Open access peer-reviewed chapter

The One-Health Approach to Infectious Disease Outbreaks Control

Written By

Sima Ernest Rugarabamu

Submitted: 13 July 2020 Reviewed: 23 December 2020 Published: 14 September 2021

DOI: 10.5772/intechopen.95759

From the Edited Volume

Current Perspectives on Viral Disease Outbreaks - Epidemiology, Detection and Control

Edited by David Claborn

Chapter metrics overview

905 Chapter Downloads

View Full Metrics

Abstract

Close contact between people, animals, plants, and their shared environment provides more disease transmission opportunities. Host characteristics, environmental conditions, and habitat disruption can provide new opportunities for disease to occur. These changes may lead to the spread of existing and new diseases. Bacteria, viruses, fungi, protozoans, sporozoans, worms, and others cause infectious diseases. Some of these diseases may be prone to explosive outbreaks and may constitute deadly epidemic threats that could rapidly reach pandemic proportions. Drugs and vaccines can successfully control many infectious diseases; however, this is challenged by the lack of facilities and resources. In all parts of the world, infectious disease is an essential constraint to increased human, animal, and environmental interactions. Identifying hot-spot and interventions for prevention while considering the heterogeneity of target diseases to places, population time, or situation is essential. Therefore, successful infectious disease control measures must be based on understanding disease transmission pathways, strengthening surveillance systems, and intervention. Application of the One Health method is a responsive approach to infectious disease control. Much of the One-Health based approach to managing an infectious disease has been utilized with a promising effect on controlling current outbreaks. More deliberate efforts should encourage understanding of disease determinants to analyze infectious disease issues through a One-Health lens. Only through the extensive participation of all related field stakeholders can One-Health truly reach its potential to mitigate infectious disease outbreaks. This chapter reviews utilization of the One Health approach to infectious disease outbreak control.

Keywords

  • virus
  • one health
  • infectious disease
  • outbreak
  • control

1. Introduction

Infectious disease remains responsible for a large part of the world’s premature death and disability burden [1]. Bacteria, viruses, and protozoa, among other agents, cause infectious diseases to humans [2]. Usually, infected cases are present in numbers at an expected level, but an outbreak may occur every once in a while [3]. A new strain of the disease agent can significantly impact either the local or global levels [4]. For example, global pandemics of smallpox, cholera, and influenza periodically threatened populations before developing improved living conditions, especially in high-income countries [5]. To date, safe, effective, and affordable vaccines and the increasing availability of antibiotics have reduced the burden of many such diseases in high-income countries, though there is a lack of adequate control in many middle and low-income countries [5, 6].

Close contact between people, animals, plants, and their shared environment provides an increased risk of disease transmission [7]. Host characteristics, environmental conditions, and habitat disruption can provide new disease opportunities [6, 7]. These changes may lead to the spread of new or emerging diseases. For instance, viral diseases have become as common as emerging and re-emerging diseases [8]. They occur worldwide, in a variety of ecological settings [9]. Others are found only in limited ecologic and geographic foci. Over 60 percent of infectious diseases and 70 percent of humans’ emerging infections are zoonotic, with two-thirds originating in wildlife or domestic animals [9, 10]. Notable examples of globally emerging and re-emerging infectious diseases include Ebola hemorrhagic fever, dengue, chikungunya, yellow fever, and other respiratory viral infections such as pandemic influenza H1N1 2009, SARS, Avian Influenza (H5N1) and (H7N9) [11, 12]. These diseases are prone to explosive outbreaks and constitute deadly epidemic threats that could rapidly reach pandemic proportions, affecting people’s lives [13]. On the other hand, a global increase in antibiotic-resistant bacteria has resulted in pathogens resistant to most or essentially all of the available antimicrobials [14]. Problems facing scientists today include the emergence of diseases from the fast-changing human-animal ecosystems due to multiple factors [15]. Increasing human-environment interaction provides ingredients for outbreaks of infectious diseases [16].

In a world of increased frequency of interaction, animal food production for human consumption, increased use of transportation, and increased movement of people across national borders, these factors act as determinants for infectious diseases by directly or indirectly influencing the occurrence and distribution of infectious diseases [17]. Measures for successful disease control must be based on understanding disease transmission pathways, strengthening surveillance systems, and intervention [18, 19]. This is possible by the application of the One Health approach. Much of the One-Health based approach to managing outbreaks of infectious disease has been utilized with a promising effect to control current outbreaks. More deliberate efforts should encourage understanding of disease determinants to analyze infectious disease issues through a One-Health lens. Only through the extensive participation of all related field stakeholders can One-Health truly reach its potential to mitigate infectious disease outbreaks.

One Health is a collaborative, multisector, and transdisciplinary approach — working from local to global levels — to achieve optimal health outcomes by recognizing the interconnection between people, animals, plants, and their shared environment [20, 21, 22, 23]. Therefore, One Health is an approach that recognizes that people’s health is closely connected to animals’ health and shared environment. One Health has become more important recently because many factors have changed interactions between the environment, people, animals, and plants [24].

Human populations continue to grow and expand. They increase close contact with wild and domestic animals, both livestock and pets [25]. Animals play an important role in lives, whether for food, fiber, livelihoods, travel, sport, education, or companionship [26].

Animals also share susceptibility to environmental hazards and some diseases, which can sometimes serve as early warning signs of potential human illness. For instance, birds often die of the West Nile virus before people in the same area get sick with West Nile virus infection [24, 25, 26, 27].

The One Health issue includes a focus on zoonotic and vector-borne diseases, antimicrobial resistance, food safety, food security, environmental contamination, and other health threats shared by humans, the environment, and animals. Other fields such as chronic disease, mental health, injury, occupational health, and non-communicable diseases also benefit from a One Health approach involving collaboration across disciplines and sectors [28, 29].

One Health is gaining recognition globally as an effective way to fight health issues at the human-animal-environment interface, including zoonotic diseases.

The complexity of health and environmental challenges needs to be evaluated in an integrated and holistic manner to provide a more comprehensive understanding of problems and potential solutions [30, 31]. Concerted efforts in the paradigm shift from the silo-based health systems to the One Health approach is important [29, 30, 31, 32]. Decision-makers for disease prevention and control should utilize the One Health approach to prepare for and prevent illness, hospitalization, death, and the economic burden experienced during disease epidemics. In any public health emergency, an early warning system to combat epidemics is usually immediately implemented. Response networks, e.g., the Global Outbreak Alert and Response Network (G.O.A.R.N.), collaborate with institutions and networks to pool their human and technical resources to fight outbreaks [33]. The critical decision to initiate disease response is often reactive and urgently needed in a rapidly changing environment with little or incomplete information available and biased [3334]. Traditional surveillance systems provide regular data updates. However, these systems are inherently retrospective and delayed, limiting their utility for real-time epidemic response [35].

Additionally, a silo-based health system deals with present conditions or those immediately expected [30, 31, 32, 33, 34, 35, 36]. The One Health approach could help fill these holes by controlling the utility, scale, and timing of counteraction techniques [36]. For example, scientists recognized the link between human and animal health and its threats to Ebola epidemics’ welfare and economies. This resulted from the importance of collaborative and cross-disciplinary approaches for responding to emerging and resurging diseases, particularly the inclusion of a wildlife component for global disease prevention and control.

During infectious disease outbreaks, the coordination and communication of prevention strategies – such as vaccination and treatment – support the deployment and management of crucial public health resources [37]. However, earlier trial vaccines that are protective in animals and safe to humans are not useful because most existing beneficial trials are not standardized or validated. Patients’ safety remains unknown because the resources are usually limited and not enough to conduct conclusive trials [38]. Testing the drugs in animals is challenged by the unavailability of facilities to conduct research. Specifically, there are not enough biosafety laboratories. A significant issue surrounding this and other potential disease preventive measures is ensuring the availability and affordability of any useful drugs and vaccines [39]. The One Health approach could bring together scientists, public health officials, and researchers from academia, industry, and government in an open project and develop tools to address specific disease prevention challenges [40]. The tool could be a program to predict disease trends while addressing specific needs by engaging decision-makers and researchers in real-world scenarios. For instance, a collaborative effort that focuses on geographic risk can provide greater insight into which geographic areas emergent pathogens may be circulating in but are undetected [41]. These predictive models allow for more strategic focusing of resources for monitoring the emergence and spread of threats [42]. Continuous surveillance of wildlife and domestic livestock in these limited areas for early detection of pathogens may yield faster and more economical results than spreading resources worldwide to detect pathogens [39, 40, 41, 42]. Wildlife is a reservoir of an extraordinarily deep and diverse pool of novel microbial agents [43]. Even considering such overwhelming diversity, the actual numbers of microbial agents reported to infect humans and cause disease are probably many viral infections that remain undetected [44]. Continuous surveillance focusing on these few microbial agents for early detection of pathogens may yield faster and more economical results than spreading resources to all possible microbial agents to detect pathogens [43, 44, 45]. As a result, chance-based interceptions permit the utilization of information about the heterogeneity of hazard to target outbreak location to those spots, populaces, times, or circumstances where the danger of sickness is generally considerable and the probability of discovering high [46].

While the specific disease sources remain unknown, many pathogens are thought to be harbored in wild animals or the environment, with initial transmission to humans via contact with infected animal species or fomites, and later spread through human-to-human transmission [47]. The world-ecology provides habitats for diverse fauna [48]. Changes of the natural ecosystem because of social-cultural and environmental procedures have increased closeness between the human populace, domesticated animals, and wildlife, promoting increased contacts with the disease-causing microbe [49]. The One Health approach is expected to provide a good understanding of the drivers of spillover events from hot spot dwelling fauna to interfacing humans, which will enable disease prevention and control at the source while forecasting accuracy, visualization and communication, collaboration and partner engagement, state and local health department perspectives, pilot projects, and other issues at hand.

Using infectious disease outbreaks response as an example, we propose in this chapter utilization of the One Health concept approach, such as identifying spillover sources in both human and animal populations, designing comprehensive surveillance systems, and implementing an intervention approach to combat infectious disease outbreak.

Advertisement

2. Understanding disease transmission pathways

Complex interactions of humans with the biotic and abiotic components of the environment facilitate spillover events [3, 50]. Documenting how diseases occur is the key to understanding disease transmission pathways and different meanings attached to infectious diseases in different communities [34, 51]. This includes identifying potential sources and reservoirs of viruses in environmental, human, and animal systems.

Data collection is crucial to attaining preliminary information for the identification of sources of transmission. Numerous agencies publish data regarding clinical cases of human and animal disease, both spatially and temporally. These data can be collected and analyzed through integrated human-animal disease surveillance to assess infectious disease occurrence [51]. Additionally, spatial and temporal patterns, the likelihood of infectious disease in certain areas or periods, land use [52], human/livestock/wildlife population density [53], and other data may be collected and analyzed. Potential pathways can be prioritized to determine the most relevant regions [54]. A system of weight factors to perform prioritization using statistical models could be developed, such as the degree of human-animal interactions [55]. Regulation of environment and livestock waste could impact the importance of the potential pathways associated with human-animal ecosystem interactions.

Advertisement

3. Strengthening surveillance systems

Surveillance systems of the critical environmental reservoirs and pathways will allow for the early detection of outbreaks. It is essential to quickly identify critical times and critical locations for the onset of outbreaks by monitoring disease indicators such as the pathogen presence or burden in a particular community with associated risk factors. This can be achieved by the environment, livestock, wildlife, and human sampling [56]. Regular monitoring of critical reservoirs will identify peaks in presence or indicators related to early signals of disease outbreaks [57].

Traditional human and livestock disease detection and management systems are based on diagnostic analyses of clinical samples [58, 59, 60]. However, these systems fail to detect early warnings of public health threats at a broad population level and fail to predict outbreaks promptly [61]. An alternative to this could be using human-wildlife contaminated ecosystems such as community-based urine, fecal, and other samples to identify public, wildlife, or livestock health [62]. This kind of monitoring, together with unique sampling, allows early detection and prediction of outbreaks by understanding pathogens, including shedding rates, risks, and magnitudes, critical in disease surveillance [63]. Recently, raw sewage has been used to monitor the presence and abundance of COVID-19 in communities. An epidemiological tool developed and refined by environmental scientists over the last 20 years (Wastewater-Based Epidemiology — WBE) holds the potential to contain and mitigate Covid-19 outbreaks while also minimizing domino effects such as unnecessarily long stay-at-home policies that stress humans and economies alike. WBE measures chemical signatures in sewage, such as fragment biomarkers from the Severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2), only by applying the type of clinical diagnostic testing to the collective signature of entire communities [64]. As such, it could rapidly establish Covid-19 infections across an entire community. The One Health surveillance model aims to identify risks before clinical cases are reported [65]. Mapping pandemic potential can facilitate data collection representative of at-risk regions followed by risk mitigations [66]. Microbial source tracking can be more complicated, especially in limited-resource areas, which necessitates determining the environment to sample. Specific population, shedding rate and natural degradation and comparison, and correlation with clinical data are vital tools for getting reliable information for strengthening surveillance systems [66, 67, 68].

Intervention approaches in the One Health approach involve utilizing feasible innovation technology for human and animal ecosystems management, medical and veterinary interventions to oversee diseases, and education of local communities and governments to change human behavior, practices, and policy based on relationships between the environments or human and animal health [69].

The first intervention for vaccine-preventable infectious diseases is wildlife animal vaccination and treatment strategies [70]. Disease preventive measures act as a barrier between human-animal disease transmissions. For example, encouraging the disinfection of clean water could remove pathogens from the community. Interventions to prevent pathogens shedding are among the possible management that requires multiple strategies for accomplishment [71, 72]. For instance, rabies prevention by oral vaccination of wildlife with live vaccines has proven a powerful tool to eliminate or control rabies in multiple countries in Europe and North America. In 2012–2013 U.S. Department of Agriculture’s Animal and Plant Health Inspection Service through Wildlife Services program, conducted the field trial involving the distribution of new oral rabies live recombinant human adenovirus type 5 vector, expressing rabies virus glycoprotein (AdRG1.3) (Onrab) vaccine bait in five states [73]. Baits laden with oral rabies vaccines are essential for managing wildlife rabies to monitor human contacts and potential vaccine virus exposure. Continued surveillance like these is needed because of the potential for vaccine virus infection [74].

This approach and several others can be considered examples of the complementary policies for the permanent implementation of interventions. There is a need to regulate animal pathogen shedding in waste products, especially in rural areas and forest ecosystems, and other previously reported critical transmission pathways [75, 76, 77].

The modification of human behavior is also imperative to minimize the transmittance of infectious diseases and pathways in which interventions cannot be performed for cost, capability, or convenience [78, 79, 80, 81]. One primary health behavior-changing method is educating medical, veterinarians, and environmental professionals in the One-Health approach [82]. It is also crucial to educate the public where people are vulnerable to disease transmission [83]. Especially in impoverished, high-risk areas, robust measures should be taken to educate the public on the critical pathways of transmission of viral disease [84, 85, 86, 87, 88].

Advertisement

4. Conclusions

This chapter advocate utilizing the One Health model as part of the solution to the ultimate control of infectious disease outbreaks. Disease transmission includes complex frameworks that incorporate associations between humans, animals, and the environment. These systems have spatial and temporal variations that require a deep understanding of the interaction and the processes within. The most significant advance in understanding disease transmission is identifying reservoirs and primary transmission pathways.

Traditional infectious disease control measures such as case management, vaccination, active surveillance, case identification and isolation, and strategic community engagement have helped contain outbreaks. However, many people still die, and more epidemics are anticipated in previously affected and new geographical areas; new control approaches, including One Health, are essential. Research on the role of wildlife in disease causation should be undertaken to improve the situation. Wildlife surveillance data on the biodiversity of animal interface found in the hot spot regions and the pathogen’s activity in animals and humans should also be included in strategic interventions. Overall, infectious disease control’s success requires a balance between medicine, veterinary science, bioscience, epidemiology, health systems, socioanthropology, and political science, to facilitate early detection and response to unusual events.

Moreover, documenting how diseases occur is the key to understanding disease transmission pathways and different meanings attached to infectious diseases in various communities. A multipronged approach with data and tracking systems’ support is an equally important component in attaining national and global health security. The One-Health-based approach to managing an infectious disease has been utilized with a promising effect to control few current outbreaks; there has still more principally that needs to be grasped by the veterinary network. Increasingly purposeful endeavors ought to urge other professionals to examine infectious disease issues through a One-Health focal point. Only through the broad cooperation of all related field partners can One-Health arrive at its capability to control infectious diseases.

References

  1. 1. Holmes KK, Bertozzi S, Bloom BR, et al., editors. Major Infectious Diseases. 3rd edition. Washington (DC): The International Bank for Reconstruction and Development / the World Bank; 2017 Nov 3. Available from: https://www.ncbi.nlm.nih.gov/books/NBK525192/ doi: 10.1596/978-1-4648-0524-0
  2. 2. Janeway CA Jr, Travers P, Walport M, et al. Immunobiology: The Immune System in Health and Disease. 5th edition. New York: Garland Science; 2001. Infectious agents and how they cause disease.
  3. 3. Rugarabamu S, Mboera L, Rweyemamu M, et al. Forty-two years of responding to Ebola virus outbreaks in Sub-Saharan Africa: a review. B.M.J. Glob Health. 2020;5(3):e001955. Published 2020 March 8. Doi:10.1136/bmjgh-2019-001955A
  4. 4. Tognotti E. Lessons from the history of quarantine, from plague to influenza A. Emerg Infect Dis. 2013; 19(2):254-259. doi:10.3201/eid1902.120312
  5. 5. Schoch-Spana M, Bouri N, Rambhia KJ, Norwood A. stigma, health disparities, and therefore the 2009 H1N1 influenza pandemic: the way to protect Latino farmworkers future health emergencies. Biosecurity Bioterror. 2010;8:243-54 10.1089/bsp.2010.0021 [PubMed] [CrossRef] [Google Scholar]
  6. 6. Greenwood B. The contribution of vaccination to global Health: past, present, and future. Philos Trans R Soc Lond B Biol Sci. 2014; 369(1645):20130433. Published 2014 May 12. doi:10.1098/rstb.2013.0433
  7. 7. Wilkinson K, Grant WP, Green L.E., et al. Infectious diseases of animals and plants: an interdisciplinary approach. Philos Trans R Soc Lond B Biol Sci. 2011; 366(1573):1933-1942. doi:10.1098/rstb.2010.0415
  8. 8. Karesh WB, Cook RA, Bennett EL, Newcomb J Wildlife trade and global disease emergence. Emerg Infect Dis. 2005; 11: 1000-1002
  9. 9. Molyneux D, Hallaj Z, Keusch GT, et al. .Zoonoses and marginalized infectious diseases of poverty: where do we stand? Parasit Vectors. 2011; 4: 106
  10. 10. Reed KD. Viral Zoonoses. Reference Module in Biomedical Sciences. 2018; B978-0-12-801238-3.95729-5. doi:10.1016/B978-0-12-801238-3.95729-5
  11. 11. Blancou J et al. Emerging or re-emerging bacterial zoonoses: emergence, surveillance, and control factors. Veterinary Research. 2005; 36:507-522
  12. 12. Report of the WHO/FAO/OIE Joint Consultation on Emerging Zoonotic Diseases Geneva, Switzerland. 3-5 May 2004,). Accessed July 12, 2020
  13. 13. Hughes JM. Emerging infectious diseases: a C.D.C. perspective. Emerging Infectious Diseases. 2001; 7:494-496
  14. 14. Prestinaci F, Pezzotti P, Pantosti A. Antimicrobial resistance: a global, multifaceted phenomenon. Pathog Glob Health. 2015; 109(7):309-318. doi:10.1179/2047773215Y.0000000030
  15. 15. Anderson, P. K., A. A. Cunningham, N. G. Patel, F. J. Morales, P. R. Epstein, and P. Daszak. 2004. Emerging infectious diseases of plants: Pathogen pollution, climate change, and agrotechnology drivers. Trends Ecol Evol 19(10):535-544 [PubMed]
  16. 16. Eisenberg JN, Desai MA, Levy K, et al. Environmental determinants of infectious disease: a framework for tracking causal links and guiding public health research. Environ Health Perspect. 2007; 115(8):1216-1223. doi:10.1289/ehp.9806
  17. 17. Lindahl JF, Grace D. The consequences of human actions on risks for infectious diseases: a review. Infect Ecol Epidemiol. 2015; 5:30048. Published 2015 November 27. doi:10.3402/iee.v5.30048
  18. 18. Lourenço, C., Tatem, A.J., Atkinson, P.M. et al. Strengthening surveillance systems for malaria elimination: global landscaping of system performance, 2015-2017. Malar J 18, 315 (2019). https://doi.org/10.1186/s12936-019-2960-2
  19. 19. Degeling C, Johnson J, Kerridge I, et al. Implementing a One Health approach to emerging infectious disease: reflections on the socio-political, ethical, and legal dimensions. B.M.C. Public Health. 2015; 15:1307. Published 2015 December 29. doi:10.1186/s12889-015-2617-1
  20. 20. Zinsstag J, Schelling E, Waltner-Toews D, Tanner M. From "one medicine" to "one health" and systemic approaches to health and well-being. Prev Vet Med (2011) 101:148-56.10.1016/j.prevetmed.2010.07.003 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  21. 21. One Health Initiative Task Force. One Health: A New Professional Imperative. Schaumburg, IL: American Veterinary Medical Association; (2008) [Google Scholar]
  22. 22. Zinsstag J, Schelling E, Waltner-Toews D, Tanner M, editors. One Health, the Theory, and Practice of Integrated Health Approaches. C.A.B.I. (2015).10.1079/9781780643410.0000 [CrossRef] [Google Scholar]
  23. 23. Lebov J, Grieger K, Womack D, Zaccaro D, Whitehead N, Kowalcyk B, et al. A framework for One health research. One Health (2017) 3:44-50.10.1016/j.onehlt.2017.03.004 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  24. 24. One Health Basics. (2018, November 5). Retrieved August 9, 2020, from https://www.cdc.gov/onehealth/basics/index.html
  25. 25. Allan, B. F., F. Keesing, and R. S. Ostfeld. 2003. Effect of forest fragmentation on Lyme disease risk. Conserv Biol 17(1):267-272
  26. 26. Wood L, Martin K, Christian H, Nathan A, Lauritsen C, Houghton S, et al. (2015) The Pet Factor - Companion Animals as a Conduit for Getting to Know People, Friendship Formation and Social Support. PLoS ONE 10(4): e0122085. https://doi.org/10.1371/journal.pone.0122085
  27. 27. Reif JS. Animal sentinels for environmental and public health. Public Health Rep. 2011 May-Jun;126 Suppl 1(Suppl 1):50-7. doi: 10.1177/00333549111260S108. PMID: 21563712; PMCID: PMC3072903
  28. 28. Capua, I., and D. J. Alexander. 2008. Ecology, epidemiology, and human health implications of avian influenza viruses: Why do we need to share genetic data? Zoonoses Public Health 55(1):2-15 [PubMed]
  29. 29. CDC (U.S. Centers for Disease Control and Prevention). 1998. Preventing emerging infectious diseases: A strategy for the 21st century. Overview of the updated CDC plan. MMWR 47(15):1-14 [PubMed]
  30. 30. G. Gongal, One Health approach in the South East Asia region: opportunities and challenges Curr. Top. Microbiol. Immunol., 366 (2013), pp. 113-122
  31. 31. L.K. Allen-Scott, B. Buntain, J.M. Hatfield, A. Meisser, C.J. Thomas Academic institutions, and one Health: building capacity for transdisciplinary research approaches to address complex health issues the animal-human-ecosystem Interface Acad. Med., 90 (7) (2015), pp. 866-871
  32. 32. Mourya DT, Yadav PD, Ullas PT, et al. Emerging/re-emerging viral diseases & new viruses on the Indian horizon [published correction appears in Indian J Med Res. 2019 May;149(5):688]. Indian J Med Res. 2019; 149(4):447-467
  33. 33. Chauhan RP, Dessie ZG, Noreddin A, El Zowalaty ME. Systematic Review of Important Viral Diseases in Africa in Light of the 'One Health' Concept. Pathogens. 2020;9(4):301. Published 2020 April 20. doi:10.3390/pathogens9040301
  34. 34. Brownstein, J.S., Freifeld, C.C., and Madoff, L.C. (2009) Digital Disease Detection – Harnessing the online for Public Health Surveillance. New England Journal of drugs 360, 2153-2157
  35. 35. Jones KE, Patel N, Levy M, et al. Global trends in emerging infectious diseases. Nature 2008; 451:990-94
  36. 36. Malik MR, El Bushra H, Opoka M, Formenty P, Valayudhan R, Eremin S, et al. Strategic approach to controlling viral hemorrhagic fever outbreaks within the Eastern Mediterranean Region: Report from a regional consultation. Eastern Mediterranean Health Journal, 2013; 19 (10)
  37. 37. J.S. Towner, T.K. Sealy, M.L. Khristova, et al.Newly discovered Ebola virus related to hemorrhagic fever outbreak in UgandaPLoS Pathogens, 4 (2008), p. e1000212
  38. 38. Markoff, L. yellow jack Outbreak in Sudan." N Engl J Med 2013; 368(8): 689-691
  39. 39. Humphrey, John M., et al. "Dengue and chikungunya seroprevalence among Qatari nationals and immigrants residing in Qatar." PLoS O.N.E., vol. 14, no. 1, 2019, p. e0211574. Accessed July 26, 2020
  40. 40. Onyango, C. O., Opoka, M. L., Ksiazek, T. G., Formenty, P., Ahmed, A., Tukei, P. M., Sang, R. C., Ofula, V. O., Konongoi, S. L., Coldren, R. L., Grein, T., Legros, D., Bell, M., De Cock, K. M., Bellini, W. J., Towner, J. S., Nichol, S. T., & Rollin, P. E. (2007). Laboratory diagnosis of Ebola hemorrhagic fever during an epidemic in Yambio, Sudan, 2004. Journal of Infectious Diseases, 196(SUPPL. 2), S193-S198. https://doi.org/10.1086/520609
  41. 41. World Health Organization. Global Alert and Response (G.A.R.). Valley fever in Yemen - Update 4-26 October 2000. Available from http://www.who.int/csr/don/2000_10_26/en/index.html
  42. 42. Suk JE, Van Cangh T, Beauté J, Bartels C, Tsolova S, Pharris A, Ciotti M, Semenza JC. The interconnected and cross-border nature of risks posed by infectious diseases. Glob Health Action. 2014 October 10; 7:25287. doi: 10.3402/gha.v7.25287. Erratum in: Glob Health Action. 2015; 8:27635. Erratum in: Glob Health Action. 2015 Jan; 8(1):27635. PMID: 25308818; PMCID: PMC4195207
  43. 43. Madani TA, Al-Mazrou YY, Al-Jeffri MH, Mishkhas AA, Al-Rabeah AM, Turkistani AM, et al. valley Fever Epidemic in Saudi Arabia: Epidemiological, Clinical, and Laboratory Characteristics. Clinical Infectious Diseases. 2003 October 15, 2003; 37(8):1084-92
  44. 44. Hassan OA, Ahlm C, Sang R, Evander M. The 2007 valley fever outbreak in Sudan. PLoS Negl Trop Dis. 2011 Sep; 5(9):e1229
  45. 45. Mackenzie JS, Jeggo M. The One Health Approach-Why Is It So Important. Trop Med Infect Dis. 2019; 4(2):88. Published 2019 May 31. doi:10.3390/tropicalmed4020088 A
  46. 46. Jones K.E., Patel N.G., Levy M.A., Storeygard A., Balk D., Gittleman J.L., Daszak P. Global trends in emerging infectious diseases. Nature. 2008; 451:990-993. doi: 10.1038/nature06536 [PMC free article] [PubMed] [CrossRef] [Google Scholar]
  47. 47. Karesh W.B., Dobson A., Lloyd-Smith J.O., Lubroth J., Dixon M.A., Bennett M., Aldrich S., Harrington T., Formenty P., Loh E.H., et al. Ecology of zoonoses: natural and unnatural histories. Lancet. 2012; 380:1936-1945. doi: 10.1016/S0140-6736(12)61678-X [PMC free article] [PubMed] [CrossRef]
  48. 48. Onyango CO, Opoka ML, Ksiazek TG, Formenty P, Ahmed A, Tukei PM, et al. Laboratory Diagnosis of Ebola hemorrhagic fever during an epidemic in Yambio, Sudan, 2004. Journal of Infectious Diseases. 2007 November 15, 2007; 196(Supplement 2): S193-S8
  49. 49. World Health Organization. Global Alert and Response (G.A.R.). Valley fever in Yemen - Update 4-26 October 2000. Available from http://www.who.int/csr/don/2000_10_26/en/index.html
  50. 50. World Health Organization. Global Alert and Response. Avian influenza – a situation in Djibouti May 12, 2006. Available at http://www.who.int/csr/don/2006_05_12/en/index.html
  51. 51. Aguirre, A.A., Longcore, T., Barbieri, M. et al. The One Health Approach to Toxoplasmosis: Epidemiology, Control, and Prevention Strategies. Eco Health 16, 378-390 (2019). https://doi.org/10.1007/s10393-019-01405-7
  52. 52. Semenza JC, Lindgren E, Balkanyi L, Espinosa L, Almqvist MS, Penttinen P, Rocklöv J. Determinants and drivers of communicable disease threat events in Europe. Emerg Infect Dis. 2016; 22:581-9
  53. 53. Krishna Regmi, Ruth Gilbert & Colin Thunhurst (2015), how can health systems be strengthened to regulate and stop an Ebola outbreak? A narrative review, Infection Ecology & Epidemiology, 5:1, DOI: 10.3402/iee.v5.28877
  54. 54. Meseko CA, Egbetade AO, Fagbo S. Ebola virus disease control in West Africa: an ecological, one health approach. Pan Afr Med J. 2015; 21:6. Published 2015 May 4. doi:10.11604/pamj.2015.21.6.6587A
  55. 55. Lewis JS, Farnsworth ML, Burdett CL, Theobald DM, Gray M, Miller RS. Biotic and abiotic factors predicting the worldwide distribution and population density of an invasive large mammal. Sci Rep. 2017; 7:44152. Published 2017 March 9. doi:10.1038/srep44152
  56. 56. Okello AL, Bardosh K, Smith J, Welburn SC. One Health: past successes and future challenges in three African contexts. PLoS Negl Trop Dis. 2014; 8(5):e2884. Published 2014 May 22. doi:10.1371/journal.pntd.0002884
  57. 57. Alexander KA, Lewis BL, Marathe M, Eubank S, Blackburn JK. Modeling of wildlife-associated zoonoses: applications and caveats. Vector Borne Zoonotic Dis. 2012; 12(12):1005-1018. doi:10.1089/vbz.2012.0987A
  58. 58. Farrar, J., 2008. A threat but also a chance for collaborative international science. Presentation, First board meeting on Achieving Sustainable Global Capacity for Surveillance and Response to Emerging Diseases of Zoonotic Origin, Washington, DC, June 25-26
  59. 59. Gilson, L., 2006. Trust in health care: Theoretical perspectives and research needs. J Health Organ Manag 20(5):359-375 [PubMed]
  60. 60. C.D.C. 2009. b. Integrated disease surveillance and response. http://www.cdc.gov/ids/about.htm (accessed July 12, 2020)
  61. 61. Institute of drugs (U.S.) Forum on Microbial Threats. Global communicable disease Surveillance and Detection: Assessing the Challenges—Finding Solutions, Workshop Summary. Washington (D.C.): National Academies Press (U.S.); 2007. Summary and Assessment. Available from: https://www.ncbi.nlm.nih.gov/books/NBK52862/
  62. 62. Argudín MA, Deplano A, Meghraoui A, et al. Bacteria from Animals as a Pool of Antimicrobial Resistance Genes. Antibiotics (Basel). 2017;6(2):12. Published 2017 June 6. doi:10.3390/antibiotics6020012 A
  63. 63. Anderson RM, May RM. Infectious diseases of humans: Dynamics and control. New York: Oxford University Press; 1992
  64. 64. Daughton CG. Wastewater surveillance for population-wide Covid-19: The present and future. Sci Total Environ. 2020; 736:139631. doi:10.1016/j.scitotenv.2020.139631
  65. 65. Nii-Trebi NI. Emerging and Neglected Infectious Diseases: Insights, Advances, and Challenges. Biomed Res Int. 2017; 2017:5245021. doi:10.1155/2017/5245021
  66. 66. More D. M., Folkers G. K., Fauci A. S. Emerging infections: a perpetual challenge. The Lancet Infectious Diseases. 2008; 8(11):710-719. doi: 10.1016/S1473-3099(08)70256-1. - DOI - P.M.C. –
  67. 67. Redding, D.W., Atkinson, P.M., Cunningham, A.A., et al. Impacts of environmental and socio-economic factors on emergence and epidemic potential of Ebola in Africa. Nat Commun 10, 4531 (2019). https://doi.org/10.1038/s41467-019-12499-6A
  68. 68. Kruse H, kirkemo AM, Handeland K. Wildlife as a source of zoonotic infections. Emerg Infect Dis. 2004; 10(12):2067-2072. doi:10.3201/eid1012.040707A
  69. 69. Leroy EM, Kumulungui B, Pourrut X, et al. Fruit bats as reservoirs of Ebola virus. Nature. 2005 December 1; 438(7068):575-6 [PubMed] [Google Scholar] A
  70. 70. Jones KE, Patel N, Levy M, et al. Global trends in emerging infectious diseases. Nature. 2008 Feb 21; 451(7181):990-3 [PMC free article] [PubMed] [Google Scholar]
  71. 71. Breman JG, Johnson KM. Ebola then and now. N Engl J Med. 2014 October 30; 371(18):1663-6 [PubMed] [Google Scholar]
  72. 72. A Stangl, A.L., Earnshaw, V.A., Logie, C.H. et al. The Health Stigma and Discrimination Framework: a worldwide, crosscutting framework to tell research, intervention development, and policy on health-related stigmas. BMC Med 17, 31 (2019). https://doi.org/10.1186/s12916-019-1271-3
  73. 73. Fehlner-Gardiner C, Rudd R, Donovan D, Slate D, Kempf L, Badcock J. Comparing ONRAB® AND RABORAL V-RG® oral rabies vaccine field performance in raccoons and striped skunks, New Brunswick, Canada, and Maine, USA. J Wildl Dis. 2012;48(1):157-167. doi:10.7589/0090-3558-48.1.157
  74. 74. O'Brien E, Xagoraraki I. Understanding temporal and spatial variations of viral disease within the U.S.: the necessity for a one-health-based data collection and analysis approach. One Health. 2019; 8:100105. Published 2019, October 19doi:10.1016/j.onehlt.2019.100105
  75. 75. Maki J, Guiot AL, Aubert M, Brochier B, Cliquet F, Hanlon CA, King R, Oertli EH, Rupprecht CE, Schumacher C, Slate D, Yakobson B, Wohlers A, Lankau EW. Oral vaccination of wildlife using a vaccinia-rabies-glycoprotein recombinant virus vaccine (RABORAL V-RG®): a global review. Vet Res. 2017 Sep 22; 48(1):57. Doi: 10.1186/s13567-017-0459-9. PMID: 28938920; PMCID: PMC5610451
  76. 76. O’Brien E., Xagoraraki I. A water-focused one-health approach for early detection and prevention of viral outbreaks. One Health. 2019;7 [PMC free article] [PubMed] [Google Scholar]
  77. 77. Mauer W.A., Kaneene J.B. Integrated human-animal disease surveillance. Emerg. Infect. Dis. 2005; 11:1490-1491 [PMC free article] [PubMed] [Google Scholar]
  78. 78. Neo J.P.S., Tan BH. The utilization of animals as a surveillance tool for monitoring environmental health hazards, human health hazards, and bioterrorism. Vet Microbiol. 2017; 203:40-48. doi:10.1016/j.vetmic.2017.02.007
  79. 79. Ellwanger, Joel Henrique, Kaminski, Valéria de Lima, & Chiefs, José Artur Bogo. (, 2017). the way to detect new viral outbreaks or epidemics? We'd like to survey the circulation of viruses in humans and other animals using fast, practical, cheap, and broad-spectrum methodologies. Brazilian Journal of Infectious Diseases, 21(2), 211-212. https://doi.org/10.1016/j.bjid.2016.12.001
  80. 80. Selma Souf, Recent advances in diagnostic testing for viral infections, Bioscience Horizons: The International Journal of Student Research, Volume 9, 2016, hzw010, https://doi.org/10.1093/biohorizons/hzw010
  81. 81. Umemneku Chikere CM, Wilson K, Graziadio S, and Vale L, Allen AJ (2019) diagnostic assay evaluation methodology: a scientific review of methods employed to gauge diagnostic tests within the absence of gold standard – An update. PLoS ONE 14(10): e0223832. https://doi.org/10.1371/journal.pone.0223832
  82. 82. Vourc’h G, Bridges VE, Gibbens J, et al. Detecting emerging diseases in livestock through clinical observations [published correction appears in Emerg Infect Dis. 2006 Apr; 12(4):714]. Emerg Infect Dis. 2006; 12(2):204-210. doi:10.3201/eid1202.050498
  83. 83. National Research Council (U.S.) Committee on Climate, Ecosystems, Infectious Diseases, and Human Health. Under the Weather: Climate, Ecosystems, and communicable disease. Washington (D.C.): National Academies Press (U.S.); 2001. 7, Toward the event of Disease Early Warning Systems. Available from: https://www.ncbi.nlm.nih.gov/books/NBK222241/
  84. 84. Penakalapati G, Swarthout J, Delahoy MJ, et al. Exposure to Animal Feces and Human Health: a Scientific Review and Proposed Research Priorities. Environ Sci Technol. 2017; 51(20):11537-11552. doi:10.1021/acs.est.7b02811
  85. 85. Institute of drugs (U.S.) Forum on Emerging Infections; Burroughs T, Knobler S, Lederberg J, editors. The Emergence of Zoonotic Diseases: Understanding the Impact on Animal and Human Health: Workshop Summary. Washington (D.C.): National Academies Press (U.S.); 2002. 5, Surveillance and Management of Zoonotic Diseases Outbreaks. Available from: https://www.ncbi.nlm.nih.gov/books/NBK98092/
  86. 86. Nsubuga P, White ME, Thacker SB, et al. Public Health Surveillance: A Tool for Targeting and Monitoring Interventions. In: Jamison DT, Breman J.G., Measham AR, et al., editors
  87. 87. Disease Control Priorities in Developing Countries. 2nd edition. Washington (D.C.): The International Bank for Reconstruction and Development / The World Bank; 2006. Chapter 53. Available from: https://www.ncbi.nlm.nih.gov/books/NBK11770/ Co-published by Oxford University Press, New York
  88. 88. Madhav N, Oppenheim B, Gallivan M, et al. Pandemics: Risks, Impacts, and Mitigation. In: Jamison DT, Gelband H, Horton S, et al., editors. Disease Control Priorities: Improving Health and Reducing Poverty. 3rd edition. Washington (D.C.): The International Bank for Reconstruction and Development / The World Bank; 2017 November 27. Chapter 17. Available from: https://www.ncbi.nlm.nih.gov/books/NBK525302/ doi: 10.1596/978-1-4648-0527-1_ch17

Written By

Sima Ernest Rugarabamu

Submitted: 13 July 2020 Reviewed: 23 December 2020 Published: 14 September 2021