Next Article in Journal
EⅡB Mutation Reduces the Pathogenicity of Listeria monocytogenes by Negatively Regulating Biofilm Formation Ability, Infective Capacity, and Virulence Gene Expression
Previous Article in Journal
Putative mRNA Biomarkers for the Eradication of Infection in an Equine Experimental Model of Septic Arthritis
Previous Article in Special Issue
Retrospective Analysis of the Detection of Pathogens Associated with the Porcine Respiratory Disease Complex in Routine Diagnostic Samples from Austrian Swine Stocks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Reproduction Number of Swine Viral Respiratory Diseases: A Systematic Review

by
Dana C. Pittman Ratterree
1,
Sapna Chitlapilly Dass
2 and
Martial L. Ndeffo-Mbah
1,*
1
Department of Veterinary Integrative Biosciences, School of Veterinary Medicine and Biomedical Sciences, Texas A&M University, College Station, TX 77843, USA
2
Department of Animal Science, College of Agriculture and Life Sciences, Texas A&M University, College Station, TX 77843, USA
*
Author to whom correspondence should be addressed.
Vet. Sci. 2024, 11(7), 300; https://doi.org/10.3390/vetsci11070300
Submission received: 14 May 2024 / Revised: 26 June 2024 / Accepted: 30 June 2024 / Published: 2 July 2024
(This article belongs to the Special Issue Emerging and Re-emerging Swine Viruses)

Abstract

:

Simple Summary

Diseases in the swine industry can cause major economic and health issues. This review examines R0 estimates, which measure how efficiently respiratory diseases spread among pigs and compares transmission risks within and between farms. Seven diseases were studied: Aujeszky’s disease, PPRSV, Circovirus, Influenza A, Encephalomyocarditis Virus, Classical Swine Fever, and African Swine Fever. Data from research articles showed varying R0 values, with higher transmission rates within herds for some diseases. Effective disease control requires prompt detection, response to outbreaks, and continuous monitoring to reduce its impact on the swine industry.

Abstract

Diseases in the swine industry can cause significant economic and health impacts. This review examines R0 estimates for respiratory diseases in pigs, assessing variations and comparing transmission risks within and between farms. A literature search of three databases aggregated peer-reviewed research articles on swine viral respiratory diseases’ R0 values. The study focused on seven diseases: Aujeszky’s disease (AD), Porcine Reproductive and Respiratory Syndrome Virus (PRRSV), Circovirus, Influenza A (IA), Encephalomyocarditis Virus (EV), Classical Swine Fever (CSF), and African Swine Fever (ASF). R0 values were estimated for transmission within and between herds/farms using various methods, from complex mathematical models to simple calculations. Data primarily came from disease surveillance and laboratory experiments. The median R0 for within-herd and between-herd transmission was 10 and 3.31 for AD, 2.78 and 1.14 for PRRSV, 5.9 and 0.89 for Circovirus, 1.75 and 1.6 for CSF, and 3.94 and 3.15 for ASF. For IA and EV, only within-herd R0 values were estimated at 8.65 and 1.3, respectively. Diseases with high R0 values highlight the need for prompt detection and response to outbreaks. Continuous monitoring and evaluation of pathogen transmissibility are crucial for enhancing disease surveillance and reducing the impact of livestock diseases.

1. Introduction

Pork is a leading source of protein consumption worldwide, making up a third of all meat production globally [1]. Therefore, diseases in the swine industry are likely to have significant economic and indirect health impacts in both pork-producing and pork-consuming communities. A disease is considered economically important if it leads to death, wasting, decreased efficiency of gain, reduced rate of gain, immune suppression/increased susceptibility to other disease conditions, or infertility in livestock species used as food [2,3]. Respiratory diseases are one of the most important causes of economic losses in pork production [2,3,4].
In this study, we review seven viral diseases with various respiratory clinical presentations and transmission routes. Respiratory symptoms include but are not limited to coughing, sneezing, shortness of breath (dyspnea), and inability to grow up to production standards [1,5,6]. Aujeszky’s disease, caused by suid herpesvirus 1, is a respiratory syndromic disease that can be spread by various transmission routes, including direct contact, airborne, and fomite transmission [7]. While the disease has been mostly eradicated in many countries in Europe and North America, it remains endemic in areas like south-eastern Europe and Latin America [8]. Circoviruses, such as Porcine Circovirus 2 and 3 (PCV2 and PCV3), are emerging pathogens in the pork industry that impact multiple organ systems, including the respiratory system [9,10]. Studies indicate that PCV2 has been detected in air samples, and evidence indicates that there is short-distance transmission [11,12]. Encephalomyocarditis in pigs is best categorized as a respiratory syndrome, as transmission occurs primarily through contaminated feed and water [13,14]. African Swine Fever (ASF) was believed to be primarily transmitted through secretion and only has a clinical respiratory manifestation, but recent research has evidence of aerosol transmission through short distances [15,16]. Like ASF, Classical Swine Fever has evidence of short-distance aerosol transmission especially high-virulence strains [17,18,19]. Both Influenza A Virus (IAV) and Porcine Reproductive and Respiratory Syndrome Virus (PRRSV) spread through aerosol transmission, and the disease impacts gain, often leading to death in piglets [20,21,22,23,24]. While acute and dramatic clinical signs may resolve relatively quickly, this does not mean that their subclinical effects and costs are self-limiting or resolve quickly. Because pork makes up a significant portion of global meat production, research on swine diseases has been primarily driven by the needs of the food industry [1].
The basic reproductive number (R0) is the expected number of secondary infections generated by an infectious individual in an otherwise susceptible population. When R0 is less than one, the disease outbreak dies off as each infectious individual produces less than one new infection on average. Conversely, when R0 is above the threshold value, each infection generates multiple new infections, allowing the disease to spread through a population. This epidemiological metric has been widely used to inform the risk and potential magnitude of a disease outbreak, the critical herd size required for pathogen persistence, and the efficacy or magnitude of control efforts [1,25]. This review summarizes R0 estimates of respiratory diseases in pigs, as reported in the literature, evaluates variations in reported R0 values, and compares estimates of disease transmission risk within and between pig farms.

2. Materials and Methods

A literature search was conducted in the PubMed, Web of Science, and EMBASE (OVID) databases to aggregate peer-reviewed original research articles on R0 of pig viral respiratory diseases. The PRISMA guidelines were utilized to identify, screen, and determine the eligibility of the articles. A list of diseases was identified by searching the Merck Veterinary Manual and CABI Compendium by combining the terms “respiratory disease” and “pigs” [6,26]. Search terms were aggregated into disease and reproductive number categories. The terms “swine”, “pig”, and “porcine” were used to narrow the study populations. Specific viral diseases and general terminology for respiratory diseases were combined to make a disease query. The full list of search queries is provided in Table 1. All citations were imported into EndNote 21 (EndNote Version 21.2, Clarivate, Philadelphia, PA, USA).
The search returned a combination of modeling exercises, laboratory experiments, and observed outbreaks. The exclusion criteria were as follows: bacterial diseases, studies that only reported the reproductive number for intervention strategies (i.e., vaccination), research not classified as original research, and articles not published in English. To be included, the paper had research that applied to domestic pig production. From the remaining studies, information on herd classification, modeling approach, and method of calculating R0 was extracted. The herd classifications were within- and between-herd to specify whether a paper-modeled disease spread within a single farm population or across multiple farms. A brief statement about the type of model used in each paper was created from their method section.

3. Results

The total number of papers identified was 1469, of which 631 were duplicates, leaving 838 studies. After screening the titles and abstracts, 148 papers were selected for full-text screening. Among these, 49 papers covering seven diseases were deemed eligible for inclusion in the review (Figure 1). R0 values were estimated for disease transmission within herds/farms and between herds/farms (Figure 2). Various methods were used in these papers to calculate the R0 values, ranging from complex mathematical approaches like the Next-Generation method to simple direct calculations of the number of secondary infections from simulation models. The most commonly used method to calculate R0 for a specific disease model was the product of the disease transmission rate and duration of infection (Table 2). These parameters were estimated through model fitting to reported case data or directly inferred from laboratory experiments.
The data used in these studies for calibrating the models primarily consist of surveillance data and laboratory experiments. Surveillance data, as defined in this review, refers to data from a secondary source, such as a government institution in a specific geographic region. In contrast, laboratory experiments were conducted to investigate disease transmission in a controlled environment. Nine studies did not use empirical data to parameterize their models or did not explicitly specify the sources of the data or values. A phylogenetic study utilized genetic data from previous outbreaks to calculate the reproductive number. The counts and distribution of the data types are shown in Figure 3.

3.1. Aujeszky’s Disease

Three publications evaluated the R0 value of Aujeszky’s disease (Table 2). Out of these three publications, two developed a generic pig population to model Aujeszky’s disease transmission, and one developed a model for Aujeszky’s disease outbreaks in the Netherlands [27,28,29]. All three studies used stochastic compartmental models of disease transmission [27,28,29]. The two generic population models estimated the R0 value to be 6 and 4.12, using the Next-Generation Matrix method and direct calculation from model simulations, respectively [27,29]. In the Netherlands study, the Next-Generation Matrix method was used to compute the between-herd R0 value, and a survival function method was used for the within-herd R0 value [28]. The between-herd R0 value was estimated to range between 1 and 2.5, and the within-herd R0 value was equal to 10 [28]. The within-herd R0 value of 10 was three times as high as the median between-herd R0 value of 3.31 (Table 2).

3.2. Porcine Circovirus 3

Only one study met the inclusion criteria for Porcine Circovirus 3. The study by Li et al. identified two major genotypes for PCV3 and estimated the R0 value to be 3.08 (95% credible interval (CrI): 0.82–7.96) and 1.82 (95% CrI: 0.63–4.82) for PCV3a and PCV3b, respectively [10]. The phylogenetic analysis used for R0 value calculations was based on a birth-death skyline model, and R0 was calculated as the product of the transmission rate and duration of infection [10].

3.3. Porcine Circovirus 2

Three studies from the same research team estimated R0 for Porcine Circovirus 2 (PCV2). Andraud et al. calculated the R0 value within and between pens for both types of within-herd transmission [30]. They calculated R0 under two different assumptions about infectiousness: (1) Infectiousness ends at seroconversion, and (2) Infectiousness ends at seroconversion, and viral load decline occurs. Under assumption 1, the between-pen R0 was estimated to be 0.58 (95% Confidence Interval (CI): 0.23–1.47), and under assumption 2, between-pen R0 increased to 1.2 (95% CI: 0.5, 2.9). The within-pen R0 was estimated to be 5.5 (95% CI: 3.3–9.0) under assumption 1 and 8.9 (95% CI: 5.1–15.4) under assumption 2 [30]. The study developed a stochastic compartmental model for PCV2 transmission parameterized using animal experiment data. R0 was calculated as the product of the transmission rate and duration of infection [30]. Another study by Andraud et al. used a deterministic model, parameterized using animal experiment data, to estimate the within-herd transmission of PCV2 [31]. The R0 value was 5.9 (95% CI: 1.8–10.1), estimated using a survival function method [31]. The last study used two experimental datasets to estimate the within-herd R0 value using a final-size algorithm [32]. These were time-series individual-level virological data collected from laboratory experiments to investigate PCV2 transmission in controlled environments. The first study was conducted by Rose et al., and the second by Andraud et al. [30,32]. Using the data produced by their experiment, Rose et al. calculated R0 as 6.94 (95% CI: 0.42–15.01) [32]. By merging their data with Andruad et al., a previous experiment by the same research group, R0 was estimated as 5.1 (95% CI: 2.5–8.2) [30,32].

3.4. Influenza A

Three publications estimated the within-herd R0 value of swine Influenza A (Table 2). Rose et al. used case data from outbreaks on swine farms in Brittany, France, and deterministic compartmental models to investigate the dynamics of within-farm swine Influenza outbreaks [5]. They used Poisson regression to fit their models to case incidence data and an exponential growth method to calculate the R0 value. The within-herd R0 value was estimated to range from 2.5 to 6.9 [5]. The remaining two papers used similar study designs and generated comparable results. These similarities include using data from animal experimental trial studies, stochastic compartmental models for disease transmission, and calculating R0 as the product of the transmission rate times the duration of infection (Table 2). Their R0 values of 10.4 and 10.66, respectively [33,34], were significantly lower than those reported by Rose et al. [5].

3.5. Porcine Reproductive and Respiratory Syndrome Virus

For Porcine Reproductive and Respiratory Syndrome Virus (PRRSV), six publications estimated the within-herd R0 to range from 2.6 to 5.42 (Table 2). Four studies used a generic pig population to model PRRSV outbreaks, and two modeled PRRSV outbreaks in the United States (Table 2). Of the generic population studies, three used laboratory experiments to inform parameter values of compartmental models and calculated R0 as the product of the transmission rate times the duration of the infection. The fourth study used the survival function method to calculate within-herd R0.
Two studies estimated the between-herd R0 in the United States using data from the Morrison Swine Health Monitoring Project (MSHMP) and transmission trees to model disease spread [35,36]. Pamornchianavakul et al. used phylogenetic analysis, while Arruda et al. utilized the Wallinga and Teunis method to calculate R0 values [35,36]. Both estimated between-herd R0 values of around 1 despite using different calculation methods (Table 2 [35,36]).

3.6. Encephalomyocarditis

Two publications estimated the within-herd R0 for encephalomyocarditis. Both studies used a stochastic compartmental model to investigate within-farm disease transmission and the White and Pagano method to compute R0 (Table 2). The first study by Kluivers et al. used serological data from infected pens on a pig farm during an autumn 2001 outbreak in Belgium to inform their model [13,14], and the second study by Maurice et al. used data from animal laboratory experiments. Kluivers et al. estimated R0 as 1.36 (95% CI: 0.93, 2.23), and Maurice et al. estimated it as 1.24 (95% CI: 0.39, 4.35) [13,14].

3.7. Classic Swine Fever

Twelve publications evaluated the R0 value of Classical Swine Fever (CSF) (Table 2). Eight of the twelve papers used generic populations to model CSF transmission; two papers used case data from the Netherlands, one used data from India, and one used data from Bangladesh (Table 2). Six studies calculated R0 as the product of transmission rate times duration of infection; two used an attack rate method, two used an exponential growth method, and two used a martingale estimator method (Table 2).
A machine learning model was developed by Suresh et al. to create an early warning system for potential outbreaks related to climate change [42]. The model was trained on geotagged incidence data for Guwahati City, Assam, from 2005 to 2021 [42]. Risk factors, such as humidity and vegetation, were added to the model to ascertain the climate-disease relationship on a spatial grid. R0 was calculated using the attack rate for each grid, producing values ranging from 1.04 to 2.07 [42]. Chowdhuary et al. used an exponential growth function to model surveillance data for a 2015 outbreak in the Kurigram district of Bangladesh [41]. They used two methods for calculating R0: the attack rate and the exponential growth rate (Table 2). The attack rate method estimated R0 as 1.6 (95% CI: 1.5–1.7) and exponential growth estimated as 1.5 (95% CI: 1.3–1.7) (Table 2).
Stegeman et al. published two papers using data from the 1997 to 1998 outbreak of CSF in the Netherlands to estimate the within- and between-herd R0 (Table 2). The between-herd R0 was estimated to be 6.8 [43]. Data were obtained from 21,500 herds in the Netherlands, of which 429 experienced an outbreak. In the second paper, a simple deterministic compartmental model was fitted to the data, and the within-herd R0 was calculated as the product of the transmission rate times the duration of infection. To calculate the within-herd R0, Stegeman et al. used data from 82 of the 429 infected herds because these outbreaks could be identified from a single known contact [44]. The within-herd R0 was estimated at 2.9, using the exponential growth method applied to a stochastic compartmental model [44].
Durand et al. investigated the potential use of clinical data (quantitative and qualitative), considered as infectivity markers, to reliably estimate CSF R0 value. The clinical-qualitative dataset reported the number of infected animals with clinical scores greater than 0, and the clinical-quantitative dataset reported the sum of clinical scores of infected animals [48]. They used viremia-based data (virus in the bloodstream) as the standard dataset for R0 calculation [48]. R0 values computed from the clinical datasets were higher than the viremia-based ones, but the confidence intervals all overlapped, indicating that these values were not significantly different (Table 2). The within-herd values varied from 4.0 (95% CI: 1.9–8.4) to 12.2 (95% CI: 5.5–27.3), and the between-herd values varied from 1.1 (95% CI: 0.4–2.9) to 1.8 (95% CI: 0.7–4.8) (Table 2) [48].
Laevans et al. performed two animal experimental studies using pigs of different ages to evaluate the within-pen risk of CSF transmission [25,26]. They used experimental data with different final outbreak size methods for R0 calculation to estimate within-herd R0 values [18,19]. The weaner pigs had an estimated R0 of 81.3, whereas the slaughter pigs had a significantly smaller estimate of 13.7 [18,19]. Klinkenberg et al. used data collected in the Laevans et al. experiments to parameterize a stochastic compartmental model of CSF transmission [18,19,45]. Calculating R0 as the product of the transmission rate times duration of infection, they estimated the within-herd R0 to be 100 for weaner pigs and 15.5 for the slaughter pigs; these estimates were much higher than those of Laevans et al. (Table 2). These results clearly illustrate the high sensitivity of R0 to the calculation method.
Weesendrop et al. reported the highest within-herd R0 values of 143 for a high transmissibility strain and 23 for a low transmissibility strain [49]. Bouma et al. reported the lowest within-herd R0 value, 2.3 (95% CI: 0–5.0) [46]. This disease had the largest range of R0 estimates across all diseases, likely due to the diversity in the R0 calculation methods and the type of data used. Studies using laboratory experimental data were more likely to generate higher R0 values than those using surveillance data (Table 2).

3.8. African Swine Fever

Eighteen publications estimated R0 for African Swine Fever (ASF), the most common disease identified in this search (Table 2). Among the articles extracted, there was an even split between studies using generic populations to model ASF transmission and those using laboratory experiment data and location-specific surveillance data. Nine studies utilized surveillance data to calibrate their models, four studies used data from Vietnam, two studies used data from China, two studies used data from Russia, one study used data from Ukraine, and one study used data from Uganda (Table 2). Fourteen studies estimated within-herd R0, and five studies estimated between-herd R0. Using genotyping information from surveillance data, R0 results were stratified by genotype.

3.8.1. Genotype IX

Only Barongo et al. investigated genotype IX, a strain mostly found in eastern Africa [51]. This Uganda study focused on small free-range pig production in the Gulu district, whose economy is rooted in small-scale agriculture [51]. The data used in this study were part of a larger effort to understand the transmission of ASF in Gulu District [51]. The period of interest was from April 2010 to November 2011 [51]. Herds were defined as all pigs in a pig-keeping household [51]. Barongo et al. used three separate models to depict transmission: a spatial simulation, an epidemic doubling time method, and a simple SI (Susceptible-Infected) deterministic model. The spatial simulation, referred to as the nearest infectious neighbor by Barongo et al., had the highest estimate of R0 as 3.24 (95% CI: 3.21–3.27) [51]. The doubling time method produced an R0 of 1.63. For the SI model, R0 was computed as the product of the transmission rate and duration of infection, and the model was fitted to the data in three ways: linear regression, curve fitting, and bootstrapping, leading to slightly different R0 values. The within-herd R0 was estimated as 1.58 for curve fitting, 1.77 for bootstrapping, and 1.9 for linear regression [51].

3.8.2. Genotype I

R0 estimates for genotype I were calculated in two laboratory experiments and in one study using surveillance data from Ukraine. Eble et al. used data from a previous laboratory experiment for the Netherlands/86 strain to calculate the R0 for carrier animals, finding that transmission does occur but at low levels with R0 less than 1 (R0 = 0.3) [52]. De Carvalho Ferreira et al. generated R0 estimates of 18.0 (95% CI: 6.90–46.9) for the Malta/78 isolate and 4.92 (95% CI: 1.45–16.6) for Netherlands/86 [53]. The Ukraine study investigated the transmission of ASF in the Odesa region of Ukraine [54]. It used historical data from the 1977 outbreak collected by the USSR Ministry of Agriculture [54]. The study estimated the between and within-herd R0 to be 1.65 (95% CI: 1.42–1.88) and 7.46 (95% CI: 5.68–9.21), respectively [54].

3.8.3. Genotype II

The most recent R0 estimates come from outbreaks in Asia. Of the 11 studies of genotype II ASF, three used data from outbreaks in China and four studies used data from Vietnam. All the studies out of Asia pertain to the introduction of the disease to the country, which occurred from late 2018 to early 2019 [55,57,58,59,61,62,63].
All studies on ASF in China used deterministic compartmental models and the Next-Generation Matrix method to estimate R0. Haung et al. used incidence data from the Ministry of Rural Agriculture of China and estimated R0 to be 0.6 [55]. Song et al. used data from the August 2018 ASF outbreak at an Aiyuan farm of Jiangsu Jiahua Breeding Pig Company in Siyang County, China [58]. The R0 value was estimated to be 4.18, which was smaller than that reported by Li et al. (Table 2) [58,59]. Li et al. used data from outbreaks in Xuanzhou at Guquan, Jinba, and Liacheng facilities during September 2018 and estimated R0 as 4.82 (95% CI: 3.84, 6.11) in Guquan, 7.94 (95% CI: 7.26, 8.72) in Jinba, and 11.90 (95% CI: 10.71, 12.91) in Liancheng [59].
All four studies pertaining to Vietnam used data from the 2019 outbreak of ASF in the northern region, the introduction of the disease into the country. Three of the four used compartmental models, and one used a serial interval model assuming a Poisson distribution to model daily incidence. Mai et al. was the only study that calculated the between-herd R0 [57]. The between-herd R0 was calculated for 15, 19, and 30-day infectious periods, yielding a range of values from 1.41 to 10.8 [57]. Table 2 shows the between-herd R0 values at different locations for the 15-day infectious period. Mai et al. used three methods to estimate the within-herd R0 using data from seven farms in Hung Yen and three farms in Ninh Binh and Ha Nam in Vietnam [62]. They estimated R0 as 1.49 using the exponential growth method, 1.58 using the White and Pagano method, and 1.46 using the attack rate method [62]. These data were also used to parameterize a SIR (Susceptible-Infected-Recovered) model under two farm sizes and lengths of the infectious period. They estimated R0 for a 100–299 pig farm to be 1.66 (95% CI: 0.88–2.84) and 1.4 (95% CI: 1.01–1.90) for a 300–999 pig farm [62]. Another study used only two mid-sized farms from Hung Yen using a statistics-based model [57]. These facilities were chosen for commercial design, strict biosecurity measures, and the way workers were assigned to only work in one barn with no visitors allowed on the premises [63]. Le et al. collected data from 15 March through 15 November 2019, on a small family-owned farm in Hanoi, Vietnam, that only had seventeen pigs [61]. They estimated the within-herd R0 to be 10.4 (95% CrI: 1.1–30.4) [61]. The remaining two studies estimated R0 to range from 1.55–1.78 [57,62].
The two studies from Russia used compartmental models and data from the Caucasian region and calculated R0 as the product of the transmission rate and the duration of the infection. Guinat et al. used data from the Federal Research Center for Virology and Microbiology and selected data from nine industrial herds between 2010 and 2014 [60]. A stochastic modeling approach was implemented to estimate the within-herd transmission of the infection [60]. The smallest within-herd R0 median value was estimated as 4.4 (95% CrI: 2.09–13.4) and the largest as 17.3 (95% 3.5–45.5) [60]. Gulenkin et al. obtained data from the World Animal Health Information System Office International Epizootic (WAHID OIE) database for the period between 2007 and 2010 [56]. They used a deterministic model for both between and within-herd disease transmission. Using a curve fitting method to estimate the model parameters, the within-herd R0 was estimated to range from 8 to 11. Alternatively, using a general linear model for model fitting R0, the value was estimated as 9.8 (95% CI: 3.9–15.6) [56]. Estimates for the between-herd R0 were limited to data from the Republic of North Ossetia in 2008, and its value ranged from 2 to 3 [56].
Two studies related to ASF genotype II used laboratory experiments to model within-herd ASF transmission. Guinet et al. utilized compartmental models, and the product of the transmission rate times duration of infection was the most used approach to calculate R0. Guinet et al. conducted an experiment on the Georgia 2007/1 strain and used a stochastic model considering different latent periods [64]. They calculated R0 for within and between pens for 3-, 4-, and 5-day latent periods. Oh et al. inoculated pigs using ASF from a 2020 outbreak in Thanh Hóa province, Vietnam, using survival analysis to model transmission between pigs [65]. R0 was calculated using the exponential growth method and White and Pagano method, producing the values 2.91 (95% CI: 1.51–7.06) and 4.015 (95% CI; 1.13–9.8), respectively.

3.8.4. Genotype Not Specified

Five papers produced mathematical models developed to analyze the dynamics of ASF outbreaks but did not specify the genotype used in parameterization. Parameters for these models come from the literature; therefore, these models are proof of concept rather than applied to specific outbreaks [68,69,70]. Four of the five studies estimated within-herd R0 ranging from 3.77 to 13.02. The fifth study developed a model to analyze the spread of ASF caused by a contaminated human vector between farms, calculating R0 as 18.57 [66].

4. Discussion

4.1. R0 Methodologies

The methods used for calculating R0 varied greatly between studies. This review found 16 different ways R0 was calculated, with each disease having at least two different calculation methods. Calculation methods ranged from simple, like the product of the transmission rate and the duration of infection, to more complex methods, like the Next-Generation Matrix. The measure of R0 itself is situation-specific. Studies using data from animal experimental trials tended to use the product of the transmission rate times the duration of the infection, while studies using surveillance data employed the largest range of calculation methods, with most studies using more than one method to estimate R0. For example, Chowdhuary et al. used both attack rate and exponential growth methods to estimate the R0 for CSF and found no significant differences between the estimates [41]. On the other hand, Barongo et al. used four methods for calculating R0, with one of the estimates above 3 and the rest between 1.5 and 2 [51]. Modeling studies using complex compartmental models for disease transmission have tended to use the Next-Generation Matrix method for R0 calculation. Differences in R0 estimates for the same pathogen may vary with R0 calculation methods, pathogen variants [37,71], the data source used for model calibration (surveillance vs. laboratory data), and data location.

4.2. Comparison of R0 Estimates from Laboratory and Surveillance Data

Laboratory experiments often yield higher R0 estimates compared to surveillance data for several reasons. It must be noted that laboratory experiments are limited in scope, as they can primarily be used to calculate within-herd transmission for a relatively small number of animals. In contrast, surveillance data are broad and can encompass not only the incidence on one farm but an entire geographic area, covering hundreds or thousands of animals. Due to the magnitude of farms, surveillance data may underreport cases as animals are not continuously tested, whereas the controlled environment of laboratory experiments allows for animals to be continuously tested. Furthermore, the environment of laboratory experiments lends itself to an increased transmission risk. Experiments are often conducted in relatively small pens or rooms, potentially enhancing disease transmission compared to actual farms that generally have larger pens or rooms. Laevans et al. attributed differences in pen size for weaner and slaughter pigs as the primary factor leading to a significant magnitude difference between the two estimates, 81.3 for weaner pigs and 13.7 for slaughter pigs [18,19]. Additionally, in experimental trials, the control group is not subject to any disease mitigation measures, whereas farms, especially industrial ones, continuously employ some level of biosecurity or mitigation measures. Some experiments in this review utilized high-virulence strains, which may not reflect the strains currently circulating on farms, potentially leading to an overestimation of R0 and misrepresentation of actual transmission risks [49,50,53].

4.3. Comparison of within and Between-Herd R0

Understanding the transmission dynamics of viruses involves examining two main aspects: within- and between-herd R0. Between-herd R0 tends to show lower values compared to within-herd R0, partly due to how it is measured. The between-herd R0 estimates the virus’s ability to spread from one farm to another, essentially tracking its movement between different populations in distinct areas. The unit of measurement here is a farm, regardless of its pig population size. Conversely, within-herd R0 focuses on the virus’s transmission from pig to pig within a single farm. This often yields higher estimates, possibly because there are more opportunities for interactions among animals within the same farm compared to interactions between animals from different farms, such as through animal movement. Consequently, diseases like viral respiratory illnesses are more likely to spread rapidly within a single farm than between different farms. When calculating within-herd R0, we typically rely on data about the incidence of infection within a specific farm, while between-herd R0 is based on information about the number of farms infected over time.

4.4. Disease Eradication and R0

The basic reproduction number is a key epidemiological metric that can be used to measure the transmission risk of infectious disease, predict the expected size of an outbreak, and estimate the control effort needed to prevent or interrupt disease transmission, among others [72]. However, the R0 value is not necessarily a direct indication of the potential eradication of an animal infectious disease. Disease eradication depends on several factors, such as disease characteristics (e.g., host reservoirs, clinical symptoms, infectious period, and immunity duration), economic factors, political will, and control tools (e.g., vaccine, depopulation, culling, movement restrictions) [73,74]. For example, an animal disease with a low R0 value, several host species, and/or wildlife reservoirs, such as Circovirus infection, may be harder to eradicate than a disease with a higher R0 value but limited to a single host species with no persistence in wildlife, such as AD.

4.5. Limitations of R0 Calculation for Complex Diseases

Diseases such as PRRSV have a complex infectious period, during which infectious hosts experience both acute and subclinical/chronic infectious phases with different levels of infectivity. Although this characteristic is central to PRRSV disease progression and transmission risk, most PRRSV R0 models have not explicitly considered the differential role of these infectious phases in disease transmission, and none has estimated their contribution to R0 value [7,35,36,37,38,39,40,75,76]. Such information would be pivotal for accurately evaluating the timing and effectiveness of disease control strategies. Further studies should investigate the differential contribution of acute and subclinical/chronic infectious phases to R0 and develop a general framework to deal with these situations in computing R0 and other disease transmission risk metrics.

5. Conclusions

Despite the limitations of R0, it is a useful tool to begin quantifying the risk of disease outbreaks and the magnitude of an outbreak and investigate intervention strategies’ effectiveness for disease control or prevention. For example, understanding the R0 for within-herd transmission is crucial for determining the vaccination efficacy required for preventing or eradicating disease on a farm. Between-herd R0 is paramount for understanding the vulnerability of farm networks to disease outbreaks and the vaccination coverage required to prevent large-scale disease outbreaks. Between-herd transmission is mostly driven by interactions between farms through commerce, supply chains, and the movement of infected livestock and related products. To improve the estimate of between-herd R0, future studies should explicitly incorporate data on interactions between farms. However, such interaction and mobility datasets are not readily available for most countries.
Swine diseases with a larger R0 emphasize the urgency of promptly detecting and responding to outbreaks. Addressing outbreaks can be challenging, regardless of resource availability. Through timely case surveillance, forecasting tools, and rapid dissemination of information to farmers, veterinarians, and government officials, animals and the food supply can be better protected against infectious disease outbreaks. Preventing swine respiratory diseases from spreading and entering the global supply chain is imperative, as they pose an imminent threat to food supply. Continuous monitoring and evaluation of pathogen transmissibility are paramount to enhance disease surveillance efficiency and decrease the impact of livestock diseases.

Author Contributions

Conceptualization, M.L.N.-M. and S.C.D.; methodology, D.C.P.R.; software, D.C.P.R.; validation, D.C.P.R. and M.L.N.-M.; formal analysis, D.C.P.R. and M.L.N.-M.; investigation, D.C.P.R.; resources, D.C.P.R.; data curation, D.C.P.R.; writing—original draft preparation, D.C.P.R. and M.L.N.-M.; writing—review and editing, D.C.P.R., M.L.N.-M. and S.C.D.; visualization, D.C.P.R.; supervision, M.L.N.-M. and S.C.D.; project administration, M.L.N.-M. and S.C.D.; funding acquisition, M.L.N.-M. and S.C.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the United States Department of Agriculture, grant number: APHIS-USDA AP23OA000000C013 to S.C.D. and M.L.N.-M. The funders had no role in the study design, data collection and analysis, decision to publish, or preparation of the manuscript.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. VanderWaal, K.; Deen, J. Global Trends in Infectious Diseases of Swine. Proc. Natl. Acad. Sci. USA 2018, 115, 11495–11500. [Google Scholar] [CrossRef] [PubMed]
  2. Renken, C.; Nathues, C.; Swam, H.; Fiebig, K.; Weiss, C.; Eddicks, M.; Ritzmann, M.; Nathues, H. Application of an Economic Calculator to Determine the Cost of Porcine Reproductive and Respiratory Syndrome at Farm-Level in 21 Pig Herds in Germany. Porc. Health Manag. 2021, 7, 3. [Google Scholar] [CrossRef] [PubMed]
  3. Calderón Díaz, J.A.; Fitzgerald, R.M.; Shalloo, L.; Rodrigues da Costa, M.; Niemi, J.; Leonard, F.C.; Kyriazakis, I.; García Manzanilla, E. Financial Analysis of Herd Status and Vaccination Practices for Porcine Reproductive and Respiratory Syndrome Virus, Swine Influenza Virus, and Mycoplasma Hyopneumoniae in Farrow-to-Finish Pig Farms Using a Bio-Economic Simulation Model. Front. Vet. Sci. 2020, 7, 556674. [Google Scholar] [CrossRef] [PubMed]
  4. Boeters, M.; Garcia-Morante, B.; van Schaik, G.; Segalés, J.; Rushton, J.; Steeneveld, W. The Economic Impact of Endemic Respiratory Disease in Pigs and Related Interventions-a Systematic Review. Porc. Health Manag. 2023, 9, 45. [Google Scholar] [CrossRef]
  5. Rose, N.; Hervé, S.; Eveno, E.; Barbier, N.; Eono, F.; Dorenlor, V.; Andraud, M.; Camsusou, C.; Madec, F.; Simon, G. Dynamics of Influenza A Virus Infections in Permanently Infected Pig Farms: Evidence of Recurrent Infections, Circulation of Several Swine Influenza Viruses, and Reassortment Events. Vet. Res. 2013, 44, 72. [Google Scholar] [CrossRef]
  6. Respiratory Diseases of Pigs. Available online: https://www.cabidigitallibrary.org/doi/10.1079/cabicompendium.66893 (accessed on 2 May 2024).
  7. Wittmann, G.; Hall, S.A. Aujeszky’s Disease. In Springer Science & Business Media; Martinus Nijhoff Publisher: Leiden, The Netherlands, 2012; ISBN 9789400975538. [Google Scholar]
  8. Ukhovskyi, V.V.; Korniienko, L.Y.; Pyskun, A.V.; Chechet, O.M.; Drozhzhe, Z.M.; Polupan, I.M.; Aliekseieva, G.B.; Moroz, O.A.; Romanov, O.M. Spread of Aujeszky’s Disease among Wild Boars and Domestic Pigs in Ukraine. Regul. Mech. Biosyst. 2022, 13, 46–52. [Google Scholar] [CrossRef]
  9. Chen, S.; Zhang, L.; Li, X.; Niu, G.; Ren, L. Recent Progress on Epidemiology and Pathobiology of Porcine Circovirus 3. Viruses 2021, 13, 1944. [Google Scholar] [CrossRef]
  10. Li, G.; He, W.; Zhu, H.; Bi, Y.; Wang, R.; Xing, G.; Zhang, C.; Zhou, J.; Yuen, K.-Y.; Gao, G.F.; et al. Origin, Genetic Diversity, and Evolutionary Dynamics of Novel Porcine Circovirus 3. Adv. Sci. 2018, 5, 1800275. [Google Scholar] [CrossRef] [PubMed]
  11. Verreault, D.; Létourneau, V.; Gendron, L.; Massé, D.; Gagnon, C.A.; Duchaine, C. Airborne Porcine Circovirus in Canadian Swine Confinement Buildings. Vet. Microbiol. 2010, 141, 224–230. [Google Scholar] [CrossRef]
  12. López-Lorenzo, G.; López-Novo, C.; Prieto, A.; Díaz, P.; Panadero, R.; Rodríguez-Vega, V.; Morrondo, P.; Fernández, G.; Díaz-Cao, J.M. Monitoring of Porcine Circovirus Type 2 Infection through Air and Surface Samples in Vaccinated and Unvaccinated Fattening Farms. Transbound. Emerg. Dis. 2022, 69, 1108–1117. [Google Scholar] [CrossRef]
  13. Kluivers, M.; Maurice, H.; Vyt, P.; Koenen, F.; Nielen, M. Transmission of Encephalomyocarditis Virus in Pigs Estimated from Field Data in Belgium by Means of R0. Vet. Res. 2006, 37, 757–766. [Google Scholar] [CrossRef] [PubMed]
  14. Maurice, H.; Nielen, M.; Stegeman, J.A.; Vanderhallen, H.; Koenen, F. Transmission of Encephalomyocarditis Virus (EMCV) among Pigs Experimentally Quantified. Vet. Microbiol. 2002, 88, 301–314. [Google Scholar] [CrossRef]
  15. Li, X.; Hu, Z.; Fan, M.; Tian, X.; Wu, W.; Gao, W.; Bian, L.; Jiang, X. Evidence of Aerosol Transmission of African Swine Fever Virus between Two Piggeries under Field Conditions: A Case Study. Front. Vet. Sci. 2023, 10, 1201503. [Google Scholar] [CrossRef]
  16. Olesen, A.S.; Lohse, L.; Boklund, A.; Halasa, T.; Gallardo, C.; Pejsak, Z.; Belsham, G.J.; Rasmussen, T.B.; Bøtner, A. Transmission of African Swine Fever Virus from Infected Pigs by Direct Contact and Aerosol Routes. Vet. Microbiol. 2017, 211, 92–102. [Google Scholar] [CrossRef] [PubMed]
  17. Weesendorp, E.; Landman, W.J.M.; Stegeman, A.; Loeffen, W.L.A. Detection and Quantification of Classical Swine Fever Virus in Air Samples Originating from Infected Pigs and Experimentally Produced Aerosols. Vet. Microbiol. 2008, 127, 50–62. [Google Scholar] [CrossRef] [PubMed]
  18. Laevens, H.; Koenen, F.; Deluyker, H.; Berkvens, D.; de Kruif, A. An Experimental Infection with Classical Swine Fever Virus in Weaner Pigs. I. Transmission of the Virus, Course of the Disease, and Antibody Response. Vet. Q. 1998, 20, 41–45. [Google Scholar] [CrossRef]
  19. Laevens, H.; Koenen, F.; Deluyker, H.; de Kruif, A. Experimental Infection of Slaughter Pigs with Classical Swine Fever Virus: Transmission of the Virus, Course of the Disease and Antibody Response. Vet. Rec. 1999, 145, 243–248. [Google Scholar] [CrossRef]
  20. Dee, S.A.; Deen, J.; Jacobson, L.; Rossow, K.D.; Mahlum, C.; Pijoan, C. Laboratory Model to Evaluate the Role of Aerosols in the Transport of Porcine Reproductive and Respiratory Syndrome Virus. Vet. Rec. 2005, 156, 501–504. [Google Scholar] [CrossRef]
  21. Kristensen, C.S.; Bøtner, A.; Takai, H.; Nielsen, J.P.; Jorsal, S.E. Experimental Airborne Transmission of PRRS Virus. Vet. Microbiol. 2004, 99, 197–202. [Google Scholar] [CrossRef]
  22. La, A.; Zhang, Q.; Cicek, N. Modelling Aerosol Transmission of Porcine Reproductive and Respiratory Syndrome Virus between Buildings Using Computational Fluid Dynamics. Biosyst. Eng. 2023, 236, 175–192. [Google Scholar] [CrossRef]
  23. Andraud, M.; Hervé, S.; Gorin, S.; Barbier, N.; Quéguiner, S.; Paboeuf, F.; Simon, G.; Rose, N. Evaluation of Early Single Dose Vaccination on Swine Influenza A Virus Transmission in Piglets: From Experimental Data to Mechanistic Modelling. Vaccine 2023, 41, 3119–3127. [Google Scholar] [CrossRef] [PubMed]
  24. Prost, K.; Kloeze, H.; Mukhi, S.; Bozek, K.; Poljak, Z.; Mubareka, S. Bioaerosol and Surface Sampling for the Surveillance of Influenza A Virus in Swine. Transbound. Emerg. Dis. 2019, 66, 1210–1217. [Google Scholar] [CrossRef] [PubMed]
  25. Vynnycky, E.; White, R. An Introduction to Infectious Disease Modelling; Oxford University Press: Oxford, UK, 2010; ISBN 9780198565765. [Google Scholar]
  26. Overview of Respiratory Diseases of Pigs. Available online: https://www.merckvetmanual.com/respiratory-system/respiratory-diseases-of-pigs/overview-of-respiratory-diseases-of-pigs (accessed on 2 May 2024).
  27. Houben, E.H.P.; Dijkhuizen, A.A.; de Jong, M.C.M.; Kimman, T.G.; van der Valk, P.C.; Verheijden, J.H.M.; Nieuwenhuis, H.U.R.; Hunneman, W.A.; Huysman, C.N. Control Measures Directed at Aujeszky’s Disease Virus: A Theoretical Evaluation of between-Farm Effects. Prev. Vet. Med. 1993, 15, 35–52. [Google Scholar] [CrossRef]
  28. Buijtels, J.; Huirne, R.; Dijkhuizen, A.; de Jong, M.; van Nes, A. Computer Simulation to Support Policy Making in the Control of Pseudorabies. Vet. Microbiol. 1997, 55, 181–185. [Google Scholar] [CrossRef] [PubMed]
  29. Van Nes, A.; De Jong, M.C.; Buijtels, J.A.; Verheijden, J.H. Implications Derived from a Mathematical Model for Eradication of Pseudorabies Virus. Prev. Vet. Med. 1998, 33, 39–58. [Google Scholar]
  30. Andraud, M.; Grasland, B.; Durand, B.; Cariolet, R.; Jestin, A.; Madec, F.; Rose, N. Quantification of Porcine Circovirus Type 2 (PCV-2) within- and between-Pen Transmission in Pigs. Vet. Res. 2008, 39, 43. [Google Scholar] [CrossRef] [PubMed]
  31. Andraud, M.; Grasland, B.; Durand, B.; Cariolet, R.; Jestin, A.; Madec, F.; Pierre, J.S.; Rose, N. Modelling the Time-Dependent Transmission Rate for Porcine Circovirus Type 2 (PCV2) in Pigs Using Data from Serial Transmission Experiments. J. R. Soc. Interface 2009, 6, 39–50. [Google Scholar] [CrossRef]
  32. Rose, N.; Andraud, M.; Bigault, L.; Jestin, A.; Grasland, B. A Commercial PCV2a-Based Vaccine Significantly Reduces PCV2b Transmission in Experimental Conditions. Vaccine 2016, 34, 3738–3745. [Google Scholar] [CrossRef] [PubMed]
  33. Allerson, M.; Deen, J.; Detmer, S.E.; Gramer, M.R.; Joo, H.S.; Romagosa, A.; Torremorell, M. The Impact of Maternally Derived Immunity on Influenza A Virus Transmission in Neonatal Pig Populations. Vaccine 2013, 31, 500–505. [Google Scholar] [CrossRef]
  34. Romagosa, A.; Allerson, M.; Gramer, M.; Joo, H.S.; Deen, J.; Detmer, S.; Torremorell, M. Vaccination of Influenza a Virus Decreases Transmission Rates in Pigs. Vet. Res. 2011, 42, 120. [Google Scholar] [CrossRef]
  35. Pamornchainavakul, N.; Makau, D.N.; Paploski, I.A.D.; Corzo, C.A.; VanderWaal, K. Unveiling Invisible Farm-to-Farm PRRSV-2 Transmission Links and Routes through Transmission Tree and Network Analysis. Evol. Appl. 2023, 16, 1721–1734. [Google Scholar] [CrossRef] [PubMed]
  36. Arruda, A.G.; Alkhamis, M.A.; VanderWaal, K.; Morrison, R.B.; Perez, A.M. Estimation of Time-Dependent Reproduction Numbers for Porcine Reproductive and Respiratory Syndrome across Different Regions and Production Systems of the US. Front. Vet. Sci. 2017, 4, 46. [Google Scholar] [CrossRef]
  37. Chase-Topping, M.; Xie, J.; Pooley, C.; Trus, I.; Bonckaert, C.; Rediger, K.; Bailey, R.I.; Brown, H.; Bitsouni, V.; Barrio, M.B.; et al. New Insights about Vaccine Effectiveness: Impact of Attenuated PRRS-Strain Vaccination on Heterologous Strain Transmission. Vaccine 2020, 38, 3050–3061. [Google Scholar] [CrossRef] [PubMed]
  38. Charpin, C.; Mahé, S.; Keranflec’h, A.; Belloc, C.; Cariolet, R.; Le Potier, M.-F.; Rose, N. Infectiousness of Pigs Infected by the Porcine Reproductive and Respiratory Syndrome Virus (PRRSV) Is Time-Dependent. Vet. Res. 2012, 43, 69. [Google Scholar] [CrossRef]
  39. Rose, N.; Renson, P.; Andraud, M.; Paboeuf, F.; Le Potier, M.F.; Bourry, O. Porcine Reproductive and Respiratory Syndrome Virus (PRRSv) Modified-Live Vaccine Reduces Virus Transmission in Experimental Conditions. Vaccine 2015, 33, 2493–2499. [Google Scholar] [CrossRef]
  40. Pileri, E.; Gibert, E.; Soldevila, F.; García-Saenz, A.; Pujols, J.; Diaz, I.; Darwich, L.; Casal, J.; Martín, M.; Mateu, E. Vaccination with a Genotype 1 Modified Live Vaccine against Porcine Reproductive and Respiratory Syndrome Virus Significantly Reduces Viremia, Viral Shedding and Transmission of the Virus in a Quasi-Natural Experimental Model. Vet. Microbiol. 2015, 175, 7–16. [Google Scholar] [CrossRef] [PubMed]
  41. Chowdhuary, M.G.A.; Islam, S.S.; Khair, A.; Hossain, M.M.; Ahmed, G.; Brum, E.; Debnath, N.C.; Badhy, S.C.; Habib, M.A.; Rumi, T.B.; et al. A Further Outbreak of Classical Swine Fever in Indigenous Pigs in Kurigram District, Bangladesh. Transbound. Emerg. Dis. 2020, 67, 1922–1929. [Google Scholar] [CrossRef]
  42. Suresh, K.P.; Barman, N.N.; Bari, T.; Jagadish, D.; Sushma, B.; Darshan, H.V.; Patil, S.S.; Bora, M.; Deka, A. Application of Machine Learning Models for Risk Estimation and Risk Prediction of Classical Swine Fever in Assam, India. Virusdisease 2023, 34, 514–525. [Google Scholar] [CrossRef]
  43. Stegeman, A.; Elbers, A.R.; Smak, J.; de Jong, M.C. Quantification of the Transmission of Classical Swine Fever Virus between Herds during the 1997–1998 Epidemic in The Netherlands. Prev. Vet. Med. 1999, 42, 219–234. [Google Scholar] [CrossRef]
  44. Stegeman, A.; Elbers, A.R.; Bouma, A.; de Smit, H.; de Jong, M.C. Transmission of Classical Swine Fever Virus within Herds during the 1997–1998 Epidemic in The Netherlands. Prev. Vet. Med. 1999, 42, 201–218. [Google Scholar]
  45. Klinkenberg, D.; de Bree, J.; Laevens, H.; de Jong, M.C.M. Within- and between-Pen Transmission of Classical Swine Fever Virus: A New Method to Estimate the Basic Reproduction Ratio from Transmission Experiments. Epidemiol. Infect. 2002, 128, 293–299. [Google Scholar] [CrossRef] [PubMed]
  46. Bouma, A.; De Smit, A.J.; De Jong, M.C.; De Kluijver, E.P.; Moormann, R.J. Determination of the Onset of the Herd-Immunity Induced by the E2 Sub-Unit Vaccine against Classical Swine Fever Virus. Vaccine 2000, 18, 1374–1381. [Google Scholar] [CrossRef]
  47. Dewulf, J.; Laevens, H.; Koenen, F.; Mintiens, K.; De Kruif, A. An Experimental Infection with Classical Swine Fever Virus in Pregnant Sows: Transmission of the Virus, Course of the Disease, Antibody Response and Effect on Gestation. J. Vet. Med. B Infect. Dis. Vet. Public Health 2001, 48, 583–591. [Google Scholar] [CrossRef]
  48. Durand, B.; Davila, S.; Cariolet, R.; Mesplède, A.; Le Potier, M.-F. Comparison of Viraemia- and Clinical-Based Estimates of within- and between-Pen Transmission of Classical Swine Fever Virus from Three Transmission Experiments. Vet. Microbiol. 2009, 135, 196–204. [Google Scholar] [CrossRef]
  49. Weesendorp, E.; Backer, J.; Stegeman, A.; Loeffen, W. Transmission of Classical Swine Fever Virus Depends on the Clinical Course of Infection Which Is Associated with High and Low Levels of Virus Excretion. Vet. Microbiol. 2011, 147, 262–273. [Google Scholar] [CrossRef]
  50. Weesendorp, E.; Backer, J.; Stegeman, A.; Loeffen, W. Effect of Strain and Inoculation Dose of Classical Swine Fever Virus on within-Pen Transmission. Vet. Res. 2009, 40, 59. [Google Scholar] [CrossRef] [PubMed]
  51. Barongo, M.B.; Ståhl, K.; Bett, B.; Bishop, R.P.; Fèvre, E.M.; Aliro, T.; Okoth, E.; Masembe, C.; Knobel, D.; Ssematimba, A. Estimating the Basic Reproductive Number (R0) for African Swine Fever Virus (ASFV) Transmission between Pig Herds in Uganda. PLoS ONE 2015, 10, e0125842. [Google Scholar] [CrossRef] [PubMed]
  52. Eblé, P.L.; Hagenaars, T.J.; Weesendorp, E.; Quak, S.; Moonen-Leusen, H.W.; Loeffen, W.L.A. Transmission of African Swine Fever Virus via Carrier (survivor) Pigs Does Occur. Vet. Microbiol. 2019, 237, 108345. [Google Scholar] [CrossRef] [PubMed]
  53. de Carvalho Ferreira, H.C.; Backer, J.A.; Weesendorp, E.; Klinkenberg, D.; Stegeman, J.A.; Loeffen, W.L.A. Transmission Rate of African Swine Fever Virus under Experimental Conditions. Vet. Microbiol. 2013, 165, 296–304. [Google Scholar] [CrossRef]
  54. Korennoy, F.I.; Gulenkin, V.M.; Gogin, A.E.; Vergne, T.; Karaulov, A.K. Estimating the Basic Reproductive Number for African Swine Fever Using the Ukrainian Historical Epidemic of 1977. Transbound. Emerg. Dis. 2017, 64, 1858–1866. [Google Scholar] [CrossRef]
  55. Huang, Y.; Li, J.; Zhang, J.; Jin, Z. Dynamical Analysis of the Spread of African Swine Fever with the Live Pig Price in China. Math. Biosci. Eng. 2021, 18, 8123–8148. [Google Scholar] [CrossRef] [PubMed]
  56. Gulenkin, V.M.; Korennoy, F.I.; Karaulov, A.K.; Dudnikov, S.A. Cartographical Analysis of African Swine Fever Outbreaks in the Territory of the Russian Federation and Computer Modeling of the Basic Reproduction Ratio. Prev. Vet. Med. 2011, 102, 167–174. [Google Scholar] [CrossRef] [PubMed]
  57. Mai, T.N.; Nguyen, T.T.; Vu, V.A.; Vu, T.N.; Huynh, T.M.L. Estimation of the Herd-Level Basic Reproduction Number for African Swine Fever in Vietnam, 2019. Vet. World 2022, 15, 2850–2855. [Google Scholar] [CrossRef] [PubMed]
  58. Song, H.; Guo, L.; Jin, Z.; Liu, S. Modelling and Stability Analysis of ASFV with Swill and the Virus in the Environment. Math. Biosci. Eng. 2022, 19, 13028–13049. [Google Scholar] [CrossRef] [PubMed]
  59. Li, J.; Jin, Z.; Wang, Y.; Sun, X.; Xu, Q.; Kang, J.; Huang, B.; Zhu, H. Data-Driven Dynamical Modelling of the Transmission of African Swine Fever in a Few Places in China. Transbound. Emerg. Dis. 2022, 69, E646–E658. [Google Scholar] [CrossRef]
  60. Guinat, C.; Porphyre, T.; Gogin, A.; Dixon, L.; Pfeiffer, D.U.; Gubbins, S. Inferring within-Herd Transmission Parameters for African Swine Fever Virus Using Mortality Data from Outbreaks in the Russian Federation. Transbound. Emerg. Dis. 2018, 65, e264–e271. [Google Scholar] [CrossRef]
  61. Le, V.P.; Lan, N.T.; Canevari, J.T.; Villanueva-Cabezas, J.P.; Padungtod, P.; Trinh, T.B.N.; Nguyen, V.T.; Pfeiffer, C.N.; Oberin, M.V.; Firestone, S.M.; et al. Estimation of a Within-Herd Transmission Rate for African Swine Fever in Vietnam. Animals 2023, 13, 571. [Google Scholar] [CrossRef]
  62. Mai, T.N.; Sekiguchi, S.; Huynh, T.M.L.; Cao, T.B.P.; Le, V.P.; Dong, V.H.; Vu, V.A.; Wiratsudakul, A. Dynamic Models of Within-Herd Transmission and Recommendation for Vaccination Coverage Requirement in the Case of African Swine Fever in Vietnam. Vet. Sci. China 2022, 9, 292. [Google Scholar] [CrossRef]
  63. Mai, N.T.A.; Trinh, T.B.N.; Nguyen, V.T.; Lai, T.N.H.; Le, N.P.; Nguyen, T.T.H.; Nguyen, T.L.; Ambagala, A.; Do, D.L.; Van Phan, L. Estimation of Basic Reproduction Number (R0) of African Swine Fever (ASF) in Mid-Size Commercial Pig Farms in Vietnam. Front. Vet. Sci. 2022, 9, 918438. [Google Scholar] [CrossRef]
  64. Guinat, C.; Gubbins, S.; Vergne, T.; Gonzales, J.L.; Dixon, L.; Pfeiffer, D.U. Experimental Pig-to-Pig Transmission Dynamics for African Swine Fever Virus, Georgia 2007/1 Strain. Epidemiol. Infect. 2016, 144, 25–34. [Google Scholar] [CrossRef]
  65. Oh, S.-I.; Bui, N.A.; Bui, V.N.; Dao, D.T.; Cho, A.; Lee, H.G.; Jung, Y.-H.; Do, Y.J.; Kim, E.; Bok, E.-Y.; et al. Pathobiological Analysis of African Swine Fever Virus Contact-Exposed Pigs and Estimation of the Basic Reproduction Number of the Virus in Vietnam. Porc. Health Manag. 2023, 9, 30. [Google Scholar] [CrossRef] [PubMed]
  66. Chuchard, P.; Prathumwan, D.; Trachoo, K.; Maiaugree, W.; Chaiya, I. The SLI-SC Mathematical Model of African Swine Fever Transmission among Swine Farms: The Effect of Contaminated Human Vector. Axioms 2022, 11, 329. [Google Scholar] [CrossRef]
  67. Kouidere, A.; Balatif, O.; Rachik, M. Analysis and Optimal Control of a Mathematical Modeling of the Spread of African Swine Fever Virus with a Case Study of South Korea and Cost-Effectiveness. Chaos Solitons Fractals 2021, 146, 110867. [Google Scholar] [CrossRef]
  68. Shi, R.; Li, Y.; Wang, C. Analysis of a Fractional-Order Model for African Swine Fever with Effect of Limited Medical Resources. Fractal Fract. 2023, 7, 430. [Google Scholar] [CrossRef]
  69. Shi, R.; Zhang, Y.; Wang, C. Dynamic Analysis and Optimal Control of Fractional Order African Swine Fever Models with Media Coverage. Animals 2023, 13, 2252. [Google Scholar] [CrossRef]
  70. Shi, R.; Li, Y.; Wang, C. Stability Analysis and Optimal Control of a Fractional-Order Model for African Swine Fever. Virus. Res. 2020, 288, 198111. [Google Scholar] [CrossRef] [PubMed]
  71. Liu, Y.; Lillepold, K.; Semenza, J.C.; Tozan, Y.; Quam, M.B.M.; Rocklöv, J. Reviewing Estimates of the Basic Reproduction Number for Dengue, Zika and Chikungunya across Global Climate Zones. Environ. Res. 2020, 182, 109114. [Google Scholar] [CrossRef] [PubMed]
  72. Inaba, H.; Nishiura, H. The Basic Reproduction Number of an Infectious Disease in a Stable Population: The Impact of Population Growth Rate on the Eradication Threshold. Math. Model. Nat. Phenom. 2008, 3, 194–228. [Google Scholar] [CrossRef]
  73. Thomson, G.R.; Penrith, M.-L. Eradication of Transboundary Animal Diseases: Can the Rinderpest Success Story Be Repeated? Transbound. Emerg. Dis. 2017, 64, 459–475. [Google Scholar] [CrossRef]
  74. Elbers, A.R.; Braamskamp, J.; Dekkers, L.J.; Voets, R.; Duinhof, T.; Hunneman, W.A.; Stegeman, J.A. Aujeszky’s Disease Virus Eradication Campaign Successfully Heading for Last Stage in The Netherlands. Vet. Q. 2000, 22, 103–107. [Google Scholar] [CrossRef]
  75. Phoo-ngurn, P.; Kiataramkul, C.; Chamchod, F. Modeling the Spread of Porcine Reproductive and Respiratory Syndrome Virus (PRRSV) in a Swine Population: Transmission Dynamics, Immunity Information, and Optimal Control Strategies. Adv. Differ. Equ. 2019, 2019, 432. [Google Scholar] [CrossRef]
  76. Evans, C.M.; Medley, G.F.; Creasey, S.J.; Green, L.E. A Stochastic Mathematical Model of the within-Herd Transmission Dynamics of Porcine Reproductive and Respiratory Syndrome Virus (PRRSV): Fade-out and Persistence. Prev. Vet. Med. 2010, 93, 248–257. [Google Scholar] [CrossRef] [PubMed]
Figure 1. PRISMA chart of Pig Viral Respiratory diseases.
Figure 1. PRISMA chart of Pig Viral Respiratory diseases.
Vetsci 11 00300 g001
Figure 2. Distribution of R0 values by disease for within- and between-herd transmission.
Figure 2. Distribution of R0 values by disease for within- and between-herd transmission.
Vetsci 11 00300 g002
Figure 3. Distribution of data types extracted for all swine respiratory diseases.
Figure 3. Distribution of data types extracted for all swine respiratory diseases.
Vetsci 11 00300 g003
Table 1. Terms by category used for generating results.
Table 1. Terms by category used for generating results.
CategoryTerms
Mathematical“Reproduction number” OR “reproductive number” OR “reproductive rate” OR “reproductive ratio” OR “reproduction ratio” OR “R0” OR “force of infection” OR “secondary infection” OR “secondary case” OR “attack rate” OR “attack ratio” OR “population dynamics” OR “infection dynamics” OR “transmission rate” or “R0 Structural Parameters”
Animallivestock OR farm * OR pig OR porcine OR swine
Disease“Respiratory virus” OR “Respiratory disease” OR paramyxovirus OR herpesvirus OR alphaherpesvirus OR “adenovirus” OR “viral pneumonia” OR “porcine respiratory disease” OR “porcine infectious disease” OR “porcine coronavirus” OR “porcine respirovirus 1” OR “swine orthopneuovirus” OR “swine influenza” OR “Porcine reproductive and respiratory virus” OR PRRSV OR “syndrome PRRSV Virus” OR “swine influenza” OR SIV OR “African Swine Fever” OR ASF OR ASFV OR “suid herpesvirus 1” OR SHV1 OR “Porcine Epidemic Abortion and Respiratory Syndrome” OR “classical swine fever” OR “CSF” OR “H1N1”OR “H3N2” OR “H1N2”OR “PRCV” OR “PCV2” OR “porcine cytomegalovirus” OR “PCMV”OR “Aujeszky’s disease virus” OR “ADV”OR “Encephalmyocarditis Virus” OR “hemagglutinating encephalomyelitis virus”
The asterisk is a wildcard search operator to include different ending of the word like “farming” and “farms”.
Table 2. Extraction results stratified by disease.
Table 2. Extraction results stratified by disease.
StudyR0Herd TypeLocationMethodModelData Source
Aujeszky’s Disease/Pseudorabies
(Houben et al., 1993) [27]6BetweenGenericNext-Generation MethodStochastic compartmental modelNo empirical data used
(Buijtels et al., 1997) [28]4.12BetweenGenericEstimated directly from model simulationStochastic compartmental modelNo empirical data used
(Van Nes et al., 1998) [29] 1–2.5BetweenThe NetherlandsNext-Generation MethodStochastic compartmental modelEmpirical case data from one of the largest breeding companies in the Netherlands and the Central Bureau of Statistics.
(Van Nes et al., 1998) [29]10WithinThe NetherlandsSurvival Function MethodStochastic compartmental modelEmpirical case data from one of the largest breeding companies in the Netherlands and the Central Bureau of Statistics.
Porcine Circovirus 3
(Li et al., 2018) [10]3.08
(CrI: 0.82–7.96)
Not specifiedGlobalProduct of the transmission rate and duration of infectionBirth-death skyline modelGene sequences data
(Li et al., 2018) [10]1.82
(CrI: 0.63–4.86)
Not specifiedGlobalProduct of transmission rate and duration of infectionBirth-death skyline modelGene sequences data
Porcine Circovirus 2
(Andraud et al., 2008) [30]Between pen 1:
0.58
(CI: 0.23–1.47)
Between pen 2:
1.2 (CI: 0.5–2.9)
Within pen 1
5.5 (CI: 3.3–9.0)
Within pen 2
8.9 (CI:5.1–15.4)
WithinGenericProduct of transmission rate and duration of infectionStochastic compartmental modelAnimal experiments conducted by the authors (France)
(Andraud et al., 2009) [31]5.9 (CI: 1.8–10.1)WithinGenericSurvival Function MethodDeterministic compartmental modelAnimal experiments conducted by the authors (France)
(Rose et al., 2016) [32]5.1 (CrI: 2.5–8.2)
6.94 (CrI: 0.42–15.05)
WithinGenericFinal-Size AlgorithmExponential decay function to model the individual probability of escaping infectionAnimal experiments conducted by the authors (France)
Influenza A
(Allerson et al., 2013) [33]10.4
(CI: 6.6–15.8)
WithinGenericProduct of transmission rate and duration of infectionStochastic compartmental modelAnimal experiments conducted by the authors (United States)
(Rose et al., 2013) [5]2.5 (CI: 1.9–2.9)
to
6.9 (CI: 4.1–10.5)
WithinFranceExponential growth MethodDeterministic compartmental modelCases data of outbreaks in swine farms located in Brittany, France
(Romagosa et al., 2011) [34]10.66
(CI: 6.57–16.46)
WithinGenericProduct of transmission rate and duration of infectionStochastic compartmental modelAnimal experiments conducted by the authors (United States)
PRRSV
(Pamornchainavakul et al., 2023) [35]1 (IQR: 1–2). Highest value of 5BetweenUnited StatesPhylogenetic analysisTransmission tree modelMorrison Swine Health Monitoring Project (MSHMP) database, which was established to track progress on PRRS control in the U.S. and aimed to be the national hub for voluntary data sharing between swine veterinarians from different production systems
(Arruda et al., 2017) [36]Southeast: 1.14,
South: 1.14,
Upper Midwest East: 1.30,
Upper Midwest West: 1.10.
BetweenUnited StatesWallinga & Teunis MethodTransmission tree modelMSHMP database, which was established to track progress on PRRS control in the U.S. and aimed to be the national hub for voluntary data sharing between swine veterinarians from different production systems. Four areas across the US were chosen, including farms located in the states of North Carolina [Southeast (SE)], Oklahoma [South (S)], Minnesota/Iowa [Upper Midwest East (UME)], and Nebraska/South Dakota [Upper Midwest West (UMW)]. Those regions represented areas within the US characterized by high (SE and UME) and low (S and UMW) swine density
(Chase-Topping et al., 2020) [37]5.93–32.3WithinGenericProduct of transmission rate and duration of infectionStochastic compartmental modelData from an animal experiment conducted by authors (Belgium)
(Charpin et al., 2012) [38]2.6 (CI: 1.8–3.3)WithinGenericSurvival Function MethodDeterministic compartmental modelData from an animal experiment conducted by authors (France)
(Rose et al., 2015) [39]5.42 (CI: 2.94–9.04)WithinGenericProduct of transmission rate and duration of infectionStochastic compartmental modelAnimal experiments conducted by the authors (France)
(Pileri et al., 2015) [40]2.78 (CI: 2.13–3.43)WithinGenericProduct of transmission rate and duration of infectionStochastic compartmental modelAnimal experiments conducted by the authors (Spain)
Encephalomyocarditis
(Kluivers et al., 2006) [13]1.36 (0.93–2.23)WithinBelgiumWhite and Pagano MethodStochastic compartmental modelSerology data from infected pens in a pig farm during an autumn 2001 outbreak in Belgium
(Maurice et al., 2002) [14]1.24 (0.39–4.35)WithinGenericWhite and Pagano MethodStochastic compartmental modelAnimal experiments conducted by the authors (Belgium)
Classical Swine Fever
(Chowdhuary et al., 2020) [41]1.6 (CI:1.5–1.7)BetweenBangladeshAttack Rate MethodExponential growth functionPrimary data of identified cases of CSF during the Nov–Dec 2015 outbreak in Bangladesh
(Chowdhuary et al., 2020) [41]1.5 (CI: 1.3–1.7)BetweenBangladeshExponential Growth MethodExponential growth functionPrimary data of identified cases of CSF during the Nov–Dec 2015 outbreak in Bangladesh
(Suresh et al., 2023) [42]1.0–2.07BetweenIndiaAttack Rate MethodDiverse machine Learning Algorithms such as Random Forest, Classification tree analysis, and gradient boostingIncidence data from 2005 to 2021 from surveillance activities in India.
(Stegeman et al., 1999) [43]6.8BetweenThe NetherlandsProduct of the transmission rate and duration of infectionDeterministic compartmental modelSurveillance by the authors of 21,500 herds that reported the infection between 1997 and 1998
(Stegeman et al., 1999) [44]2.9WithinThe NetherlandsExponential Growth MethodStochastic compartmental modelSurveillance by the authors of 82 herds that reported the infection between 1997 and 1998
(Laevens et al., 1998) [18]81.3 (S.E. 109.54)WithinGenericMartingale estimator using final outbreak size approachKaplan-Meier Survival AnalysisAnimal experiments conducted by the authors (Belgium)
(Laevens et al., 1999) [19]13.7 (S.E. 13.7)WithinGenericMartingale estimator using final outbreak size approachCox Proportional Hazard AnalysisAnimal experiments conducted by the authors (Belgium)
(Klinkenberg et al., 2002) [45]Within Pen:
Weaner: 100 (CI: 54.4–186)
Slaughter 15.5 (CI: 6.20–38.7)
Between Pen:
Weaner 7.77 (CI: 4.68–12.9
Slaughter 3.39 (CI: 1.54–7.45)
WithinGenericProduct of the transmission and duration of infectionStochastic compartmental modelAnimal experiments by Laevens et al. [19]
(Bouma et al., 2000) [46]2.3 (CI:0–5.0)WithinGenericProduct of the transmission rate and duration of infectionStochastic compartmental modelAnimal experiments conducted by the authors (The Netherlands)
(Dewulf et al., 2001) [47]13.0WithinGenericMartingale estimator using final outbreak size approach Stochastic compartmental modelAnimal experiments conducted by the authors (Belgium)
(Durand et al., 2009) [48]Within Pen:
(VB) dataset 4.0 (CI: 1.9–8.4), (CB) dataset qualitative 7.6 (CI: 3.4–17.0) quantitative 12.2 (CI: 5.5–27.3)
Between Pen:
(VB) dataset 1.1 (CI: 0.4–2.9), (CB) dataset qualitative 1.4 (CI: 0.5–3.7), quantitative 1.8 (CI: 0.7–4.8)
WithinGenericProduct of the transmission and duration of infectionDeterministic compartmental modelAnimal experiments conducted by the authors. Viraemia-based (VB), Clinical-based (CB) qualitative or quantitative. (France)
(Weesendorp et al., 2011) [49]Low: 23.1 (CI: 11.5–45.0)
High: 143 (CI: 56.3–373)
WithinGenericProduct of the transmission rate and the duration of infectionStochastic compartmental modelAnimal experiments conducted by the authors (The Netherlands)
(Weesendorp et al., 2009) [50]Zoelen: 0 (CI: 0–0.92) Paderborn Middle: 148 (CI: 53.8–382) Paderborn High: 35.9 (CI: 14.5–77.6) Brescia: 17.5 (CI: 7.13–36.9)WithinGenericProduct of the transmission rate and the duration of infectionStochastic compartmental modelAnimal experiments conducted by the authors (The Netherlands)
African Swine Fever
Genotype IX
(Barongo et al., 2015) [51]1.63 (CI:1.56–1.72)BetweenUgandaExponential Growth MethodEpidemic doubling methodData was from previous studies from villages in Gulu District performed by other researchers. Outbreaks included in this study occurred between April 2010 and November 2011
(Barongo et al., 2015) [51]3.24
(CI: 3.21–3.27)
BetweenUgandaEstimated directly from model simulationSpatial simulation modelData was from previous studies from villages in Gulu District performed by other researchers. Outbreaks included in this study occurred between April 2010 and November 2011
(Barongo et al., 2015) [51]1.90
(CI: 1.87–1.94)
BetweenUgandaProduct of transmission rate and duration of infection Deterministic compartmental model (model fitting using linear regression)Data was from previous studies from villages in Gulu District performed by other researchers. Outbreaks included in this study occurred between April 2010 and November 2011
(Barongo et al., 2015) [51]1.58BetweenUgandaProduct of transmission rate and duration of infectionDeterministic compartmental model (model fitted using curve fitting)Data was from previous studies from villages in Gulu District performed by other researchers. Outbreaks included in this study occurred between April 2010 and November 2011
(Barongo et al., 2015) [51]1.77 (CI: 1.74–1.81)BetweenUgandaProduct of transmission rate and duration of infectionDeterministic compartmental model (model calibrated using bootstrapping)Data was from previous studies from villages in Gulu District performed by other researchers. Outbreaks included in this study occurred between April 2010 and November 2011
Genotype I
(Eblé et al., 2019) [52]0.3WithinGenericProduct of the transmission rate and the duration of infectionDeterministic compartmental modelData from lab conducted by other researcher (The Netherlands)
(de Carvalho Ferreira et al., 2013) [53]18.0
(CI: 6.90–46.9).
WithinGenericProduct of the transmission rate and the duration of infectionReconstructed compartmental modelAnimal experiments conducted by the authors (The Netherlands)
(Korennoy et al., 2017) [54]7.46
(CI: 5.68–9.21)
WithinUkraineExponential Growth MethodThe epidemic curve is modeled as an exponential growing functionGeneral Office of Veterinary, Ministry of Agriculture,
USSR, Moscow during the period from 10 February to 2 July 1977 for Odessa.
(Korennoy et al., 2017) [54] 1.65
(CI: 1.42–1.88)
BetweenUkraineExponential Growth MethodThe epidemic curve is modeled as an exponential growing functionGeneral Office of Veterinary, Ministry of Agriculture,
USSR, Moscow during the period from 10 February to 2 July 1977 for Odessa.
Genotype II
(Huang et al., 2021) [55]0.6Not specificChinaNext-Generation MethodDeterministic compartmental modelMinistry of Rural Agriculture of China: The cumulative number of reported infected pigs by ASF is the cumulative number of African Swine Fever cases that comes from the Official Veterinary Bulletin
(Gulenkin et al., 2011) [56]2–3BetweenRussiaProduct of the transmission rate and the duration of infection.Deterministic compartmental modelData for this study was acquired from WAHID (OIE) [World Animal Health Information System Office International Epizootic] database. R0 between herds was calculated with data from the 2008 Republic of North Ossetia.
(Mai et al. 2022) [57]2.48 (CI: 2.39–2.58) and 4.5 (CI: 4.35–4.65), 5.4 (CI: 5.1–5.55), 1.58 (CI: 1.56–1.61), 1.41 (CI: 1.38–1.44), and 1.94 (CI: 1.92–1.97) at national and Hung Yen, Thai Binh, Thai Nguyen, Quang Ninh, Hai Duong ProvincesBetweenVietnamProduct of the transmission rate and the duration of infection.Stochastic compartmental modelData from early outbreak February to August 2019. Local data of ASF outbreaks in Thai Binh, Hung Yen, Thai Nguyen, Quang Ninh, and Hai Duong Provinces were reported to their respective Animal Husbandry and Veterinary sub-departments under the Department of Animal Health, Vietnam. National data for this study were collected directly from the Department of Animal Health
(Song et al., 2022) [58]4.1865WithinChinaNext-Generation MethodDeterministic compartmental modelAugust 2018 ASF outbreak at an Aiyuan farm of Jiangsu Jiahua Breeding Pig Company in Siyang County, China. There are 14,929 pigs in the 13 pigpens
(Li et al., 2022) [59]4.82–11.90WithinChinaNext-Generation MethodCompartmental model that explicitly accounts for both direct and indirect transmission between pigs. Indirect transmission driven by contaminated swills and fomitesData collected by China Animal Health and Epidemiology Center
pertaining to four outbreaks in September 2018 in Xuan Zhou
(Gulenkin et al., 2011) [56]8–11WithinRussiaProduct of the transmission rate and the duration of infectionDeterministic compartmental model (model fitted using curve fitting)Data acquired from WAHID (OIE) [World Animal Health Information System Office International Epizootic] database. Reported ASF outbreaks in the territory of the Russian Federation during the period 2007–2010.
(Gulenkin et al., 2011) [56]9.8 (CI: 3.9–15.6)WithinRussiaProduct of the transmission rate and the duration of infectionDeterministic compartmental model (model calibrated using generalized linear modeling)Data acquired from WAHID (OIE) database. Reported ASF outbreaks in the territory of the Russian Federation during the period 2007–2010.
(Guinat et al., 2018) [60]4.4
(CrI: 2.0–13.4)
17.3
(CrI: 3.5–45.5)
WithinRussian FederationProduct of the transmission rate and the duration of infection.Stochastic compartmental modelIncidence and testing data for ASFV in the RF are routinely collected by the Federal Research Center for Virology and Microbiology (FRCVM). Data on pig mortality were obtained for nine pig herds in which ASFV was detected through routine surveillance between 2010 and 2014
(Le et al., 2023) [61]10.4 (CrI: 1.1–30.4)WithinVietnamProduction of the transmission rate and the duration of infectionStochastic compartmental modelSurveillance outbreak data from a small, 17 pigs, family-owned farm in Thái Bình province.
(Mai et al., 2022) [62]1.49 (CI: 1.05–2.2)WithinVietnamExponential growth MethodDeterministic compartmental modelData collected from ten private farms categorized in different scales (seven farms from 100 to 299 and the rest from 300 to 999) were purposely included after these farms were confirmed with ASF-infected status by real-time PCR technique. Farms were selected based on the availability and quality of the data collected on morbidity. All provinces were in Northern Vietnam.
(Mai et al., 2022) [62]1.58
(CI: 0.92–2.56)
WithinVietnamWhite and Pagano MethodDeterministic compartmental modelData collected from ten private farms categorized in different scales (seven farms from 100 to 299 and the rest from 300 to 999) were purposely included after these farms were confirmed with ASF-infected status by real-time PCR technique. Farms were selected based on the availability and quality of the data collected on morbidity. All provinces were in Northern Vietnam.
(Mai et al., 2022) [62]1.46
(CI: 1.38–1.57)
WithinVietnamAttack Rate MethodDeterministic compartmental modelData collected from ten private farms categorized in different scales (seven farms from 100 to 299 and the rest from 300 to 999) were purposely included after these farms were confirmed with ASF-infected status by real-time PCR technique. Farms were selected based on the availability and quality of the data collected on morbidity. All provinces were in Northern Vietnam.
(Mai et al., 2022) [63]1.55–1.78WithinVietnamProduct of the sum of the infection incidence and the weighted infectivity function.A serial interval model based on daily incidence assuming a Poisson distribution.Two commercial farrow-to-finish pig farms (HY1 and HY2) located in two different districts in Hung Yen province, Vietnam were selected immediately after ASF outbreaks were confirmed in the two farms.
(Guinat et al., 2016) [64]Model 1: 4.0 (CI: 1.2–8.5) Model 2: 5.3 (CI: 1.7–10.3) Model 3: 7.2 (CI: 2.1–14.2)
Between Pen:
Model 1: 2.0 (CI: 1.2–8.5)
Model 2: 2.5 (CI: 0.8–5.2)
Model 3: 3.5 (CI: 1.2–7.0)
WithinGenericProduct of the transmission rate and the duration of infectionDeterministic compartmental modelAnimal experiments conducted by the authors. (United Kingdom)
(Oh et al., 2023) [65]2.91
(CI: 1.51–7.06)
WithinGenericExponential growth MethodSurvival AnalysisAnimal experiments conducted by the authors (Vietnam)
(Oh et al., 2023) [65]4.015 (CI:1.13–9.8)WithinGenericWhite and Pagano MethodSurvival AnalysisAnimal experiments conducted by the authors (Vietnam)
Genotype Not Specified
(Chuchard et al., 2022) [66]18.57BetweenGenericNext-Generation MethodDeterministic compartmental modelNone
(Kouidere et al., 2021) [67]5.71WithinGenericNext-Generation MethodDeterministic compartmental modelNone
(Shi et al., 2023) [68]3.88–9.92WithinGenericNext-Generation MethodFractional-order Compartmental modelNone
(Shi et al., 2023) [69]13.02WithinGenericNext-Generation MethodFractional-order Compartmental modelNone
(Shi et al., 2020) [70]3.77WithinGenericNext-Generation MethodFractional-order Compartmental modelNone
IQR: Interquartile range; CI: 95% Confidence Interval; CrI: 95% Credible Interval.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pittman Ratterree, D.C.; Chitlapilly Dass, S.; Ndeffo-Mbah, M.L. The Reproduction Number of Swine Viral Respiratory Diseases: A Systematic Review. Vet. Sci. 2024, 11, 300. https://doi.org/10.3390/vetsci11070300

AMA Style

Pittman Ratterree DC, Chitlapilly Dass S, Ndeffo-Mbah ML. The Reproduction Number of Swine Viral Respiratory Diseases: A Systematic Review. Veterinary Sciences. 2024; 11(7):300. https://doi.org/10.3390/vetsci11070300

Chicago/Turabian Style

Pittman Ratterree, Dana C., Sapna Chitlapilly Dass, and Martial L. Ndeffo-Mbah. 2024. "The Reproduction Number of Swine Viral Respiratory Diseases: A Systematic Review" Veterinary Sciences 11, no. 7: 300. https://doi.org/10.3390/vetsci11070300

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

Article Metrics

Back to TopTop