whereas remdesivir was only tested in 5 trials but these trials had the highest median number of planned inclusions per trial (400, IQR 394-453). Most of the agents used in the different trials were chosen based on preclinical assessments of antiviral activity against SARS CoV and MERS Cov corona viruses.

The primary outcomes of the studies were clinical (66%); virological (23%); radiological (8%); or immunological (3%). The trials were classified as those that included patients with severe disease only; trials that included patients with moderate disease; and trials that included patients with severe or moderate disease.

12.70.3 Limitations

The trials evaluated provided incomplete information: 23% of these were phase IV trials but the bulk of the trials (54%) did not describe the phase of the study. Only 52% of the trials (n=60) reported treatment dose and only 34% (n=39) reported the duration. A lot of the trials included a small number of patients and the trials are still ongoing, therefore no insight was provided on the outcome of the trials.

12.70.4 Significance

Nonetheless, this review serves as framework for identifying COVID-19 related trials, which can be expanded upon as new trials begin at an accelerated rate as the disease spreads around the world.

12.70.5 Credit

This review was undertaken by K Alexandropoulos as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.71 ACE-2 Expression in the Small Airway Epithelia of Smokers and COPD Patients: Implications for COVID-19

Leung et al. medRxiv. [987]

12.71.1 Keywords

12.71.2 Main Findings

In bronchial epithelial samples from 3 different cohorts of individuals, ACE-2 gene expression was found to be significantly increased in both COPD patients and smokers relative to healthy controls. Across all test subjects, ACE-2 gene expression was also highly correlated with decreased forced expiratory volume in 1 second (FEV1), which may explain the increased COVID-19 disease severity in COPD patients. Former smokers were also found to show decreased ACE2 expression relative to current smokers and had no significant difference when compared to non-smokers.

12.71.3 Limitations

While the upregulation of ACE-2 is an interesting hypothesis for COVID-19 disease severity in COPD patients, this study leaves many more unanswered questions than it addresses. Further studies are required to show whether the specific cell type isolated in these studies is relevant to the pathophysiology of COVID-19. Furthermore, there is no attempt to show whether that increased ACE-2 expression contributes to greater disease severity. Does the increased ACE-2 expression lead to greater infectivity with SARS-CoV-2? There is no mechanistic explanation for why ACE-2 levels are increased in COPD patients. The authors could also have considered the impact of co-morbidities and interventions such as corticosteroids or bronchodilators on ACE-2 expression. Finally, given the extensive sequencing performed, the authors could have conducted significantly more in-depth analyses into gene signature differences.

12.71.4 Significance

This study attempts to address an important clinical finding that both smokers and COPD patients show increased mortality from COVID-19. The novel finding that ACE-2 expression is induced in smokers and COPD patients suggests not only a mechanism for the clinical observation, but also highlights the potential benefit of smoking cessation in reducing the risk of severe COVID-19 disease.

12.71.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.72 Dynamic profile of severe or critical COVID-19 cases

[988]

12.72.1 Keywords

12.72.2 Main Findings

Authors evaluate clinical correlates of 10 patients (6 male and 4 female) hospitalized for severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2). All patients required oxygen support and received broad spectrum antibiotics and 6 patients received anti-viral drugs. Additionally, 40% of patients were co-infected with influenza A. All 10 patients developed lymphopenia, two of which developed progressive lymphopenia (PLD) and died. Peripheral blood (PB) lymphocytes were analyzed – low CD4 and CD8 counts were noted in most patients, though CD4:CD8 ratio remained normal.

12.72.3 Limitations

The authors evaluated a small cohort of severe SARS-CoV-2 cases and found an association between T cell lymphopenia and adverse outcomes. However, this is an extremely small and diverse cohort (40% of patients were co-infected with influenza A). These findings need to be validated in a larger cohort. Additionally, the value of this data would be greatly increased by adding individual data points for each patient as well as by adding error bars to each of the figures.

12.72.4 Significance

This study provides a collection of clinical data and tracks evolution of T lymphocyte in 10 patients hospitalized for SARS-CoV-2, of which 4 patients were co-infected with influenza A.

12.72.5 Credit

This review was undertaken by Katherine Lindblad and Tamar Plitt as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.73 Association between Clinical, Laboratory and CT Characteristics and RT-PCR Results in the Follow-up of COVID-19 patients

Fu et al. medRxiv. [989]

12.73.1 Keywords

12.73.2 Study description

Data analyzed from 52 COVID-19 patients admitted and then discharged with COVID-19. Clinical, laboratory, and radiological data were longitudinally recorded with illness timecourse (PCR + to PCR-) and 7 patients (13.5%) were readmitted with a follow up positive test (PCR+) within two weeks of discharge.

12.73.3 Main Findings

12.73.4 Limitations

Patients sampled in this study were generally younger (65.4% < 50 yrs) and less critically ill/all discharged. Small number of recovered patients (N=18). Time of follow up was relatively short.. Limited clinical information available about patients with re-positive test (except CRP and lymph tracking).

12.73.5 Extended Results

NOTE: Patients sampled in this study were generally younger (65.4% < 50 yrs) and less critically ill/all discharged. After two consecutive negative PCR tests, patients were discharged.

Clinical Results at Admission

Change in Clinical Results following Negative Test

Patients Readmitted with PCR+ test

12.73.6 Significance

Study tracked key clinical features associated with disease progression, recovery, and determinants of clinical diagnosis/management of COVID-19 patients.

12.73.7 Credit

This review was undertaken by Natalie Vaninov as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.74 An orally bioavailable broad-spectrum antiviral inhibits SARS-CoV-2 and multiple 2 endemic, epidemic and bat coronavirus

Sheahan et al. bioRxiv. [990]

12.74.1 Keywords

12.74.2 Main Findings

β-D-N4 30 –hydroxycytidine (NHC, EIDD-1931) is an orally bioavailable ribonucleoside with antiviral activity against various RNA viruses including Ebola, Influenza and CoV. NHC activity introduceds mutations in the viral (but not cellular) RNA in a dose dependent manner that directly correlated with a decrease in viral titers. Authors show that NHC inhibited multiple genetically distinct Bat-CoV viruses in human primary epithelial cells without affecting cell viability even at high concentrations (100 µM). Prophylactic oral administration of NHC in C57BL/6 mice reduce lung titers of SARS-CoV and prevented weight loss and hemorrhage. Therapeutic administration of NHC in C57BL/6 mice 12 hours post infected with SARS-CoV reduced acute lung injury, viral titer, and lung hemorrhage. The degree of clinical benefit was dependent on the time of treatment initiation post infection. The authors also demonstrate that NHC reduces MERS-CoV infection titers, pathogenesis, and viral RNA in prophylactic and therapeutic settings.

12.74.3 Limitations

Most of the experiments were conducted using MERS-CoV, and SARS-CoV and a few experiments were conducted using other strains of CoV as opposed to SARS-CoV-2. The authors note the core residues that make up the RNA interaction sites (which constitutes the NHC interaction sites) are highly conserved among CoV and because of this conservation their understanding is that NHC can inhibit a broad-spectrum of CoV including SARS-CoV-2.

The increased viral mutation rates associated with NHC activity may have adverse effects if mutations cause the virus to become drug resistant, more infectious or speed-up immune evasion. In addition, the temporal diminishing effectiveness of NHC on clinical outcome when NHC was used therapeutically is concerning. However, the longer window (7-10 days) for clinical disease onset in human patients from the time of infection compared to that of mice (24-48 hours), may associate with increased NHC effectiveness in the clinic.

12.74.4 Significance

Prophylactic or therapeutic oral administration of NHC reduces lung titers and prevents acute lung failure in C57B\6 mice infected with CoV. Given its broad-spectrum antiviral activity, NHC could turn out to be a useful drug for treating current, emerging and future corona virus outbreaks.   #### Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.75 Identification of antiviral drug candidates against SARS-CoV-2 from FDA-approved drugs

Sangeun Jeon et al. [991]

12.75.1 Keywords

12.75.2 Main Findings

A panel of ~3,000 FDA- and IND-approved antiviral drugs were previously screened for inhibitory efficacy against SARS CoV, a coronavirus related to the novel coronavirus SARS CoV-2 (79.5%) homology. 35 of these drugs along with another 15 (suggested by infectious disease specialists) were tested in vitro for their ability to inhibit SARS CoV-2 infectivity of Vero cells while preserving cell viability. The infected cells were scored by immunofluorescence analysis using an antibody against the N protein of SARS CoV-2. Chloroquine, lopinavir and remdesivir were used as reference drugs.

Twenty four out of 50 drugs exhibited antiviral activity with IC50 values ranging from 0.1-10µM. Among these, two stood ou: 1) the-anti helminthic drug niclosamide which exhibited potent antiviral activity against SARS CoV-2 (IC50=0.28 µM). The broad-spectrum antiviral effect of niclosamide against SARS and MERS-CoV have been previously documented and recent evidence suggests that in may inhibit autophagy and reduce MERS C0V replication. 2) Ciclesonide, a corticosteroid used to treat asthma and allergic rhinitis, also exhibited antiviral efficacy but with a lower IC50 (4.33µM) compared to niclosamide. The antiviral effects of ciclesonide were directed against NSP15, a viral riboendonuclease which is the molecular target of this drug.

12.75.3 Limitations

The drugs were tested against SARS CoV-2 infectivity in vitro only, therefore preclinical studies in animals and clinical trials in patients will be need for validation of these drugs as therapeutic agents for COVID-19. In addition, niclosamide exhibits low adsorption pharmatokinetically which could be alleviated with further development of drug formulation to increase effective delivery of this drug to target tissues. Nonetheless, niclosamide and ciclesonide represent promising therapeutic agents against SARS CoV-2 given that other compounds tested in the same study including favipiravir (currently used in clinical trials) and atazanavir (predicted as the most potent antiviral drug by AI-inference modeling) did not exhibit antiviral activity in the current study.

12.75.4 Credit

This review was undertaken by K Alexandropoulos as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.76 Respiratory disease and virus shedding in rhesus macaques inoculated with SARS-CoV-2

Munster et al. bioRxiv. [992]

12.76.1 Keywords

animal model, pulmonary infiltrates, dynamic of antibody response, cytokine

12.76.2 Main Findings

Inoculation of 8 Resus macaques with SARS-Cov-2 , which all showed clinical signs of infection (respiratory pattern, reduced appetite, weight loss, elevated body temperature) resulting in moderate, transient disease. Four animals were euthanized at 3dpi, the 4 others at 21 dpi. Study of viral loads in different organs showed that nose swab and throat swabs were the most sensitive, with bronchio-alveolar lavage. Interstitial pneumonia was visible in radiographies and at the histological scale too. Clinically, the macaques had similar symptoms as described in human patients with moderate disease.

Viral shedding was consistently detected in nose swabs and throat swabs immediately after infection but less consistent thereafter which could reflect virus administration route (intranasal, oral). Bronchoalveolar lavages performed as a measure of virus replication in the lower respiratory tract on animals maintained for 21 days, contained high viral loads in 1 and 3dpi. The majority of the animals exhibited pulmonary edema and mild to moderate interstitial pneumonia on terminal bronchioles. In addition to the lung, viral RNA could also be detected throughout the respiratory track where viral replication mainly occurred.

Immunologic responses included leukocytosis, neutrophilia, monocytosis and lymphopenia in the majority of the animals at 1dpi. Lymphocytes and monocytes re-normalized at 2dpi. Neutrophils declined after 3dpi and through 10dpi after which they started to recover. After infection, serum analysis revealed significant increases in IL1ra, IL6, IL10, IL15, MCP-1, MIP-1b, but quick normalization (3dpi). Antibody response started around 7dpi, and the antibody titers stayed elevated until 21dpi (day of animal euthanasia).

12.76.3 Limitations

The macaques were inoculated via a combination of intratracheal, intranasal, ocular and oral routes, which might not reproduce how humans get infected. Maybe this can lead to different dynamics in the host immune response. Also, the authors noted that the seroconversion was not directly followed by a decline in viral loads, as observed in covid19 patients.

12.76.4 Significance

This work confirms that rhesus macaques can be a good model to study Covid-19, as it has been shown by other groups [967,980,993]. While these experiments recapitulate moderate COVID-19 in humans, the mode of inoculation via a combination of intratracheal, intranasal, ocular and oral routes, might not reproduce how humans get infected and may lead to different dynamics in the host immune response. For example, the authors noted that the seroconversion was not directly followed by a decline in viral loads, as observed in COVID-19 patients. Therefore, it will be interesting to follow their antibody titers longer and further assess the possibility/effect of reinfection in these macaques. It is essential to be able to understand the dynamic of the disease and associated immune responses, and to work on vaccine development and antiviral drug testing.

12.76.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.77 ACE2 Expression is Increased in the Lungs of Patients with Comorbidities Associated with Severe COVID-19

[994]

12.77.1 Keywords

12.77.2 Main Findings

12.77.3 Limitations:

12.77.4 Significance

12.77.5 Credit

Review by Samarth Hegde as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.78 Meplazumab treats COVID-19 pneumonia: an open-labelled, concurrent controlled add-on clinical trial

Bian et al. medRxiv. [995]

12.78.1 Keywords

12.78.2 Main Findings

This work is based on previous work by the same group that demonstrated that SARS-CoV-2can also enter host cells via CD147 (also called Basigin, part of the immunoglobulin superfamily, is expressed by many cell types) consistent with their previous work with SARS-CoV-1. 1 A prospective clinical trial was conducted with 17 patients receiving Meplazumab, a humanized anti-CD147 antibody, in addition to all other treatments. 11 patients were included as a control group (non-randomized).

They observed a faster overall improvement rate in the Meplazumab group (e.g. at day 14 47% vs 17% improvement rate) compared to the control patients. Also, virological clearance was more rapid with median of 3 days in the Meplazumab group vs 13 days in control group. In laboratory values, a faster normalization of lymphocyte counts in the Meplazumab group was observed, but no clear difference was observed for CRP levels.

12.78.3 Limitations

While the results from the study are encouraging, this study was non-randomized, open-label and on a small number of patients, all from the same hospital. It offers evidence to perform a larger scale study. Selection bias as well as differences between treatment groups (e.g. age 51yo vs 64yo) may have contributed to results. The authors mention that there was no toxic effect to Meplazumab injection but more patient and longer-term studies are necessary to assess this.

12.78.4 Significance

These results seem promising as for now there are limited treatments for Covid-19 patients, but a larger cohort of patient is needed. CD147 has already been described to facilitate HIV [996], measles virus [997], and malaria [998] entry into host cells. This group was the first to describe the CD147-spike route of SARS-Cov-2 entry in host cells [969] p147. Indeed, they had previously shown in 2005 that SARS-Cov could enter host cells via this transmembrane protein [999]. Further biological understanding of how SARS-CoV-2 can enter host cells and how this integrates with ACE2R route of entry is needed. Also, the specific cellular targets of the anti-CD147 antibody need to be assessed, as this protein can be expressed by many cell types and has been shown to involved in leukocytes aggregation [1000]. Lastly, Meplazumab is not a commercially-available drug and requires significant health resources to generate and administer which might prevent rapid development and use.

12.78.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.79 Potent human neutralizing antibodies elicited 1 by SARS-CoV-2 infection

Ju et al. bioRxiv. [150]

12.79.1 Keywords

12.79.2 Main Findings

In this study the authors report the affinity, cross reactivity (with SARS-CoV and MERS-CoV virus) and viral neutralization capacity of 206 monoclonal antibodies engineered from isolated IgG memory B cells of patients suffering from SARS-CoV-2 infection in Wuhan, China. All patients but one recovered from disease. Interestingly, the patient that did not recover had less SARS-CoV-2 specific B cells circulating compared to other patients.

Plasma from all patients reacted to trimeric Spike proteins from SARS-CoV-2, SARS-CoV and MERS-CoV but no HIV BG505 trimer. Furthermore, plasma from patients recognized the receptor binding domain (RBD) from SARS-CoV-2 but had little to no cross-reactivity against the RBD of related viruses SARS-CoV and MERS-CoV, suggesting significant differences between the RBDs of the different viruses. Negligible levels of cross-neutralization using pseudoviruses bearing Spike proteins of SARS-CoV-2, SARS-CoV or MERS-CoV, were observed, corroborating the ELISA cross-reactivity assays on the RBDs.

SARS-CoV-2 RBD specific B cells constituted 0.005-0.065% of the total B cell population and 0.023-0.329% of the memory subpopulation. SARS-CoV specific IgG memory B cells were single cell sorted to sequence the antibody genes that were subsequently expressed as recombinant IgG1 antibodies. From this library, 206 antibodies with different binding capacities were obtained. No discernible patterns of VH usage were found in the 206 antibodies suggesting immunologically distinct responses to the infection. Nevertheless, most high-binding antibodies were derived by clonal expansion. Further analyses in one of the patient derived clones, showed that the antibodies from three different timepoints did not group together in phylogenetic analysis, suggesting selection during early infection.

Using surface plasmon resonance (SPR) 13 antibodies were found to have 10-8 tp 10-9 dissociation constants (Kd). Of the 13 antibodies, two showed 98-99% blocking of SARS-CoV-2 RBD-ACE2 receptor binding in competition assays. Thus, low Kd values alone did not predict ACE2 competing capacities. Consistent with competition assays the two antibodies that show high ACE2 blocking (P2C-2F6 and P2C-1F11) were the most capable of neutralizing pseudoviruses bearing SARS-CoV-2 spike protein (IC50 of 0.06 and 0.03 µg/mL, respectively). Finally, using SPR the neutralizing antibodies were found to recognize both overlapping and distinct epitopes of the RBD of SARS-CoV-2.

12.79.3 Limitations

  1. Relatively low number of patients

    1. No significant conclusion can be drawn about the possible > correlation between humoral response and disease severity
  2. In vitro Cytopathic Effect Assay (CPE) for neutralization activity

    1. Huh7 cells were used in neutralization assays with > pseudoviruses, and they may not entirely mimic what happens in > the upper respiratory tract

    2. CPE assay is not quantitative

  3. Duplicated panel in Figure 4C reported (has been fixed in version 2)

12.79.4 Significance

This paper offers an explanation as to why previously isolated antibodies against SARS-CoV do not effectively block SARS-CoV-2. Also, it offers important insight into the development of humoral responses at various time points during the first weeks of the disease in small but clinically diverse group of patients. Furthermore, it provides valuable information and well characterized antibody candidates for the development of a recombinant antibody treatment for SARS-CoV-2. Nevertheless, it also shows that plasmapheresis might have variability in its effectiveness, depending on the donor’s antibody repertoire at the time of donation.

12.79.5 Credit

Review by Jovani Catalan-Dibene as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.80 Characterisation of the transcriptome and proteome of SARS-CoV-2 using direct RNA sequencing and tandem mass spectrometry reveals evidence for a cell passage induced in-frame deletion in the spike glycoprotein that removes the furin-like cleavage site.

Davidson et al. [1001]

12.80.1 Keywords

12.80.2 Main Findings

The authors performed long read RNA sequencing using an Oxford Nanopore MinION as well as tandem mass spec (MS) on Vero cells (a cell line derived from kidney cells of the African green monkey that is deficient in interferon) infected with SARS-CoV-2.

The authors found that passage of the virus in Vero cells gave rise to a spontaneous 9 amino acid deletion (679-NSPRRARSV-687 to I) in the spike (S) protein. The deleted sequence overlaps a predicted furin cleavage site at the S1 / S2 domain boundary that is present in SARS-CoV-2 but not SARS-CoV or the closely related bat coronavirus RaTG13, which are cleaved at S1 / S2 by other proteases [51]. Furin cleavage sites at similar positions in other viruses have been linked to increased pathogenicity and greater cell tropism [1002]. Loss of this site in SARS-CoV-2 has also already been shown to increase viral entry into Vero but not BHK cells (which are also interferon deficient) [63]. The authors therefore make an important contribution in demonstrating that passage in Vero cells may lead to spontaneous loss of a key pathogenicity-conferring element in SARS-CoV-2.

12.80.3 Limitations

As the authors note, a similar study posted earlier by Kim et al., which also passaged SARS-CoV-2 in Vero cells, did not identify any loss in the S protein furin cleavage site [1003]. It therefore remains to be determined how likely it is that this mutation spontaneously arises. A quantitative investigation using multiple experimental replicas to understand the spontaneous viral mutation rate at this site and elsewhere would be informative. Also, the mechanistic basis for the higher viral fitness conferred by loss of the furin cleavage site in Vero cells – but, evidently, not in vivo in humans, as this site is maintained in all currently sequenced circulating isolates - remains to be understood.

Due to the high base-call error rate of MinION sequencing, the authors’ bioinformatic pipeline required aligning transcripts to a reference to correct sequencing artifacts. This presumably made it difficult or impossible to identify other kinds of mutations, such as single nucleotide substitutions, which may occur even more frequently than the deletions identified in this work. Pairing long read sequencing with higher-accuracy short-read sequencing may be one approach to overcome this issue.

12.80.4 Significance

As the authors suggest, animal studies using live virus challenge may need to periodically verify the genomic integrity of the virus, or potentially risk unknowingly using a likely less-pathogenic variant of the virus.

More broadly, the results emphasize the complexity and plasticity of the SARS-CoV-2 viral transcriptome and proteome. For example, the authors found multiple versions of transcripts encoding the nucleocapsid (N) protein, each with different small internal deletions, some of which were verified for translation by MS. A number of peptides arising from translation of unexpected rearrangements of transcripts were also detected. Additionally, the authors identified phosphorylation of a number of viral proteins (N, M, ORF 3a, nsp3, nsp9, nsp12 and S). For any cases where these have functional consequences, targeting the kinases responsible could be an avenue for drug development. Understanding the functional consequences of the mutations, transcript variations, and post translational modifications identified in this study will be important future work.

12.80.5 Credit

This review was undertaken by Tim O’Donnell, Maria Kuksin as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.81 A SARS-CoV-2-Human Protein-Protein Interaction Map Reveals Drug Targets and Potential Drug- Repurposing

Gordon et al. bioRxiv [153]

12.81.1 Keywords

12.81.2 Main Findings

Gordon et al cloned, tagged and expressed 26 of the 29 SARS-CoV-2 proteins individually in HEK293T cells and used mass spectrometry to identify protein-protein interactions. They identified 332 viral-host protein-protein interactions. Furthermore, they used these interactions to identify 66 existing drugs known to target host proteins or host pathways (eg SARS-CoV-2 N and Orf8 proteins interact with proteins regulated by the mTOR pathway, so mTOR inhibitors Silmitasertib and Rapamycin are possible drug candidates).

12.81.3 Limitations

The main limitation of the study stems from the reductionist model: overexpression of plasmids encoding individual viral proteins in HEK293T cells. This precludes any interactions between the viral proteins, or the combined effects of multiple proteins on the host, as they are expressed individually. Moreover, HEK293T cells come from primary embryonic kidney and therefore might not reflect how SARS-CoV-2 interacts with its primary target, the lung. However, the authors found that the proteins found to interact with viral proteins in their experiments are enriched in lung tissue compared to HEK293Ts.

12.81.4 Significance

The authors provide a “SARS-CoV-2 interaction map,” which may provide potential hypotheses as to how the virus interacts with the host. Further, they identified existing drugs that could disrupt these host-viral interactions and curb SARS-CoV-2 infection. Although these interactions have not been validated, this paper acts as a valuable resource.

12.81.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.82 First Clinical Study Using HCV Protease Inhibitor Danoprevir to Treat Naïve and Experienced COVID-19 Patients

Chen et al. medRxiv. [1004]

12.82.1 Keywords

12.82.2 Main Findings

The authors treated 11 Covid-19 patients with Danoprevir, a commercialized HCV protease inhibitor [1005](p4), boosted by ritonavir [1006], a CYP3A4 inhibitor (which enhances the plasma concentration and bioavailabilty of Danoprevir). Two patients had never received anti-viral therapy before (=naïve), whereas nine patients were on Lopinavir/Ritonavir treatment before switching to Danoprevir/Ritonavir (=experienced). The age ranged from 18 to 66yo.

Naïve patients that received Danoprevir/Ritonavir treatment had a decreased hospitalization time. Patients treated with Lopinavir/Ritonavir did not have a negative PCR test, while after switching to Danoprevir/Ritonavir treatment, the first negative PCR test occurred at a median of two days.

12.82.3 Limitations

The results of the study are very hard to interpret as there is no control group not receiving Danoprevir/Ritonavir treatment. This was especially true in naïve patients who seemed to have more mild symptoms before the start of the study and were younger (18 and 44yo) compared to the experienced patients (18 to 66yo). The possibility that the patients would have recovered without Danoprevir/Ritonavir treatment cannot be excluded.

12.82.4 Significance

The authors of the study treated patients with Danoprevir, with the rational to that this is an approved and well tolerated drug for HCV patients [1006], and that it could also target the protease from SARS-CoV-2 (essential for viral replication and transcription). Indeed, homology modelling data indicated that HCV protease inhibitors have the highest binding affinity to Sars-Cov2 protease among other approved drugs [1007].

While this study shows that the combination of Danoprevir and Ritonavir might be beneficial for Covid-19 patients, additional clinical trials with more patients and with better methodology (randomization and control group) are needed to make further conclusions.

12.82.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.83 Efficacy of hydroxychloroquine in patients with COVID-19: results of a randomized clinical trial

[379]

12.83.1 Keywords

12.83.2 Study Description

This is a randomized clinical trial of hydroxychloroquine (HCQ) efficacy in the treatment of COVID-19. From February 4 – February 28, 2020 142 COVID-19 positive patients were admitted to Renmin Hospital of Wuhan University. 62 patients met inclusion criteria and were enrolled in a double blind, randomized control trial, with 31 patients in each arm.

Inclusion criteria:

  1. Age ≥ 18 years

  2. Positive diagnosis COVID-19 by detection of SARS-CoV-2 by RT-PCR

  3. Diagnosis of pneumonia on chest CT

  4. Mild respiratory illness, defined by SaO2/SPO2 ratio > 93% or PaO2/FIO2 ratio > 300 mmHg in hospital room conditions (Note: relevant clinical references described below.)

    1. Hypoxia is defined as an SpO2 of 85-94%; severe hypoxia < 85%.

    2. The PaO2/FIO2 (ratio of arterial oxygen tension to fraction of inspired oxygen) is used to classify the severity of acute respiratory distress syndrome (ARDS). Mild ARDS has a PaO2/FIO2 of 200-300 mmHg, moderate is 100-200, and severe < 100.

  5. Willing to receive a random assignment to any designated treatment group; not participating in another study at the same time

Exclusion criteria:

  1. Severe or critical respiratory illness (not explicitly defined, presumed to be respiratory function worse than outlined in inclusion criteria); or participation in trial does not meet patient’s maximum benefit or safe follow up criteria

  2. Retinopathy or other retinal diseases

  3. Conduction block or other arrhythmias

  4. Severe liver disease, defined by Child-Pugh score ≥ C or AST > twice the upper limit

  5. Pregnant or breastfeeding

  6. Severe renal failure, defined by eGFR ≤ 30 mL/min/1.73m2, or on dialysis

  7. Potential transfer to another hospital within 72h of enrollment

  8. Received any trial treatment for COVID-19 within 30 days before the current study

All patients received the standard of care: oxygen therapy, antiviral agents, antibacterial agents, and immunoglobulin, with or without corticosteroids. Patients in the HCQ treatment group received additional oral HCQ 400 mg/day, given as 200 mg 2x/day. HCQ was administered from days 1-5 of the trial. The primary endpoint was 5 days post enrollment or a severe adverse reaction to HCQ. The primary outcome evaluated was time to clinical recovery (TTCR), defined as return to normal body temperature and cough cessation for > 72h. Chest CT were imaged on days 0 and 6 of the trial for both groups; body temperature and patient reports of cough were collected 3x/day from day 0 – 6. The mean age and sex distribution between the HCQ and control arms were comparable.

12.83.3 Main Findings

There were 2 patients showing mild secondary effects of HCQ treatment. More importantly, while 4 patients in the control group progressed to severe disease, none progressed in the treatment group.

TTCR was significantly decreased in the HCQ treatment arm; recovery from fever was shortened by one day (3.2 days control vs. 2.2 days HCQ, p = 0.0008); time to cessation of cough was similarly reduced (3.1 days control vs. 2.0 days HCQ, p = 0.0016).

Overall, it appears that HCQ treatment of patients with mild COVID-19 has a modest effect on clinical recovery (symptom relief on average 1 day earlier) but may be more potent in reducing the progression from mild to severe disease.

12.83.4 Limitations

This study is limited in its inclusion of only patients with mild disease, and exclusion of those on any treatment other than the standard of care. It would also have been important to include the laboratory values of positive RT-PCR detection of SARS-CoV-2 to compare the baseline and evolution of the patients’ viral load.

12.83.5 Limitations

Despite its limitations, the study design has good rigor as a double blind RCT and consistent symptom checks on each day of the trail. Now that the FDA has approved HCQ for treatment of COVID-19 in the USA, this study supports the efficacy of HCQ use early in treatment of patients showing mild symptoms, to improve time to clinical recovery, and possibly reduce disease progression. However, most of the current applications of HCQ have been in patients with severe disease and for compassionate use, which are out of the scope of the findings presented in this trial. Several additional clinical trials to examine hydroxychloroquine are now undergoing; their results will be critical to further validate these findings.

12.83.6 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

Structure-based modeling of SARS-CoV-2 peptide/HLA-A02 antigens

https://doi.org/10.1101/2020.03.23.004176

Immunology keywords:

CoVID-19, 2019-nCoV, SARS-CoV-2, comparative, homology, peptide, modeling, simulation, HLA-A, antigen

Summary of Findings:

Limitations:

Importance/Relevance:

References:

  1. Bender, B. J., Cisneros, A., 3rd, Duran, A. M., Finn, J. A., Fu, D., Lokits, A. D., . . . Moretti, R. (2016). Protocols for Molecular Modeling with Rosetta3 and RosettaScripts. Biochemistry, 55(34), 4748-4763. doi:10.1021/acs.biochem.6b00444

  2. Alford, R. F., Leaver-Fay, A., Jeliazkov, J. R., O’Meara, M. J., DiMaio, F. P., Park, H., . . . Gray, J. J. (2017). The Rosetta All-Atom Energy Function for Macromolecular Modeling and Design. J Chem Theory Comput, 13(6), 3031-3048. doi:10.1021/acs.jctc.7b00125

Review by Jonathan Chung as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn school of medicine, Mount Sinai.

12.84 Serology characteristics of SARS-CoV-2 infection since the exposure and post symptoms onset

Lou et al. medRxiv. [1008]

12.84.1 Keywords

12.84.2 Main Findings

Currently, the diagnosis of SARS-CoV-2 infection entirely depends on the detection of viral RNA using polymerase chain reaction (PCR) assays. False negative results are common, particularly when the samples are collected from upper respiratory. Serological detection may be useful as an additional testing strategy. In this study the authors reported that a typical acute antibody response was induced during the SARS-CoV-2 infection, which was discuss earlier1. The seroconversion rate for Ab, IgM and IgG in COVID-19 patients was 98.8% (79/80), 93.8% (75/80) and 93.8% (75/80), respectively. The first detectible serology marker was total antibody followed by IgM and IgG, with a median seroconversion time of 15, 18 and 20 days-post exposure (d.p.e) or 9, 10- and 12-days post-onset (d.p.o). Seroconversion was first detected at day 7d.p.e in 98.9% of the patients. Interestingly they found that viral load declined as antibody levels increased. This was in contrast to a previous study [949], showing that increased antibody titers did not always correlate with RNA clearance (low number of patient sample).

12.84.3 Limitations

Current knowledge of the antibody response to SAR-CoV-2 infection and its mechanism is not yet well elucidated. Similar to the RNA test, the absence of antibody titers in the early stage of illness could not exclude the possibility of infection. A diagnostic test, which is the aim of the authors, would not be useful at the early time points of infection but it could be used to screen asymptomatic patients or patients with mild disease at later times after exposure.

12.84.4 Significance

Understanding the antibody responses against SARS-CoV2 is useful in the development of a serological test for the diagnosis of COVID-19. This manuscript discussed acute antibody responses which can be deducted in plasma for diagnostic as well as prognostic purposes. Thus, patient-derived plasma with known antibody titers may be used therapeutically for treating COVID-19 patients with severe illness.

12.84.5 Credit

This review was undertaken and edited by Konstantina A as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.85 SARS-CoV-2 launches a unique transcriptional signature from in vitro, ex vivo, and in vivo systems

Blanco-Melo et al. bioRxiv. [1009]

12.85.1 Keywords

12.85.2 Main Findings

Given the high mortality rate of SARS-CoV-2 relative to other respiratory viruses such seasonal IAV and RSV, there may be underlying host-pathogen interactions specific to SARS-CoV-2 that predispose to a worse clinical outcome. Using in vivo, ex vivo, and in vitro systems, the authors profiled host cell transcriptional responses to SARS-CoV-2 and to other common respiratory viruses (seasonal IAV and RSV). SARS-CoV-2 infection in vitro led to an induction of type I interferon response signaling and the upregulation of cytokine/chemokines transcripts. In comparison with IAV and RSV infection, SARS-CoV-2 in vitro appears to uniquely induce less type I and type III interferon expression and higher levels of two cytokines previously implicated in respiratory inflammation. Lastly, in vivo data from ferrets showed a reduced induction of cytokines and chemokines by SARS-CoV-2 infection relative to IAV infection.

12.85.3 Limitations

While these results are promising, there are several key weaknesses of this paper. 1) As the authors point out, there is an undetectable level of SARS-CoV-2 putative receptor (ACE2) and protease (TMPRSS2) expression in the lung epithelial cell line used for the in vitro studies. This raises the important question of whether viral replication actually occurs in any of the models used, which may explain the lack of interferon production observed in vitro in SARS-CoV-2 treated cells. Further studies characterizing viral titers across timepoints are needed. 2) Furthermore, these studies only characterize the host response at a single dose and timepoint per virus, and it is unclear why these doses/timepoints were chosen. This leaves open the possibility that the observed differences between viruses could be due to differences in dose, timing, host response, or a combination of all of these. 3) It is unclear whether ferrets are productively infected, which cell types are infected, and the extent/timing of the clinical course of infection. Moreover, the in vitro and in vivo data do not strongly correlate and the reasons for this are unclear.

12.85.4 Significance

This paper describes potentially unique transcriptional signatures of host cells exposed to SARS-CoV-2. If validated, these findings may help explain clinical outcomes and could be targeted in future therapeutic interventions.

12.85.5 Potential Conflicts of Interest Disclosure

The reviewers are also researchers at the Icahn School of Medicine at Mount Sinai.

12.85.6 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.86 A New Predictor of Disease Severity in Patients with COVID-19 in Wuhan, China

Zhou et al. bioRxiv. [1010]

12.86.1 Keywords

12.86.2 Main Findings

377 hospitalized patients were divided into two groups: severe and non-severe pneumonia. The laboratory results of their first day of admission were retrospectively analyzed to identify predictors of disease severity.

After adjusting for confounding factors from chronic comorbidities (such as high blood pressure, type 2 diabetes, coronary heart disease, and chronic obstructive pulmonary disease), the independent risk factors identified for severe pneumonia were age, the ratio of neutrophil/lymphocytes counts, CRP and D-dimer levels.

To further increase the specificity and sensibility of these markers, they showed that their multiplication [(Neutrophil/lymphocyte count) * CRP * D-dimer] was a better predictor of disease severity, with higher sensitivity (95.7%) and specificity (63.3%), with a cutoff value of 2.68.

12.86.3 Limitations

This study included 377 hospitalized patients. Among them, 45.6% patients tested positive for SARS-Cov-2 nucleic acid test results, and others were included in the study based on clinically diagnosis even if the molecular diagnosis was negative. Thus, additional studies are needed to verify this on a larger number of covid-19 certified patients and the cutoff value might be adjusted. Also, all the patients that did not have the clinical characteristics of severe pneumonia were included in the non-severe pneumonia group, but usually patients are also divided into moderate and mild disease.

Also, studying different subset of lymphocytes could lead to a more specific predictor. Another study showed that the neutrophils to CD8+ T cells ratio was a strong predictor of disease severity [919]. Another more precise study showed that the percentage of helper T cells and regulatory T cells decrease but the percentage of naïve helper T cells increases in severe cases [912]. Taking these subpopulations into account might make the predictor more powerful.

Other studies also noted an inverse correlation between disease severity and LDH [953] or IL6 [962] levels, but the authors here do not discuss LDH nor IL6 levels, although this could help to strengthen the predictor.

The study is based on the results obtained on the first day of admission, studying the dynamic of the changes in patients might also be interesting to better predict disease severity.

12.86.4 Significance

This study confirms that the neutrophil to lymphocyte ratio can be a predictor of disease severity as shown by many others [911,912,926]. The novelty here is that they show that a combination with other markers can enhance the specificity and sensibility of the predictor, although the study could be improved by taking into account sub-populations of lymphocytes and more biological factors from patients such as LDH and IL6.

12.86.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.87 Metabolic disturbances and inflammatory dysfunction predict severity of coronavirus disease 2019 (COVID-19): a retrospective study

Shuke Nie et al. medRxiv. [1011]

12.87.1 Keywords

12.87.2 Main Findings

Retrospective Study on 97 COVID-19 hospitalized patients (25 severe and 72 non-severe) analyzing clinical and laboratory parameter to predict transition from mild to severe disease based on more accessible indicators (such as fasting blood glucose, serum protein or blood lipid) than inflammatory indicators. In accordance with other studies, age and hypertension were risk factors for disease severity, and lymphopenia and increased IL-6 was observed in severe patients. The authors show that fasting blood glucose (FBG) was altered and patients with severe disease were often hyperglycemic. Data presented support that hypoproteinaemia, hypoalbuminemia, and reduction in high-densitylipoprotein (HDL-C) and ApoA1 were associated with disease severity.

12.87.3 Limitations

In this study non-severe patients were divided in two groups based on average course of the disease: mild group1 (14 days, n=28) and mild group 2 (30 days, n=44). However mild patients with a longer disease course did not show an intermediate phenotype (between mild patients with shorter disease course and severe patients), hence it is unclear whether this was a useful and how it impacted the analysis. Furthermore, the non-exclusion of co-morbidity factors in the analysis may bias the results (e.g. diabetic patients and glucose tests) It is not clear at what point in time the laboratory parameters are sampled. In table 3, it would have been interesting to explore a multivariate multiple regression. The correlation lacks of positive control to assess the specificity of the correlation to the disease vs. correlation in any inflammatory case. The dynamic study assessing the predictability of the laboratory parameter is limited to 2 patients. Hence there are several associations with disease severity, but larger studies are necessary to test the independent predictive value of these potential biomarkers.

12.87.4 Significance

As hospital are getting overwhelmed a set of easily accessible laboratory indicators (such as serum total protein) would potentially provide a triage methodology between potentially severe cases and mild ones. This paper also opens the question regarding metabolic deregulation and COVID-19 severity.

12.87.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.88 Viral Kinetics and Antibody Responses in Patients with COVID-19

[1012]

12.88.1 Keywords

12.88.2 Main Findings

12.88.3 Limitations

Specific for immune monitoring.

12.88.4 Significance

12.88.5 Credit

Review by Samarth Hegde as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

[1014]

12.89.1 Keywords

12.89.2 Main Findings

This study is based on flow cytometry immunophenotyping of PBMCs from 28 patients diagnosed positive for SARS-Cov2 (COVID19). The authors identify a population of abnormally large (FSC-hi) monocytes, present in COVID19 patients, but absent in PBMCs of healthy volunteers (n=16) or patients with different infections (AIDS, malaria, TB). This FSC-hi monocytic population contains classical, intermediate and non-classical (monocytes (based on CD14 and CD16 expression) that produce inflammatory cytokines (IL-6, TNF and IL-10). The authors suggest an association of FSC-hi monocytes with poor outcome and correlate a high percentage of FSC-low monocytes, or higher ratio of FSC-low/hi monocytes, with faster hospital discharge.

12.89.3 Limitations

While identification of the monocytic population based on FSC is rather robust, the characterization of these cells remains weak. A comprehensive comparison of FSC-hi monocytes with FSC-low monocytes from patients and healthy controls would be of high value. It is unclear if percentages in blood are among CD45+ cells. Furthermore, it would have been important to include absolute numbers of different monocytic populations (in table 1 there are not enough samples and it is unclear what the authors show).

The authors show expression of the ACE2 receptor on the surface of the monocytes, and highlight these cells as potential targets of SARS-Cov2. However, appropriate controls are needed. CD16 has high affinity to rabbit IgG and it is unclear whether the authors considered unspecific binding of rabbit anti-ACE2 to Fc receptors. Gene expression of ACE-2 on monocytes needs to be assessed. Furthermore, it would be important to confirm infection of monocytes by presence of viral proteins or viral particles by microscopy.

Considering the predictive role of FSC-hi monocytes on the development of the disease and its severity, some data expected at this level are neither present nor addressed. Although the cohort is small, it does include 3 ICU patients. What about their ratio of FSC-low vs FSC-hi monocytes in comparison to other patients? Was this apparent early in the disease course? Does this population of FSC-hi monocytes differ between ICU patients and others in terms of frequency, phenotype or cytokine secretion?

In general, figures need to revised to make the data clear. For example, in Fig. 5, according to the legend it seems that patients with FSC-high monocytes are discharged faster from the hospital. However according to description in the text, patients were grouped in high or low levels of FSC-low monocytes.

12.89.4 Significance

Despite the limitations of this study, the discovery of a FSC-high monocyte population in COVID-19 patients is of great interest. With similar implication, a the recent study by Zhou et al. [917] identified a connection between an inflammatory CD14+CD16+ monocyte population and pulmonary immunopathology leading to deleterious clinical manifestations and even acute mortality after SARS-CoV-2 infections. Although the presence of these monocytes in the lungs has yet to be demonstrated, such results support the importance of monocytes in the critical inflammation observed in some COVID19 patients.

12.89.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.90 Correlation between universal BCG vaccination policy and reduced morbidity and mortality for COVID-19: an epidemiological study

Miller et al. medRxiv. [1015]

12.90.1 Keywords

12.90.2 Main Findings

The authors compared middle and high income countries that never had a universal BCG vaccination policy (Italy, Lebanon, Nederland, Belgium) and countries with a current policy (low income countries were excluded from the analysis as their number of cases and deaths might be underreported for the moment). Countries that never implement BCG vaccination have a higher mortality rate than countries which have a BCG vaccination policy (16.38 deaths per million people vs 0.78). Next, the authors show that an earlier start of vaccination correlates with a lower number of deaths per million inhabitants. They interpret this as the vaccine protecting a larger fraction of elderly people, which are usually more affected by COVID-19. Moreover, higher number of COVID-19 cases were presented in countries that never implemented a universal BCG vaccination policy.

12.90.3 Limitations

While this study aims to test an intriguing hypothesis unfortunately, the data is not sufficient at this time to accurately make any determinations. Several caveats must be noted including: not all countries are in the same stage of the pandemic, the number of cases/deaths is still changing very rapidly in a lot of countries and thus the association may only reflect exposure to the virus. This analysis would need to be re-evaluated when all the countries are passed the pandemic and more accurate numbers are available. Additionally, very few middle and high-income countries ever implemented universal BCG vaccination, which can be a source of bias (5 countries, vs 55 that have a BCG vaccine policy). Effective screening and social isolation policies also varied considerable across the countries tested and may reflect another important confounder. The authors could consider analyzing the Case Fatality Rate (CFR, % of patients with COVID-19 that die), to more correct for exposure although testing availability will still bias this result. Variability in mortality within countries or cities with variable vaccination and similar exposure could also be appropriate although confounders will still be present.

12.90.4 Significance

BCG vaccine is a live attenuated strain derived from Mycobacterium bovis and used for a vaccine for tuberculosis (TB). This vaccine has been proven to be efficient in preventing childhood meningitis TB, but doesn’t prevent adult TB as efficiently. For this reason, several countries are now only recommending this vaccine for at-risk population only.

This study shows that there is a correlation between BCG vaccination policy and reduced mortality for Covid-19. Indeed, BCG vaccine has been shown to protect against several viruses and enhance innate immunity [1016], which could explain why it could protect against SARS-CoV-2 infection, but the exact mechanism is still unknown. Moreover, the efficiency of adult/older people vaccination and protection against Covid-19 still needs to be assessed. Regarding this, Australian researchers are starting a clinical trial of BCG vaccine for healthcare workers [1017], to assess if it can protect them against Covid-19.

12.90.5 Credit

This review was undertaken as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.91 Non-neural expression of SARS-CoV-2 entry genes in the olfactory epithelium

[1018]

12.91.1 Keywords

12.91.2 Main Findings

12.91.3 Limitations

12.91.4 Significance

12.91.5 Credit

Review by Samarth Hegde as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

Title:

SARS-CoV-2 proteome microarray for mapping COVID-19 antibody interactions at amino acid resolution

Immunology keywords: SARS-CoV-2, COVID-19, high throughput, peptide microarray, antibody epitope screening

The main finding of the article:

This study screened the viral protein epitopes recognized by antibodies in the serum of 10 COVID-19 patients using a new SARS-CoV-2 proteome peptide microarray. The peptide library was constructed with 966 linear peptides, each 15 amino acids long with a 5 amino acid overlap, based on the protein sequences encoded by the genome of the Wuhan-Hu-1 strain.

To investigate crossreactivity between SARS-CoV-1 and SARS-CoV-2, they tested rabbit monoclonal and polyclonal antibodies against SARS-CoV-1 nucleocapsid (N) in the microarray. Antibodies against SARS-CoV-1 N displayed binding to the SARS-CoV-2 nucleocapsid (N) peptides. Polyclonal antibodies showed some crossreactivity to other epitopes from membrane (M), spike (S), ORF1ab and ORF8. This suggests that previous exposure to SARS-CoV-1 may induced antibodies recognizing both viruses.

Screening of IgM and IgG antibodies from 10 COVID-19 patients showed that many antibodies targeted peptides on M, N, S, Orf1ab, Orf3a, Orf7a, and Orf8 from SARS-CoV-2, while immunodominant epitopes with antibodies in more than 80 % COVID-19 patients were present in N, S and Orf3. It is shown that the receptor binding domain (RBD) resides on S protein and RBD is important for SARS-CoV-2 to enter the host cells via ACE2. Among six epitopes on S protein, structural analysis predicted that three epitopes were located at the surface and three epitopes were located inside of the protein. Furthermore, some IgM antibodies from 1 patient and IgG antibodies from 2 patients bound to the same epitope (residue 456-460, FRKSN) which resided within the RBD, and structural analysis determined that this epitope was located in the region of the RBD loop that engages with ACE2.

Critical analysis of the study:

In addition to the limitations mentioned in the manuscript, it would have been informative to do the analysis over the course of the disease. The pattern of antibody recognition, especially on S protein, and the course of antibodies of different isotypes recognizing the same peptide might correlate to the clinical course in these patients. It would alos have been informative to analyze the presence of cross-reactive antibodies from pateints previously exposed to SARS-CoV-1.

The importance and implications for the current epidemics:

This study identified linear immunodominant epitopes on SARS-CoV-2, Wuhan-Hu-1 strain. This is a valuable information to design vaccines that will elicit desirable immune responses.

The Novel Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) Directly Decimates Human Spleens and Lymph Nodes

Review by Matthew D. Park

Revised by Miriam Merad

Keywords: COVID-19, SARS-CoV-2, spleen, lymph node, ACE2, macrophage

Main findings

It has been previously reported that COVID-19 patients exhibit severe lymphocytopenia, but the mechanism through which this depletion occurs has not been described. In order to characterize the cause and process of lymphocyte depletion in COVID-19 patients, the authors performed gross anatomical and in situ immune-histochemical analyses of spleens and lymph nodes (hilar and subscapular) obtained from post-mortem autopsies of 6 patients with confirmed positive viremia and 3 healthy controls (deceased due to vehicle accidents).

Primary gross observations noted significant splenic and LN atrophy, hemorrhaging, and necrosis with congestion of interstitial blood vessels and large accumulation of mononuclear cells and massive lymphocyte death. They found that CD68+ CD169+ cells in the spleens, hilar and subscapular LN, and capillaries of these secondary lymphoid organs expressed the ACE2 receptor and stain positive for the SARS-CoV-2 nucleoprotein (NP) antigen, while CD3+ T cells and B220+ B cells lacked both the ACE2 receptor and SARS-CoV-2 NP antigen. ACE2+ NP+ CD169+ macrophages were positioned in the splenic marginal zone (MZ) and in the marginal sinuses of LN, which suggests that these macrophages were positioned to encounter invading pathogens first and may contribute to virus dissemination.

Since SARS-CoV-2 does not directly infect lymphocytes, the authors hypothesized that the NP+ CD169+ macrophages are responsible for persistent activation of lymphocytes via Fas::FasL interactions that would mediate activation-induced cell death (AICD). Indeed, the expression of Fas was significantly higher in virus-infected tissue than that of healthy controls, and TUNEL staining showed significant lymphocytic apoptosis. Since pro-inflammatory cytokines like IL-6 and TNF-α can also engage cellular apoptosis and necrosis, the authors interrogated the cytokine expression of the secondary lymphoid organs from COVID-19 patients; IL-6, not TNF-α, was elevated in virus-infected splenic and lymph node tissues, compared to those of healthy controls, and immunofluorescent staining showed that IL-6 is primarily produced by the infected macrophages. In vitro infection of THP1 cells with SARS-CoV-2 spike protein resulted in selectively increased Il6 expression, as opposed to Il1b and Tnfa transcription. Collectively, the authors concluded that a combination of Fas up-regulation and IL-6 production by NP+ CD169+ macrophages induce AICD in lymphocytes in secondary lymphoid organs, resulting in lymphocytopenia.

In summary, this study reports that CD169+ macrophages in the splenic MZ, subscapular LN, and the lining capillaries of the secondary lymphoid tissues express ACE2 and are susceptible to SARS-CoV-2 infection. The findings point to the potential role of these macrophages in viral dissemination, immunopathology of these secondary lymphoid organs, hyperinflammation and lymphopenia.

Limitations

Technical

A notable technical limitation is the small number of samples (n=6); moreover, the analysis of these samples using multiplexed immunohistochemistry and immunofluorescence do not necessarily provide the depth of unbiased interrogation needed to better identify the cell types involved.

Biological

The available literature and ongoing unpublished studies, including single-cell experiments of spleen and LN from organ donors, do not indicate that ACE2 is expressed by macrophages; however, it remains possible that ACE2 expression may be triggered by type I IFN in COVID-19 patients. Importantly, the SARS-CoV-2 NP staining of the macrophages does not necessarily reflect direct infection of these macrophages; instead, positive staining only indicates that these macrophages carry SARS-CoV-2 NP as antigen cargo, which may have been phagocytosed. Direct viral culture of macrophages isolated from the secondary lymphoid organs with SARS-CoV-2 is required to confirm the potential for direct infection of macrophages by SARS-CoV-2. Additionally, it is important to note that the low to negligible viremia reported in COVID-19 patients to-date does not favor a dissemination route via the blood, as suggested by this study, which would be necessary to explain the presence of virally infected cells in the spleen.

Relevance

Excess inflammation in response to SARS-CoV-2 infection is characterized by cytokine storm in many COVID-19 patients. The contribution of this pathology to the overall fatality rate due to COVID-19, not even necessarily directly due to SARS-CoV-2 infection, is significant. A better understanding of the full effect and source of some of these major cytokines, like IL-6, as well as the deficient immune responses, like lymphocytopenia, is urgently needed. In this study, the authors report severe tissue damage in spleens and lymph nodes of COVID-19 patients and identify the role that CD169+ macrophages may play in the hyperinflammation and lymphocytopenia that are both characteristic of the disease. It may, therefore, be important to note the effects that IL-6 inhibitors like Tocilizumab and Sarilumab may specifically have on splenic and LN function. It is important to note that similar observations of severe splenic and LN necrosis and inflammation in patients infected with SARS-CoV-1 further support the potential importance and relevance of this study.

12.92 Cigarette smoke triggers the expansion of a subpopulation of respiratory epithelial cells that express the SARS-CoV-2 receptor ACE2

[1019]

12.92.1 Keywords

12.92.2 Main Findings

12.92.3 Limitations:

12.92.4 Significance

12.92.5 Credit

Review by Samarth Hegde as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

12.93 The comparative superiority of IgM-IgG antibody test to real-time reverse transcriptase PCR detection for SARS-CoV-2 infection diagnosis

Liu et al. medRxiv. [1020]

12.93.1 Keywords

12.93.2 Main Findings

The study compares IgM and IgG antibody testing to RT-PCR detection of SARS-CoV-2 infection. 133 patients diagnosed with SARS-CoV-2 in Renmin Hospital (Wuhan University, China) were analyzed. The positive ratio was 78.95% (105/133) in IgM antibody test (SARS-CoV-2 antibody detection kit from YHLO Biotech) and 68.42% (91/133) in RT-PCR (SARS-CoV-2 ORF1ab/N qPCR detection kit). There were no differences in the sensitivity of SARS-CO-V2 diagnosis in patients grouped according to disease severity. For example, IgG responses were detected in 93.18% of moderate cases, 100% of severe cases and 97.3% of critical cases. In sum, positive ratios were higher in antibody testing compared to RT-PCR detection, demonstrating a higher detection sensitivity of IgM-IgG testing for patients hospitalized with COVID-19 symptoms.

12.93.3 Limitations

This analysis only included one-time point of 133 hospitalized patients, and the time from symptom onset was not described. There was no discussion about specificity of the tests and no healthy controls were included. It would be important to perform similar studies with more patients, including younger age groups and patients with mild symptoms as well as asymptomatic individuals. It is critical to determine how early after infection/symptom onset antibodies can be detected and the duration of this immune response.

12.93.4 Significance

The IgM-IgG combined testing is important to improve clinical sensitivity and diagnose COVID-19 patients. The combined antibody test shows higher sensitivity than individual IgM and IgG tests or nucleic acid-based methods, at least in patients hospitalized with symptoms.

12.93.5 Credit

Review by Erica Dalla as part of a project by students, postdocs and faculty at the Immunology Institute of the Icahn School of Medicine, Mount Sinai.

Title: Lectin-like Intestinal Defensin Inhibits 2019-nCoV Spike binding to ACE2

Immunology keywords: defensins, spike protein, intestinal Paneth cells, ACE2 binding

Main Findings:

Human ACE2 was previously identified as the host receptor for SARS-CoV-2. Despite ACE2 being expressed in both lung alveolar epithelial cells and small intestine enterocytes, respiratory problems are the most common symptom after viral infection while intestinal symptoms are much less frequent. Thus, the authors here investigate the biology behind the observed protection of the intestinal epithelium from SARS-CoV-2. Human defensin 5 (HD5), produced by Paneth cells in the small intestine, was shown to interact with human ACE2, with a binding affinity of 39.3 nM by biolayer interferometry (BLI). A blocking experiment using different doses of HD5 coating ACE2 showed that HD5 lowered viral spike protein S1 binding to ACE2. Further, a molecular dynamic simulation demonstrated a strong intermolecular interaction between HD5 and the ACE2 ligand binding domain. To test HD5 inhibitory effect on S1 binding to ACE2, human intestinal epithelium Caco-2 cells were preincubated with HD5. Preincubation strongly reduced adherence of S1 to surface of cells. HD5 was effective at a concentration as low as 10 µg/mL, comparable to the concentration found in the intestinal fluid.

Limitations:

The study focuses exclusively on intestinal cells. However, HD5 could have been tested to block ACE2-S1 binding in human lung epithelial cells as a potential treatment strategy. It would be useful to know whether HD5 could also prevent viral entry in lung cells.

Relevance:

This work provides the first understanding of the different efficiency of viral entry and infection among ACE2-expressing cells and tissues. Specifically, the authors show that human defensin 5 produced in the small intestine is able to block binding between S1 and ACE2 necessary for viral entry into cells. The study provides a plausible explanation on why few patients show intestinal symptoms and suggests that patients with intestinal disease that decrease defensins’ production may be more susceptible to SARS-CoV-2. It also indicates that HD5 could be used as a molecule to be exogenously administered to patients to prevent viral infection in lung epithelial cells.

Title:

Susceptibility of ferrets, cats, dogs and different domestic animals to SARS-coronavirus-2

Immunology keywords: SARS-CoV-2, ferret, cat, laboratory animal, domestic animals

The main finding of the article:

This study evaluated the susceptibility of different model laboratory animals (ferrets), as well as companion (cats and dogs), and domestic animals (pigs, chickens and ducks) to SARS-CoV-2. They tested infection with two SARS-CoV2 isolates, one from an environmental sample collected in the Huanan Seafood Market in Wuhan (F13-E) and the other from a human patient in Wuhan (CTan-H).

Ferrets were inoculated with either of the two viruses by intranasal route with 105 pfu, and the viral replication was evaluated. Two ferrets from each group were euthanized on day 4 post infection (p.i.). AT day 4 p.i., viral RNA and infectious viruses were detected only in upper respiratory tract (nasal turbinate, upper palate, tonisls, but not in the trachea, lungs or other tissues. Viral RNA and virus titer in the remaining ferrets were monitored in nasal washes and rectal swabs on days 2, 4, 6, 8 and 10 p.i. Viral RNA and infectious viruses were detected in nasal washes until day 8 p.i. One ferret in each group developed fever and loss of appetite on days 10 and 12 p.i., however, viral RNA was practically undetactable. These two ferrets showed severe lymphoplasmacytic perivasculitis and vasculitis in the lungs and lower antibody titers compare to other 4 ferrets.

Cats. Five subadult 8-month-old domestic cats were inoculated with CTan-h virus and three uninfected cats were placed in a cage adjacent to each of the infected cats to monitor respiratory droplet transmission. Viral RNA was detected in the upper respiratory organs from all infected cats and in one out of three exposed cats. All infected (inoculated and exposed) cats developed elevated antibodies against SARS-CoV2. Viral replication studies with juvenile cats (70-100 days) revealed massive lesions in the nasal and tracheal mucosa epithelium and lungs of two inoculated cats which died or were euthanized on day 3 p.i., and infection in one out of three exposed cats. These results indicated SARS-CoV2 could replicate in cats, that juvenile cats were more susceptible that adults, and theat SARS-CoV2 could be transmit via respiratory droplets between cats.

Dogs and others. Five 3-month-old beagle dogs were inoculated and housed with two uninoculated beagles in a room. Two virus inoculated dogs seroconverted, but others including two contact dogs were all seronegative for SARS-CoV2 and infectious virus was not detected in any swabs collected. Viral RNA was not detected in swabs from pigs, chickens, and ducks inoculated or contacted. These results indicated that dogs, pigs, chickens, and ducks might have low or no susceptibility to SARS-CoV2.

Critical analysis of the study:

This manuscript describes the viral replication and clinical symptoms of SARS-CoV2 infection in ferrets, and the SARS-CoV2 infection and transmission in cats. Clinical and pathological analysis was not performed in cats, therefore the correlation of virus titer with symptoms severity in the adult and juvenile cats could not be determined.

The importance and implications for the current epidemics:

SARS-CoV-2 transmission to tigers, cats and dogs has been previously reported. It should be noted that this manuscript did not evaluate the transmission from cats to human. Nevertheless, it clearly showed higher susceptibility of ferrets and domestic cats to SARS-CoV-2. This data strongly indicates the need for surveillance of possible infection and transmission of SARS-CoV-2 by domestic cats.

12.94 Virus-host interactome and proteomic survey of PMBCs from COVID-19 patients reveal potential virulence factors influencing SARS-CoV-2 pathogenesis

Li et al. bioRxiv. [155]

12.94.1 Keywords

12.94.2 Main findings

The authors identified intra-viral protein-protein interactions (PPI) with two different approaches: genome wide yeast-two hybrid (Y2H) and co-immunoprecipitation (co-IP). A total of 58 distinct PPI were characterized. A screen of viral-host PPI was also established by overexpressing all the SARS-CoV-2 genes with a Flag epitope into HEK293 cells and purifying each protein complex. Interacting host proteins were then identified by liquid chromatography and tandem mass spectrometry. 251 cellular proteins were identified, such as subunits of ATPase, 40S ribosomal proteins, T complex proteins and proteasome related proteins, for a total of 631 viral-host PPI. Several interactions suggesting protein-mediated modulation of the immune response were identified, highlighting the multiple ways SARS-CoV-2 might reprogram infected cells.

Subsequently, the authors compared global proteome profiles of PBMCs from healthy donors (n=6) with PBMC from COVID-19 patients with mild (n=22) or severe (n=13) symptoms. 220 proteins were found to be differentially expressed between healthy donors and mild COVID-19 patients, and a pathway analysis showed a general activation of the innate immune response. 553 proteins were differentially expressed between the PBMC of mild and severe COVID-19 patients, most of them (95%) being downregulated in severe patients. Functional pathway analysis indicated a defect of T cell activation and function in severe COVID-19. There was also evidence suggesting reduced antibody secretion by B cells. Together, these results suggest a functional decline of adaptive immunity. A FACS analysis of PBMC from severe patients indicated higher levels of IL6 and IL8 but not IL17 compared to mild patients.

Finally, the authors focused on NKRF, an endogenous repressor of IL8/IL6 synthesis that was previously identified as interacting with SARS-Cov-2 nsp9,10,12,13 and 15. Individually expressed nsp9 and nsp10 (but not nsp12, nsp13, nsp15) induced both IL6 and IL8 in lung epithelial A459 cells, indicating that nsp9 and nsp10 may be directly involved in the induction of these pro-inflammatory cytokines. The authors finally argue that nsp9 and nsp10 represent potential drug targets to prevent over-production of IL6 and IL8 in infected cells, and reducing the over-activation of neutrophils.

12.94.3 Limitations

First, the authors seem to have forgotten to include the extended data in the manuscript, and their proteomic data does not seem to be publicly available for the moment, which limits greatly our analysis of their results.

While this work provides important data on host and viral PPI, only 19 interactions were identified by Y2H system but 52 with co-IP. The authors do not comment about what could lead to such differences between the two techniques and they don’t specify whether they detected the same interactions using the two techniques.

Moreover, the PBMC protein quantification was performed comparing bulk PBMC. Consequently, protein differences likely reflect differences in cell populations rather than cell-intrinsic differences in protein expression. While this analysis is still interesting, a similar experiment performed on pre-sorted specific cell populations would allow measuring proteome dynamics at a higher resolution.

Finally, the authors did not discussed their results in regards to another SARS-CoV-2 interactome of host-viral PPI that had been published previously1. This study reported 332 host-virus PPI, but no interaction of viral proteins with NKRF was found. Some interactions were found in both studies (eg. N and G3BP1, Orf6 and RAE1). However, the time point used to lyse the cells were different (40h previously vs 72h here), which could explain some of the differences.

12.94.4 Relevance

The identification of many interactions between intra-viral and host-virus PPI provides an overview of host protein and pathways that are modulated by SARS-CoV-2, which can lead to the identification of potential targets for drug development.

In the model proposed by the authors, nsp9 and nsp10 from SARS-Cov-2 induce an over-expression of IL6 and IL8 by lung epithelial cells, which recruits neutrophils and could lead to an excess in lung infiltration. Nsp9 has been shown to be essential for viral replication for SARS-Cov-12, and shares a 97% homology with nsp9 from SARS-Cov-23. Further, nsp9 crystal structure was recently solved3, which can help to develop drug inhibitors if this protein is further confirmed as being important for the virulence of SARS-Cov-2.

1. Gordon DE, Jang GM, Bouhaddou M, et al. A SARS-CoV-2-Human Protein-Protein Interaction Map Reveals Drug Targets and Potential Drug-Repurposing. bioRxiv. March 2020:2020.03.22.002386. doi:10.1101/2020.03.22.002386

2. Miknis ZJ, Donaldson EF, Umland TC, Rimmer RA, Baric RS, Schultz LW. Severe acute respiratory syndrome coronavirus nsp9 dimerization is essential for efficient viral growth. J Virol. 2009;83(7):3007-3018. doi:10.1128/JVI.01505-08

3. Littler DR, Gully BS, Colson RN, Rossjohn J. Crystal Structure of the SARS-CoV-2 Non-Structural Protein 9, Nsp9. Molecular Biology; 2020. doi:10.1101/2020.03.28.013920

Title: Prediction and Evolution of B Cell Epitopes of Surface Protein in SARS-CoV-2

Keywords: SARS-CoV-2; Epitopes; Bioinformatics; Evolution

Summary/Main findings:

Lon et al. used a bioinformatic analysis of the published SARS-CoV-2 genomes in order to identify conserved linear and conformational B cell epitopes found on the spike (S), envelope (E), and membrane (M) proteins. The characterization of the surface proteins in this study began with an assessment of the peptide sequences in order to identify hydrophilicity indices and protein instability indices using the Port-Param tool in ExPASy. All three surface proteins were calculated to have an instability score under 40 indicating that they were stable. Linear epitopes were identified on the basis of surface probability and antigenicity, excluding regions of glycosylation. Using BepiPred 2.0 (with a cutoff value of 0.35) and ABCpred (with a cutoff value of 0.51), 4 linear B cell epitopes were predicted for the S protein, 1 epitope for the E protein, and 1 epitope for the M protein. For structural analysis, SARS-CoV assemblies published in the Protein Data Bank (PDB) acting as scaffolds for the SARS-CoV-2 S and E amino acid sequences were used for input into the SWISS-MODEL server in order to generate three-dimensional structural models for the assessment of conformational epitopes. Using Ellipro (cutoff value of 0.063) and SEPPA (cutoff value of 0.5), 1 conformational epitope was identified for the S protein and 1 epitope was identified for the E protein, both of which are accessible on the surface of the virus. Finally, the Consurf Server was used to assess the conservation of these epitopes. All epitopes were conserved across the published SARS-CoV-2 genomes and one epitope of the spike protein was predicted to be the most stable across coronavirus phylogeny.

Critical Analysis/Limitations:

While this study provides a preliminary identification of potential linear and conformational B cell epitopes, the translational value of the epitopes described still needs extensive experimental validation to ascertain whether these elicit a humoral immune response. The conformational epitope analyses are also limited by the fact that they are based off of predicted 3D structure from homology comparisons and not direct crystal structures of the proteins themselves. Additionally, since there was not a published M protein with a high homology to SARS-CoV-2, no conformational epitopes were assessed for this protein. Finally, while evolutionary conservation is an important consideration in understanding the biology of the virus, conservation does not necessarily imply that these sites neutralize the virus or aid in non-neutralizing in vivo protection.

Relevance/Implications:

With further experimental validation that confirms that these epitopes induce effective antibody responses to the virus, the epitopes described can be used for the development of treatments and vaccines as well as better characterize the viral structure to more deeply understand pathogenesis.

13 Appendix B

Contributors were asked to complete this template to summarize and evaluate new papers related to diagnostics.

Title: Please edit the title to add the name of the paper after the colon

Please paste a link to the paper or a citation here:

Link:

What is the paper’s Manubot-style citation?

Citation:

Please list some keywords (3-10) that help identify the relevance of this paper to COVID-19

Please note the publication / review status

Which areas of expertise are particularly relevant to the paper?

Questions to answer about each paper:

Please provide 1-2 sentences introducing the study and its main findings

Study question(s) being investigated:

What type of testing scenario is being considered?

Is it a screening test (used for individuals with no symptoms), diagnostic test (used for individuals with symptoms), or definitive test (used for individuals who have had previous positive test results on diagnostic or screening tests)?

Study population:

What is the model system (e.g., human study, animal model, cell line study)?

What is the sample size?

What is the “pre-test” probability of disease in the study population (i.e., what is the anticipated prevalence of the disease?)

For human studies, the following are related to the pre-test probability:

What countries/regions are considered?

What is the age range, gender, other relevant characteristics?

What is the setting of the study (e.g., random sample of school children, retirement communities, etc.)?

What other specific inclusion-exclusion criteria are considered?

Reference test:

What reference test is considered as a “gold standard” comparator for the test under investigation?

Test assignment:

How are the new and reference tests assigned?

Examples of assignment could include: Recruited individuals have initially undergone neither the new nor the reference test; individuals tested as positive or negative by the reference test undergo the new test; individuals who have undertaken the new test are assessed by the standard test.

Are there any other relevant details about the study design?

Depending on how individuals are chosen, the test may be biasing towards more sick or less sick individuals or very clear-cut positive/negative cases. Any factors that would influence this bias should be included here.

Test conduct:

How were tests performed?

Describe technical details of assays used, when measurements were taken and by whom, etc. for both the new and standard tests.

Test Assessment

Describe how individuals are classified as positive or negative, e.g. if a threshold is used.

Is there evidence that the test is precise/reproducible when repeated more than once?

Are measurements complete?

For example: Do some participants undergo just one test (the new or the reference test)? Are there individuals with inconclusive results?

Results summary:

What are the estimated sensitivity, specificity, positive predictive value (PPV), and negative predicted value (NPV)?

Note that the PPV and NPV represent “post-test” probabilities of disease and are generally more meaningful than sensitivity and specificity. Sometimes the post-test odds will be given instead.

What are the confidence bounds around these intervals?

Interpretation of results for study population:

How good is the test at ruling in or ruling out a disease based on the post-test probabilities?

Are there identified side affects of the test?

Is patient adherence to the test likely to be an issue?

Extrapolation of conclusions to other groups of individuals

How well is the test likely to work in populations with different pretest odds?

For example, if the prevalence is lower, then the PPV will also be lower, but the NPV will be higher.

How costly is the test?

How difficult is it to perform the test in different settings?

Could the test be combined with other existing tests?

Summary of reliability

1-2 sentences on concluding remarks, including summary of strengths, weaknesses, limitations.

Progress

Check off the components as they are completed. If the component is not applicable, check the box as well.

14 Appendix C

Contributors were asked to complete this template to summarize and evaluate new papers related to therapeutics.

Title: Please edit the title to add the name of the paper after the colon

Please paste a link to the paper or a citation here:

Link:

What is the paper’s Manubot-style citation?

Citation:

Please list some keywords (3-10) that help identify the relevance of this paper to COVID-19

Please note the publication / review status

Which areas of expertise are particularly relevant to the paper?

Questions to answer about each paper:

Please provide 1-2 sentences introducing the study and its main findings

Study question(s) being investigated:

How many/what drugs/combinations are being considered?

What are the main hypotheses being tested?

Study population:

What is the model system (e.g., human study, animal model, cell line study)?

What is the sample size? If multiple groups are considered, give sample size for each group (including controls).

For human studies:

What countries/regions are considered?

What is the age range, gender, other relevant characteristics?

What is the setting of the study (random sample of school children, inpatient, outpatient, etc)?

What other specific inclusion-exclusion criteria are considered?

For example, do the investigators exclude patients with diagnosed neoplasms or patients over/under a certain age?

Treatment assignment:

How are treatments assigned?

For example, is it an interventional or an observational study?

Is the study randomized?

A study can be interventional but not randomized (e.g., a phase I or II clinical trial is interventional but often not randomized).

Provide other relevant details about the design.

This includes possible treatment stratification (e.g., within litters for animal studies, within hospitals for human studies), possible confounding variables (e.g., having a large age range of individuals), possible risks of bias and how they are addressed (e.g., is there masking in a clinical trial? how are individuals chosen in an observational study?).

Outcome Assessment:

Describe the outcome that is assessed and whether it is appropriate.

For example: Is the outcome assessed by a clinician or is it self-reported? Is the outcome based on viral load or a functional measurement (e.g., respiratory function, discharge from hospital)? What method is used to measure the outcome? How long after a treatment is the outcome measured?

Are outcome measurements complete?

For example, are there individuals lost to follow up?

Are outcome measurements subject to various kinds of bias?

For example, a lack of masking in randomized clinical trials.

Statistical Methods Assessment:

What methods are used for inference?

For example, logistic regression, nonparametric methods.

Are the methods appropriate for the study?

For example, are clustered data treated independently or are clusters adjusted for, such as different hospitals or litters?

Are adjustments made for possible confounders?

For example, adjustment for age, sex, or comorbidities.

Results Summary:

What is the estimated association?

For example, is it an estimated odds ratio, a median difference in detected cases, etc?

What measures of confidence or statistical significance are provided?

For example, confidence intervals, p-values, and/or Bayes factors.

Interpretation of results for study population:

Can we make a causal interpretation for the individuals in the study of drug -> outcome, such as “taking drug A improves likelihood of survival twofold over taking drug B.”

For example, with a well-performed animal study or randomized trial it is often possible to infer causality. If is an observational study, does it match up with some of the Bradford Hill criteria? https://www.edwardtufte.com/tufte/hill https://en.wikipedia.org/wiki/Bradford_Hill_criteria

Are there identified side effects or interactions with other drugs?

For example, is the treatment known to cause liver damage or to not be prescribed for individuals with certain comorbities?

Are there specific subgroups with different findings?

For example, do individuals with a specific baseline seem to do particularly well? Be particularly cautious with respect to multiple testing here.

Extrapolation of conclusions to other groups of individuals not specifically included in the study:

If the study is an animal study, which animal and how relevant is that model?

Is the model system appropriate? Is there evidence from past use that it’s highly-relevant to therapeutic design in this context?

If it is a human study, what characteristics of the study population may support/limit extrapolation?

Summary of reliability

1-2 sentences on concluding remarks, including summary of strengths, weaknesses, limitations.

Progress

Check off the components as they are completed. If the component is not applicable, check the box as well.

15 Appendix D

Contributors were asked to complete this template to summarize and evaluate new papers related to topics besides therapeutics and diagnostics.

Title: Please edit the title to add the name of the paper after the colon.

General Information Please paste a link to the paper or a citation here:

Link:

What is the paper’s Manubot-style citation?

Citation:

Is this paper primarily relevant to Background or Pathogenesis?

Please list some keywords (3-10) that help identify the relevance of this paper to COVID-19

Please note the publication / review status

Which areas of expertise are particularly relevant to the paper?

Summary

Suggested questions to answer about each paper: - What did they analyze? - What methods did they use? - Does this paper study COVID-19, SARS-CoV-2, or a related disease and/or virus? - What is the main finding (or a few main takeaways)? - What does this paper tell us about the background and/or diagnostics/therapeutics for COVID-19 / SARS-CoV-2? - Do you have any concerns about methodology or the interpretation of these results beyond this analysis?

Any comments or notes?