Παρασκευή 20 Μαρτίου 2020

51.
 2020 Feb 24;9. pii: e55570. doi: 10.7554/eLife.55570.

Estimated effectiveness of symptom and risk screening to prevent the spread of COVID-19.

Abstract

Traveller screening is being used to limit further spread of COVID-19 following its recent emergence, and symptom screening has become a ubiquitous tool in the global response. Previously, we developed a mathematical model to understand factors governing the effectiveness of traveller screening to prevent spread of emerging pathogens (Gostic et al., 2015). Here, we estimate the impact of different screening programs given current knowledge of key COVID-19 life history and epidemiological parameters. Even under best-case assumptions, we estimate that screening will miss more than half of infected people. Breaking down the factors leading to screening successes and failures, we find that most cases missed by screening are fundamentally undetectable, because they have not yet developed symptoms and are unaware they were exposed. Our work underscores the need for measures to limit transmission by individuals who become ill after being missed by a screening program. These findings can support evidence-based policy to combat the spread of COVID-19, and prospective planning to mitigate future emerging pathogens.

KEYWORDS:

COVID-19; SARS-CoV-2; emerging infectious disease; epidemic containment; epidemic control; epidemiology; global health; human; travel screening; virus
PMID:
 
32091395
 
PMCID:
 
PMC7060038
 
DOI:
 
10.7554/eLife.55570
[Indexed for MEDLINE] 
Free PMC Article
Icon for eLife Sciences Publications, LtdIcon for PubMed Central
52.
 2020 Jan 31;14(1):3-17. doi: 10.3855/jidc.12425.

2019-nCoV (Wuhan virus), a novel Coronavirus: human-to-human transmission, travel-related cases, and vaccine readiness.

Abstract

On 31 December 2019 the Wuhan Health Commission reported a cluster of atypical pneumonia cases that was linked to a wet market in the city of Wuhan, China. The first patients began experiencing symptoms of illness in mid-December 2019. Clinical isolates were found to contain a novel coronavirus with similarity to bat coronaviruses. As of 28 January 2020, there are in excess of 4,500 laboratory-confirmed cases, with > 100 known deaths. As with the SARS-CoV, infections in children appear to be rare. Travel-related cases have been confirmed in multiple countries and regions outside mainland China including Germany, France, Thailand, Japan, South Korea, Vietnam, Canada, and the United States, as well as Hong Kong and Taiwan. Domestically in China, the virus has also been noted in several cities and provinces with cases in all but one provinence. While zoonotic transmission appears to be the original source of infections, the most alarming development is that human-to-human transmission is now prevelant. Of particular concern is that many healthcare workers have been infected in the current epidemic. There are several critical clinical questions that need to be resolved, including how efficient is human-to-human transmission? What is the animal reservoir? Is there an intermediate animal reservoir? Do the vaccines generated to the SARS-CoV or MERS-CoV or their proteins offer protection against 2019-nCoV? We offer a research perspective on the next steps for the generation of vaccines. We also present data on the use of in silico docking in gaining insight into 2019-nCoV Spike-receptor binding to aid in therapeutic development. Diagnostic PCR protocols can be found at https://www.who.int/health-topics/coronavirus/laboratory-diagnostics-for-novel-coronavirus.

KEYWORDS:

2019-nCoV; Wuhan; coronavirus; human-to-human transmission; vaccine readiness
PMID:
 
32088679
 
DOI:
 
10.3855/jidc.12425
[Indexed for MEDLINE] 
Free full text
Icon for The Journal of Infection in Developing Countries
53.
 2020 Jan 31;14(1):1-2. doi: 10.3855/jidc.12496.

Fear of the novel coronavirus.

KEYWORDS:

2019-nCoV; china; outbreak
PMID:
 
32088678
 
DOI:
 
10.3855/jidc.12496
[Indexed for MEDLINE] 
Free full text
Icon for The Journal of Infection in Developing Countries
54.
 2020 Mar;22(2):86-91. doi: 10.1016/j.micinf.2020.02.004. Epub 2020 Feb 20.

Lessons learned from the 2019-nCoV epidemic on prevention of future infectious diseases.

Pan X1Ojcius DM2Gao T3Li Z4Pan C5Pan C6.

Abstract

Only a month after the outbreak of pneumonia caused by 2019-nCoV, more than forty-thousand people were infected. This put enormous pressure on the Chinese government, medical healthcare provider, and the general public, but also made the international community deeply nervous. On the 25th day after the outbreak, the Chinese government implemented strict traffic restrictions on the area where the 2019-nCoV had originated-Hubei province, whose capital city is Wuhan. Ten days later, the rate of increase of cases in Hubei showed a significant difference (p = 0.0001) compared with the total rate of increase in other provinces of China. These preliminary data suggest the effectiveness of a traffic restriction policy for this pandemic thus far. At the same time, solid financial support and improved research ability, along with network communication technology, also greatly facilitated the application of epidemic prevention measures. These measures were motivated by the need to provide effective treatment of patients, and involved consultation with three major groups in policy formulation-public health experts, the government, and the general public. It was also aided by media and information technology, as well as international cooperation. This experience will provide China and other countries with valuable lessons for quickly coordinating and coping with future public health emergencies.

KEYWORDS:

2019-nCoV; Government; Public health emergency; Traffic restriction
PMID:
 
32088333
 
DOI:
 
10.1016/j.micinf.2020.02.004
[Indexed for MEDLINE]
Icon for Elsevier Science
55.
 2020 Mar;22(2):80-85. doi: 10.1016/j.micinf.2020.02.002. Epub 2020 Feb 20.

The epidemic of 2019-novel-coronavirus (2019-nCoV) pneumonia and insights for emerging infectious diseases in the future.

Li JY1You Z2Wang Q3Zhou ZJ4Qiu Y5Luo R6Ge XY7.

Abstract

At the end of December 2019, a novel coronavirus, 2019-nCoV, caused an outbreak of pneumonia spreading from Wuhan, Hubei province, to the whole country of China, which has posed great threats to public health and attracted enormous attention around the world. To date, there are no clinically approved vaccines or antiviral drugs available for these human coronavirus infections. Intensive research on the novel emerging human infectious coronaviruses is urgently needed to elucidate their route of transmission and pathogenic mechanisms, and to identify potential drug targets, which would promote the development of effective preventive and therapeutic countermeasures. Herein, we describe the epidemic and etiological characteristics of 2019-nCoV, discuss its essential biological features, including tropism and receptor usage, summarize approaches for disease prevention and treatment, and speculate on the transmission route of 2019-nCoV.

KEYWORDS:

2019-nCoV; ACE2; Bat; Pneumonia; SARS-CoV; Spike
PMID:
 
32087334
 
DOI:
 
10.1016/j.micinf.2020.02.002
[Indexed for MEDLINE]
Icon for Elsevier Science
56.
57.
59.
 2020 Mar 16;14(1):72-73. doi: 10.5582/bst.2020.01047. Epub 2020 Feb 19.

Breakthrough: Chloroquine phosphate has shown apparent efficacy in treatment of COVID-19 associated pneumonia in clinical studies.

Abstract

The coronavirus disease 2019 (COVID-19) virus is spreading rapidly, and scientists are endeavoring to discover drugs for its efficacious treatment in China. Chloroquine phosphate, an old drug for treatment of malaria, is shown to have apparent efficacy and acceptable safety against COVID-19 associated pneumonia in multicenter clinical trials conducted in China. The drug is recommended to be included in the next version of the Guidelines for the Prevention, Diagnosis, and Treatment of Pneumonia Caused by COVID-19 issued by the National Health Commission of the People's Republic of China for treatment of COVID-19 infection in larger populations in the future.

KEYWORDS:

2019-nCoV; COVID-19; SARS-CoV-2; chloroquine; pneumonia
PMID:
 
32074550
 
DOI:
 
10.5582/bst.2020.01047
[Indexed for MEDLINE] 
Free full text
Icon for J-STAGE, Japan Science and Technology Information Aggregator, Electronic
60.
 2020 May;92(5):461-463. doi: 10.1002/jmv.25711. Epub 2020 Mar 1.

The course of clinical diagnosis and treatment of a case infected with coronavirus disease 2019.

Han W1Quan B2Guo Y3Zhang J4Lu Y5Feng G1Wu Q1Fang F1Cheng L1Jiao N6Li X1Chen Q7.

KEYWORDS:

COVID-19; SARS-CoV-2; lopinavir and ritonavir tablets; severe respiratory syndrome
PMID:
 
32073161
 
DOI:
 
10.1002/jmv.25711
[Indexed for MEDLINE]
Icon for Wiley
61.
63.
 2020 Feb;25(6). doi: 10.2807/1560-7917.ES.2020.25.6.2000094.

First cases of coronavirus disease 2019 (COVID-19) in France: surveillance, investigations and control measures, January 2020.

Abstract

A novel coronavirus (severe acute respiratory syndrome coronavirus 2, SARS-CoV-2) causing a cluster of respiratory infections (coronavirus disease 2019, COVID-19) in Wuhan, China, was identified on 7 January 2020. The epidemic quickly disseminated from Wuhan and as at 12 February 2020, 45,179 cases have been confirmed in 25 countries, including 1,116 deaths. Strengthened surveillance was implemented in France on 10 January 2020 in order to identify imported cases early and prevent secondary transmission. Three categories of risk exposure and follow-up procedure were defined for contacts. Three cases of COVID-19 were confirmed on 24 January, the first cases in Europe. Contact tracing was immediately initiated. Five contacts were evaluated as at low risk of exposure and 18 at moderate/high risk. As at 12 February 2020, two cases have been discharged and the third one remains symptomatic with a persistent cough, and no secondary transmission has been identified. Effective collaboration between all parties involved in the surveillance and response to emerging threats is required to detect imported cases early and to implement adequate control measures.

KEYWORDS:

COVID-19, SARS-CoV-2, 2019-nCov; France; Surveillance; contact tracing; containment; coronavirus
PMID:
 
32070465
 
PMCID:
 
PMC7029452
 
DOI:
 
10.2807/1560-7917.ES.2020.25.6.2000094
[Indexed for MEDLINE] 
Free PMC Article
Icon for Ingenta plcIcon for PubMed Central
64.
 2020 Feb 17;9(1):5. doi: 10.1186/s40249-020-0621-x.

The association between internal migration and pulmonary tuberculosis in China, 2005-2015: a spatial analysis.

Liao WB1Ju K1Gao YM2Pan J3,4.

Abstract

BACKGROUND:

Internal migration places individuals at high risk of contracting tuberculosis (TB). However, there is a scarcity of national-level spatial analyses regarding the association between TB and internal migration in China. In our research, we aimed to explore the spatial variation in cases of sputum smear-positive pulmonary TB (SS + PTB) in China; and the associations between SS + PTB, internal migration, socioeconomic factors, and demographic factors in the country between 2005 and 2015.

METHODS:

Reported cases of SS + PTB were obtained from the national PTB surveillance system database; cases were obtained at the provincial level. Internal migration data were extracted from the national population sampling survey and the census. Spatial autocorrelations were explored using the global Moran's statistic and local indicators of spatial association. The spatial temporal analysis was performed using Kulldorff's scan statistic. Fixed effects regression was used to explore the association between SS + PTB and internal migration.

RESULTS:

A total of 4 708 563 SS + PTB cases were reported in China between 2005 and 2015, of which 3 376 011 (71.7%) were male and 1 332 552 (28.3%) were female. There was a trend towards decreasing rates of SS + PTB notifications between 2005 and 2015. The result of global spatial autocorrelation indicated that there were significant spatial correlations between SS + PTB rate and internal migration each year (2005-2015). Spatial clustering of SS + PTB cases was mainly located in central and southern China and overlapped with the clusters of emigration. The proportions of emigrants and immigrants were significantly associated with SS + PTB. Per capita GDP and education level were negatively associated with SS + PTB. The internal migration flow maps indicated that migrants preferred neighboring provinces, with most migrating for work or business.

CONCLUSIONS:

This study found a significant spatial autocorrelation between SS + PTB and internal migration. Both emigration and immigration were statistically associated with SS + PTB, and the association with emigration was stronger than that for immigration. Further, we found that SS + PTB clusters overlapped with emigration clusters, and the internal migration flow maps suggested that migrants from SS + PTB clusters may influence the TB epidemic characteristics of neighboring provinces. These findings can help stakeholders to implement effective PTB control strategies for areas at high risk of PTB and those with high rates of internal migrants.

KEYWORDS:

China; Internal migration; Pulmonary tuberculosis; Spatial analysis
PMID:
 
32063228
 
PMCID:
 
PMC7025414
 
DOI:
 
10.1186/s40249-020-0621-x
[Indexed for MEDLINE] 
Free PMC Article
Icon for BioMed CentralIcon for PubMed Central
65.
 2020 Mar 16;14(1):3-8. doi: 10.5582/bst.2020.01043. Epub 2020 Feb 17.

Challenges to the system of reserve medical supplies for public health emergencies: reflections on the outbreak of the severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) epidemic in China.

Abstract

On December 31, 2019, the Wuhan Municipal Health Commission announced an outbreak of severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), China is now at a critical period in the control of the epidemic. The Chinese Government has been taking a series of rapid, comprehensive, and effective prevention and control measures. As the pandemic has developed, a fact has become apparent: there is a serious dearth of emergency medical supplies, and especially an extreme shortage of personal protective equipment such as masks and medical protective clothing. This is one of the major factors affecting the progress of epidemic prevention and control. Although China has made great efforts to strengthen the ability to quickly respond to public health emergencies since the SARS outbreak in 2003 and it has clarified requirements for emergency supplies through legislation, the emergency reserve supplies program has not been effectively implemented, and there are also deficiencies in the types, quantity, and availability of emergency medical supplies. A sound system of emergency reserve supplies is crucial to the management of public health emergencies. Based on international experiences with pandemic control, the world should emphasize improving the system of emergency reserve medical supplies in the process of establishing and improving public health emergency response systems, and it should promote the establishment of international cooperative programs to jointly deal with public health emergencies of international concern in the future.

KEYWORDS:

COVID-19; SARS-CoV-2; medical supplies; public health emergency
PMID:
 
32062645
 
DOI:
 
10.5582/bst.2020.01043
[Indexed for MEDLINE] 
Free full text
Icon for J-STAGE, Japan Science and Technology Information Aggregator, Electronic
66.
 2020 May;92(5):491-494. doi: 10.1002/jmv.25709. Epub 2020 Feb 21.

Overlapping and discrete aspects of the pathology and pathogenesis of the emerging human pathogenic coronaviruses SARS-CoV, MERS-CoV, and 2019-nCoV.

Liu J1,2Zheng X1,2Tong Q1Li W1Wang B1,2Sutter K3,2Trilling M3,2Lu M3,2Dittmer U3,2Yang D1,2.

Abstract

First reported from Wuhan, The People's Republic of China, on 31 December 2019, the ongoing outbreak of a novel coronavirus (2019-nCoV) causes great global concerns. Based on the advice of the International Health Regulations Emergency Committee and the fact that to date 24 other countries also reported cases, the WHO Director-General declared that the outbreak of 2019-nCoV constitutes a Public Health Emergency of International Concern on 30 January 2020. Together with the other two highly pathogenic coronaviruses, the severe acute respiratory syndrome coronavirus (SARS-CoV) and Middle East respiratory syndrome coronavirus (MERS-CoV), 2019-nCov and other yet to be identified coronaviruses pose a global threat to public health. In this mini-review, we provide a brief introduction to the pathology and pathogenesis of SARS-CoV and MERS-CoV and extrapolate this knowledge to the newly identified 2019-nCoV.

KEYWORDS:

coronavirus; immnopathology; immune responses; pathogenesis; respiratory tract; virus classification
PMID:
 
32056249
 
DOI:
 
10.1002/jmv.25709
[Indexed for MEDLINE]
Icon for Wiley
67.
 2020 May;92(5):476-478. doi: 10.1002/jmv.25708. Epub 2020 Feb 24.

Does SARS-CoV-2 has a longer incubation period than SARS and MERS?

Jiang X1,2Rayner S3,4,5Luo MH1,2,5,6,7.

Abstract

The outbreak of a novel coronavirus (SARS-CoV-2) since December 2019 in Wuhan, the major transportation hub in central China, became an emergency of major international concern. While several etiological studies have begun to reveal the specific biological features of this virus, the epidemic characteristics need to be elucidated. Notably, a long incubation time was reported to be associated with SARS-CoV-2 infection, leading to adjustments in screening and control policies. To avoid the risk of virus spread, all potentially exposed subjects are required to be isolated for 14 days, which is the longest predicted incubation time. However, based on our analysis of a larger dataset available so far, we find there is no observable difference between the incubation time for SARS-CoV-2, severe acute respiratory syndrome coronavirus (SARS-CoV), and middle east respiratory syndrome coronavirus (MERS-CoV), highlighting the need for larger and well-annotated datasets.

KEYWORDS:

coronavirus; incubation; local infection/replication/spread; pandemic; virulence
PMID:
 
32056235
 
DOI:
 
10.1002/jmv.25708
[Indexed for MEDLINE]
Icon for Wiley
68.
 2020 Mar;25(3):278-280. doi: 10.1111/tmi.13383. Epub 2020 Feb 16.

The COVID-19 epidemic.

Velavan TP1,2,3Meyer CG1,2,3.

KEYWORDS:

2019-nCoV; COVID-19; Epidemic; SARS-CoV2; Wuhan
PMID:
 
32052514
 
DOI:
 
10.1111/tmi.13383
[Indexed for MEDLINE] 
Free full text
Icon for Wiley
69.
 2020 May;92(5):479-490. doi: 10.1002/jmv.25707. Epub 2020 Mar 3.

Potential interventions for novel coronavirus in China: A systematic review.

Abstract

An outbreak of a novel coronavirus (COVID-19 or 2019-CoV) infection has posed significant threats to international health and the economy. In the absence of treatment for this virus, there is an urgent need to find alternative methods to control the spread of disease. Here, we have conducted an online search for all treatment options related to coronavirus infections as well as some RNA-virus infection and we have found that general treatments, coronavirus-specific treatments, and antiviral treatments should be useful in fighting COVID-19. We suggest that the nutritional status of each infected patient should be evaluated before the administration of general treatments and the current children's RNA-virus vaccines including influenza vaccine should be immunized for uninfected people and health care workers. In addition, convalescent plasma should be given to COVID-19 patients if it is available. In conclusion, we suggest that all the potential interventions be implemented to control the emerging COVID-19 if the infection is uncontrollable.

KEYWORDS:

2019-CoV; COVID-19; MERS; SARS; coronavirus; potential interventions
PMID:
 
32052466
 
DOI:
 
10.1002/jmv.25707
[Indexed for MEDLINE]
Icon for Wiley
70.
 2020 Feb;61(2):55-57. doi: 10.11622/smedj.2020018. Epub 2020 Feb 13.

Outbreak of COVID-19 - an urgent need for good science to silence our fears?

PMID:
 
32052064
 
PMCID:
 
PMC7052000
 
DOI:
 
10.11622/smedj.2020018
[Indexed for MEDLINE] 
Free PMC Article
Icon for Singapore Medical AssociationIcon for PubMed Central
71.
 2020 May;92(5):473-475. doi: 10.1002/jmv.25706. Epub 2020 Feb 18.

Economic impacts of Wuhan 2019-nCoV on China and the world.

PMID:
 
32048740
 
DOI:
 
10.1002/jmv.25706
[Indexed for MEDLINE]
Icon for Wiley
72.
 2020 Feb;25(5). doi: 10.2807/1560-7917.ES.2020.25.5.2000062.

Incubation period of 2019 novel coronavirus (2019-nCoV) infections among travellers from Wuhan, China, 20-28 January 2020.

Abstract

A novel coronavirus (2019-nCoV) is causing an outbreak of viral pneumonia that started in Wuhan, China. Using the travel history and symptom onset of 88 confirmed cases that were detected outside Wuhan in the early outbreak phase, we estimate the mean incubation period to be 6.4 days (95% credible interval: 5.6-7.7), ranging from 2.1 to 11.1 days (2.5th to 97.5th percentile). These values should help inform 2019-nCoV case definitions and appropriate quarantine durations.

KEYWORDS:

2019-nCoV; Wuhan; exposure; incubation period; novel coronavirus; symptom onset
PMID:
 
32046819
 
PMCID:
 
PMC7014672
 
DOI:
 
10.2807/1560-7917.ES.2020.25.5.2000062
[Indexed for MEDLINE] 
Free PMC Article
Icon for Ingenta plcIcon for PubMed Central
73.
 2020 Feb;25(5). doi: 10.2807/1560-7917.ES.2020.25.5.2000080.

Effectiveness of airport screening at detecting travellers infected with novel coronavirus (2019-nCoV).

Erratum in

Abstract

We evaluated effectiveness of thermal passenger screening for 2019-nCoV infection at airport exit and entry to inform public health decision-making. In our baseline scenario, we estimated that 46% (95% confidence interval: 36 to 58) of infected travellers would not be detected, depending on incubation period, sensitivity of exit and entry screening, and proportion of asymptomatic cases. Airport screening is unlikely to detect a sufficient proportion of 2019-nCoV infected travellers to avoid entry of infected travellers.

KEYWORDS:

2019-nCoV; airport screening; effectiveness; emerging infections; interventions; surveillance; thermal scanning
PMID:
 
32046816
 
PMCID:
 
PMC7014668
 
DOI:
 
10.2807/1560-7917.ES.2020.25.5.2000080
[Indexed for MEDLINE] 
Free PMC Article
Icon for Ingenta plcIcon for PubMed Central
74.
 2020 Feb;25(6). doi: 10.2807/1560-7917.ES.2020.25.6.2000110. Epub 2020 Feb 11.

Epidemiological research priorities for public health control of the ongoing global novel coronavirus (2019-nCoV) outbreak.

KEYWORDS:

2019-nCoV, coronavirus, emerging infections, response, air-borne infections; emerging or re-emerging diseases; viral infections
PMID:
 
32046814
 
PMCID:
 
PMC7029449
 
DOI:
 
10.2807/1560-7917.ES.2020.25.6.2000110
[Indexed for MEDLINE] 
Free PMC Article
Icon for Ingenta plcIcon for PubMed Central
75.
 2020 Jan - Feb;33:101578. doi: 10.1016/j.tmaid.2020.101578. Epub 2020 Feb 8.

Going global - Travel and the 2019 novel coronavirus.

KEYWORDS:

2019-nCoV; Coronaviruses; Epidemic; Travelers
PMID:
 
32044389
 
DOI:
 
10.1016/j.tmaid.2020.101578
[Indexed for MEDLINE]
Icon for Elsevier Science
76.
77.
 2020 Mar;20(3):280. doi: 10.1016/S1473-3099(20)30068-2. Epub 2020 Feb 7.

Pandemic potential of 2019-nCoV.

PMID:
 
32043978
 
DOI:
 
10.1016/S1473-3099(20)30068-2
[Indexed for MEDLINE]
Icon for Elsevier Science
78.
 2020 Feb;46(2):315-328. doi: 10.1007/s00134-020-05943-5. Epub 2020 Feb 10.

Critical care management of adults with community-acquired severe respiratory viral infection.

Arabi YM1,2,3Fowler R4,5,6Hayden FG7.

Abstract

With the expanding use of molecular assays, viral pathogens are increasingly recognized among critically ill adult patients with community-acquired severe respiratory illness; studies have detected respiratory viral infections (RVIs) in 17-53% of such patients. In addition, novel pathogens including zoonotic coronaviruses like the agents causing Severe Acute Respiratory Syndrome (SARS), Middle East Respiratory Syndrome (MERS) and the 2019 novel coronavirus (2019 nCoV) are still being identified. Patients with severe RVIs requiring ICU care present typically with hypoxemic respiratory failure. Oseltamivir is the most widely used neuraminidase inhibitor for treatment of influenza; data suggest that early use is associated with reduced mortality in critically ill patients with influenza. At present, there are no antiviral therapies of proven efficacy for other severe RVIs. Several adjunctive pharmacologic interventions have been studied for their immunomodulatory effects, including macrolides, corticosteroids, cyclooxygenase-2 inhibitors, sirolimus, statins, anti-influenza immune plasma, and vitamin C, but none is recommended at present in severe RVIs. Evidence-based supportive care is the mainstay for management of severe respiratory viral infection. Non-invasive ventilation in patients with severe RVI causing acute hypoxemic respiratory failure and pneumonia is associated with a high likelihood of transition to invasive ventilation. Limited existing knowledge highlights the need for data regarding supportive care and adjunctive pharmacologic therapy that is specific for critically ill patients with severe RVI. There is a need for more pragmatic and efficient designs to test different therapeutics both individually and in combination.

KEYWORDS:

Acute respiratory distress syndrome; Antiviral therapy; Coronavirus; Influenza; Neuraminidase inhibitor
PMID:
 
32040667
 
DOI:
 
10.1007/s00134-020-05943-5
[Indexed for MEDLINE]
Icon for Springer
79.
 2020 Mar 16;14(1):64-68. doi: 10.5582/bst.2020.01030. Epub 2020 Feb 9.

Clinical characteristics and therapeutic procedure for four cases with 2019 novel coronavirus pneumonia receiving combined Chinese and Western medicine treatment.

Abstract

Pneumonia associated with the 2019 novel coronavirus (2019-nCoV) is continuously and rapidly circulating at present. No effective antiviral treatment has been verified thus far. We report here the clinical characteristics and therapeutic procedure for four patients with mild or severe 2019-nCoV pneumonia admitted to Shanghai Public Health Clinical Center. All the patients were given antiviral treatment including lopinavir/ritonavir (Kaletra®), arbidol, and Shufeng Jiedu Capsule (SFJDC, a traditional Chinese medicine) and other necessary support care. After treatment, three patients gained significant improvement in pneumonia associated symptoms, two of whom were confirmed 2019-nCoV negative and discharged, and one of whom was virus negative at the first test. The remaining patient with severe pneumonia had shown signs of improvement by the cutoff date for data collection. Results obtained in the current study may provide clues for treatment of 2019-nCoV pneumonia. The efficacy of antiviral treatment including lopinavir/ritonavir, arbidol, and SFJDC warrants further verification in future study.

KEYWORDS:

2019-nCoV; Shufeng Jiedu Capsule; arbidol; lopinavir; ritonavir
PMID:
 
32037389
 
DOI:
 
10.5582/bst.2020.01030
[Indexed for MEDLINE] 
Free full text
Icon for J-STAGE, Japan Science and Technology Information Aggregator, Electronic
80.
 2020;42:e2020007. doi: 10.4178/epih.e2020007. Epub 2020 Feb 9.

Epidemiologic characteristics of early cases with 2019 novel coronavirus (2019-nCoV) disease in Korea.

Abstract

In about 20 days since the diagnosis of the first case of the 2019 novel coronavirus (2019-nCoV) in Korea on January 20, 2020, 28 cases have been confirmed. Fifteen patients (53.6%) of them were male and median age of was 42 years (range, 20-73). Of the confirmed cases, 16, 9, and 3 were index (57.2%), first-generation (32.1%), and second-generation (10.7%) cases, respectively. All first-generation and second-generation patients were family members or intimate acquaintances of the index cases with close contacts. Fifteen among 16 index patients had entered Korea from January 19 to 24, 2020 while 1 patient had entered Korea on January 31, 2020. The average incubation period was 3.9 days (median, 3.0), and the reproduction number was estimated as 0.48. Three of the confirmed patients were asymptomatic when they were diagnosed. Epidemiological indicators will be revised with the availability of additional data in the future. Sharing epidemiological information among researchers worldwide is essential for efficient preparation and response in tackling this new infectious disease.

KEYWORDS:

2019-nCoV; Epidemiology; Isolation; Korea; Outbreak; Quarantine
PMID:
 
32035431
 
DOI:
 
10.4178/epih.e2020007
[Indexed for MEDLINE] 
Free full text
Icon for Publishing M2Community
82.
 2020 Mar;22(2):69-71. doi: 10.1016/j.micinf.2020.01.004. Epub 2020 Feb 4.

Pathogenicity and transmissibility of 2019-nCoV-A quick overview and comparison with other emerging viruses.

Abstract

A zoonotic coronavirus, tentatively labeled as 2019-nCoV by the World Health Organization (WHO), has been identified as the causative agent of the viral pneumonia outbreak in Wuhan, China, at the end of 2019. Although 2019-nCoV can cause a severe respiratory illness like SARS and MERS, evidence from clinics suggested that 2019-nCoV is generally less pathogenic than SARS-CoV, and much less than MERS-CoV. The transmissibility of 2019-nCoV is still debated and needs to be further assessed. To avoid the 2019-nCoV outbreak turning into an epidemic or even a pandemic and to minimize the mortality rate, China activated emergency response procedures, but much remains to be learned about the features of the virus to refine the risk assessment and response. Here, the current knowledge in 2019-nCoV pathogenicity and transmissibility is summarized in comparison with several commonly known emerging viruses, and information urgently needed for a better control of the disease is highlighted.

KEYWORDS:

ACE2; Basic reproduction number (R(0)); Case fatality rate; Novel coronavirus infection; SARS
PMID:
 
32032682
 
DOI:
 
10.1016/j.micinf.2020.01.004
[Indexed for MEDLINE] 
Free full text
Icon for Elsevier Science
83.
 2020 May;92(5):522-528. doi: 10.1002/jmv.25700. Epub 2020 Feb 19.

Genomic variance of the 2019-nCoV coronavirus.

Abstract

There is a rising global concern for the recently emerged novel coronavirus (2019-nCoV). Full genomic sequences have been released by the worldwide scientific community in the last few weeks to understand the evolutionary origin and molecular characteristics of this virus. Taking advantage of all the genomic information currently available, we constructed a phylogenetic tree including also representatives of other coronaviridae, such as Bat coronavirus (BCoV) and severe acute respiratory syndrome. We confirm high sequence similarity (>99%) between all sequenced 2019-nCoVs genomes available, with the closest BCoV sequence sharing 96.2% sequence identity, confirming the notion of a zoonotic origin of 2019-nCoV. Despite the low heterogeneity of the 2019-nCoV genomes, we could identify at least two hypervariable genomic hotspots, one of which is responsible for a Serine/Leucine variation in the viral ORF8-encoded protein. Finally, we perform a full proteomic comparison with other coronaviridae, identifying key aminoacidic differences to be considered for antiviral strategies deriving from previous anti-coronavirus approaches.

KEYWORDS:

CLUSTAL analysis; biostatistics & bioinformatics; coronavirus; data visualization; virus classification
PMID:
 
32027036
 
DOI:
 
10.1002/jmv.25700
[Indexed for MEDLINE]
Icon for Wiley
84.
 2020 May;92(5):501-511. doi: 10.1002/jmv.25701. Epub 2020 Feb 14.

Transmission dynamics and evolutionary history of 2019-nCoV.

Abstract

To investigate the time origin, genetic diversity, and transmission dynamics of the recent 2019-nCoV outbreak in China and beyond, a total of 32 genomes of virus strains sampled from China, Thailand, and the USA with sampling dates between 24 December 2019 and 23 January 2020 were analyzed. Phylogenetic, transmission network, and likelihood-mapping analyses of the genome sequences were performed. On the basis of the likelihood-mapping analysis, the increasing tree-like signals (from 0% to 8.2%, 18.2%, and 25.4%) over time may be indicative of increasing genetic diversity of 2019-nCoV in human hosts. We identified three phylogenetic clusters using the Bayesian inference framework and three transmission clusters using transmission network analysis, with only one cluster identified by both methods using the above genome sequences of 2019-nCoV strains. The estimated mean evolutionary rate for 2019-nCoV ranged from 1.7926 × 10-3 to 1.8266 × 10-3 substitutions per site per year. On the basis of our study, undertaking epidemiological investigations and genomic data surveillance could positively impact public health in terms of guiding prevention efforts to reduce 2019-nCOV transmission in real-time.

KEYWORDS:

2019-nCoV; TMRCA; evolutionary rate; phylogenetic cluster; time to most recent common ancestor; transmission cluster
PMID:
 
32027035
 
DOI:
 
10.1002/jmv.25701
[Indexed for MEDLINE]
Icon for Wiley
85.
 2020 Mar 2;21(5):730-738. doi: 10.1002/cbic.202000047. Epub 2020 Feb 25.

Learning from the Past: Possible Urgent Prevention and Treatment Options for Severe Acute Respiratory Infections Caused by 2019-nCoV.

Abstract

With the current trajectory of the 2019-nCoV outbreak unknown, public health and medicinal measures will both be needed to contain spreading of the virus and to optimize patient outcomes. Although little is known about the virus, an examination of the genome sequence shows strong homology with its better-studied cousin, SARS-CoV. The spike protein used for host cell infection shows key nonsynonymous mutations that might hamper the efficacy of previously developed therapeutics but remains a viable target for the development of biologics and macrocyclic peptides. Other key drug targets, including RNA-dependent RNA polymerase and coronavirus main proteinase (3CLpro), share a strikingly high (>95 %) homology to SARS-CoV. Herein, we suggest four potential drug candidates (an ACE2-based peptide, remdesivir, 3CLpro-1 and a novel vinylsulfone protease inhibitor) that could be used to treat patients suffering with the 2019-nCoV. We also summarize previous efforts into drugging these targets and hope to help in the development of broad-spectrum anti-coronaviral agents for future epidemics.

KEYWORDS:

2019-nCoV; 3CLpro; RdRp; SARS; antiviral agents; coronavirus; spike proteins
PMID:
 
32022370
 
DOI:
 
10.1002/cbic.202000047
[Indexed for MEDLINE]
Icon for Wiley
86.
 2020 May;92(5):495-500. doi: 10.1002/jmv.25698. Epub 2020 Mar 3.

Immunoinformatics-aided identification of T cell and B cell epitopes in the surface glycoprotein of 2019-nCoV.

Abstract

The 2019 novel coronavirus (2019-nCoV) outbreak has caused a large number of deaths with thousands of confirmed cases worldwide, especially in East Asia. This study took an immunoinformatics approach to identify significant cytotoxic T lymphocyte (CTL) and B cell epitopes in the 2019-nCoV surface glycoprotein. Also, interactions between identified CTL epitopes and their corresponding major histocompatibility complex (MHC) class I supertype representatives prevalent in China were studied by molecular dynamics simulations. We identified five CTL epitopes, three sequential B cell epitopes and five discontinuous B cell epitopes in the viral surface glycoprotein. Also, during simulations, the CTL epitopes were observed to be binding MHC class I peptide-binding grooves via multiple contacts, with continuous hydrogen bonds and salt bridge anchors, indicating their potential in generating immune responses. Some of these identified epitopes can be potential candidates for the development of 2019-nCoV vaccines.

KEYWORDS:

2019-nCoV; COVID-19; SARS-CoV-2; coronavirus; epitope prediction; immunoinformatics
PMID:
 
32022276
 
DOI:
 
10.1002/jmv.25698
[Indexed for MEDLINE]
Icon for Wiley
87.
 2020 May;92(5):518-521. doi: 10.1002/jmv.25699. Epub 2020 Feb 12.

The first two cases of 2019-nCoV in Italy: Where they come from?

Abstract

A novel Coronavirus, 2019-nCoV, has been identified as the causal pathogen of an ongoing epidemic, with the first cases reported in Wuhan, China, last December 2019, and has since spread to other countries worldwide, included Europe and very recently Italy. In this short report, phylogenetic reconstruction was used to better understand the transmission dynamics of the virus from its first introduction in China focusing on the more recent evidence of infection in a couple of Chinese tourists arrived in Italy on 23rd January 2020 and labeled as Coronavirus Italian cases. A maximum clade credibility tree has been built using a dataset of 54 genome sequences of 2019-nCoV plus two closely related bat strains (SARS-like CoV) available in GenBank. Bayesian time-scaled phylogenetic analysis was implemented in BEAST 1.10.4. The Bayesian phylogenetic reconstruction showed that 2019-2020 nCoV firstly introduced in Wuhan on 25 November 2019, started epidemic transmission reaching many countries worldwide, including Europe and Italy where the two strains isolated dated back 19 January 2020, the same that the Chinese tourists arrived in Italy. Strains isolated outside China were intermixed with strains isolated in China as evidence of likely imported cases in Rome, Italy, and Europe, as well. In conclusion, this report suggests that further spread of 2019-nCoV epidemic was supported by human mobility and that quarantine of suspected or diagnosed cases is useful to prevent further transmission. Viral genome phylogenetic analysis represents a useful tool for the evaluation of transmission dynamics and preventive action.

KEYWORDS:

engineering and technology; epidemiology; infection; macromolecular design
PMID:
 
32022275
 
DOI:
 
10.1002/jmv.25699
[Indexed for MEDLINE]
Icon for Wiley
88.
 2020 Jan;25(4). doi: 10.2807/1560-7917.ES.2020.25.4.2000058.

Pattern of early human-to-human transmission of Wuhan 2019 novel coronavirus (2019-nCoV), December 2019 to January 2020.

Erratum in

Abstract

Since December 2019, China has been experiencing a large outbreak of a novel coronavirus (2019-nCoV) which can cause respiratory disease and severe pneumonia. We estimated the basic reproduction number R0 of 2019-nCoV to be around 2.2 (90% high density interval: 1.4-3.8), indicating the potential for sustained human-to-human transmission. Transmission characteristics appear to be of similar magnitude to severe acute respiratory syndrome-related coronavirus (SARS-CoV) and pandemic influenza, indicating a risk of global spread.

KEYWORDS:

2019-nCoV; Wuhan; coronavirus; emerging infectious disease; mathematical modelling
PMID:
 
32019669
 
PMCID:
 
PMC7001239
 
DOI:
 
10.2807/1560-7917.ES.2020.25.4.2000058
[Indexed for MEDLINE] 
Free PMC Article
Icon for Ingenta plcIcon for PubMed Central
89.
 2020 Jan;25(4). doi: 10.2807/1560-7917.ES.2020.25.4.2000057.

Novel coronavirus (2019-nCoV) early-stage importation risk to Europe, January 2020.

Abstract

As at 27 January 2020, 42 novel coronavirus (2019-nCoV) cases were confirmed outside China. We estimate the risk of case importation to Europe from affected areas in China via air travel. We consider travel restrictions in place, three reported cases in France, one in Germany. Estimated risk in Europe remains high. The United Kingdom, Germany and France are at highest risk. Importation from Beijing and Shanghai would lead to higher and widespread risk for Europe.

KEYWORDS:

2019-nCoV; Europe; importation risk; travel ban
PMID:
 
32019667
 
PMCID:
 
PMC7001240
 
DOI:
 
10.2807/1560-7917.ES.2020.25.4.2000057
[Indexed for MEDLINE] 
Free PMC Article
Icon for Ingenta plcIcon for PubMed Central
91.
 2020 Mar;22(2):74-79. doi: 10.1016/j.micinf.2020.01.003. Epub 2020 Feb 1.

Measures for diagnosing and treating infections by a novel coronavirus responsible for a pneumonia outbreak originating in Wuhan, China.

Abstract

On 10 January 2020, a new coronavirus causing a pneumonia outbreak in Wuhan City in central China was denoted as 2019-nCoV by the World Health Organization (WHO). As of 24 January 2020, there were 887 confirmed cases of 2019-nCoV infection, including 26 deaths, reported in China and other countries. Therefore, combating this new virus and stopping the epidemic is a matter of urgency. Here, we focus on advances in research and development of fast diagnosis methods, as well as potential prophylactics and therapeutics to prevent or treat 2019-nCoV infection.

KEYWORDS:

2019-nCoV; Diagnosis; Preventive; Therapeutic
PMID:
 
32017984
 
DOI:
 
10.1016/j.micinf.2020.01.003
[Indexed for MEDLINE]
Icon for Elsevier Science
92.
 2020 Apr;79:104211. doi: 10.1016/j.meegid.2020.104211. Epub 2020 Jan 30.

Novel coronavirus: From discovery to clinical diagnostics.

Abstract

A novel coronavirus designated as 2019-nCoV first appeared in Wuhan, China in late December 2019. Dozens of people died in China, and thousands of people infected as 2019-nCoV continues to spread around the world. We have described the discovery, emergence, genomic characteristics, and clinical diagnostics of 2019-nCoV.

KEYWORDS:

Bat; China; Pneumonia; coronavirus
PMID:
 
32007627
 
DOI:
 
10.1016/j.meegid.2020.104211
[Indexed for MEDLINE]
Icon for Elsevier Science
95.
 2020 Feb 15;201(4):P7-P8. doi: 10.1164/rccm.2014P7.

Novel Wuhan (2019-nCoV) Coronavirus.

PMID:
 
32004066
 
DOI:
 
10.1164/rccm.2014P7
[Indexed for MEDLINE]
Icon for Atypon
96.
97.
 2020 Mar 16;14(1):69-71. doi: 10.5582/bst.2020.01020. Epub 2020 Jan 28.

Drug treatment options for the 2019-new coronavirus (2019-nCoV).

Lu H1,2,3.

Abstract

As of January 22, 2020, a total of 571 cases of the 2019-new coronavirus (2019-nCoV) have been reported in 25 provinces (districts and cities) in China. At present, there is no vaccine or antiviral treatment for human and animal coronavirus, so that identifying the drug treatment options as soon as possible is critical for the response to the 2019-nCoV outbreak. Three general methods, which include existing broad-spectrum antiviral drugs using standard assays, screening of a chemical library containing many existing compounds or databases, and the redevelopment of new specific drugs based on the genome and biophysical understanding of individual coronaviruses, are used to discover the potential antiviral treatment of human pathogen coronavirus. Lopinavir /Ritonavir, Nucleoside analogues, Neuraminidase inhibitors, Remdesivir, peptide (EK1), abidol, RNA synthesis inhibitors (such as TDF, 3TC), anti-inflammatory drugs (such as hormones and other molecules), Chinese traditional medicine, such ShuFengJieDu Capsules and Lianhuaqingwen Capsule, could be the drug treatment options for 2019-nCoV. However, the efficacy and safety of these drugs for 2019- nCoV still need to be further confirmed by clinical experiments.

KEYWORDS:

2019-nCoV; Coronaviruses; pneumonia
PMID:
 
31996494
 
DOI:
 
10.5582/bst.2020.01020
[Indexed for MEDLINE] 
Free full text
Icon for J-STAGE, Japan Science and Technology Information Aggregator, Electronic
98.
 2020;1251:115-121. doi: 10.1007/5584_2019_462.

Epidemic Influenza Seasons from 2008 to 2018 in Poland: A Focused Review of Virological Characteristics.

Abstract

The objective of this review was to elaborate on changes in the virological characteristics of influenza seasons in Poland in the past decade. The elaboration was based on the international influenza surveillance system consisting of Sentinel and non-Sentinel programs, recently adopted by Poland, in which professionals engaged in health care had reported tens of thousands of cases of acute upper airway infections. The reporting was followed by the provision of biological specimens collected from patients with suspected influenza and influenza-like infection, in which the causative contagion was then verified with molecular methods. The peak incidence of influenza infections has regularly been in January-March each epidemic season. The number of tested specimens ranged from 2066 to 8367 per season from 2008/2009 to 2017/2018. Type A virus predominated in nine out of the ten seasons and type B virus of the Yamagata lineage in the 2017/2018 season. Concerning the influenza-like infectionrespiratory syncytial virus predominated in all the seasons. There was a sharp increase in the proportion of laboratory confirmations of influenza infection from season to season in relation to the number of specimens examined, from 3.2% to 42.4% over the decade. The number of confirmations, enabling a prompt commencement of antiviral treatment, related to the number of specimens collected from patients and on the virological situation in a given season. Yet influenza remains a health scourge, with a dismally low yearly vaccination rate, which recently reaches just about 3.5% of the general population in Poland.

KEYWORDS:

Diagnostics; Epidemic seasons; Influenza; Sentinel program; Surveillance; Virological characteristics
PMID:
 
31989546
 
DOI:
 
10.1007/5584_2019_462
[Indexed for MEDLINE]
Icon for Springer
99.
 2020 Mar;92:228-233. doi: 10.1016/j.ijid.2020.01.018. Epub 2020 Jan 22.

The utility of serial procalcitonin measurements in addition to pneumonia severity scores in hospitalised community-acquired pneumonia: A multicentre, prospective study.

Abstract

OBJECTIVES:

The usefulness of serial procalcitonin (PCT) measurements for predicting the prognosis and treatment efficacy for hospitalised community-acquired pneumonia (CAP) patients was investigated.

METHODS:

This prospective, multicentre, cohort study enrolled consecutive CAP patients who were hospitalised at 10 hospitals in western Japan from September 2013 to September 2016. PCT and C-reactive protein (CRP) were measured on admission (PCT D1 and CRP D1), within 48-72 h after admission (PCT D3 and CRP D3), and within 144-192 h after admission. CURB-65 and the Pneumonia Severity Index (PSI) were assessed on admission. The primary outcome was 30-day mortality; secondary outcomes were early and late treatment failure rates.

RESULTS:

A total of 710 patients were included. The 30-day mortality rate was 3.1%. On multivariate analysis, only PCT D3/D1 ratio >1 [odds ratio (95% confidence interval): 4.33 (1.46-12.82),P = 0.008] and PSI [odds ratio (95% confidence interval): 2.32 (1.07-5.03), P = 0.03] were significant prognostic factors. Regarding treatment efficacy, PCT D3/D1 >1 was a significant predictor of early treatment failure on multivariate analysis. PCT D3/D1 with the PSI significantly improved the prognostic accuracy over that of the PSI alone.

CONCLUSIONS:

PCT should be measured consecutively, not only on admission, to predict the prognosis and treatment efficacy in CAP.

KEYWORDS:

Biomarker; C-reactive protein; Pneumonia; Procalcitonin; Prognosis
PMID:
 
31981766
 
DOI:
 
10.1016/j.ijid.2020.01.018
[Indexed for MEDLINE] 
Free full text
Icon for Elsevier Science
100.
 2020 Mar;92:208-213. doi: 10.1016/j.ijid.2020.01.017. Epub 2020 Jan 21.

Influenza A-associated severe pneumonia in hospitalized patients: Risk factors and NAI treatments.

Zou Q1Zheng S2Wang X1Liu S3Bao J1Yu F1Wu W4Wang X5Shen B6Zhou T7Zhao Z8Wang Y9Chen R10Wang W11Ma J12Li Y13Wu X14Shen W15Xie F16Vijaykrishna D17Chen Y18.

Abstract

OBJECTIVE:

The risk factors and the impact of NAI treatments in patients with severe influenza A-associated pneumonia remain unclear.

METHODS:

A multicenter, retrospective, observational study was conducted in Zhejiang, China during a severe influenza epidemic in August 2017-May 2018. Clinical records of patients (>14 y) hospitalized with laboratory-confirmed influenza A virus infection and who developed severe pneumonia were compared to those with mild-to-moderate pneumonia. Risk factors related to pneumonia severity and effects of NAI treatments (monotherapy and combination of peramivir and oseltamivir) were analyzed.

RESULTS:

202 patients with influenza A-associated severe pneumonia were enrolled, of whom 84 (41.6%) had died. Male gender (OR = 1.782; 95% CI: 1.089-2.917; P = 0.022), chronic pulmonary disease (OR = 2.581; 95% CI: 1.447-4.603; P = 0.001) and diabetes mellitus (OR = 2.042; 95% CI: 1.135-3.673; P = 0.017) were risk factors related to influenza A pneumonia severity. In cox proportional hazards model, severe pneumonia patients treated with double dose oseltamivir (300mg/d) had a better survival rate compared to those receiving a single dose (150 mg/d) (HR = 0.475; 95%CI: 0.254-0.887; P = 0.019). However, different doses of peramivir (300 mg/d vs. 600 mg/d) and combination therapy (oseltamivir-peramivir vs. monotherapy) showed no differences in 60-day mortality (P = 0.392 and P = 0.658, respectively).

CONCLUSIONS:

Patients with male gender, chronic pulmonary disease, or diabetes mellitus were at high risk of developing severe pneumonia after influenza A infection. Double dose oseltamivir might be considered in treating influenza A-associated severe pneumonia.

KEYWORDS:

Double dose; Influenza; Oseltamivir; Severe pneumonia
PMID:
 
31978583
 
DOI:
 
10.1016/j.ijid.2020.01.017
[Indexed for MEDLINE] 
Free full text
Icon for Elsevier Science

Δεν υπάρχουν σχόλια:

Δημοσίευση σχολίου

Αρχειοθήκη ιστολογίου