Severe Acute Respiratory Syndrome–Associated Coronavirus 2 Infection and Organ Dysfunction in the ICU: Opportunities for Translational Research : Critical Care Explorations

Secondary Logo

Journal Logo

Review Article

Severe Acute Respiratory Syndrome–Associated Coronavirus 2 Infection and Organ Dysfunction in the ICU: Opportunities for Translational Research

Verhoef, Philip A. MD, PhD1,2; Kannan, Sujatha MD3; Sturgill, Jamie L. PhD4; Tucker, Elizabeth W. MD3; Morris, Peter E. MD4; Miller, Andrew C. MD5; Sexton, Travis R. PhD6; Koyner, Jay L. MD7; Hejal, Rana MD8; Brakenridge, Scott C. MD9; Moldawer, Lyle L. PhD9; Hotchkiss, Richard S. MD10–12; Blood, Teresa M. BS11; Mazer, Monty B. MD11; Bolesta, Scott PharmD13; Alexander, Sheila A. PhD, RN14; Armaignac, Donna Lee PhD15; Shein, Steven L. MD16; Jones, Christopher MD10; Hoemann, Caroline D. PhD17; Doctor, Allan MD18; Friess, Stuart H. MD19; Parker, Robert I. MD20; Rotta, Alexandre T. MD21; Remy, Kenneth E. MD, MHSC, MSCI10,11,19

Author Information
Critical Care Explorations 3(3):p e0374, March 2021. | DOI: 10.1097/CCE.0000000000000374
  • Open
  • COVID-19

Abstract

Objectives: 

Since the beginning of the coronavirus disease 2019 pandemic, hundreds of thousands of patients have been treated in ICUs across the globe. The severe acute respiratory syndrome–associated coronavirus 2 virus enters cells via the angiotensin-converting enzyme 2 receptor and activates several distinct inflammatory pathways, resulting in hematologic abnormalities and dysfunction in respiratory, cardiac, gastrointestinal renal, endocrine, dermatologic, and neurologic systems. This review summarizes the current state of research in coronavirus disease 2019 pathophysiology within the context of potential organ-based disease mechanisms and opportunities for translational research.

Data Sources: 

Investigators from the Research Section of the Society of Critical Care Medicine were selected based on expertise in specific organ systems and research focus. Data were obtained from searches conducted in Medline via the PubMed portal, Directory of Open Access Journals, Excerpta Medica database, Latin American and Caribbean Health Sciences Literature, and Web of Science from an initial search from December 2019 to October 15, 2020, with a revised search to February 3, 2021. The medRxiv, Research Square, and clinical trial registries preprint servers also were searched to limit publication bias.

Study Selection: 

Content experts selected studies that included mechanism-based relevance to the severe acute respiratory syndrome–associated coronavirus 2 virus or coronavirus disease 2019 disease.

Data Extraction: 

Not applicable.

Data Synthesis: 

Not applicable.

Conclusions: 

Efforts to improve the care of critically ill coronavirus disease 2019 patients should be centered on understanding how severe acute respiratory syndrome–associated coronavirus 2 infection affects organ function. This review articulates specific targets for further research.

The coronavirus disease 2019 (COVID-19) pandemic caused by the beta-coronavirus severe acute respiratory syndrome–associated coronavirus (SARS-CoV) 2 virus has resulted in more than 106 million cases and more than 2.32 million deaths (as of February 7, 2021) since emerging in December 2019 (1–3). Approximately, 20% of patients require hospitalization, and 25% of those require ICU admission due to refractory hypoxemia, shock, or multiple organ failure (MOF) (4–7). Numerous descriptive studies of COVID-19 pathophysiology, mechanism, signaling, and potential immunologic consequences have been published, but many reports lack integration and context. COVID-19 victims do not succumb from the SARS-CoV-2 infection per se; rather, mortality occurs due to pneumonia, renal failure, thrombosis, and MOF that reflect the innate immune signaling response to the infection. The international community has been much less focused on the ramifications of innate immune signaling activation in COVID-19 (8).

In addition, analysis of COVID-19 outcomes further exposes existing health disparities: Black and Hispanic Americans are disproportionately impacted in terms of morbidity and mortality. Additionally, males have demonstrated lower survival rates (9–13). Although these differences clearly indicate the need for evaluation of social determinants of health in COVID-19, they may also reveal previously unrecognized mechanistic and signaling pathways that could be targeted for potential therapeutics. This review summarizes the current mechanisms and reported pathophysiology of COVID-19 disease to characterize and understand the illness and treatment. Since many patients present with single organ disruption and/or current level of evidence has been explored in specific organs, we present the known data in this fashion. We also explore why organ disruption in the setting of COVID-19 (e.g., pneumonia, acute respiratory distress syndrome [ARDS], renal failure) and the response to COVID-19 therapy appear to frequently differ from classical forms of these pathologies.

OVERVIEW OF DISEASE TRANSMISSION AND CLINICAL FEATURES

SARS-CoV-2 is predominantly transmitted via respiratory droplets released by coughing or sneezing (14). Similar to its predecessor SARS-CoV-1 (2003 SARS outbreak), SARS-CoV-2 binds to the angiotensin-converting enzyme (ACE) 2 receptor via the virus Spike protein (15). Host organs expressing the ACE2 receptor that are targets for the virus Spike protein are presented in Figure 1. After binding, the furin domain of the Spike protein is cleaved by host proteases (e.g., serine protease transmembrane protease, serine [TMPRSS]2), which enables viral and cellular membrane fusion and subsequent internalization and release of the viral RNA (15). The ACE2 protein is widely expressed on respiratory epithelium, which is the presumed entry point, although ACE2 protein expression has been detected in many different organs (16).

F1
Figure 1.:
Tissues expressing angiotensin-2 receptor (angiotensin-converting enzyme [ACE] 2) and related COVID-19 symptoms. ARDS = acute respiratory distress syndrome, GI = gastrointestinal.

Infected individuals may remain asymptomatic or may develop a wide range of symptoms including fever, malaise, anosmia, hypogeusia, sore throat, headache, cough, shortness of breath, chest pain, nausea, abdominal pain, diarrhea, or cognitive changes. COVID-19 patients often present with lymphopenia, elevated nonspecific inflammatory markers (C-reactive protein, lactic acid dehydrogenase [LDH], d-dimer, ferritin, aspartate aminotransferase [AST]), and patchy/bilateral infiltrates on chest radiography. Patients who proceed to critical illness typically develop shortness of breath and hypoxemia within 5–8 days after symptom onset and are more likely to be older, blood type A, and have multiple comorbidities (7,17–19). Critically ill COVID-19 patients often require prolonged respiratory support and have high risk for MOF, thrombotic coagulopathies, acute kidney injury (AKI), sudden myocardial dysfunction, and prolonged hospitalization (20,21).

REPORTED HOST IMMUNE RESPONSES IN COVID-19

There are likely several SARS-CoV-2–induced inflammatory responses depending on host-pathogen interaction and disease evolution (summarized in Fig. 2). Similar to other viral infections, viral replication and cellular damage directly activate the host immune response, which contribute to the initial inflammatory response (22). Viral replication induces host cell death and the release of multiple danger-associated molecular patterns (DAMPs), which increase localized and systemic inflammation via proinflammatory cytokine and chemokine secretion (23). SARS-CoV-1 causes pyroptosis and activates the nucleotide-binding oligomerization domain, leucine rich repeat and pyrin domain containing-3 inflammasome, further activating a proinflammatory cytokine cascade (24). Thus, inflammasome activation may be one mechanism by which SARS-CoV-2 recruits host immune cells to the infection site, causing capillary leak, inflammatory infiltrates, and pulmonary edema. Recent studies also implicated neutrophil extracellular traps (NETs) in disease progression, although their impact as a driver of downstream inflammation remains unclear (25–27).

F2
Figure 2.:
Proposed host immune responses secondary to severe acute respiratory syndrome–associated coronavirus 2 (SARS-CoV-2) infection. Although several host immune responses are activated by SARS-CoV-2 infection, all mechanisms appear to activate janus kinase/signal transducer and activator of transcription (JAK/STAT), p38 mitogen-activated protein kinase (MAPK), and/or nuclear factor–κB (NF-κB) pathways. This leads to release of proinflammatory cytokines (tumor necrosis factor [TNF] α, interleukin [IL]–6, IL-1β) and chemokines (C-X-C motif ligand [CXCL] 1, CXCL10), and immune cell recruitment. Current evidence implicates three potential mechanisms. 1) Direct viral infection induces host cell death and the release of multiple damage/danger-associated molecular patterns (DAMPs), which increase proinflammatory cytokine and chemokine secretion. 2) Activation of NLRP3 inflammasome and pyroptosis triggers a cascade of proinflammatory cytokines such as IL-1β. 3) Viral infection dysregulates the renin-angiotensin system (RAS) by down-regulating the ACE2 receptor, which leads to decreased angiotensin (Ang) 1–7, increased AngII, and subsequent AngII-receptor type 1 (AT1R) activation. This ultimately activates JAK/STAT, p38 mitogen-activated protein kinase (MAPK), and NF-κB pathways and induces a proinflammatory state. CRP = C-reactive protein, NK = natural killer, NLRP3 = nucleotide-binding oligomerization domain, leucine rich repeat and pyrin domain containing-3.

Consensus is lacking on how to characterize the severity and nature of the inflammatory response. Several early case studies in COVID-19 reported markedly elevated levels of interleukin (IL)–1β, IL-6, IL-10, tumor necrosis factor (TNF)–α, and other mediators, leading many to characterize it as a “cytokine storm.” However, not every perturbation in a disease setting is maladaptive, and response severity measured at indiscriminate time points does not necessarily correlate with pathogenicity. Distinguishing between appropriate and dysregulated inflammatory responses in critically ill patients remains challenging. Most cytokines induce pleiotropic downstream effects with interdependent biological activities, and interactions among these mediators are neither linear nor uniform. The term “cytokine storm” implies that the elevated cytokine levels are necessarily injurious to host cells, and widespread acceptance of this term fueled the repurposing of many immunotherapy drugs to suppress various inflammatory pathways. One approach to understanding these responses has been to use the first 72 hours of temperature data from hospitalized COVID-19 patients as a surrogate for the inflammatory response, thereby allowing subphenotyping of patients (28,29). Perhaps not surprisingly, a subset of COVID-19 patients manifests a dysfunctional “hyperinflammatory” response with persistent fevers and elevated inflammatory markers, whereas a separate subset manifests a dysfunctional “hypoinflammatory” response, with an absence of fevers and worse MOF (29). Thus, any discussion of the immune response in COVID-19 must consider that although some patients mount an appropriate response and resolve the infection, others develop variably different dysfunctional responses which manifest as different COVID-19 subphenotypes.

Consistent with a hypoinflammatory subphenotype, a significant proportion of COVID-19 patients manifest a disease similar to immunoparalysis in sepsis, which involves decreased human leukocyte antigen (HLA)-DR expression and profound, persistent lymphopenia, including reductions in both CD4 and CD8 T cells. In addition, the remaining immune cells in many COVID-19 patients are functionally impaired: monocytes have been shown to release less TNF-α, whereas T cells release less IFN-γ and demonstrate increased PD-1 expression, consistent with an exhausted phenotype (30–33). The observed down-regulation of IL-2 and IL-7 in COVID-19 patients indicates an impairment in the development, differentiation, and homeostatic expansion of T cells (34). Such patients may benefit from immune stimulation, rather than immune suppression. Indeed, recent work suggests that treatment with the inhaled antiviral cytokine interferon-α2b results in accelerated viral clearance and reduced lung injury (35).

In contrast, the hyperinflammatory response subphenotype involves elevated levels of IL-6, IL-10, IL-8, and chemokines (e.g., C-X-C motif ligand [CXCL]–8, CXCL1, CXCL10, and C-C motif chemokine ligand-5) (21,34,36–40). CXCL10 reflects recent IFN-γ activity within 14 days, suggesting that the elevated CXCL10 levels at the time of ICU admission might be indicative of increased T helper 1, natural killer (NK), or natural killer T cell activity early in the illness, even though IFN-γ levels were not elevated among these patients (34). TNF-α levels range from low to normal or increased, indicating that the hyperinflammatory immune response is distinct from the classical cytokine storm associated with diseases like macrophage-activation syndrome in which TNF-α is both a lead cytokine and a therapeutic target (34). IL-1β levels are variable, reported as increased or decreased, and there are currently trials using anakinra, which blocks IL-1β activity, for treatment of the hyperinflammatory phase of COVID-19 (34).

Despite these overall patterns, IL-6 levels are orders of magnitude lower in patients with severe or critical COVID-19 disease (median 26–210 pg/mL) than in patients with non–COVID-19 ARDS (median 578–1,618 pg/mL) (21,36–43). Elevated IL-6 levels are needed to activate and potentiate the adaptive immune response and promote T cell regulation. By contrast, excessive IL-6 levels can block lymphopoiesis and induce lymphocyte death (44). All distinct lymphocyte subsets (NK cells, B cells, and T cells) may be affected by this innate overactivation (34). The degree of IL-6 elevation has been correlated with adverse outcomes in COVID-19 patients and has led to trials of anti–IL-6 therapy for COVID-19 patients, with variable success (45,46).

Other contributors to the hyperinflammatory state that are observed in many patients include dysregulation of the renin-angiotensin system (RAS) and its interaction with the host immune response. ACE2 regulates RAS homeostasis by cleaving angiotensin II (proinflammatory/profibrotic) into angiotensin 1–7, which leads to an anti-inflammatory/antifibrotic/antioxidant response (47–59). Angiotensin II is elevated in preclinical ARDS models and elicits proinflammatory effects by activating angiotensin II receptor type 1 (AT1R), nuclear factor–κB, janus kinase/signal transducer and activator of transcription, and p38 mitogen-activated protein kinase pathways; activation of these signaling pathways has stimulated interest in trials of kinase inhibitors for COVID-19. The cleaved peptide (angiotensin 1–7) reduces inflammation by binding the Mas receptor to antagonize AT1R (56). Previous studies report that SARS-CoV-1 infection down-regulates ACE2 in mice, which correlated with severe acute lung pathologies mediated by increased angiotensin II and AT1R activation (57). Angiotensin II was elevated in a small cohort of COVID-19 patients and appeared to correlate with viral load and lung injury, thus supporting the hypothesized role of RAS-immune axis dysregulation (60).

The role of ACE2 in COVID-19 pathophysiology led to hypotheses regarding the therapeutic use of ACE inhibitors (ACEIs) and angiotensin receptor blockers (ARBs) in patients with COVID-19. One hypothesis suggests that ACEI-mediated ACE2 inhibition could reduce morbidity and mortality by down-regulating RAS (57). By contrast, ACEIs, and ARBs could hypothetically increase ACE2 expression, leading to increased host susceptibility to viral invasion of target cells (52). Several small observational studies investigated the clinical outcomes of these hypotheses in patients with COVID-19, and current data suggest no benefit or harm from ACEIs and ARBs on risk of developing COVID-19 or clinical outcomes (61,62).

The primary site of SARS-CoV-2 infection in the lungs is type 2 alveolar epithelial cells (AT2 cells). Other cell types also express the ACE2 receptor and may contribute to the multiple organ inflammatory response observed in patients with COVID-19 disease. ACE2 expression is essentially ubiquitous and has been detected in the brain, heart, oral and nasal mucosa, nasopharynx, liver, kidney, lung, stomach, small intestine, colon, skin, lymph nodes, thymus, bone marrow, and spleen (16,58). ACE2 expression has been reported on lymphocytes, and SARS-CoV-2 viral particles can be detected in lymphocytes (63,64). Therefore, direct lymphocyte infection and cell death may be responsible for lymphopenia in COVID-19 patients and explain the mechanism by which cytokine release syndrome can develop in severe COVID-19 infections (65).

SARS-CoV-1 and SARS-CoV-2 have been isolated from macrophages, and viral infection of these cells in vitro appears to be associated with increased cytokine release (66,67). However, it remains difficult to distinguish these inflammatory mechanisms from direct cellular infection effects due to ubiquitous ACE2 expression on immune cells (i.e., macrophages and dendritic cells) and in multiple organs. Alternatively, MOF may be due to damage resulting from a combination of the systemic inflammatory response or ongoing immune suppression and persistent viral replication. Finally, the emerging postinfectious inflammatory syndrome in children (termed “multisystem inflammatory syndrome in children”) is likely mediated by inflammatory cytokines, although this remains an area of active exploration and will not be the subject of this review (68).

METHODS

We performed a narrative review of the literature retrieved from searches of Medline via the PubMed portal, Directory of Open Access Journals, Excerpta Medica database, Latin American and Caribbean Health Sciences Literature, and Web of Science originally from December 2019 to October 15, 2020, and revised twice to December 10, 2020, and February 3, 2021. We included data prior to December 2019 when relevant for the reported or proposed mechanisms in each organ system. Searches were not limited by date, language, or publication status. Publication bias was limited by searching clinical trial registries including ClinicalTrials.gov, World Health Organization International Clinical Trials Registry Platform, and the Australian New Zealand Clinical Trials Registry. We also searched the medRxiv and Research Square preprint servers to limit publication bias. Grey literature was eligible for inclusion if the authors responded affirmatively to correspondence with the requested information.

We used the following National Library of Medicine Mesh search terms: SARS-CoV-2 [Mesh], COVID-19 [Mesh], Middle East Respiratory Syndrome Coronavirus [Mesh], SARS virus [Mesh]; Respiratory Distress Syndrome [Mesh], and Pneumonia [Mesh]. Additional search terms included 2020 pandemic and individual organ system terms. Articles were not assessed using standardized and validated scales to assess risk-of-bias or evidence quality. The data were not pooled, so statistical analyses were not performed.

COVID-19 DISEASE MECHANISMS IN ORGAN FAILURE

Respiratory Manifestations and Failure

SARS-CoV-2 was first identified via a bronchoalveolar lavage sample from a patient with severe ARDS in Wuhan, China, indicating that the predominant location of viral replication and shedding is the respiratory tract (69). Patients with COVID-19 develop a wide spectrum of pulmonary diseases with varying severities, suggesting multiple mechanisms of disease propagation. Early evaluation of TMPRSS2 and ACE2 expression reported that both receptors were localized in nasal epithelial cells and AT2 cells (16,70). AT2 cells have a critical role in normal pulmonary physiology because they secrete surfactant and produce alveolar type 1 cells during acute lung injury (ALI). Children have lower illness severity than their adult counterparts, which may be partly explained by an age-dependent decrease in ACE2 expression in nasal epithelium (71). The exact mechanism of SARS-CoV-2 infection of AT2 cells is under investigation, although it is clear that AT2 viral infection leads to significant lung injury and may partly explain pulmonary sequela observed in patients.

SARS-CoV-2 is a promiscuous virus with multiple mammalian hosts and hijacks a variety of different receptors and enzymes to invade cells. SARS-CoV-2 enters cells (by exploiting Spike protein sites) via ACE2, dipeptidyl peptidase 4 (DPP4), CD147 (a highly glycosylated cell surface protein with wide tissue expression), or the cathepsin pathway to cleave and facilitate viral entry (72–75). Blocking DPP4 attenuates ALI in a murine model, and DPP4 has been studied in airway remodeling and fibrosis, whereas CD147 has been implicated in mucus hypersecretion (76–78). Thus, identifying all receptors/proteases involved in SARS-CoV-2 respiratory infection will enhance our mechanistic insight into disease pathogenesis.

Autopsy of COVID-19 patient lungs reveals several mechanistic clues for respiratory failure. First, there is evidence of increased angiogenesis based on imaging and gene expression data (79). Second, COVID-19 patients with ARDS display significant pulmonary fibrosis in the later disease stages, which impairs oxygenation (80). Third, significant deposition of microthrombi is common and may partly explain the proposed “L” (low lung weight, low elastance, and low inspiratory driving pressures) and “H” (high lung weight, high elastance, and high inspiratory driving pressures) ARDS phenotypes (81), although some experts have refuted the presence of the L phenotype. Fourth, COVID-19 patients are reported to have increased susceptibility to ventilator-associated pneumonia, for as yet undefined reasons compared with non–COVID-19 patients receiving invasive ventilation (82). Respiratory effects also occur along with MOF, profound alterations in coagulation, and hyperinflammatory cytokine profiles that are observed in other critical illnesses. Inflammation and tissue injury do not necessarily associate with the distribution of SARS-CoV-2 RNA and protein in post mortem examinations of lungs from COVID-19 patients, indicating that the inflammatory response itself is responsible for much of the respiratory failure (83).The extrapulmonary mechanisms by which SARS-CoV-2 infection triggers respiratory failure are discussed in subsequent sections.

Hematologic Dysfunction

COVID-19 disease significantly impacts the hematologic system and hemostasis. Prothrombin time (PT) and activated partial thromboplastin time are prolonged and d-dimer levels are elevated, suggesting the presence of a consumptive coagulopathy such as disseminated intravascular coagulation. Diffuse alveolar hemorrhage has been reported in COVID-19 patients, although a hypercoagulable phenotype with micro/macrothrombi formation in venous and arterial sites is more prevalent (12,64,65). Thrombi have been documented in the pulmonary, renal, brain, hepatic, and cardiac vasculature. Marked d-dimer elevation accompanied by normal or increased platelet count and fibrinogen levels (84–86) is predictive of severe disease in COVID-19 infection (62). This pattern, along with increased clot tensile strength without increased fibrinolysis on thromboelastomeric assays in COVID-19 patients (66), suggests that d-dimer elevation likely reflects the patient’s inflammatory response rather than consumptive coagulopathy. This is supported by differences in thrombi types between COVID-19 patients and other critically ill patients, and the reduced effectiveness of anticoagulation therapy (either prophylactic or therapeutic) in COVID-19–related thrombosis (87,88).

It remains unclear why COVID-19–related coagulopathy generates a thrombotic phenotype rather than a hemorrhagic phenotype. Recent data suggest that the thrombotic phenotype likely results from specific interactions between the host inflammatory response (e.g., complement) and activation of coagulation, platelets, and endothelial cells. SARS-CoV-2 infection promotes cytokine release through DAMPs, viral nucleic acid recognition, and RAS axis disruption. Reduced ACE2 levels may increase bradykinin levels (89), which may promote coagulopathy in at least two distinct ways: 1) enhancing complement activation (27,87) and 2) promoting neutrophil activation and NET formation. Inflammation generally activates the complement system, and subsequent depletion of plasma complement is associated with increasing disease severity and a prothrombotic state (Fig. 3) (90–94). Separately, bradykinin-induced neutrophil recruitment initiates a positive feedback pathway of thrombin-mediated platelet activation promoting NET formation, which further enhances inflammation and platelet activation and leads to NET colocalization with microthrombi (26,95). Inhibition of NET formation reduces ARDS progression and thrombus formation in non–COVID-19 ARDS (96), whereas increased NET formation is associated with ARDS and COVID-19 severity, supporting a role for NETs in increased thrombi risk in COVID-19 patients (88,97–99).

F3
Figure 3.:
Proposed mechanism for severe acute respiratory syndrome–associated coronavirus 2 (SARS-CoV-2)–induced coagulopathy and thrombotic microangiopathy. The SARS-CoV-2 virus directly injures cells/tissues and induces local and systemic inflammatory cascades, which induce the release of cytokines and damage/danger-associated molecular patterns (DAMPs) and activate three interconnected procoagulation pathways. Coagulation factor XII is activated during the contact phase of coagulation, thereby activating complement, thrombin, and a positive feedback loop for inflammation. Inflammation damages endothelial cells, further activating thrombin and polymorphonucleated cells (PMNs) via tumor necrosis factors (TNFs) and tissue factor (TF). Cytokines and DAMPs also directly activate PMNs, which initiates the development of NETosis and activates platelets. Activated platelets, neutrophil extracellular traps (NETs), and fibrin combine to form clinically significant clots in patients with coronavirus disease 2019 infections. CLOT = clot formation, EC = endothelial cell, FXII = factor XII, FXIIa = activated factor XII, HMWK = high molecular weight kininogen, NETosis = neutrophil extracellular traps, PK = prekallikrein, Plt = platelet.

SARS-CoV2 infection also elicits endotheliopathy, which likely has a role in thrombotic events (100). Several markers of platelet and endothelial activation/injury, including soluble P-selectin, soluble thrombomodulin, and von Willebrand factor (vWf), were higher in critically ill COVID-19 patients than in noncritically ill patients and healthy controls, suggesting that platelets and endothelial cells are involved in the infection pathophysiology. The frequency of thrombotic events is approximately nine-fold higher in COVID-19 patients with dyslipidemia than in those without (101,102). A recent report linked dyslipidemia, hypertension, and endotheliopathy and demonstrated that lipids (enhanced by low-density lipoproteins), rapid blood flow, and a disintegrin and metalloproteinase with a thrombospondin type 1 motif, member 13 (ADAMTS13) modulate the formation of secreted vWf into long fibrils tethered to endothelial cells (103,104). These studies suggest causative links between platelets, endothelial cells, and lipids in the venous and thrombotic events observed in COVID-19 patients and may explain the need for therapeutic anticoagulation.

Another relevant hematologic perturbation is an association between blood type and COVID-19 frequency and severity. A study of 2,173 COVID-19 patients reported that more COVID-19 cases had type A blood (38% vs 32% community control) than type O blood (26% vs 35% community control) (105). Although there may be a protective effect of anti-A antibodies, the type A antigen itself is likely to be involved in stabilizing vWf levels in blood, as blood group type A-vWf is more resistant to proteolysis than blood group type O-vWF (106,107). African Americans with type A blood have higher baseline vWf levels than Caucasians (108), which could increase the risk of thrombotic disorders. A/B blood types are similar across ethnic groups and do not fully account for ethnic disparities in COVID-19 disease. Ethnic groups most affected by COVID-19 are more likely to inherit null traits for three lesser-known blood antigens. 1) The Duffy antigen receptor for chemokines (DARC) is a glycoprotein receptor for Plasmodium falciparum and is expressed on erythrocytes (109). 2) The DARC-null phenotype confers resistance to malaria and is carried by 68% of African Americans and 5% of Hispanics (110) compared with 0% of Caucasians and 0% Chinese (109). DARC serves as a decoy receptor and is a “sink” for several chemokines; loss of the DARC receptor in COVID-19 could enable the cytokine storm and ARDS (111). 3) Lewis-null/Secretor-null individuals lack blood group system (AB) glycosylated antigens along epithelial linings in airways, gut, and kidney (109). Lewis antigens are fucosyl groups that are up-regulated in bronchial epithelium after inflammation (112) and have discrete roles in dendritic cell activation (113) and neutrophil transepithelial migration (110). Secretor status promotes some viral infections and inhibits others (109), although the effect on SARS-CoV-2 infection remains unclear. Lewis-null/Secretor-null is a rare phenotype that is three- to four-fold more frequent in African Americans and Brazilians than in Whites and Chinese (109). Thus, in addition to socioeconomic factors (e.g., healthcare access, nutrition, crowded housing, employment conditions) (114), DARC and Lewis-null/Secretor-null alleles may contribute biological factors that drive ethnic disparities in severe COVID-19.

Future research on complement, coagulation, platelet activation, NETs, thrombotic microangiopathy (TMA), vWf, ADAMTS13, RAS axis, bradykinin, and blood type in COVID-19 disease will enhance our understanding of how SARS-CoV-2 viral infection affects numerous tissues and organs through its effect on the vasculature. In addition, given recent reports about the potential protective effects of aspirin therapy, the effects of antiplatelet agents (including aspirin, colchicine or P2Y12 inhibitors) in preventing thrombotic events observed in COVID-19 patients should be further investigated (115).

Cardiac Manifestations

Early population studies on COVID-19 detected a disproportionate number of deaths in patients with preexisting cardiovascular disease (36,116). Patients with hypertension likely have RAS axis dysregulation, which may increase morbidity and mortality. There is evidence that elevated troponin and N-terminal pro-brain natriuretic peptide (NT-proBNP) levels are mortality risk factors in patients with and without preexisting cardiovascular disease (117). COVID-19 patients without preexisting cardiovascular disease can develop cardiovascular manifestations such as arrhythmias, myocarditis, and heart failure (117), and sudden cardiac failure can occur 1–3 weeks after initial infection, even while patients’ MOF is improving (116). These combined results indicate that SARS-CoV-2 infection impacts the cardiovascular system during both acute and convalescent phases.

A simple hypothesis of SARS-CoV-2–induced cardiac dysfunction is through direct infection of the myocardium and/or coronary endothelium. ACE2 and TMPRSS2 are expressed in both tissues; although the myocardium could theoretically act as a host for SARS-CoV-2, there are no documented reports myocardial infection (118,119). Viral RNA has been detected in the bloodstream, but infectious viral particles have not been isolated from blood, which may limit the potential for viral infection of myocardium or cardiac endothelium (69,120–124). There are reports of viral detection in myocardial tissue samples with possible myocarditis, although this could be explained by migration of infected macrophages from the lung into the myocardium (125–127).

An alternative hypothesis of SARS-CoV-2–induced cardiac dysfunction is that the proposed hypercoagulable state predisposes patients to develop coronary artery thromboses and subsequent sudden heart failure, with elevated troponin and NT-proBNP levels reflecting direct myocardial injury. Only two studies reported ST-segment elevation in COVID-19 patients (128,129), thus limiting current data in support of this hypothesis. However, myocardial infarctions are unlikely to be a common occurrence in spite of the high prevalence of hypercoagulability, as myocardial biomarkers are not detected at high frequency.

The prominent hyperinflammatory response may be responsible for the cardiac manifestations of COVID-19 disease. Proinflammatory cytokines decrease left ventricular function and lead to left ventricle dilation (130). Immune complexes and autoantibodies generated during hyperinflammatory states may precipitate acute cardiomyopathy (131). Hyperinflammatory states associated with recruitment of macrophage and T cells to the heart may cause substantial cardiac injury and the development of fulminant myocarditis (121,125,131,132). Research investigating the underlying mechanisms of hyperinflammation in COVID-19 should focus on its impact on cardiac function and the development of myocarditis.

Gastrointestinal and Hepatic Involvement

Various symptoms and laboratory abnormalities indicative of gastrointestinal and hepatobiliary involvement have been reported in COVID-19 patients. Reports of nausea, vomiting, diarrhea (up to 10%), and abdominal discomfort are common, and SARS-CoV-2 RNA is detectable in the stool (133–135). Up to 31% of patients with COVID-19-associated ARDS were found to have elevated lipase levels without pancreatitis, suggesting possible impairment in pancreatic microcirculation (136). Liver injury is more common in those with gastrointestinal symptoms and in the critically ill (133,137). Gastrointestinal symptoms correlate with more severe COVID-19 illness (133). Possible routes for COVID-19 involvement in gastrointestinal include trachea-esophagus-ileum-colon and disruption of the gut-liver and gut-pulmonary axes (133,138). ACE2 and TMPRSS are expressed by absorptive enterocytes in the small and large bowel and likely have mechanistic roles in gastrointestinal involvement (133,138). Once infected, enterocyte malabsorption may lead to increased gastrointestinal wall permeability and enteric symptoms such as diarrhea (139).

In normal liver tissue, ACE2 expression in bile duct epithelial cells (cholangiocytes) is higher than that in hepatocytes (16,133,140). Hepatocyte ACE2 expression increases in chronic liver diseases and during hypoxia (125,133). Isolated severe liver injury is rare and generally accompanies shock, respiratory failure, cardiac insufficiency, and/or renal insufficiency (137). Hepatocyte injury biomarkers (AST, ALT, LDH) may be increased, whereas liver synthetic function may be impaired as manifested by decreased albumin and increased PT (137).

Pathologic examination of COVID-19 patients may reveal moderate microvascular steatosis, mild sinusoidal dilatation, minimal lymphocytic infiltration, mild lobular and portal activity, multifocal hepatic necrosis, and rarely canalicular cholestasis (126,141–143). Intranuclear or intracytoplasmic viral inclusions have not been reported (126,141). These results suggest that COVID-19–mediated hepatic injury has multifactorial etiology that may differ in individual patients (144). Potential mechanisms include: 1) direct viral effects; 2) treatment-associated toxicity; 3) underlying disease predisposing to secondary injury; and 4) inflammatory response and other effects of severe illness (e.g., parenteral nutrition, hypotension, hypoxia, mechanical ventilation with high positive end-expiratory pressure [18–20 cm H2O]) (133,144,145). Identifying the subset of mechanisms that are operating for a given patient may enable personalization of therapies to optimize outcomes.

Renal Manifestations

The frequency of AKI in severe COVID-19 disease appears variable. Cohort studies using consensus definitions report AKI rates of 0–39%, and 5–19% of critically ill patients require renal replacement therapy (146–151). The underlying pathophysiology of COVID-19–associated AKI may share similarities with other forms of multifactorial AKI in the setting of critical illness (152,153). Current evidence from postmortem samples and limited biopsy series suggests a myriad of renal injuries (154–156), predominantly loss of proximal tubular brush border and acute tubular injury or the presence of inflammatory cells/infiltrates, viral particles in the parenchyma, and fibrin thrombi (in approximately 10% of samples). These results suggest a TMA/endothelial injury process (157) but do not identify a specific cause of AKI. In many cases, injury may be from hypotension-induced ischemia and hemodynamic sequelae of positive pressure ventilation that reduces renal perfusion, reduces cardiac output, and increases renal venous pressures (158–162). Tubular injury may stem from nephrotoxin exposure, which traditionally accounts for up to 25% of all hospital-based AKI. Given the COVID-19 polypharmacy in select scenarios, nephrotoxins undoubtedly complicate this issue (163).

There are conflicting reports regarding the virus in urine, although SARS-CoV-2 RNA/particles have been identified in urine (47,154,164,165). In early AKI with a slightly injured tubule, the virus Spike protein may bind the ACE2 receptor and enter renal tubular epithelial cells, where it can then replicate (55). Epithelial cell entry may explain some of the pathologically observed renal injury patterns (16,148). Genetic variations in the ACE2 receptor may contribute to the disproportionate impact of COVID-19 in certain subpopulations. Several reports suggested associations between COVID-19–mediated AKI and collapsing glomerulopathy (157,166–168). Collapsing glomerulopathy is characterized by tuft collapse with podocyte hypertrophy and intracytoplasmic protein resorption droplets. It has been associated with high-risk APOL1 alleles in individuals of African descent, thus providing a biologic factor that may explain ethnic disparities in COVID-19 outcomes (168,169). Collapsing glomerulopathy has been associated with other viral infections including HIV, Zika, and Dengue. The exact mechanisms and links of collapsing glomerulopathy to the APOL1 gene remain under intense investigation. Future investigations into COVID-19–associated AKI should examine the specific impact of ACE2 and apolipoprotein L1 on the development of severe AKI.

Neurologic Manifestations

CNS involvement of human coronavirus infections has been extensively documented (170–172). CNS manifestations of SARS-CoV-2 infection have been reported in 14–36% of hospitalized patients, including altered mental status, temporary anosmia, seizures, encephalitis, and ischemic stroke (173–175). These CNS pathologies may be related to direct viral neurotropism, CNS migration of infected peripheral immune cells, or secondary to systemic infection (176,177). Autopsy of COVID-19 patients revealed virus in brain tissue, primarily in neurons (64,178). Neurons and the capillary endothelium express ACE2 receptors, providing two possible routes for brain invasion (59). Intranasal inoculation of SARS-CoV-2 in transgenic mice expressing human ACE2 resulted in rapid and widespread infection of primarily neurons and astrocytes in the brain (179). Anosmia is reported in patients infected with SARS-CoV-1 and SARS-CoV-2, thus supporting this route of entry, but olfactory tract abnormalities have not been observed in MRI (174,180,181). Coronaviruses infect macrophages and glial cells, but there is currently little evidence that peripheral immune cell infection leads to CNS access (182,183). COVID-19–induced MOF may result in brain tissue hypoxia/ischemia and blood-brain barrier disruption, thereby potentially allowing SARS-CoV-2 access to the CNS via the blood stream, although further data are needed to support this hypothesis. Further work also is needed to optimize testing modalities and procedures as the duration of SARS-CoV-2 in cerebrospinal fluid after symptom onset is unknown.

Up to 31% of critically ill COVID-19 patients have thromboembolic complications (184,185). The frequency of ischemic stroke in COVID-19 patients is currently unknown, although one retrospective study reported an frequency as high as 2.8% in hospitalized COVID-19 patients with severe disease (173). A case series of six patients with COVID-19 and ischemic stroke reported that patients had moderate to critical disease and an 8–24-day interval from COVID-19 symptom onset to stroke symptoms (186). Possible indications and patient selection criteria for systemic anticoagulation therapy to reduce stroke risk in COVID-19 patients remain to be determined, along with any beneficial changes in the approach to systemic thrombolysis or endovascular therapy. Several studies reported that mechanically ventilated COVID-19 patients require unusually high levels of sedation with a markedly higher frequency of delirium (187–198). Increased delirium rates may be due to direct virus or inflammation-mediated effects, greater sedation requirements, or other yet unidentified factors.

Influenza pandemics before the 20th century were historically followed by increased reports of neuropsychiatric symptoms including psychosis, depression, anxiety, insomnia, and mania (199,200). Recent viral pandemics and epidemics (influenza H1N1, SARS-CoV-1, and Middle East respiratory syndrome–associated coronavirus) have been associated with neuropsychiatric sequelae including narcolepsy, seizures, and demyelinating processes (201–203). It will be crucial to institute longitudinal studies to assess COVID-19 association with long-term neuropsychiatric sequelae to determine the disease burden and identify possible therapeutic interventions, particularly given the emerging recognition of post-ICU syndrome and long-haul disease among COVID-19 survivors.

Endocrine Manifestations

Obesity and type 2 diabetes mellitus (DM) are important risk factors associated with COVID-19 severity, morbidity, and mortality (204–207). Adiposity promotes chronic low-grade inflammation and immune dysregulation, leading to a more robust inflammatory response, profound adaptive immune suppression, and reduced viral clearance during infection with SARS-CoV-2. Patients with obesity or type 2 DM may have a chronic endotheliopathy and prothrombotic state associated with dyslipidemia and high levels of circulating vWf, as noted above (208); when combined with COVID-19–mediated coagulation perturbations, the risk of cardiovascular and CNS thrombotic events increases (209).

Uncontrolled hyperglycemia is associated with poor outcomes in COVID-19 (209,210). This may result from hyperglycemia-associated glycosylation of ACE2 and viral Spike proteins that facilitate entry into target host cells, thus increasing viral loads and suppressing viral clearance. Studies on SARS-CoV-1 suggest that the virus may directly infiltrate ACE2-expressing pancreatic islet cells, thereby reducing insulin production and causing hyperglycemia (210). It is currently unknown whether islet cell damage is transient or permanent in COVID-19 survivors. Growing evidence suggests that SARS-CoV2 infection may trigger new onset diabetes, either via direct pancreatic invasion or due to the immune response to the virus (211–213).

Adrenal insufficiency is an uncommon early finding in COVID-19, and higher serum cortisol levels have been associated with increased mortality (214). Persistent fatigue and malaise following SARS-CoV-2 infection may be related to postinfection adrenal insufficiency. Direct viral CNS infiltration may enable infection of pituitary gland cells, leading to transient suppression of the hypothalamic-pituitary-adrenal axis. Alternatively, adrenal insufficiency may be caused by immune-mediated hypophysitis (215). Bilateral adrenal hemorrhage secondary to renal vein and adrenal plexus thrombosis was reported in a COVID-19 patient who had positive antiphospholipid antibodies (216). Systemic corticosteroid therapy has been quite successful in treating patients with severe COVID-19, which may reflect both the anti-inflammatory effects of steroids but also partial treatment of adrenal insufficiency.

Patients often develop low levels of thyroid-stimulating hormone and triiodothyronine during the early phases of critical COVID-19 illness, which typically normalize during recovery (215). Although this may represent euthyroid sick syndrome, there is evidence that SARS-CoV-1 virus can directly invade and affect thyroid tissue (205,217,218). This could explain late-onset subacute thyroiditis observed in COVID-19 patients, possibly secondary to immune-mediated processes that cause thyrotoxicosis to hypothyroidism. However, these mechanisms have yet to be elucidated (217,218).

Dermatologic Manifestations

Cutaneous manifestations have been reported in 5–20% of symptomatic COVID-19 patients (219–221). Erythematous macules or papules over the trunk and extremities, comparable with those observed in other viral exanthems, occur early in the infection and are the most common cutaneous manifestation. The rash may be pruritic, lasts for a median of 10 days, and generally is associated with a more favorable clinical course (222). The exanthem results from perivascular dermatitis and vasculitis with neutrophilic and lymphocytic infiltration as a response to the virus (223). Urticaria, with or without angioedema, comprises up to 20% of the rashes observed in COVID-19, particularly in those with more severe disease (224,225). Vesicular eruptions (similar to those of chickenpox) are associated with moderate disease severity, exhibit acantholysis with dyskeratosis, and may be due to a cytopathic effect of viral invasion (226). Purpura or petechiae are uncommon and have been reported in patients with thrombocytopenia and coagulopathy (221). Livedo racemosa resembles livedo reticularis but is more diffuse and likely secondary to small or medium vessel vasculopathy with occlusive microthrombi or immune complexes, complement deposition, and reperfusion injury. The term “COVID toes” is used to describe a chilblain-like lesion in COVID-19 patients. Unlike their idiopathic counterpart, COVID-related chilblains occur in the absence of exposure to cold. These acral lesions are often the only presenting symptom of COVID-19 infection and are likely caused by immune complex deposition or microthrombi (227–231). Two recent studies present comprehensive depictions of the most common cutaneous manifestations of COVID-19 (230,231).

FUTURE AREAS FOR RESEARCH AND CONCLUSIONS

The COVID-19 pandemic will persist well in to 2021 and beyond. Further investigations on the mechanisms underlying the frequency, prevention, and treatment of COVID-19–related organ failure are urgently needed. The current hypotheses of hyper- and hypoinflammation driving the clinical disease course do not facilitate consensus views for therapeutic approaches. Future research should use multiomic methods to refine descriptive findings and identify relevant signaling pathways. These studies should include patients with different age, sex, and race, as potential therapies may not be efficacious across heterogeneous populations. Although a vaccine holds the most promise, it is likely that adjunctive immunorestoration or antiviral therapies are needed to improve outcomes. These approaches should be based on specific immune signaling pathways rather than observed changes in nonspecific biomarkers. Given that there are several distinct strategies for vaccine development (messenger RNA, viral vector, adjuvanted protein, and inactivated virus), the impact of these different strategies on the immune response, as well as the impact on organ dysfunction among immunized individuals who become infected, remains an unanswered question. For example, given the now known early induction of cellular immune exhaustion by SARS-CoV-2 infection that limits T and B cell responses, vaccine development must include induction of both of these responses including antiviral CD8+ T cells to provide robust protection from severe disease (232). However, it is unclear how patients that have already acquired the disease when the optimal time for vaccination should occur to maximize protection? In addition, the recent emergence of SARS-CoV-2 variants raises additional translational research questions: what is the nature of the immune response triggered by such variants, and is their increased transmissibility a reflection of a distinct immune response? Do these variants result in differential organ dysfunction among infected patients? The critical care research community is well poised to merge basic science with translational, clinical, and big data approaches to tackle this devastating illness.

ACKNOWLEDGMENTS

We would like to thank the Society of Critical Care Medicine and the Research Section for assistance in developing this article.

REFERENCES

1. Park M, Cook AR, Lim JT, et al. A systematic review of COVID-19 epidemiology based on current evidence. J Clin Med. 2020; 9:967
2. Onder G, Rezza G, Brusaferro S. Case-fatality rate and characteristics of patients dying in relation to COVID-19 in Italy. JAMA. 2020; 323:1775–1776
3. Abdollahi E, Champredon D, Langley JM, et al. Temporal estimates of case-fatality rate for COVID-19 outbreaks in Canada and the United States. CMAJ. 2020; 192:E666–E670
4. Ahn DG, Shin HJ, Kim MH, et al. Current status of epidemiology, diagnosis, therapeutics, and vaccines for novel Coronavirus disease 2019 (COVID-19). J Microbiol Biotechnol. 2020; 30:313–324
5. Ciotti M, Angeletti S, Minieri M, et al. COVID-19 outbreak: An overview. Chemotherapy. 2020; 64:1–9
6. Du RH, Liu LM, Yin W, et al. Hospitalization and critical care of 109 decedents with COVID-19 pneumonia in Wuhan, China. Ann Am Thorac Soc. 2020; 17:839–846
7. Huang C, Wang Y, Li X, et al. Clinical features of patients infected with 2019 novel coronavirus in Wuhan, China. Lancet. 2020; 395:497–506
8. Barrett CD, Yaffe MB. COVID-19: All the wrong moves in all the wrong places. Sci Signal. 2020; 13:eabe4242
9. Yancy CW. COVID-19 and African Americans. JAMA. 2020; 323:1891–1892
10. Petrilli CM, Jones SA, Yang J, et al. Factors associated with hospital admission and critical illness among 5279 people with coronavirus disease 2019 in New York City: Prospective cohort study. BMJ. 2020; 369:m1966
11. Kullar R, Marcelin JR, Swartz TH, et al. Racial disparity of coronavirus disease 2019 in African American communities. J Infect Dis. 2020; 222:890–893
12. Holmes L Jr, Enwere M, Williams J, et al. Black-White risk differentials in COVID-19 (SARS-COV2) transmission, mortality and case fatality in the United States: Translational epidemiologic perspective and challenges. Int J Environ Res Public Health. 2020; 17:4322
13. Wortham JM, Lee JT, Althomsons S, et al. Characteristics of persons who died with COVID-19 - United States, February 12-May 18, 2020. MMWR Morb Mortal Wkly Rep. 2020; 69:923–929
14. Liu Y, Ning Z, Chen Y, et al. Aerodynamic analysis of SARS-CoV-2 in two Wuhan hospitals. Nature. 2020; 582:557–560
15. Hoffmann M, Kleine-Weber H, Schroeder S, et al. SARS-CoV-2 cell entry depends on ACE2 and TMPRSS2 and is blocked by a clinically proven protease inhibitor. Cell. 2020; 181:271–280.e8
16. Hamming I, Timens W, Bulthuis ML, et al. Tissue distribution of ACE2 protein, the functional receptor for SARS coronavirus. A first step in understanding SARS pathogenesis. J Pathol. 2004; 203:631–637
17. Guan WJ, Ni ZY, Hu Y, et al.; China Medical Treatment Expert Group for Covid-19. Clinical characteristics of coronavirus disease 2019 in China. N Engl J Med. 2020; 382:1708–1720
18. Grasselli G, Zangrillo A, Zanella A, et al.; COVID-19 Lombardy ICU Network. Baseline characteristics and outcomes of 1591 patients infected with SARS-CoV-2 admitted to ICUs of the Lombardy Region, Italy. JAMA. 2020; 323:1574–1581
19. Marinosci A, Landis BN, Calmy A. Possible link between anosmia and COVID-19: Sniffing out the truth. Eur Arch Otorhinolaryngol. 2020; 277:2149–2150
20. Bhatraju PK, Ghassemieh BJ, Nichols M, et al. Covid-19 in critically ill patients in the Seattle Region - case series. N Engl J Med. 2020; 382:2012–2022
21. Cummings MJ, Baldwin MR, Abrams D, et al. Epidemiology, clinical course, and outcomes of critically ill adults with COVID-19 in New York City: A prospective cohort study. Lancet. 2020; 395:1763–1770
22. Yuki K, Fujiogi M, Koutsogiannaki S. COVID-19 pathophysiology: A review. Clin Immunol. 2020; 215:108427
23. Tay MZ, Poh CM, Rénia L, et al. The trinity of COVID-19: Immunity, inflammation and intervention. Nat Rev Immunol. 2020; 20:363–374
24. Chen IY, Moriyama M, Chang MF, et al. Severe acute respiratory syndrome Coronavirus viroporin 3a activates the NLRP3 inflammasome. Front Microbiol. 2019; 10:50
25. Schönrich G, Raftery MJ, Samstag Y. Devilishly radical NETwork in COVID-19: Oxidative stress, neutrophil extracellular traps (NETs), and T cell suppression. Adv Biol Regul. 2020; 77:100741
26. Zuo Y, Yalavarthi S, Shi H, et al. Neutrophil extracellular traps in COVID-19. JCI Insight. 2020; 5:e138999
27. de Bont CM, Boelens WC, Pruijn GJM. NETosis, complement, and coagulation: a triangular relationship. Cell Mol Immunol. 2019; 16:19–27
28. Bhavani SV, Wolfe KS, Hrusch CL, et al. Temperature trajectory subphenotypes correlate with immune responses in patients with sepsis. Crit Care Med. 2020; 48:1645–1653
29. Bhavani SV, Huang ES, Verhoef PA, et al. Novel temperature trajectory subphenotypes in COVID-19. Chest. 2020; 158:2436–2439
30. Remy KE, Brakenridge SC, Francois B, et al. Immunotherapies for COVID-19: Lessons learned from sepsis. Lancet Respir Med. 2020; 8:946–949
31. Mudd PA, Crawford JC, Turner JS, et al. Distinct inflammatory profiles distinguish COVID-19 from influenza with limited contributions from cytokine storm. Sci Adv. 2020; 6:eabe3024
32. Remy KE, Mazer M, Striker DA, et al. Severe immunosuppression and not a cytokine storm characterizes COVID-19 infections. JCI Insight. 2020; 5:e140329
33. Jamilloux Y, Henry T, Belot A, et al. Should we stimulate or suppress immune responses in COVID-19? Cytokine and anti-cytokine interventions. Autoimmun Rev. 2020; 19:102567
34. Notz Q, Schmalzing M, Wedekink F, et al. Pro- and anti-inflammatory responses in severe COVID-19-induced acute respiratory distress syndrome-an observational pilot study. Front Immunol. 2020; 11:581338
35. Zhou Q, MacArthur MR, He X, et al. Interferon-alpha2b treatment for COVID-19 is associated with improvements in lung abnormalities. Viruses. 2020; 13:44
36. Zhou F, Yu T, Du R, et al. Clinical course and risk factors for mortality of adult inpatients with COVID-19 in Wuhan, China: A retrospective cohort study. Lancet. 2020; 395:1054–1062
37. Wu C, Chen X, Cai Y, et al. Risk factors associated with acute respiratory distress syndrome and death in patients with Coronavirus disease 2019 pneumonia in Wuhan, China. JAMA Intern Med. 2020; 180:934–943
38. Esen F, Ozcan PE, Orhun G, et al. Effects of adjunct treatment with intravenous Octagam on the course of severe COVID-19: Results from a retrospective cohort study. Curr Med Res Opin. 2020; 1:1–14
39. Mo P, Xing Y, Xiao Y, et al. Clinical characteristics of refractory COVID-19 pneumonia in Wuhan, China. Clin Infect Dis. 2020. Mar 16. [online ahead of print]
40. Qin C, Zhou L, Hu Z, et al. Dysregulation of immune response in patients with coronavirus 2019 (COVID-19) in Wuhan, China. Clin Infect Dis. 2020; 71:762–768
41. Calfee CS, Delucchi K, Parsons PE, et al.; NHLBI ARDS Network. Subphenotypes in acute respiratory distress syndrome: Latent class analysis of data from two randomised controlled trials. Lancet Respir Med. 2014; 2:611–620
42. Famous KR, Delucchi K, Ware LB, et al.; ARDS Network. Acute respiratory distress syndrome subphenotypes respond differently to randomized fluid management strategy. Am J Respir Crit Care Med. 2017; 195:331–338
43. Sinha P, Delucchi KL, Thompson BT, et al.; NHLBI ARDS Network. Latent class analysis of ARDS subphenotypes: A secondary analysis of the statins for acutely injured lungs from sepsis (SAILS) study. Intensive Care Med. 2018; 44:1859–1869
44. Maeda K, Baba Y, Nagai Y, et al. IL-6 blocks a discrete early step in lymphopoiesis. Blood. 2005; 106:879–885
45. Zhang P, Shi L, Xu J, et al. Elevated interleukin-6 and adverse outcomes in COVID-19 patients: A meta-analysis based on adjusted effect estimates. Immunogenetics. 2020; 72:431–437
46. Gupta S, Wang W, Hayek SS, et al.; STOP-COVID Investigators. Association between early treatment with tocilizumab and mortality among critically ill patients with COVID-19. JAMA Intern Med. 2021; 181:41–51
47. South AM, Tomlinson L, Edmonston D, et al. Controversies of renin-angiotensin system inhibition during the COVID-19 pandemic. Nat Rev Nephrol. 2020; 16:305–307
48. Sun P, Lu X, Xu C, et al. CD-sACE2 inclusion compounds: An effective treatment for coronavirus disease 2019 (COVID-19). J Med Virol. 2020; 92:1721–1723
49. Meng J, Xiao G, Zhang J, et al. Renin-angiotensin system inhibitors improve the clinical outcomes of COVID-19 patients with hypertension. Emerg Microbes Infect. 2020; 9:757–760
50. Chow JH, Mazzeffi MA, McCurdy MT. Angiotensin II for the treatment of COVID-19-related vasodilatory shock. Anesth Analg. 2020; 131:102–105
51. Danser AHJ, Epstein M, Batlle D. Renin-angiotensin system blockers and the COVID-19 pandemic: At present there is no evidence to abandon renin-angiotensin system blockers. Hypertension. 2020; 75:1382–1385
52. Patel AB, Verma A. COVID-19 and angiotensin-converting enzyme inhibitors and angiotensin receptor blockers: What is the evidence? JAMA. 2020; 323:1769–1770
53. Roncati L, Gallo G, Manenti A, et al. Renin-angiotensin system: The unexpected flaw inside the human immune system revealed by SARS-CoV-2. Med Hypotheses. 2020; 140:109686
54. Qiu Y, Zhao YB, Wang Q, et al. Predicting the angiotensin converting enzyme 2 (ACE2) utilizing capability as the receptor of SARS-CoV-2. Microbes Infect. 2020; 22:221–225
55. Othman H, Bouslama Z, Brandenburg JT, et al. Interaction of the spike protein RBD from SARS-CoV-2 with ACE2: Similarity with SARS-CoV, hot-spot analysis and effect of the receptor polymorphism. Biochem Biophys Res Commun. 2020; 527:702–708
56. Wang D, Chai XQ, Magnussen CG, et al. Renin-angiotensin-system, a potential pharmacological candidate, in acute respiratory distress syndrome during mechanical ventilation. Pulm Pharmacol Ther. 2019; 58:101833
57. Kuba K, Imai Y, Rao S, et al. A crucial role of angiotensin converting enzyme 2 (ACE2) in SARS coronavirus-induced lung injury. Nat Med. 2005; 11:875–879
58. Ferrario CM, Jessup J, Chappell MC, et al. Effect of angiotensin-converting enzyme inhibition and angiotensin II receptor blockers on cardiac angiotensin-converting enzyme 2. Circulation. 2005; 111:2605–2610
59. Harmer D, Gilbert M, Borman R, et al. Quantitative mRNA expression profiling of ACE 2, a novel homologue of angiotensin converting enzyme. FEBS Lett. 2002; 532:107–110
60. Liu Y, Yang Y, Zhang C, et al. Clinical and biochemical indexes from 2019-nCoV infected patients linked to viral loads and lung injury. Sci China Life Sci. 2020; 63:364–374
61. Mackey K, King VJ, Gurley S, et al. Risks and impact of angiotensin-converting enzyme inhibitors or angiotensin-receptor blockers on SARS-CoV-2 infection in adults: A living systematic review. Ann Intern Med. 2020; 173:195–203
62. Mancia G, Rea F, Ludergnani M, et al. Renin-angiotensin-aldosterone system blockers and the risk of Covid-19. N Engl J Med. 2020; 382:2431–2440
63. Xu H, Zhong L, Deng J, et al. High expression of ACE2 receptor of 2019-nCoV on the epithelial cells of oral mucosa. Int J Oral Sci. 2020; 12:1–5
64. Gu J, Gong E, Zhang B, et al. Multiple organ infection and the pathogenesis of SARS. J Exp Med. 2005; 202:415–424
65. Moore JB, June CH. Cytokine release syndrome in severe COVID-19. Science. 2020; 368:473–474
66. Cheung CY, Poon LL, Ng IH, et al. Cytokine responses in severe acute respiratory syndrome coronavirus-infected macrophages in vitro: Possible relevance to pathogenesis. J Virol. 2005; 79:7819–7826
67. Feng Z, Diao B, Wang R, et al. The novel severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) directly decimates human spleens and lymph nodes. medRxiv. 2020
68. Diorio C, Henrickson SE, Vella LA, et al. Multisystem inflammatory syndrome in children and COVID-19 are distinct presentations of SARS-CoV-2. J Clin Invest. 2020; 130:5967–5975
69. Zhu N, Zhang D, Wang W, et al.; China Novel Coronavirus Investigating and Research Team. A novel coronavirus from patients with pneumonia in China, 2019. N Engl J Med. 2020; 382:727–733
70. Sungnak W, Huang N, Bécavin C, et al.; HCA Lung Biological Network. SARS-CoV-2 entry factors are highly expressed in nasal epithelial cells together with innate immune genes. Nat Med. 2020; 26:681–687
71. Bunyavanich S, Do A, Vicencio A. Nasal gene expression of angiotensin-converting enzyme 2 in children and adults. JAMA. 2020; 323:2427–2429
72. Li Y, Zhang Z, Yang L, et al. The MERS-CoV receptor DPP4 as a candidate binding target of the SARS-CoV-2 spike. iScience. 2020; 23:101160
73. Strollo R, Pozzilli P. DPP4 inhibition: Preventing SARS-CoV-2 infection and/or progression of COVID-19? Diabetes Metab Res Rev. 2020; 36:e3330
74. Ulrich H, Pillat MM. CD147 as a target for COVID-19 treatment: Suggested effects of azithromycin and stem cell engagement. Stem Cell Rev Rep. 2020; 16:434–440
75. Ou X, Liu Y, Lei X, et al. Characterization of spike glycoprotein of SARS-CoV-2 on virus entry and its immune cross-reactivity with SARS-CoV. Nat Commun. 2020; 11:1620
76. Kawasaki T, Chen W, Htwe YM, et al. DPP4 inhibition by sitagliptin attenuates LPS-induced lung injury in mice. Am J Physiol Lung Cell Mol Physiol. 2018; 315:L834–L845
77. Sun J, Chu S, Lu M, et al. The roles of dipeptidyl peptidase-4 and its inhibitors in the regulation of airway epithelial-mesenchymal transition. Exp Lung Res. 2020; 46:163–173
78. Yu Q, Yang D, Chen X, et al. CD147 increases mucus secretion induced by cigarette smoke in COPD. BMC Pulm Med. 2019; 19:29
79. Ackermann M, Verleden SE, Kuehnel M, et al. Pulmonary vascular endothelialitis, thrombosis, and angiogenesis in Covid-19. N Engl J Med. 2020; 383:120–128
80. George PM, Wells AU, Jenkins RG. Pulmonary fibrosis and COVID-19: The potential role for antifibrotic therapy. Lancet Respir Med. 2020; 8:807–815
81. Bos LDJ, Sinha P, Dickson RP. The perils of premature phenotyping in COVID-19: A call for caution. Eur Respir J. 2020; 56:2001768
82. Maes M, Higginson E, Pereira-Dias J, et al. Ventilator-associated pneumonia in critically ill patients with COVID-19. Crit Care. 2021; 25:25
83. Dorward DA, Russell CD, Um IH, et al. Tissue-specific immunopathology in fatal COVID-19. Am J Respir Crit Care Med. 2021; 203:192–201
84. Terpos E, Ntanasis-Stathopoulos I, Elalamy I, et al. Hematological findings and complications of COVID-19. Am J Hematol. 2020; 95:834–847
85. Al-Samkari H, Karp Leaf RS, Dzik WH, et al. COVID-19 and coagulation: Bleeding and thrombotic manifestations of SARS-CoV-2 infection. Blood. 2020; 136:489–500
86. Mackman N, Antoniak S, Wolberg AS, et al. Coagulation abnormalities and thrombosis in patients infected with SARS-CoV-2 and other pandemic viruses. Arterioscler Thromb Vasc Biol. 2020; 40:2033–2044
87. Colling ME, Kanthi Y. COVID-19-associated coagulopathy: An exploration of mechanisms. Vasc Med. 2020; 25:471–478
88. Jayarangaiah A, Kariyanna PT, Chen X, et al. COVID-19-associated coagulopathy: An exacerbated immunothrombosis response. Clin Appl Thromb Hemost. 2020; 26:1076029620943293
89. van de Veerdonk FL, Netea MG, van Deuren M, et al. Kallikrein-kinin blockade in patients with COVID-19 to prevent acute respiratory distress syndrome. Elife. 2020; 9:e57555
90. Del Conde I, Crúz MA, Zhang H, et al. Platelet activation leads to activation and propagation of the complement system. J Exp Med. 2005; 201:871–879
91. Dzik S. Complement and coagulation: Cross talk through time. Transfus Med Rev. 2019; 33:199–206
92. Foley JH, Walton BL, Aleman MM, et al. Complement activation in arterial and venous thrombosis is mediated by plasmin. EBioMedicine. 2016; 5:175–182
93. Cugno M, Meroni PL, Gualtierotti R, et al. Complement activation in patients with COVID-19: A novel therapeutic target. J Allergy Clin Immunol. 2020; 146:215–217
94. Gavriilaki E, Brodsky RA. Severe COVID-19 infection and thrombotic microangiopathy: Success does not come easily. Br J Haematol. 2020; 189:e227–e230
95. Laridan E, Martinod K, De Meyer SF. Neutrophil extracellular traps in arterial and venous thrombosis. Semin Thromb Hemost. 2019; 45:86–93
96. Jiménez-Alcázar M, Rangaswamy C, Panda R, et al. Host DNases prevent vascular occlusion by neutrophil extracellular traps. Science. 2017; 358:1202–1206
97. Middleton EA, He XY, Denorme F, et al. Neutrophil extracellular traps contribute to immunothrombosis in COVID-19 acute respiratory distress syndrome. Blood. 2020; 136:1169–1179
98. Skendros P, Mitsios A, Chrysanthopoulou A, et al. Complement and tissue factor-enriched neutrophil extracellular traps are key drivers in COVID-19 immunothrombosis. J Clin Invest. 2020; 130:6151–6157
99. Leppkes M, Knopf J, Naschberger E, et al. Vascular occlusion by neutrophil extracellular traps in COVID-19. EBioMedicine. 2020; 58:102925
100. Goshua G, Pine AB, Meizlish ML, et al. Endotheliopathy in COVID-19-associated coagulopathy: Evidence from a single-centre, cross-sectional study. Lancet Haematol. 2020; 7:e575–e582
101. Mestre-Gómez B, Lorente-Ramos RM, Rogado J, et al.; Infanta Leonor Thrombosis Research Group. Incidence of pulmonary embolism in non-critically ill COVID-19 patients. Predicting factors for a challenging diagnosis. J Thromb Thrombolysis. 2021; 51:40–46
102. Franco-Moreno A, Muñoz-Rivas N, Mestre-Gómez B, et al. Pulmonary embolism and COVID-19: A paradigm change. Rev Clin Esp. 2020; 220:459–461
103. Zhou Z, Bernardo A, Zhu Q, et al. A G-quartet oligonucleotide blocks glycoprotein Ib-mediated platelet adhesion and aggregation under flow conditions. Thromb Haemost. 2009; 102:529–537
104. Shrimpton CN, Borthakur G, Larrucea S, et al. Localization of the adhesion receptor glycoprotein Ib-IX-V complex to lipid rafts is required for platelet adhesion and activation. J Exp Med. 2002; 196:1057–1066
105. Zhao J, Yang Y, Huang H, et al. Relationship between the ABO blood group and the COVID-19 susceptibility. Clin Infect Dis. 2020. Aug 4. [online ahead of print]
106. O’Donnell J, Boulton FE, Laffan MA. The relationship between plasma concentration of alpha2-macroglobulin and ABO blood group. Thromb Haemost. 2002; 88:167–168
107. Bowen DJ. An influence of ABO blood group on the rate of proteolysis of von Willebrand factor by ADAMTS13. J Thromb Haemost. 2003; 1:33–40
108. Miller CH, Haff E, Platt SJ, et al. Measurement of von Willebrand factor activity: Relative effects of ABO blood type and race. J Thromb Haemost. 2003; 1:2191–2197
109. Cooling L. Blood groups in infection and host susceptibility. Clin Microbiol Rev. 2015; 28:801–870
110. Brazil JC, Liu R, Sumagin R, et al. α3/4 Fucosyltransferase 3-dependent synthesis of Sialyl Lewis A on CD44 variant containing exon 6 mediates polymorphonuclear leukocyte detachment from intestinal epithelium during transepithelial migration. J Immunol. 2013; 191:4804–4817
111. Calfee CS, Ware LB, Glidden DV, et al.; National Heart, Blood, and Lung Institute Acute Respiratory Distress Syndrome Network. Use of risk reclassification with multiple biomarkers improves mortality prediction in acute lung injury. Crit Care Med. 2011; 39:711–717
112. Delmotte P, Degroote S, Lafitte JJ, et al. Tumor necrosis factor alpha increases the expression of glycosyltransferases and sulfotransferases responsible for the biosynthesis of sialylated and/or sulfated Lewis x epitopes in the human bronchial mucosa. J Biol Chem. 2002; 277:424–431
113. Fehres CM, van Beelen AJ, Bruijns SCM, et al. In situ delivery of antigen to DC-SIGN(+)CD14(+) dermal dendritic cells results in enhanced CD8(+) T-cell responses. J Invest Dermatol. 2015; 135:2228–2236
114. Ramasamy R, Milne K, Bell D, et al. Molecular mechanisms for thrombosis risk in Black people: A role in excess mortality from COVID-19. Br J Haematol. 2020; 190:e78–e80
115. Chow JH, Khanna AK, Kethireddy S, et al. Aspirin use is associated with decreased mechanical ventilation, ICU admission, and in-hospital mortality in hospitalized patients with COVID-19. Anesth Analg. 2020. Oct 21. [online ahead of print]
116. Ruan Q, Yang K, Wang W, et al. Clinical predictors of mortality due to COVID-19 based on an analysis of data of 150 patients from Wuhan, China. Intensive Care Med. 2020; 46:846–848
117. Guo T, Fan Y, Chen M, et al. Cardiovascular implications of fatal outcomes of patients with Coronavirus disease 2019 (COVID-19). JAMA Cardiol. 2020; 5:811–818
118. Chen L, Li X, Chen M, et al. The ACE2 expression in human heart indicates new potential mechanism of heart injury among patients infected with SARS-CoV-2. Cardiovasc Res. 2020; 116:1097–1100
119. Bertram S, Heurich A, Lavender H, et al. Influenza and SARS-coronavirus activating proteases TMPRSS2 and HAT are expressed at multiple sites in human respiratory and gastrointestinal tracts. PLoS One. 2012; 7:e35876
120. Akhmerov A, Marbán E. COVID-19 and the heart. Circ Res. 2020; 126:1443–1455
121. Bansal M. Cardiovascular disease and COVID-19. Diabetes Metab Syndr. 2020; 14:247–250
122. Hulot JS. COVID-19 in patients with cardiovascular diseases. Arch Cardiovasc Dis. 2020; 113:225–226
123. Fried JA, Ramasubbu K, Bhatt R, et al. The variety of cardiovascular presentations of COVID-19. Circulation. 2020; 141:1930–1936
124. Zheng YY, Ma YT, Zhang JY, et al. COVID-19 and the cardiovascular system. Nat Rev Cardiol. 2020; 17:259–260
125. Tavazzi G, Pellegrini C, Maurelli M, et al. Myocardial localization of coronavirus in COVID-19 cardiogenic shock. Eur J Heart Fail. 2020; 22:911–915
126. Xu Z, Shi L, Wang Y, et al. Pathological findings of COVID-19 associated with acute respiratory distress syndrome. Lancet Respir Med. 2020; 8:420–422
127. López-Boado MA, Colomer J, Targarona EM, et al. Free radical ablation prevents ischemic injury after long periods of cold storage in rat pancreas transplantation. Transplant Proc. 1990; 22:2241–2242
128. Dehghani P, Davidson LJ, Grines CL, et al. North American COVID-19 ST-segment-elevation myocardial infarction (NACMI) registry: Rationale, design, and implications. Am Heart J. 2020; 227:11–18
129. Gramegna M, Baldetti L, Beneduce A, et al. ST-segment-elevation myocardial infarction during COVID-19 pandemic: Insights from a regional public service healthcare hub. Circ Cardiovasc Interv. 2020; 13:e009413
130. Mann DL. Innate immunity and the failing heart: The cytokine hypothesis revisited. Circ Res. 2015; 116:1254–1268
131. Maisch B. Cardio-immunology of myocarditis: Focus on immune mechanisms and treatment options. Front Cardiovasc Med. 2019; 6:48
132. Carrillo-Salinas FJ, Ngwenyama N, Anastasiou M, et al. Heart inflammation: Immune cell roles and roads to the heart. Am J Pathol. 2019; 189:1482–1494
133. Kukla M, Skonieczna-Zydecka K, Kotfis K, et al. COVID-19, MERS and SARS with concomitant liver injury-systematic review of the existing literature. J Clin Med. 2020; 9:1420
134. Zippi M, Fiorino S, Occhigrossi G, et al. Hypertransaminasemia in the course of infection with SARS-CoV-2: Incidence and pathogenetic hypothesis. World J Clin Cases. 2020; 8:1385–1390
135. Xiao F, Tang M, Zheng X, et al. Evidence for gastrointestinal infection of SARS-CoV-2. Gastroenterology. 2020; 158:1831–1833.e3
136. Rasch S, Herner A, Schmid RM, et al. High lipasemia is frequent in Covid-19 associated acute respiratory distress syndrome. Pancreatology. 2021; 21:306–311
137. Hart SA, Tanel RE, Kipps AK, et al. Intensive care unit and acute care unit length of stay after congenital heart surgery. Ann Thorac Surg. 2020; 110:1396–1403
138. He Y, Wang Z, Li F, et al. Public health might be endangered by possible prolonged discharge of SARS-CoV-2 in stool. J Infect. 2020; 80:e18–e19
139. Gu J, Han B, Wang J. COVID-19: Gastrointestinal manifestations and potential fecal-oral transmission. Gastroenterology. 2020; 158:1518–1519
140. Guan GW, Gao L, Wang JW, et al. [Exploring the mechanism of liver enzyme abnormalities in patients with novel coronavirus-infected pneumonia]. Zhonghua Gan Zang Bing Za Zhi. 2020; 28:100–106
141. Zhang Y, Zheng L, Liu L, et al. Liver impairment in COVID-19 patients: A retrospective analysis of 115 cases from a single centre in Wuhan city, China. Liver Int. 2020; 40:2095–2103
142. Yao XH, Li TY, He ZC, et al. [A pathological report of three COVID-19 cases by minimal invasive autopsies]. Zhonghua Bing Li Xue Za Zhi. 2020; 49:411–417
143. Tian S, Xiong Y, Liu H, et al. Pathological study of the 2019 novel coronavirus disease (COVID-19) through postmortem core biopsies. Mod Pathol. 2020; 33:1007–1014
144. Li Y, Xiao SY. Hepatic involvement in COVID-19 patients: Pathology, pathogenesis, and clinical implications. J Med Virol. 2020; 92:1491–1494
145. Cardoso FS, Pereira R, Germano N. Liver injury in critically ill patients with COVID-19: A case series. Crit Care. 2020; 24:190
146. Li X, Wang L, Yan S, et al. Clinical characteristics of 25 death cases with COVID-19: A retrospective review of medical records in a single medical center, Wuhan, China. Int J Infect Dis. 2020; 94:128–132
147. Wang R, Liao C, He H, et al. COVID-19 in hemodialysis patients: A report of 5 cases. Am J Kidney Dis. 2020; 76:141–143
148. Pan XW, Xu D, Zhang H, et al. Identification of a potential mechanism of acute kidney injury during the COVID-19 outbreak: A study based on single-cell transcriptome analysis. Intensive Care Med. 2020; 46:1114–1116
149. Henry BM, Lippi G. Chronic kidney disease is associated with severe Coronavirus disease 2019 (COVID-19) infection. Int Urol Nephrol. 2020; 52:1193–1194
150. Shi S, Qin M, Shen B, et al. Association of cardiac injury with mortality in hospitalized patients with COVID-19 in Wuhan, China. JAMA Cardiol. 2020; 5:802–810
151. Chan L, Chaudhary K, Saha A, et al. Acute kidney injury in hospitalized patients with COVID-19. medRxiv. 2020
152. Murray PT, Mehta RL, Shaw A, et al.; ADQI 10 workgroup. Potential use of biomarkers in acute kidney injury: Report and summary of recommendations from the 10th acute dialysis quality initiative consensus conference. Kidney Int. 2014; 85:513–521
153. Endre ZH, Kellum JA, Di Somma S, et al. Differential diagnosis of AKI in clinical practice by functional and damage biomarkers: Workgroup statements from the tenth acute dialysis quality initiative consensus conference. Contrib Nephrol. 2013; 182:30–44
154. Su H, Yang M, Wan C, et al. Renal histopathological analysis of 26 postmortem findings of patients with COVID-19 in China. Kidney Int. 2020; 98:219–227
155. Wu H, Larsen CP, Hernandez-Arroyo CF, et al. AKI and collapsing glomerulopathy associated with COVID-19 and APOL1 high-risk genotype. J Am Soc Nephrol. 2020; 31:1688–1695
156. Kudose S, Batal I, Santoriello D, et al. Kidney biopsy findings in patients with COVID-19. J Am Soc Nephrol. 2020; 31:1959–1968
157. Sharma P, Uppal NN, Wanchoo R, et al.; Northwell Nephrology COVID-19 Research Consortium. COVID-19-associated kidney injury: A case series of kidney biopsy findings. J Am Soc Nephrol. 2020; 31:1948–1958
158. Mullins RJ, Dawe EJ, Lucas CE, et al. Mechanisms of impaired renal function with PEEP. J Surg Res. 1984; 37:189–196
159. Kaukinen S, Eerola R. Positive end expiratory pressure ventilation, renal function and renin. Ann Clin Res. 1979; 11:58–62
160. Qvist J, Pontoppidan H, Wilson RS, et al. Hemodynamic responses to mechanical ventilation with PEEP: The effect of hypervolemia. Anesthesiology. 1975; 42:45–55
161. Doty JM, Saggi BH, Blocher CR, et al. Effects of increased renal parenchymal pressure on renal function. J Trauma. 2000; 48:874–877
162. Koyner JL, Murray PT. Mechanical ventilation and lung-kidney interactions. Clin J Am Soc Nephrol. 2008; 3:562–570
163. Mehta RL, Awdishu L, Davenport A, et al. Phenotype standardization for drug-induced kidney disease. Kidney Int. 2015; 88:226–234
164. Wölfel R, Corman VM, Guggemos W, et al. Virological assessment of hospitalized patients with COVID-2019. Nature. 2020; 581:465–469
165. Sun J, Zhu A, Li H, et al. Isolation of infectious SARS-CoV-2 from urine of a COVID-19 patient. Emerg Microbes Infect. 2020; 9:991–993
166. Peleg Y, Kudose S, D’Agati V, et al. Acute kidney injury due to collapsing glomerulopathy following COVID-19 infection. Kidney Int Rep. 2020; 5:940–945
167. Larsen CP, Bourne TD, Wilson JD, et al. Collapsing glomerulopathy in a patient with COVID-19. Kidney Int Rep. 2020; 5:935–939
168. Nasr SH, Kopp JB. COVID-19-associated collapsing glomerulopathy: An emerging entity. Kidney Int Rep. 2020; 5:759–761
169. Genovese G, Friedman DJ, Ross MD, et al. Association of trypanolytic ApoL1 variants with kidney disease in African Americans. Science. 2010; 329:841–845
170. Arbour N, Day R, Newcombe J, et al. Neuroinvasion by human respiratory coronaviruses. J Virol. 2000; 74:8913–8921
171. Lau KK, Yu WC, Chu CM, et al. Possible central nervous system infection by SARS Coronavirus. Emerg Infect Dis. 2004; 10:342–344
172. Yeh EA, Collins A, Cohen ME, et al. Detection of coronavirus in the central nervous system of a child with acute disseminated encephalomyelitis. Pediatrics. 2004; 113:e73–e76
173. Mao L, Jin H, Wang M, et al. Neurologic manifestations of hospitalized patients with Coronavirus disease 2019 in Wuhan, China. JAMA Neurol. 2020; 77:683–690
174. Beltrán-Corbellini Á, Chico-García JL, Martínez-Poles J, et al. Acute-onset smell and taste disorders in the context of COVID-19: A pilot multicentre polymerase chain reaction based case-control study. Eur J Neurol. 2020; 27:1738–1741
175. Frontera JA, Sabadia S, Lalchan R, et al. A prospective study of neurologic disorders in hospitalized patients with COVID-19 in New York city. Neurology. 2021; 96:e575–e586
176. Bohmwald K, Gálvez NMS, Ríos M, et al. Neurologic alterations due to respiratory virus infections. Front Cell Neurosci. 2018; 12:386
177. Desforges M, Le Coupanec A, Dubeau P, et al. Human Coronaviruses and other respiratory viruses: Underestimated opportunistic pathogens of the central nervous system? Viruses. 2019; 12:14
178. Ding Y, He L, Zhang Q, et al. Organ distribution of severe acute respiratory syndrome (SARS) associated coronavirus (SARS-CoV) in SARS patients: Implications for pathogenesis and virus transmission pathways. J Pathol. 2004; 203:622–630
179. Netland J, Meyerholz DK, Moore S, et al. Severe acute respiratory syndrome coronavirus infection causes neuronal death in the absence of encephalitis in mice transgenic for human ACE2. J Virol. 2008; 82:7264–7275
180. Galougahi MK, Ghorbani J, Bakhshayeshkaram M, et al. Olfactory bulb magnetic resonance imaging in SARS-CoV-2-induced anosmia: The first report. Acad Radiol. 2020; 27:892–893
181. Hwang CS. Olfactory neuropathy in severe acute respiratory syndrome: Report of A case. Acta Neurol Taiwan. 2006; 15:26–28
182. Li YC, Bai WZ, Hashikawa T. Response to commentary on “The neuroinvasive potential of SARS-CoV-2 may play a role in the respiratory failure of COVID-19 patients”. J Med Virol. 2020; 92:707–709
183. Qin C, Zhou L, Hu Z, et al. Clinical characteristics and outcomes of COVID-19 patients with a history of stroke in Wuhan, China. Stroke. 2020; 51:2219–2223
184. Klok FA, Kruip MJHA, van der Meer NJM, et al. Incidence of thrombotic complications in critically ill ICU patients with COVID-19. Thromb Res. 2020; 191:145–147
185. Lodigiani C, Iapichino G, Carenzo L, et al.; Humanitas COVID-19 Task Force. Venous and arterial thromboembolic complications in COVID-19 patients admitted to an academic hospital in Milan, Italy. Thromb Res. 2020; 191:9–14
186. Beyrouti R, Adams ME, Benjamin L, et al. Characteristics of ischaemic stroke associated with COVID-19. J Neurol Neurosurg Psychiatry. 2020; 91:889–891
187. Hanidziar D, Bittner EA. Sedation of mechanically ventilated COVID-19 patients: Challenges and special considerations. Anesth Analg. 2020; 131:e40–e41
188. Khan SH, Lindroth H, Hendrie K, et al. Time trends of delirium rates in the intensive care unit. Heart Lung. 2020; 49:572–577
189. Debnath M, Berk M, Maes M. Changing dynamics of psychoneuroimmunology during the COVID-19 pandemic. Brain Behav Immun Health. 2020; 5:100096
190. Garg RK, Paliwal VK, Gupta A. Encephalopathy in patients with COVID-19: A review. J Med Virol. 2021; 93:206–222
191. Fotuhi M, Mian A, Meysami S, et al. Neurobiology of COVID-19. J Alzheimers Dis. 2020; 76:3–19
192. Cipriani G, Danti S, Nuti A, et al. A complication of coronavirus disease 2019: Delirium. Acta Neurol Belg. 2020; 120:927–932
193. Dinakaran D, Manjunatha N, Naveen Kumar C, et al. Neuropsychiatric aspects of COVID-19 pandemic: A selective review. Asian J Psychiatr. 2020; 53:102188
194. Bianchetti A, Rozzini R, Guerini F, et al. Clinical presentation of COVID19 in dementia patients. J Nutr Health Aging. 2020; 24:560–562
195. Beach SR, Praschan NC, Hogan C, et al. Delirium in COVID-19: A case series and exploration of potential mechanisms for central nervous system involvement. Gen Hosp Psychiatry. 2020; 65:47–53
196. Meagher D, Adamis D, Timmons S, et al. Developing a guidance resource for managing delirium in patients with Covid-19. Ir J Psychol Med. 2020. May 28: 1–6. [online ahead of print]
197. O’Hanlon S, Inouye SK. Delirium: Amissing piece in the COVID-19 pandemic puzzle. Age Ageing. 2020; 49:497–498
198. Kotfis K, Williams Roberson S, Wilson JE, et al. COVID-19: ICU delirium management during SARS-CoV-2 pandemic. Crit Care. 2020; 24:176
199. Menninger KA. Influenza and schizophrenia - An analysis of post influenzal “dementia precox” as of 1918, and five years later - Further studies of the psychiatric aspects of influenza. Am J Psychiatry. 1926; 5:469–529
200. Honigsbaum M. “An inexpressible dread”: Psychoses of influenza at fin-de-siècle. Lancet. 2013; 381:988–989
201. Manjunatha N, Math SB, Kulkarni GB, et al. The neuropsychiatric aspects of influenza/swine flu: A selective review. Ind Psychiatry J. 2011; 20:83–90
202. Kim JE, Heo JH, Kim HO, et al. Neurological complications during treatment of Middle East respiratory syndrome. J Clin Neurol. 2017; 13:227–233
203. Wu H, Zhuang J, Stone WS, et al. Symptoms and occurrences of narcolepsy: A retrospective study of 162 patients during a 10-year period in eastern China. Sleep Med. 2014; 15:607–613
204. Rottoli M, Bernante P, Belvedere A, et al. How important is obesity as a risk factor for respiratory failure, intensive care admission and death in hospitalised COVID-19 patients? Results from a single Italian centre. Eur J Endocrinol. 2020; 183:389–397
205. Marazuela M, Giustina A, Puig-Domingo M. Endocrine and metabolic aspects of the COVID-19 pandemic. Rev Endocr Metab Disord. 2020; 21:495–507
206. Sattar N, McInnes IB, McMurray JJV. Obesity is a risk factor for severe COVID-19 infection: Multiple potential mechanisms. Circulation. 2020; 142:4–6
207. de Lucena TMC, da Silva Santos AF, de Lima BR, et al. Mechanism of inflammatory response in associated comorbidities in COVID-19. Diabetes Metab Syndr. 2020; 14:597–600
208. Amraei R, Rahimi N. COVID-19, renin-angiotensin system and endothelial dysfunction. Cells. 2020; 9:1652
209. Bansal R, Gubbi S, Muniyappa R. Metabolic syndrome and COVID 19: Endocrine-immune-vascular interactions shapes clinical course. Endocrinology. 2020; 161:bqaa112
210. Pal R, Bhansali A. COVID-19, diabetes mellitus and ACE2: The conundrum. Diabetes Res Clin Pract. 2020; 162:108132
211. Ebekozien OA, Noor N, Gallagher MP, et al. Type 1 diabetes and COVID-19: Preliminary findings from a Multicenter Surveillance Study in the U.S. Diabetes Care. 2020; 43:e83–e85
212. Mongioì LM, Barbagallo F, Condorelli RA, et al. Possible long-term endocrine-metabolic complications in COVID-19: Lesson from the SARS model. Endocrine. 2020; 68:467–470
213. Caruso D, Zerunian M, Polici M, et al. Chest CT features of COVID-19 in Rome, Italy. Radiology. 2020; 296:E79–E85
214. Tan T, Khoo B, Mills EG, et al. Association between high serum total cortisol concentrations and mortality from COVID-19. Lancet Diabetes Endocrinol. 2020; 8:659–660
215. Somasundaram NP, Ranathunga I, Ratnasamy V, et al. The impact of SARS-Cov-2 virus infection on the endocrine system. J Endocr Soc. 2020; 4:bvaa082
216. Frankel M, Feldman I, Levine M, et al. Bilateral adrenal hemorrhage in Coronavirus disease 2019 patient: A case report. J Clin Endocrinol Metab. 2020; 105:dgaa487
217. Bellastella G, Maiorino MI, Esposito K. Endocrine complications of COVID-19: What happens to the thyroid and adrenal glands? J Endocrinol Invest. 2020; 43:1169–1170
218. Brancatella A, Ricci D, Viola N, et al. Subacute thyroiditis after SARS-COV-2 infection. J Clin Endocrinol Metab. 2020; 105:dgaa276
219. Recalcati S. Cutaneous manifestations in COVID-19: A first perspective. J Eur Acad Dermatol Venereol. 2020; 34:e212–e213
220. Hedou M, Carsuzaa F, Chary E, et al. Comment on ‘cutaneous manifestations in COVID-19: A first perspective’ by Recalcati S. J Eur Acad Dermatol Venereol. 2020; 34:e299–e300
221. De Giorgi V, Recalcati S, Jia Z, et al. Cutaneous manifestations related to coronavirus disease 2019 (COVID-19): A prospective study from China and Italy. J Am Acad Dermatol. 2020; 83:674–675
222. Gupta A, Madhavan MV, Sehgal K, et al. Extrapulmonary manifestations of COVID-19. Nat Med. 2020; 26:1017–1032
223. Gianotti R, Veraldi S, Recalcati S, et al. Cutaneous clinico-pathological findings in three COVID-19-positive patients observed in the metropolitan area of Milan, Italy. Acta Derm Venereol. 2020; 100:adv00124
224. Kaya G, Kaya A, Saurat JH. Clinical and histopathological features and potential pathological mechanisms of skin lesions in COVID-19: Review of the literature. Dermatopathology (Basel). 2020; 7:3–16
225. Gottlieb M, Long B. Dermatologic manifestations and complications of COVID-19. Am J Emerg Med. 2020; 38:1715–1721
226. Mahé A, Birckel E, Merklen C, et al. Histology of skin lesions establishes that the vesicular rash associated with COVID-19 is not ‘varicella-like’. J Eur Acad Dermatol Venereol. 2020; 34:e559–e561
227. de Masson A, Bouaziz JD, Sulimovic L, et al.; SNDV (French National Union of Dermatologists-Venereologists). Chilblains is a common cutaneous finding during the COVID-19 pandemic: A retrospective nationwide study from France. J Am Acad Dermatol. 2020; 83:667–670
228. Freeman EE, McMahon DE, Lipoff JB, et al.; American Academy of Dermatology Ad Hoc Task Force on COVID-19. Pernio-like skin lesions associated with COVID-19: A case series of 318 patients from 8 countries. J Am Acad Dermatol. 2020; 83:486–492
229. Fernandez-Nieto D, Jimenez-Cauhe J, Suarez-Valle A, et al. Characterization of acute acral skin lesions in nonhospitalized patients: A case series of 132 patients during the COVID-19 outbreak. J Am Acad Dermatol. 2020; 83:e61–e63
230. Galván Casas C, Català A, Carretero Hernández G, et al. Classification of the cutaneous manifestations of COVID-19: A rapid prospective nationwide consensus study in Spain with 375 cases. Br J Dermatol. 2020; 183:71–77
231. Ortega-Quijano D, Jimenez-Cauhe J, Selda-Enriquez G, et al. Algorithm for the classification of COVID-19 rashes. J Am Acad Dermatol. 2020; 83:e103–e104
232. Mudd PA, Remy KE. Prolonged adaptive immune activation in COVID-19: Implications for maintenance of long-term immunity? J Clin Invest. 2021; 131:e143928
Keywords:

angiotensin-converting enzyme 2 receptor; coronavirus disease 2019; critical illness; pandemic; severe acute respiratory syndrome–associated coronavirus 2

Copyright © 2021 The Authors. Published by Wolters Kluwer Health, Inc. on behalf of the Society of Critical Care Medicine.