Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 04 May 2020

Non-vector-borne transmission of lumpy skin disease virus

  • Kononov Aleksandr 1 ,
  • Byadovskaya Olga 1 ,
  • Wallace B. David 2 , 3 ,
  • Prutnikov Pavel 1 ,
  • Pestova Yana 1 ,
  • Kononova Svetlana 1 ,
  • Nesterov Alexander 1 ,
  • Rusaleev Vladimir 1 ,
  • Lozovoy Dmitriy 1 &
  • Sprygin Alexander 1  

Scientific Reports volume  10 , Article number:  7436 ( 2020 ) Cite this article

7652 Accesses

43 Citations

5 Altmetric

Metrics details

  • Animal disease models
  • Biological models
  • Biological techniques

The transmission of “lumpy skin disease virus” (LSDV) has prompted intensive research efforts due to the rapid spread and high impact of the disease in recent years, especially in Eastern Europe and Balkan countries. In this study, we experimentally evaluate the vaccine-derived virulent recombinant LSDV strain (Saratov/2017) and provide solid evidence on the capacity of the virus for transmission in a vector-proof environment. In the 60-day long experiment, we used inoculated bulls (IN group) and two groups of in-contact animals (C1 and C2), with the former (C1) being in contact with the inoculated animals at the onset of the trial and the latter (C2) being introduced at day 33 of the experiment. The infection in both groups of contact animals was confirmed clinically, serologically and virologically, and viremia was demonstrated in blood, nasal and ocular excretions, using molecular tools. Further studies into LSDV biology are a priority to gain insights into whether the hypothesized indirect contact mode evidenced in this study is a de novo -created feature, absent from both parental stains of the novel (recombinant) LSDV isolate used, or whether it was dormant, but then unlocked by the process of genetic recombination. Author summary: In global terms, LSD has been termed a “neglected disease” due to its historic natural occurrence of being restricted to Africa and, occasionally, Israel. However, after its slow spread throughout the Middle East, the disease is now experiencing a resurgence of research interest following a recent and rapid spread into more northern latitudes. Given the dearth of solid findings on potential transmission mechanisms, no efficient or reliable control program currently exists, which does not involve the use of live attenuated vaccines or stamping out policies – both of which are controversial for implementation in non-endemic regions or countries. The vector-borne mode is the only working concept currently available, but with scarce evidence to support the aggressive spread northwards – except for human-assisted spread, including legal or illegal animal transportation. The emergence of outbreaks is not consistently linked to weather conditions, with the potential for new outbreaks to occur and spread rapidly. Here, for the first time, we provide evidence for indirect contact-mode transmission for a naturally-occurring recombinant LSDV isolated from the field. In an insect-proof facility, we obtained solid evidence that the novel LSDV strain can pass to in-contact animals. Given the recombinant nature of the virus utilised, its genetic background relating to the observed transmission pattern within the study needs to be delineated.

Similar content being viewed by others

essay on lumpy virus in english

Experimental evidence of mechanical lumpy skin disease virus transmission by Stomoxys calcitrans biting flies and Haematopota spp. horseflies

essay on lumpy virus in english

Genomic analysis of lumpy skin disease virus asian variants and evaluation of its cellular tropism

essay on lumpy virus in english

In-vitro and in-vivo study of the interference between Rift Valley fever virus (clone 13) and Sheeppox/Lumpy Skin disease viruses

Introduction.

Lumpy skin disease (LSD) virus belongs to the genus Capripoxvirus within the family Poxviridae (Buller et al ., 2005). It contains a double-stranded, covalently-linked linear DNA genome enveloped by a lipid bilayer. Mature capripox virions have an oval profile and large lateral bodies, and they are 320 ×260 nm on average in size 1 . The viral genome is approximately 151 kb in size, encoding 156 putative genes 2 . LSD virus (LSDV) genes share a high degree of co-linearity and amino acid identity (average of 65%) with genes of other known mammalian poxviruses, particularly suipoxvirus, yatapoxvirus, and leporipoxvirus. Within the genus Capripoxvirus , which also includes sheep pox and goat pox viruses, genetic identities are at least 96% 3 . Generally, based on full-genome sequences deposited in GenBank, LSDV exists as two genetic lineages: field isolates and live attenuated vaccine strains, although evidence for a naturally occurring recombinant between a vaccine strain and field strain has also been reported 4 .

Displaying a strict and narrow host range, LSDV mainly infects cattle and water buffaloes 5 , with one documented isolation from sheep 6 , 7 , although a limited number of other ruminant species, such as antelope, also appear to be susceptible. In recent decades, the disease has been documented in both Southern and Northern hemispheres due to a rapid expansion of its range, although historically, it was restricted to the African continent 8 , 9 , 10 . The unprecedented incursions into new territories and resurgence of LSD creates an objective need for joint efforts to investigate the obscure nature of LSDV transmission 11 .

Initial evidence and studies suggested that direct or indirect contact (without vectors) are ineffective routes for LSDV transmission 12 , whereas the predominant transmission pathway for sheep pox and goat pox viruses is via aerosols. However, importantly, other poxviruses are most commonly spread by direct contact rather than by transmission by arthropod vectors such as by contaminated insects during biting of their hosts. Moreover, poxvirus infection can also occur through ingestion, parenteral inoculation, or droplet or aerosol exposure to mucous membranes or broken skin. Some poxviruses can be transmitted by fomites (inanimate objects). Transmission pathways may vary among poxviruses even within a genus 13 . For orthopoxviruses that infect humans, infection occurs via exposure to aerosols or droplets (variola virus) or through close personal contact. Poxviruses from the Parapoxvirus genus (e.g. orf virus or milker’s nodule virus) can pass from one animal to another through direct or indirect contact. Unfortunately, there are no available studies addressing the issue whether closely related capripoxviruses actually employ different routes of spread.

Transmission of LSDV is surmised to occur through mechanical vector-borne spread via insect or tick bites as most outbreaks occur during the warmer (and, often, wetter) summer months when potential vector species numbers are high 11 , 14 , 15 . Unfortunately, transmission studies conducted to date using arthropod species which may be involved in transmission are mostly species which are restricted to the Southern hemisphere (e.g. Rhipicephalus appendiculatus ticks), which complicates analysis of transmission potential of local species in climatically new geographic areas in the Northern hemisphere. Moreover, LSD outbreaks are not only reported within the warmer summer period, optimal for arthropod blood meal search activity, but also outside of it.This observation thus points to a possible non-vector-borne route for spread of the virus (WAHIS, 2019). Studies using a small cohort of animals were conducted in the past to show that direct contact transmission between infected and naïve animals is possible, but at an extremely low efficiency rate 12 , 16 . New work to replicate the findings has not yet been encouraged. Nevertheless, field evidence suggests that successful transmission can be achieved when naïve animals are allowed to share a drinking trough with severely infected animals 17 , 18 .

This supports the common hypothesis that direct contact does not appear to be an effective route for LSDV transmission. In addition, recent experiments with various field strains did not result in successful contact transmission 19 , 20 , 21 . Attenuated vaccine strains have also been claimed to be devoid of transmission capacity, but recent field evidence argues to the contrary 22 , 23 .

In this environment of uncertainty for LSD spread, elucidation of the exact transmission mode/s could contribute significantly towards improving control and eradication programs.

In this paper, we report on work performed to evaluate transmission of the naturally occurring vaccine-derived virulent recombinant strain of LSDV Saratov/2017 4 , in an experimental setting and for the first time conclusively demonstrate non-vector-borne transmission of the virus.

Materials and methods

The vaccine-derived virulent recombinant LSDV strain, LSDV Saratov/2017, was isolated by FGBI ARRIAH researchers from a cow presenting with severe clinical signs of LSD 22 . The strain was obtained from the FGBI ARRIAH depository and refreshed using two rounds of passaging in goat testis cells. To prepare the final inoculum virus, the refreshed virus was subjected to polymerase chain reaction (PCR) amplification of different loci of vaccine and field strain genomes to confirm the identity of this virus strain 4 .

Ethics statement

The animal experiment, as well as the euthanasia procedure, were approved by the Ethics Committee of the Federal Center for Animal Health, Russia (Permit Number: №2/1-21082018) and conducted in strict accordance with Directive 2010/63/EU on the protection of animals used for scientific purposes.

Experimental design

The initial experimental group consisted of 10 bulls of the Russian Black Pied breed aged 6-8 months. The animals were consecutively numbered from 1 to 10 in a random fashion and maintained in Animal Biosafety Level 3 housing with a 12-hourly light-dark cycle, relative humidity of 30% to 70%, temperature of 23 to 26 °C and all animals were monitored twice daily by the veterinary staff. Water and feed were provided ad libitum . The experiment was carried out in an insect-proof facility. To detect any possible dipteran presence, indoor blood-feeding insect UV light traps and sticky traps were mounted at regular intervals on the walls of the facility. The animals were also examined for the presence of ticks.

The five animals with even numbering (2, 4, 6, 8, 10) each received 2 ml of 5 log TCID 50 /ml of the recombinant virus, LSDV Saratov/2017, intravenously (called the infected/inoculated group – IN) and the remaining five animals with odd numbers were mock inoculated (called C1 - in-contact group 1, also acting as negative controls) with phosphate-buffered saline (PBS) (Table  1 ). The animals were placed in a row along a shared trough according to their consecutive numbering. Their mobility was restricted using tethering, although contact between adjacent animals was possible. At post-inoculation (p.i.) day 33, when there were clear signs of infection in the C1 animals (e.g. crusts, shedding), another group of five bulls (C2 group) was introduced, and positioned between the clinically ill animals, including infected and in-contact animals (Table  2 ).

The animals were monitored daily for the presence of fever until day 52 and clinical signs of LSD until the end of the experiment. To evaluate the time course of virus shedding and viremia, nasal and ocular discharges, skin scabs, blood and semen were collected and tested for virus presence using PCR until day 50.

To assess the clinical picture and collect samples, all experimental animals were first stunned using a penetrative captive bolt to induce unconsciousness, followed by the administration of “Adilinum super” (Federal Center for Toxicological, Radiation and Biological Safety, Kazan, Russia) at day 61 p.i. at a recommended dose of 5 mg/kg according to the drug use instruction approved by the Russian Federal Service for Veterinary and Phytosanitary Surveillance in 2008. At the recommended dose, the Adilinum mechanism of action provides painless and rapid euthanasia: cerebral death commences first followed by circulatory collapse.

Virus isolation in cell culture

Only skin lesions were used for virus isolation and culture. They were ground finely and added to sterile saline at a mass/volume (m/v) ratio of 1:10, followed by repeated freeze–thaw cycles at −80 °C and room temperature (3 times) to disrupt cell membranes. The processed samples were clarified at 1,500 rpm for 15 minutes (min). The supernatant was removed, mixed with antibiotics (penicillin and streptomycin at final concentrations of 2000 IU/mL and 2 mg/mL, respectively) at room temperature for 90 min and then inoculated onto ovine testis or goat gonad cells (70–80% confluency for both), as follows: after removal of the growth medium from flasks containing the cell cultures and washing twice with Hanks’ medium, a 0.1 mL volume of inoculum was added. The flasks were incubated at 37 °C for 90 min to ensure virus adhesion. Once the incubation period was completed, 10 mL of maintenance medium supplemented with 1.0 mL fetal bovine serum was added. The flasks were again incubated at 37 °C. The inoculated cell cultures were observed daily for cytopathic effect (CPE). The cells were then harvested when 80% CPE was observed and lysed to release the virus, using repeated freeze- thaw cycles (3×). The presence of LSDV was confirmed using real-time PCR.

Virus neutralization

Virus neutralization in flat-bottomed microplates (96 wells) was conducted using the protocol described previously 24 , with a few modifications. The test was performed on ovine testis cells using 2 replicates. The volume of inocumlum virus was 100 µl into each well. The neutralization index was considered negative if less than or equal to1:8.

ELISA testing

A double-antigen ELISA for the detection of antibodies against capripoxviruses, including lumpy skin disease virus (LSDV), sheeppox virus (SPPV) and goatpox virus (GTPV) in serum or plasma from cattle, sheep, goats or other susceptible species (IDvet, France), was used to confirm exposure of the bulls to LSDV infection. Serum samples were collected from infected (IN) and in-contact animals (C1 and C2) (all animals per group) showing clinical signs at 0, 42 and 60 days p.i. The ELISA was performed according to the manufacturer’s recommendations. Reactions were measured as optical density (OD) readings at 450 nm using a Tecan Sunrise absorbance microplate reader (Switzerland). The antibody responses were represented as sample-to-positive (S/P) ratios (percentages), calculated as follows: S/P% ratio = (sample OD– negative control OD)/(positive control OD– negative control OD) × 100%. S/P ratios ≥ 30% were considered as being positive.

DNA extraction

The samples were aseptically handled and processed as 10% homogenates in PBS. A 200 μL aliquot was used for total nucleic acid extraction using the QIAamp DNA Mini Kit (Qiagen, Germany), following the manufacturer’s recommendations.

Real-time PCR (quantitative [q] PCR)

Sample extracts were analyzed using real-time PCR for the presence of LSDV DNA, as previously described by Sprygin et al . 25 . The fluorogenic probe was labeled at the 5′ end with the FAM reporter dye and with BHQ as a quencher at the 3′ end. Selected primers and probes were synthesized by Syntol (Moscow, Russia). PCR was performed using a Rotor-Gene Q (Qiagen, Germany) instrument and the following thermal-cycling profile: 95 °C for 10 min, followed by 45 cycles at 95 °C for 15 seconds (s) and 60 °C for 60 s. The final reaction volume was 25 μL containing 10 pmol of each primer, as well as 5 pmol of the probe, 25 mM MgCl 2 , 5 μL 5× PCR Buffer (Promega, USA), 1 μL of 10 pmol dNTPs (Invitrogen, USA), and deionized water to make up the final volume. Samples were tested and results interpreted according to the protocol, as previously described 25 .

Clinical observations

Body temperature.

The baseline average temperature for the animals before the trial was 38.4 °C. All animals infected with LSDV at the outset of the trial (group IN) had fevers up to 40.7 °C (range 39.0 °C to 40.7 °C) on p.i. day 3, indicating virus replication and circulation. By day 8 p.i., the body temperatures of all the inoculated bulls rose above 39.6 °C and their temperatures remained high until p.i. day 20, and thereafter moderate declines were measured to within the range of 38.1 to 39.6 °C in most of the infected animals up to around day 40 p.i. – thereafter a second bout of fever (bi-phasic) was measured in most of the group (except for NI-8) with temperatures rising to a high of 41.5 °C (IN-2) on day 41 p.i., tailoring off to close to normal levels from day 48 p.i. (Fig.  1 and S 1 Table ).

figure 1

Average body temperature measurements per group over the course of the trial (orange – IN bulls, blue – C1 bulls, grey – C2 bulls).

On the other hand, the control (C1) in-contact animals that did not receive virus at initial inoculation, maintained normal body temperatures of around 38.6 °C until day 19 p.i., and thereafter their temperatures also showed increases up to 41.1 °C (C1-9, day 46 p.i.), which decreased slightly on day 46 p.i., but then rising again (bi-phasic fever) in most animals (except C1-1) up until the end of the trial, i.e., day 61.

In C2 animals, which were introduced from day 33 p.i., their body temperatures increased from as early as five days post-introduction (day 38 of the trial), also showing a bi-phasic pattern of fever, with the second phase starting around day 49 p.i. (16 days post-introduction), with a high of 40.9 °C being recorded for animal C2-3.

There was a significant level of edema in the inoculated bulls (IN group) in their subscapular axillary and popliteal lymph nodes, except in bull IN-6 – this animal exhibited edema in its dewlap area and joints, accompanied by lesions on its muzzle, and general weakness (Figs. 2 and 3 ). In the C1 group, edema appeared late, between days 28–30 p.i., in the areas under the jaw, corresponding with the detectable viremia and elevated body temperatures in all the bulls in this group (up to 40.6 °C, in C1-5) (Figs. 5  and 6 ). In the C2 group, elevated temperatures were recorded in all animals within a week following their inclusion in the trial (day 33 p.i.), found to be associated with slightly enlarged lymph nodes and edema in the jaw (for C2-3), and the C2-1 bull displayed a mild edema in its dewlap.

figure 2

Lesions in the muzzle (red arrows) of IN-6 at day 21 p.i.

figure 3

Edematous joints in IN-6 bull at day 21 p.i.

figure 4

In-contact bull C1-9 with swollen submandibular space (red arrow) at day 28 p.i.

figure 5

C1-5 bull with hyperaemic muzzle epithelium at day 28 p.i.

figure 6

Viremia dynamics (averaged per group) (blue – IN, pink – C1, orange – C2).

Viremia, as assessed using qPCR on blood samples, was evident as early as p.i. day 7 in the infected animals (IN-6), peaking by p.i. days 11-13 (only detectable on day 13 in IN-8), followed by a slow decline after p.i. day 20, until p.i. day 38 in all animals in this group (with no detection in IN-8, indicating a very low level of viremia in this animal) (Fig. 6 , S1 Table ).

In the C1 control (in-contact) animals (except in C1-1), which were not inoculated, viremia was measurable using qPCR only after p.i. day 27, to a varying extent, and reached a plateau between p.i. day 38 and 41 (4 out of 5 bulls), which corresponds to days 25–28 following the peak viremia in infected (IN) bulls. Viremia in IN and C1 animals became undetectable after p.i. day 43 (except in C1-3, day 47).

For the C2 group, except for a low-level detectable viremia five days post-introduction in C2-3, no viremia was detectable in blood for the remaining animals, or at any other time point during the trial for this group, even though a low-level bi-phasic fever was measured. However, low levels of viral DNA was detectable from nasal and/or ocular secretions from all animals from this group at some point during their inclusion in the trial ( S1 Table ).

Virus shedding

Virus shedding in inoculated bulls was initially observed as early as p.i. day 11 in nasal discharges, and this lasted intermittently, at least until p.i. day 38. Ocular shedding was also observed in nos. 2, 4, 6 and 10, starting from day 13 p.i. ( S1 Table , Fig.  7 ). Nasal shedding was observed in bull nos. 2, 4, 6 and 10 by day 11 p.i. ( S1 Table and Fig.  8 ).

figure 7

Dynamics of ocular shedding (averaged per group) (blue – IN, pink – C1, yellow – C2).

figure 8

Dynamics of nasal shedding (averaged per group) (blue – IN, pink – C1, yellow – C2).

The dynamics of virus shedding in the two groups in contact with each other at the outset of the trial (IN and C1) showed no statistically significant differences (p > 0.05). The nasal shedding was more durable in IN bulls versus C1 animals with no statistical significance in virus loads. The ocular shedding was comparable in time between the two groups and did not differ statistically (Figs.  7 and 8 ).

Skin lesions

In the inoculated bulls, by p.i. day 11, there was a significant appearance of lesions (0.5 to 1 cm diameter) on the shoulders of three of them (bull nos. 4, 6 and 10). In addition, in these bulls in this group, except IN-2, the scrotum and ventral aspects of the abdomen showed typical foci of erosions. Bull IN-2 showed erosions in only the groin and scapula areas, with the temperature remaining steady at an average of 40 °C. By day 15, there was fulminating disease presentation in infected bull nos. 4, 6 and 10 with symptoms of multiple lumps, coalescing together over the entire body (up to 3 cm in diameter), erosions in the foci up to 2 cm in diameter, and lesions in the scapula and sides as large as 1.5 cm in diameter. Erosions were also observed on their muzzles and associated epithelia. Bull no. 8 displayed similar symptoms, but their appearance was delayed. By p.i. day 21, the size of the lesions had enlarged to 2.5 cm in diameter in inoculated bull nos. 4, 6 and 10, and their nasal epithelia were hyperaemic. Also, there were isolated foci of necrotic lesions in the scrota of these animals.

For the C1 bulls, by day 28 p.i., three of them exhibited small erosions in their nasal epithelia and in their muzzles (Fig.  9a ), during which period their viremia was significant, with high body temperatures (40.6 °C in C1-5). Importantly, no erosions were evident towards the rear half of their bodies (scrotum or inner sides of their legs), as opposed to small skin lesions present on their necks. Three out of five C1 animals (C1-5, C1-7, C1-9) developed characteristic signs of viral infection by day 35, with the symptoms including: enlarged lymph nodes (predominantly prescapular, paratracheal and head lymph glands) and multiple lumps over their bodies, from their heads to their tails, 0.5 to −4 cm in diameter (Fig.  9b ). However, there was no pronounced pathology in popliteal or groin lymph glands as compared to the inoculated animals, and also there were no erosions in their scrotums.

figure 9

C1 bulls with lesions. ( a ) In-contact bull (C1-3) at day 35 p.i., indicating a lesion in the muzzle (arrow). ( b ) C1-7 bull with skin lumps (lesions) on the back and ventral side.

Blood, nasal and ocular discharges were positive for LSDV DNA from these animals, as assessed using qPCR. Even though bull C1-1 was asymptomatic, it was positive for the presence of virus using qPCR analysis. The presence of virus in ocular shedding became detectable only from p.i. day 28 (nos. 3, 5, 7 and 9), while virus shedding in nasal excretions was evident mostly after 35 days p.i. (S 1 Table , Figs.  6 – 8 ). Ocular shedding was more pronounced in C1 bulls as compared to the inoculated bulls (IN group), although, in general, the inoculated bulls displayed longer and more intense shedding.

From day 38 of the experiment, i.e., five days after introducing the C2 animals, viral DNA was detectable from them in blood (viremia), ocular and nasal discharges ( S1 Table , Figs.  6 – 8 ). Shedding in C2 bulls was weak and lasted intermittently for 12 days, from day 5 to day 17 post-introduction. There were variable clinical symptoms in the C2 animals: C2-1 displayed a mild edema in its dewlap, whereas C2-5 developed a few foci of erosion in its scrotum (Table 2 ). At post-introduction day 21 (i.e., day 54 of the experiment), the C2-3 bull displayed a few small lumps in the scapular region. By post-introduction day 26, these lumps had increased in number and size, between 2.0 and 2.5 cm in diameter (Fig.  10a,b ).

figure 10

C2-3 bull: ( a ) at day 20 post-introduction - lumps developing on the neck; ( b ) at day 26 post-introduction – lump development progressing towards the tail.

Recovery from infection

Inoculated (IN) animal nos. 4, 6 and 10 had persistent clinical symptoms, with virus shedding in secretions, till p.i. day 37. But, the inoculated bulls nos. 2 and 8 were asymptomatic, without fever, despite delivering positive PCR results for the presence of virus. All the infected (inoculated) bulls stopped shedding the virus thereafter, with erosions displaying complete scarring and their edemas being cleared, but with their lymph nodes remaining enlarged (as detectable to the touch) up to p.i. day 45.

On the other hand, most of the C1 bulls (nos. 3, 5, 7, 9) displayed clinical signs until day 54. Their skin lesions became necrotic and erupted, with persistent fever between 39 and 40 °C. At p.i. day 50, these bulls developed additional skin lesions, hyperemic nasal epithelia, mild edema in their jaws and dewlaps, and enlarged lymph nodes in their necks. They also had enlarged lymph nodes that were painful to the touch. There was still detectable virus shedding, at this time.

ELISA and VNT

ELISA testing was conducted to assess the levels of seropositivity of the animals to the virus, and it was confirmed that at day 0 before the inoculation, all animals were seronegative (Table  3 ). By p.i. day 42, the five IN bulls seroconverted, whereas, only three out of the five in-contact group (C-1) showed a weak seropositive response (Table  3 ). By p.i. day 60, all IN bulls were strongly seropositive, which was also verified using virus neutralization testing (VNT), as were all C1 animals. However, C2 animals were seronegative throughout the experiment, until day 60, even though there was the detectable presence of virus in their nasal and ocular discharges (and, on one day, day 38, in the blood of bull C2-3).

Recombination is key towards molecular evolution of viruses. Many viruses appear to have diverged from a common ancestor through genetic exchanges and reassortments to expand their diversity, likely including capripoxviruses 26 . However, novel features arising in new chimeric progeny as a result of these events may include drastic phenotypic changes such as a shift in host range, pathogenicity and in transmission pathways 27 . The first field evidence of capripoxvirus recombination between a virulent and vaccine strain was published recently, with the new virus clustering outside both distinct groups of vaccine and field strains 4 . Despite the genome delineation, the degree to which the genome rearrangements contributed to new phenotypic characteristics remains to be elucidated. In this study, we follow up on the novel recombinant strain, LSDV Saratov/2017, in susceptible hosts (bulls) and demonstrate clearly for the first time non-vector in-contact transmission of this isolate with a focus on the time course of viremia and virus shedding in different groups of experimental animals.

The experiment was designed to gain an insight into whether the novel vaccine-derived virulent recombinant LSDV strain, with some genes restored to the yet unknown wild-type parental strain, such as the KSGP-like LSDV isolate, could exert new features surmised to be absent in the normal field parental strains (e.g. Dagestan/2015 20 ) – in this case, efficient non-vector-assisted transmission. Previous experiments with field strains never showed convincing proof that LSDV can transmit to susceptible hosts through air-borne or other non-vector contact modes 20 , 21 . By contrast, the findings of the current study showed that the recombinant virus was able to transmit horizontally without arthropod assistance, as neither flying insects nor ticks were detected throughout the trial period in the indoor containment facility. The first group of in-contact animals (C1) possibly became infected via indirect contact through sharing food and water troughs, while actual physical contact with adjacent animals to their left and right was also possible and may have resulted in direct contact transmission via the mucosa resulting from virus shedding. Drinking troughs have already been blamed for indirect transmission of LSDV in an insect-free setting 17 , however the complexity of transmission is far from being fully comprehended 11 , 18 .

Interestingly, the viremia in the IN bulls was observed between 7 and 22 days p.i., up to 38 days p.i., whereas Babiuk et al . (2008) previously showed that DNA in blood was detectable between 6 and 15 days p.i., only. In the same study, oral and nasal shedding was reported between 12 and 18 days p.i., whereas in our study this window was from 11 to 38 days p.i., which is much longer.

Importantly, virus loads in blood (viremia), nasal and ocular shedding did not differ significantly between IN and C1 animals (p > 0.05), but the duration of viremia and nasal shedding was twice as long in IN animals versus C1 animals, except for ocular shedding where the duration was comparable (S 1 Table ). It is likely that the intravenous route, which closely imitates an insect bite, promotes a more aggressive disease pattern as evidenced by experimental studies 20 , 21 . However, the presumed indirect contact (air-borne or alimentary [mucosa]) route as evidenced by C1 bulls in this study deserves further attention in experimental settings to delineate the complex nature of LSDV transmission. The first evidence of viremia in C1 bulls was observed at day 29 p.i. (S 1 Table ), which is eight days after the IN animals started showing the last signs of their detectable viremia (day 21 p.i.). To confirm the infection, in support with the PCR results, C1 animals did seroconvert, with three of them (C1-1, C1-3 and C1-9) showing antibodies to LSDV by day 42 p.i. (Table  3 , ELISA results), with the remaining animals having all seroconverted by day 60 p.i. (Table  3 ).

Concerning the C2 group, these animals also became infected via an in-contact mode as evidenced by clinical signs, PCR and virus isolation. Although this group had a significantly lower virus load (viremia) versus the IN and C1 groups (p < 0.05), the nasal and oral discharge loads in this group did not differ significantly from either of the other groups (p > 0.05) (Figs.  6 – 8 ). As the number of days and duration of detectable virus was also generally lower, it is surmised that there was possible environmental contamination of the swabs with virus from the IN and/or C1 animals.

Of note is that the ELISA and VNT results indicate that the C2 group did not seroconvert, even after 27 days, possibly due to the shorter time available for antibodies to be elicited in this group compared to the other groups (Table  3 ), whereas virus shedding was detected within the time frame of the experiment (S 1 Table ). The presence of the virus in nasal and ocular discharges in these animals can be explained by the profound shedding of the virus from the skin erosions in the IN and C1 groups contributing to a more rapid infection of the C2 group. In this regard, the importance of the crusts and erosions in virus transmission, be it vector-borne or in-contact, may be significant. However, as judged by the virus dynamics in nasal and ocular discharges (Figs.  7 and 8 ), the virus shedding in terms of titers did not differ statistically (p > 0.05), but, when considering the generalized disease presentation in IN bulls with fulminating skin lesions, containing high concentrations of virus, it is proposed that the skin lesions may have contributed to the virus transmission to the C1 in-contact animals, although to a lesser degree in C2 animals. This may suggest that the amount of virus shed by infected animals is key to virus spread. However, the limitation of the reported evidence is that the mechanism of virus entry cannot be conclusively established on account of the observations made during the experiment - this could be either through the alimentary and/or digestive tracts.

In addition, the localization and dissemination of skin lesions merits further discussion. From our experience, inoculated animals usually develop generalized disease with lesions breaking out evenly over the entire body, including scrotum, muzzle and epithelium, and within the same basic time period. This observation was also true for the IN bulls in this study. The C1 animals, in contrast, displayed a dramatically different pattern, where skin lesions first appeared and were initially restricted to the neck and head regions only, with no lesions in the scrotum or remainder of the body. However, eventually new lesions did develop, spreading throughout the remainder of the animals, down to the tail (Fig.  10a,b ). This is an interesting observation which may provide clues as to the gateway for LSDV infection through contact. In this scenario it is likely that the virus entered C1 animals through their lungs or other mucosa from where it travelled to the closest skin areas – head and neck - then spreading from there to the rest of the animal via the lymphatic system. Since upon intravenous inoculation the virus skips the skin barrier (such as is the case for an insect bite), it infects macrophages and settles in secondary sites, such as lymph nodes and testicles, whereas upon proposed inhalation and/or contact with conjunctiva, it needs to make its way through respiratory barriers. It is also likely that if the C1 and C2 animals had been given more time, their antibody responses would have likely been more pronounced. Moreover, C2 animals displayed no viremia, although they were in contact with virus-shedding cattle - this may also argue for the virus load factor in LSDV transmission, wherein a certain threshold needs to be surpassed.

Another important aspect of the current findings begs the question as to which genomic loci of the recombinant virus likely contributed to enhanced severity and improved capacity for contact transmission. Although identification of the particular genetic targets determining the virus phenotype was outside the scope of the current study, a number of the vaccine-derived genes within the recombinant virus, which reverted to the wild-type genotype following recombination, warrant further study 4 . Since this virus clearly showed contact transmission, it follows that all LSDVs should be capable of spreading by the same mode to varying degrees. Along with the anticipated vector-borne mode of transmission, with no single vector species yet clearly identified as being responsible for transmission, the direct or indirect contact modes have far reaching implications regarding control and eradication. The previously pursued intravenous pathway resulting from vector blood-feeding, has failed to explain outbreaks occurring when cold or dry weather conditions do not permit potential vector flight activity 1 . The likely reality is that a number of modes probably come into play, with the dominant one leading to an outbreak dependent on the specific circumstances and environmental and biological factors present immediately prior to the onset of the outbreak. This aspect warrants urgent attention and research efforts in the context of the current low-level research base on LSD, especially in non-endemic countries, to gain better insights to enable more adequate and targeted control practices to be developed and implemented.

Assimilating the evidence gained from this study, a contact mode of LSDV transmission has been clearly demonstrated for a novel recombinant LSDV recently recovered in the field. Coupled to other alternative modes like vector-assisted spread, these findings may contribute to future risk analysis and complement approaches to combating the threat this disease poses. The findings presented herein will lay a justifying basis for revisiting the control strategies currently in place, considering the danger from the emergence of hybrids, should the use of live vaccines be resorted to, especially in non-endemic countries. Further studies into LSDV biology are warranted to delineate the reasonable question if the contact mode demonstrated here is a de novo -created feature absent from both parental stains of the novel LSDV or whether it maintained a low profile, but was activated by genetic alterations following recombination.

Overall, the current approach to combat LSD and the virus, generally looks at insect vector control, besides vaccination and restriction of cattle movement, while food, water or litter are often overlooked as being ineffective sources of infection. In this scenario, control strategies need to be revisited for development of a more complex approach involving new parameters to consider in risk analyses. Moreover, contact transmission mitigates the factor of seasonality, which is linked to insect activity and widens the possibilities for spread regardless of the presence of biting insects.

Coetzer, J. A. Lumpy skin disease Infectious Diseases of Livestock. [ed.] editor Coetzer J. A. W. and editor. (eds) and Tustin R. C. s.l.: University Press Southern Africa, Oxford., 2004. pp. 1268–1276.

Tulman, E. R. et al . Genome of lumpy skin disease virus. 75 (15), 7122–30, https://doi.org/10.1128/JVI.75.15.7122-7130.2001 (2001).

Article   CAS   Google Scholar  

Tulman, E. R. et al . The genomes of sheeppox and goatpox viruses. J. Virol. 76 , 6054–6061 (2002).

Sprygin A., et al Analysis and insights into recombination signals in lumpy skin disease virus recovered in the field. PLoS One 13 (12), e0207480, https://doi.org/10.1371/journal.pone.0207480 . eCollection (2018).

El‐Nahas, E. M., El‐Habbaa, A. S., El‐Bagoury, G. F. & Radwan, M. E. I. Isolation and identification of lumpy skin disease virus from naturally infected buffaloes at Kaluobia. Egypt. Glob. Vet. 7 , 234–237 (2011).

Google Scholar  

Davies, F. G., Krauss, H., Lund, J. & Taylor, M. The laboratory diagnosis of lumpy skin disease. Res. Vet. Sci. 12 (2), 123–127 (1971).

Omoga, D. C. A. et al . Molecular based detection, validation of a LAMP assay and phylogenetic analysis of Capripoxvirus in Kenya. Journal of Advances in Biology & Biotechnology. 7 (3), 1–12 (2016).

Article   Google Scholar  

Mercier, A. et al . Spread rate of lumpy skin disease in the Balkans, 2015-2016. Transbound. Emerg. Dis. 65 (1), 240–243, https://doi.org/10.1111/tbed.12624 (2018).

Article   CAS   PubMed   Google Scholar  

Sprygin, A. et al . Epidemiological characterization of lumpy skin disease outbreaks in Russia in 2016. Transbound. Emerg. Dis. 65 (6), 1514–1521, https://doi.org/10.1111/tbed.12889 (2018).

Davies, F. G. Lumpy skin disease of cattle: a growing problem in Africa and the Near East. World Anim. Rev. 68 , 37–42 (1991).

Sprygin, A., Pestova, Y., Wallace, D. B., Tuppurainen, E. & Kononov, A. Transmission of lumpy skin disease: a short review. Virus research. 269 , 197637, https://doi.org/10.1016/j.virusres.2019.05.015 (2019).

Weiss, K. E. Lumpy skin disease virus. Virology Monographs. 3 , 111–131 (1968).

Virus Taxonomy: Classification and Nomenclature of Viruses. Eighth Report of the International Committee on Taxonomy of Viruses. Buller, R.M., Arif, B.M., Black, D.N., Dumbell, K.R., Esposito, J.J., Lefkowitz, E.J., … Tripathy D.N. Family Poxviridae. In C. M. Fauquet, M.A. Mayo, J. Maniloff, U. Desselberger, L.A. Ball (Eds). San Diego: Elsevier Academic Press.: s.n.,. pp. 117-133. (2005).

Chihota, C. M., Rennie, L. F., Kitching, R. P. & Mellor, P. S. Attempted mechanical transmission of lumpy skin disease virus by biting insects. Med. Vet. Entomol. 17 , 294–300 (2003).

Tuppurainen, E. S. et al . Mechanical transmission of lumpy skin disease virus by Rhipicephalus appendiculatus male ticks. Epidemiol Infect. 141 (2), 425–430, https://doi.org/10.1017/S0950268812000805 (2013).

Carn, V. M. & Kitching, R. P. An investigation of possible routes of transmission of lumpy skin disease virus (Neethling). Epidemiol. Infect. 114 (1), 219–226 (1995).

Haig, D. A. Lumpy skin disease. Bull. Epizoot. Dis. Afr. 5 , 421–430 (1957).

Tuppurainen ES1, Oura CA. Review: lumpy skin disease: an emerging threat to Europe, the Middle East and Asia. Transbound Emerg Dis. 59 (1), 40–8, https://doi.org/10.1111/j.1865-1682.2011.01242.x. (2012).

Irons, P. C., Tuppurainen, E. S. M. & Venter, E. H. Excretion of lumpy skin disease virus in bull semen. Theriogenology. 63 , 1290–1297 (2005).

Kononov, A. et al . Determination of lumpy skin disease virus in bovine meat and offal products following experimental infection. Transbound Emerg Dis. 66 (3), 1332–1340, https://doi.org/10.1111/tbed.13158 (2019).

Babiuk, S. et al . Quantification of lumpy skin disease virus following experimental infection in cattle. Transbound. Emerg. Dis. 55 , 299–307 (2008).

Kononov, A. et al . Detection of vaccine-like strains of lumpy skin disease virus in outbreaks in Russia in 2017. Archives of Virology. 164 (6), 1575–1585, https://doi.org/10.1007/s00705-019-04229-6 (2019).

Kolcov., A. Molecular-genetic and biological characterization of lumpy skin disease virus isolated in Saratov region of Russia in 2017. Proceedings of 12th Annual Meeting EPIZONE. Aug 27-30, (2018).

Samojlović, M. et al . detection of antibodies against lumpy skin disease virus by virus neutralization test and ELISA methods. Acta Veterinaria-Beograd. 69 (1), 47–60 (2019).

Sprygin, A. et al . A real-time PCR screening assay for the universal detection of lumpy skin disease virus DNA. BMC Res Notes. 12 (1), 371, https://doi.org/10.1186/s13104-019-4412-z (2019).

Gershon, D., Kitching, R. P., Hammond, J. M. & Black, D. N. Poxvirus genetic recombination during natural virus transmission. J. Gen. Virol. 70 , 485–489 (1989).

Ding, N. Z., Xu, D. S., Sun, Y. Y., He, H. B. & He, C. Q. A permanent host shift of rabies virus from Chiroptera to Carnivora associated with recombination. Sci Rep. 7 (1), 289, https://doi.org/10.1038/s41598-017-00395-2 (2017).

Article   ADS   CAS   PubMed   PubMed Central   Google Scholar  

Download references

Acknowledgements

The authors received no specific funding for this work.

Author information

Authors and affiliations.

Federal Center for Animal Health, Vladimir, Russia

Kononov Aleksandr, Byadovskaya Olga, Prutnikov Pavel, Pestova Yana, Kononova Svetlana, Nesterov Alexander, Rusaleev Vladimir, Lozovoy Dmitriy & Sprygin Alexander

Agricultural Research Council-Onderstepoort Veterinary Institute, P/Bag X5, 0110, Onderstepoort, South Africa

Wallace B. David

Department Veterinary Tropical Diseases, Faculty of Veterinary Science, University of Pretoria, P/Bag X4, 0110, Onderstepoort, South Africa

You can also search for this author in PubMed   Google Scholar

Contributions

K.A., B.O., P.P., P.Y., K.S. and N.A. designed the experiments and collected the data. K.A., W.D., L.D., R.V. and S.A. analyzed the data and wrote the manuscript. All authors read and approved the final version of the manuscript.

Corresponding author

Correspondence to Sprygin Alexander .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Additional information

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary information

Supplementary information., rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Aleksandr, K., Olga, B., David, W.B. et al. Non-vector-borne transmission of lumpy skin disease virus. Sci Rep 10 , 7436 (2020). https://doi.org/10.1038/s41598-020-64029-w

Download citation

Received : 30 December 2019

Accepted : 18 March 2020

Published : 04 May 2020

DOI : https://doi.org/10.1038/s41598-020-64029-w

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Molecular characterization of recombinant lsdv isolates from 2022 outbreak in indonesia through phylogenetic networks and whole-genome snp-based analysis.

  • Indrawati Sendow
  • Irene Kasindi Meki
  • Charles Euloge Lamien

BMC Genomics (2024)

First complete genome sequence of lumpy skin disease virus directly from a clinical sample in South India

  • Kalyani Putty
  • Pachineella Lakshmana Rao
  • Madhuri Subbiah

Virus Genes (2023)

Molecular characterization of a novel subgenotype of lumpy skin disease virus strain isolated in Inner Mongolia of China

  • Xiaohui Zan
  • Haibi Huang

BMC Veterinary Research (2022)

An in-depth bioinformatic analysis of the novel recombinant lumpy skin disease virus strains: from unique patterns to established lineage

  • Alena Krotova
  • Olga Byadovskaya
  • Alexander Sprygin

BMC Genomics (2022)

Full-length genome characterization of a novel recombinant vaccine-like lumpy skin disease virus strain detected during the climatic winter in Russia, 2019

  • Antoinette Van Schalkwyk
  • Aleksandr Kononov

Archives of Virology (2020)

By submitting a comment you agree to abide by our Terms and Community Guidelines . If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

essay on lumpy virus in english

  • Open access
  • Published: 19 February 2024

Unravelling the genomic origins of lumpy skin disease virus in recent outbreaks

  • Priya Yadav 1 ,
  • Ankeet Kumar 1 ,
  • Sujith S Nath 1 ,
  • Yashas Devasurmutt 1 ,
  • Geetha Shashidhar 2 ,
  • Madhvi Joshi 7 ,
  • Apurvasinh Puvar 7 ,
  • Sonal Sharma 7 ,
  • Janvi Raval 7 ,
  • Rameshchandra Pandit 7 ,
  • Priyank Chavda 7 ,
  • Sudeep Nagaraj 3 ,
  • Yogisharadhya Revanaiah 3 ,
  • Deepak Patil 4 ,
  • S K Raval 4 ,
  • Jigar Raval 4 ,
  • Amit Kanani 5 ,
  • Falguni Thakar 5 ,
  • Naveen Kumar 6 ,
  • Gundallahalli Bayyappa Manjunatha Reddy 3 ,
  • Chaitanya Joshi 7 ,
  • Baldev Raj Gulati 3 &
  • Utpal Tatu 1  

BMC Genomics volume  25 , Article number:  196 ( 2024 ) Cite this article

1778 Accesses

2 Citations

129 Altmetric

Metrics details

Lumpy skin disease virus (LSDV) belongs to the genus Capripoxvirus and family Poxviridae . LSDV was endemic in most of Africa, the Middle East and Turkey, but since 2015, several outbreaks have been reported in other countries. In this study, we used whole genome sequencing approach to investigate the origin of the outbreak and understand the genomic landscape of the virus. Our study showed that the LSDV strain of 2022 outbreak exhibited many genetic variations compared to the Reference Neethling strain sequence and the previous field strains. A total of 1819 variations were found in 22 genome sequences, which includes 399 extragenic mutations, 153 insertion frameshift mutations, 234 deletion frameshift mutations, 271 Single nucleotide polymorphisms (SNPs) and 762 silent SNPs. Thirty-eight genes have more than 2 variations per gene, and these genes belong to viral-core proteins, viral binding proteins, replication, and RNA polymerase proteins. We highlight the importance of several SNPs in various genes, which may play an essential role in the pathogenesis of LSDV. Phylogenetic analysis performed on all whole genome sequences of LSDV showed two types of variants in India. One group of the variant with fewer mutations was found to lie closer to the LSDV 2019 strain from Ranchi while the other group clustered with previous Russian outbreaks from 2015. Our study highlights the importance of genomic characterization of viral outbreaks to not only monitor the frequency of mutations but also address its role in pathogenesis of LSDV as the outbreak continues.

Peer Review reports

The Lumpy Skin Disease is caused by Lumpy Skin Disease Virus (LSDV) [ 1 ] which is a double-stranded DNA virus of genome size 150 kb and belongs to the genus Capripoxvirus , sub-family Chordopoxviridae and family Poxviridae [ 2 ]. The other members of the genus are Goat poxvirus (GTPV) and Sheep poxvirus (SPPV). LSDV is very similar to all the members of the Poxvirdae family in morphological characteristics, such as being very similar to the vaccinia virus under electron microscopy [ 3 ]. LSDV is non-zoonotic and is known to infect specific hosts, including cattle ( Bos indicus , Bos taurus ) and domestic water buffaloes ( Bubalus bubalis ) [ 4 , 5 , 6 ]. In recent years, reports have emerged that LSDV can also infect camel, giraffe, and wildebeest in the wild [ 7 , 8 , 9 ]. It spreads via contact through skin lesions, milk, and blood-sucking insects such as biting flies, ticks and mosquitoes [ 10 , 11 , 12 , 13 ]. The infection spreads more in the warm and wet periods as compared to the winters due to the increased insect population and its mobility in summers [ 14 ].

LSD virus was first reported in 2019 from India and has since caused several outbreaks. In the recent outbreak, cases started to appear in May 2022. Around 1 lakh cows have died as per recent reports in the current outbreak [ 15 ]. In a developing country like India, livestock production constitutes one of the important ways of earning a livelihood, and a deadly disease such as Lumpy Skin Disease has caused direct loss to the economy and poor production of livestock [ 15 ]. India has a cattle population of 308 million; therefore, controlling the spread of infectious diseases is important [ 16 ].The direct loss includes deaths of cattle, a decrease in milk production while the indirect losses include movement restriction of cattle across the country [ 15 ]. Earlier studies have reported pathological changes in most of the organs and tissues of infected animals such as cow mastitis, necrotic hepatitis, lymphadenitis, orchitis and also in some cases, myocardial damage [ 17 ].

World Organization of Animal Health (WOAH) has now identified lumpy skin disease as a notifiable disease [ 18 ]. Since its first discovery in Zambia in 1931 [ 19 ], this disease was initially confined to the Sub-African region until 1989 and then it started spreading across boundary to Middle East Asia [ 20 , 21 ]. LSDV was reported in 2016 in Russia and other Southeast European nations [ 21 ]. The disease first appeared in India along with other Asian countries such as China, Nepal, Thailand, Bangladesh, Bhutan in November 2019 [ 22 , 23 ]. While LSD has been reported in India since 2019, it has caused significant damage during 2022 outbreak by infecting more than 2 million cows. The symptoms of LSD vary in individual animals depending on the severity of the infection. An animal takes 1–4 weeks to develop symptoms such as high fever, ocular and nasal discharge, loss of appetite, nodular lesion on skin [ 24 ]. According to the data available, a mortality rate of 5–45% was observed [ 4 , 25 , 26 , 27 ]. The major states affected in terms of mortality and morbidity are Rajasthan, Gujarat, Uttar Pradesh, Punjab, Haryana, Karnataka, West Bengal and Maharashtra [ 28 ]. Due to the lack of treatment options, only way of preventing infection is vaccination and by separating the infected animals. The Indian government has taken several measures to control the spread of LSD, including mass vaccination campaigns (goatpox vaccine), setting up quarantine facilities and restricting the movement of infected and susceptible animals. However, the disease continues to be a challenge due to a lack of awareness about its transmission and control, as well as the difficulty in detecting infected animals in the early stages of the disease.

It is generally recognized that the LSDV may have originated from one of the earlier pox virus species and then evolved by spreading in different kind of hosts. Double-stranded DNA viruses are known to use homologous recombination for evolution towards expanding their host range and virulence [ 29 ]. In this study, we have utilized genome sequencing to identify the variants of LSDV circulating in India. Through the phylogenetic analysis we found there are two different classes of variants in India. We then performed mutation (SNP) analysis and found the groups differ significantly in the number of mutations.

Analysis of LSDV cases in the field reveal varied signs and clinical outcomes

Samples of blood, oral swab and skin scab from infected cows were collected from Karnataka, Maharashtra, Rajasthan, and Jhansi during August-November 2022 while the outbreak was going on in India. The animals were showing typical clinical signs such as reduced milk yield, loss of body mass, raised body temperature upto 40–41 °C, nasal and lachrymal discharge, lost appetite, and skin nodules observed on the body of both young and adult cows. The skin nodules burst open when the viral titre is high, exposing the internal layers of the skin which leads to open wounds and eventually secondary bacterial infection if not timely treated (Figure 1 A). The villages observed mortality rate as high as 40% for infected cows, and higher mortality was evidenced in young cows as compared to the adults.

Vectors such as flies, blood-sucking ticks, and mosquitoes were also observed in the vicinity of the animals. Due to the lack of proper ventilation and isolation facilities for cattle, LSDV managed to infect over 20 million cows in 15 different states, and the outbreak is still ongoing.

PCR-based detection strategy for LSDV

The diagnosis for Lumpy skin disease mainly depends on the typical clinical signs and differential diagnosis. The severe form of the disease has characteristic symptoms while the early signs of LSDV overlaps with other diseases like bovine herpes mammillitis, bovine papular stomatitis and foot and mouth disease [ 30 ]. Agar gel precipitation is also used to detect viral antigen in a serum or tissue sample. But this test is not specific and cannot be used for LSDV because the antigens of LSDV are shared with other poxviruses. Therefore, molecular diagnosis by using PCR is the most effective method to detect LSDV.

Samples including blood, oral swab, skin scabs were collected from different parts of India and transported to the ICAR and IISc laboratory to confirm LSDV infection. Conserved regions in the virus genome were identified using multiple sequence alignment and then primers were designed for detecting LSDV. LSDV specific primers were used for fusion gene LSDV0117, which helps the viral envelope fusion with the host membrane as described in method section. Conventional PCR was performed and the amplicon size of 472 bp was confirmed (Fig.  1 B), full image (Supplementary Fig.  1 ). All the samples collected from symptomatic cows were positive for Lumpy Skin Disease Virus.

figure 1

Clinical signs and molecular diagnostics. ( A ) Animal showing severe skin lesions and nodules on body and appearance of swollen lymph nodes ( B ) PCR-based confirmation of LSDV by using partial viral fusion gene amplification

Amplification of Indian LSDV strain genome to develop amplicon-based whole genome sequencing

DNA extracted from 22 samples (including twelve skin scab/nodule samples from cows and ten virus samples collected after cell culture passage) was used for further processing. The whole genome of LSDV is approximately 150 kb, so it was necessary to generate smaller amplicons (3 kb) to achieve full genome coverage for Oxford Nanopore Technology and overlapping primers were designed. Nanopore sequencing produces long reads, which are useful for characterizing complex genomic regions.

A total of 1,126,683 reads were obtained, ranging in size from 105 bp to 54,253 bp, with a median size of 10,000 bp. Reads smaller than 1.5 kb were filtered out to avoid non-specific reads. The coding sequences of 156 genes were obtained using GeneMarkS2 for ab initio gene prediction. The genome was assembled from the raw sequencing data, and the location of the genes was identified. Additional 12 samples collected from Jamnagar, Anand, and Surat were sequenced using Illumina sequencing, covering over 90% of the genome sequence.

Phylogenetic analysis using whole genome sequences of LSDV reveals multiple strains

Multiple sequence alignment of the LSDV whole genome sequences from different regions of the world was performed and then used to construct a maximum likelihood-based phylogenetic tree, which shows the evolutionary relationships among different LSDV strains and other pox viruses. Goat poxvirus was used as outgroup, as shown in Fig.  2 .

figure 2

Phylogenetic analysis of LSDV. Maximum-likelihood Bayesian phylogenetic tree based on whole genome sequences showing the relationship in Capripox virus family including Sheep poxvirus, Goat poxvirus and complete genome sequences of LSDV strains sequenced from many parts of India during 2022 outbreak as well as sequences available on NCBI from previous outbreaks. The genomes sequenced in this study are highlighted in red text

The genome sequences from 2022 outbreak (India) lie on two separate branches on the phylogenetic tree, which has been labelled as low-mutation and high-mutation (Fig.  2 ). Low-mutation genomes closely resemble the Ranchi sequence from 2019 outbreak and Hyderabad sequence from 2020 outbreak. This group has lesser number of variations in their genome sequences as compared to the reference Neethling LSDV sequence. The other group, high-mutation, with higher number of variations per sequence, lies separately, grouping with Russian 2015 LSDV outbreaks. This observation suggests that the current 2022 LSDV outbreak in India is a result of two different group of strains circulating together in the same region.

The Indian LSDV sequences from 2022 are distinct from the vaccine strains of LSDV, confirming that the recent outbreak in India is unlikely to be the result of vaccine spillover. All the Neethling virus-based vaccine strains and vaccine-derived recombinant strains, characterized by combined genetic sequences [ 31 ], form a separate group. The close resemblance of samples from recent outbreak to the 2019 Ranchi strain suggests that infections were occurring, although not very severe, in cows for a few years. However, the higher number of mutations in high mutation group indicates the circulation of another strain with an increased mutation rate in genes involved in host cell binding and immune evasion.

The Indian LSDV sequences are 99.98% identical to the neethling strain and 97.38% identical with Goat poxvirus. LSDV is known to code for 156 proteins, and most of the genes are common between these viruses. The differences in LSDV and goatpox sequences lies in the terminal region of the genome. These terminal regions contain genes important for virulence such as ankyrin repeat-domain containing proteins which helps the virus to bind to the host cell.

SNP analysis in the genome sequence reveals genetic variants with potential to cause severe infection

We performed mutation analysis using nucmer and found a total of 1819 variations in the 22 sequenced genomes compared to the reference sequence LSDV Neethling strain (NC003027.1). These variations at the amino acid level were dominated by silent SNPs (synonymous), followed by extragenic, SNPs (non-synonymous), deletion frameshift, and insertion frameshift mutations (Fig.  3 A). High-mutation group sequences exhibit the highest number of SNPs at the amino acid level (Fig.  3 B). Among the LSDV genes, 38 genes have more than two variations per gene, with the gene encoding Interleukin-10-like protein having the highest number of variations [ 10 ]. This is followed by the virion core protein gene with 7 variations, DNA ligase-like protein, B22R-like protein, and Ankyrin-like protein with 6 and 5 variations, respectively (Fig.  4 ). Other genes, such as Kelch-like proteins, RNA polymerase subunit, EV glycoprotein, DNA helicase transcriptional elongation factor, and early transcription factor large subunit, important for virus binding to host cells and virus replication, have three to four variations in each gene. Additionally, 25 genes have two mutations, including several hypothetical protein-coding genes, and 34 genes show one variation only. Twenty out of 22 genomes sequenced in this study showed non-synonymous changes at amino acid level (Fig.  3 C).

figure 3

The variations in 22 LSDV genome sequences from India were classified into mutation classes ( A ) All classes of variations present in a total of 22 genomes, most of the variations belong to SNP_Silent followed by extragenic mutations ( B ) Total number of variations identified in LSDV strains from 2022 outbreak. The colors in the graph used are for various regions in India ( C ) Single Nucleotide Polymorphisms (SNPs) on amino acid level across all LSDV genomes sequenced

figure 4

( A ) Graph showing number of variations in protein-coding genes in LSDV, with more than 2 variations per gene represented. ( B ) The graph represents the most common mutation on amino acid level in 22 LSDV genome sequences upon analysis

The most mutated protein-coding gene, namely LSDV005, coding for Interleukin-10-like protein is a functional viral cytokine homolog that plays a role in regulating host immune response, as inferred from UniProt. This protein has 4 helix cytokine-like core. Viral interleukins have been shown to activate cellular signalling cascades that enhance viral replication [ 32 ]. Previous gene knockout studies have shown that LSDV005 is one of the important protein-coding genes which is responsible for the virulence of the LSD virus in host cells. LSDV005 is similar to cellular IL-10 in the carboxyl terminus part and affects the immune system similarly [ 33 ].

Another highly mutated gene is LSDV094, which codes for virion core protein. LSDV genome consists of 10 virion core protein-coding genes throughout the genome which make up the core/nucleoprotein of the LSD virus. One common SNP in this gene in High-mutation leads to R578M transition, that is from basic to non-polar amino acid. Other variations inlude R581Q. Virion core proteins are one of the largest families of poxvirus proteins, many of which have been correlated as virulence factors [ 34 ].

Other genes with major variations include LSDV134, which encodes for variola virus (VV) B22R like protein. It is a VV immunomodulating gene with sequence homology to serine protease inhibitors (serpins) that possess antiapoptotic and anti-inflammatory properties. B22R has been shown to reduce the host’s immune response to the virus and rabbitpox equivalent of VV B22R has been shown to inhibit apoptosis in a caspase-independent manner and increase host range as well [ 35 ].

Another important gene which is mutated on protein level is LSDV036 which encode for RNA polymerase subunit. Poxviruses encode a multi-subunit DNA-dependent RNA-polymerase that carries out viral gene expression in host cytoplasm. There were two silent mutations, and 2 SNPs observed in High-mutation genome sequences. The SNP leads to F130S, V297A substitution and given its role in the regulating transcription, any mutations in this gene can lead to altered ability to proliferate and cause symptoms.

Gene LSDV140 encoding for Ring finger host range protein or p28-like protein was also found to be mutated with N203K transition in low-mutation while high-mutation showed SNPs T132M and S152F. P28-like protein is a part of Ub system and acts as Ubiquitin ligase. Similar mutation was reported earlier also in this protein [ 36 ]. These findings highlight the importance of consistent monitoring of genetic variations among LSDV variants. Rest of the SNPs belong to hypothetical proteins, whose functions are yet to be identified in poxviruses. Out of 156 genes of LSDV, more than 40 protein coding genes are hypothetical, and their function remains unknown. 26 of these hypothetical protein coding genes were found to have variations. A total of 399 extragenic mutations are present in 22 genome sequences, and most of the variations were observed in 3’UTR. Since 3’UTRs are involved in post-transcriptional regulation, and mRNA stability and degradation, these variations might affect gene expression and regulation in the host cell. Supplementary File 1 presents the number of mutations found in different proteins in 22 genomes reported in this study.

Many unique mutations were also identified in the genomes sequenced in the current study, such as N1 sample collected from Jamnagar, Gujarat in 2022 has K114I in gene LSDV005 coding for Interleukin-10-like protein; West Bengal sample showed unique SNP Y418F in gene LSDV116 coding for RNA polymerase subunit; Banswara, Rajasthan 2021 genomes sequence showed Y194C in LSDV057 coding for putative virion core protein. Mutations in these proteins can be one of the sophisticated mechanisms for enhancing the spread of virus and these mutations can affect the virulence and host range of the virus. These mutations can arise spontaneously or due to selection pressure from the host immune system and from the use of vaccines.

The most common mutations found in Indian LSDV 2022 sequence include changes in the viral envelope protein, which can alter the ability of virus to evade the host immune response, or in the viral polymerase gene, which can affect the replication and spread of the virus. It is important to monitor the genetic variation in LSDV in India, as it can help to understand the evolution and transmission of the virus and to design effective control strategies, including diagnostic tests.

The frequency in the occurrence of the outbreak and spread of the virus from its original geographic range has led to increased research. The first major outbreak in India happened in 2019, most of the sequences deposited for that outbreak are closely related to Neethling reference strain harbouring a few mutations. But the recent outbreak in India shows presence of at least two types of variants circulating in India.

To gain a comprehensive understanding of the variants of LSDV and their circulation patterns, samples were collected from different parts of India. The phylogenetic analysis of these sequences revealed two major groups. Analysis of the sequence composition at the nucleotide and protein levels demonstrated significant differences in the occurrence of single nucleotide polymorphisms (SNPs) between the two groups. The group that is closer to the neethling strain exhibited less variation, while the other group displayed a higher number of mutations, most of which did not result in amino acid changes. Generally, the poxviruses are not fast evolving viruses, such as Variola virus is estimated to be evolving at ˜ 0.9–1.2 × 10 − 6 substitutions/site/year [ 37 ]. But they can undergo mutations due to mechanisms such as homologous and nonhomologous recombination, gene duplications, gene loss, and the acquisition of new genes through horizontal gene transfer.

Recently, Schalkwyk et al. established the substitution rate of 7.4 × 10 − 6 substitutions/site/year in Lumpy skin disease virus [ 38 ]. Therefore, it is possible that the high severity in the recent outbreaks has increased due to the increased number of mutations, as seen in high-mutation, containing sequences from Indian regions with high-severity cases such as Rajasthan, Gujarat, and Maharashtra, which enables the virus to replicate more and cause more clinical symptoms and resulting in high mortality. Similarity in mutations of high-mutation strains with Nigeria 2021 and Russian 2015 outbreak suggests that there might have been a transboundary migration of an infected animal as both groups show higher number of similar mutations. Some SNPs caused changes in amino acids within proteins, particularly in genes like Interleukin 10-like protein, virion core-like protein, and GPCR-like protein. These mutations could be due to the evolutionary pressure on the virus resulting from the widespread use of vaccines that are not very effective, leading to mutations in important viral immune response genes.

Field analysis of LSD cases revealed a wide range of clinical signs and outcomes in infected cattle, including reduced milk production, weight loss, elevated body temperature, nasal and lachrymal discharge, loss of appetite, and the development of skin nodules. When these nodules burst open, they can lead to open wounds and bacterial infections. The high mortality rate, especially among young cattle, highlights the severity of LSDV infection. The increased severity observed in samples from Gujarat, Maharashtra, and Rajasthan may be a result of the high number of variations in those regions. Analysis of LSDV genomes identified numerous genetic variations, including silent SNPs, extragenic mutations, deletion and insertion frameshift mutations, and SNPs leading to amino acid changes. Several genes with multiple variations were found, including those encoding Interleukin-10-like protein, virion core protein, RNA polymerase subunit, and B22R-like protein. These variations can influence virulence and host range of the virus. Recently, there have been reports of LSDV infection in yaks [ 39 ], camels, free ranging gazelles in India, and giraffes in Vietnam [ 8 , 9 , 40 ], indicating that LSDV can undergo transmission across species with an increased number of mutations. A total of 230 variations were observed in the LSDV strain sequenced from camel hosts. These SNPs were similar to the strains in High-mutation sequenced from cow hosts. Mutations in proteins like B22R, coded by LSDV134, which have been shown to reduce the host’s immune response to the virus and increase host range, were also found in the LSDV strain from camels [ 35 ]. Therefore, it becomes important to identify and vaccinate other hosts which can act as reservoirs for the virus. Monitoring genetic variations in LSDV is crucial for understanding its evolution, developing control strategies, and detecting new outbreaks. The easy spread of LSDV through direct contact, vectors, and bodily fluids emphasizes the importance of proper animal management and biosecurity measures to prevent and control the disease.

Sheep poxvirus and Goat poxvirus, both belonging to the genus Capripoxvirus , have been endemic in India. However, LSDV outbreaks only started recently in Southeast Asian countries in 2019 [ 41 ]. Another poxvirus of the Poxviridae family, Smallpox virus, which caused significant damage to human health and life, has been globally eradicated in the last century through large-scale vaccination campaigns using live vaccinia vaccines [ 42 ]. Therefore, most effective way to prevent and eliminate disease is vaccination. While there are vaccines for LSD based on the Neethling strain. However, animals sometimes develop clinical symptoms such as the formation of nodules and a drop in milk yield even after vaccination. This adverse effect is known as ‘Neethling disease’ [ 43 ]. The commonly used Goat poxvirus vaccine is often employed to prevent LSDV, due to its antigenic similarity to LSDV. The fact that some outbreaks are occurring in vaccinated animals raises about the complete effectiveness of the vaccine. Possible reasons for this include the vaccination method being ineffective, flaws in the administration methods, or issues with vaccine storage.

Some studies report that the commonly used LSDV vaccine is not a pure viral culture of LSDV virus but instead contains a mix of quasi-species, which raises the chances of homologous recombination in the genome of viruses [ 31 ]. It has been reported that LSDV can also undergo recombination [ 29 ]. Therefore, while using live attenuated viruses, it is important to analyze the quality and nature of variation. Previously, several vaccine-like recombinant LSDV strains were discovered in Kazakhstan, neighbouring of Russia and China from 2017 to 2019 [ 44 ]. In our study, we analyzed the current LSDV outbreak clusters and observed that both the groups lie away from the vaccine strains as well as the recombination-derived strains. Therefore, the current outbreak is not a result of vaccine spillover. The 22 LSDV genomes sequenced in this study are separate from the recombinant LSDV strains from Southeast Asia and the Russian vaccine spillover LSDV strains, as clearly depicted in the phylogenetic analysis.

A Ranchi strain-based homogenous live-attenuated LSD vaccine, Lumpi-Pro VacInd (an experimental vaccine) has been developed to prevent the outbreaks. The ‘Neethling disease’ event in which animals develop a reaction against vaccination such as skin nodule formation was not observed in field animals and the vaccine was shown to be 100% effective till January 2023 [ 45 ]. The SNP analysis of this vaccine strain showed that it belongs to the LSDV sequences of Low-mutation in the phylogram which has lesser number of mutations. It has been shown that it neutralizes LSDV strains from 2022 outbreak [ 45 ]. Continued monitoring of vaccinated animals for extended periods will be important to assess the vaccine’s efficacy.

To prevent future outbreaks of LSDV, it is crucial to identify the possible reasons behind these outbreaks and strategize accordingly. It is important to identify all potential hosts for LSDV beyond cattle and arthropod vectors. Regulating the transboundary migrations of animals is necessary to prevent all possible pathways for LSDV introduction. Regarding vaccines, live attenuated LSDV strain vaccines should only be used after undergoing quality testing.

The analysis of LSDV cases in India during the 2022 outbreak highlighted the severe impact of the disease on cows and the significant economic losses incurred in the agricultural sector. Developing a molecular detection strategy using PCR allowed for accurate diagnosis of LSDV, overcoming the challenges posed by clinical signs overlapping with other diseases. Whole genome sequencing provided valuable information on the origin and evolution of the LSDV strains circulating in India. The phylogenetic analysis revealed the presence of two distinct groups, suggesting multiple introductions of the virus. SNP analysis identified genetic variations in essential genes, potentially affecting virulence and antigenicity.

These findings contribute to a better understanding of LSDV and provide valuable information for developing effective control strategies, such as development of diagnostic tests. Monitoring genetic variations and protein expression patterns is crucial for detecting new outbreaks, tracking the virus’s evolution, and guiding the development of targeted interventions. Overall, this research enhances our knowledge of LSDV and its impact on cow’s health, supporting efforts to mitigate its spread and minimize economic losses in the future.

Material and method

Sample collection.

Biological specimens including saliva swab, blood sample, skin lesion scabs were collected from infected cattle as per the standard practices without using any anaesthesia by veterinarians in Karnataka in India. An informed consent was taken from the cows owners before collection of samples. The animals presented symptoms such as fever, nasal discharge and characteristic pox nodular skin lesions. Blood samples were collected in EDTA vacutainers (Becton-Dickenson) and the skin lesions were collected and transported to the laboratory in 50% glycerol (Qualigens). All samples from the previous outbreaks (2019–2021) were available as cell cultures. For the 2022 samples, some were directly processed for DNA extraction, while others were first propagated in either Vero or Lamb testis cell cultures before proceeding with DNA extraction. The samples were used to extract DNA and remaining samples were stored in -80 °C for further use.

Processing of samples

For molecular diagnosis, 200 µl of Blood sample was used to extract DNA by using Nucleic acid extraction kit (HL-NA-100) from Huwel, and the DNA concentration was estimated using a nanophotometer (Implen).

A total of 22 samples of LSDV were used for sequencing (Table  1 ). Seven samples were from the outbreaks during 2020–2021, 1 sample from Ranchi outbreak in 2019, and 14 samples were from 2022 outbreak in different parts of India, as mentioned in the table below. Out of 22 samples, 12 were taken from skin nodules from cattle and 10 were taken from cell-infected viral cultures [ 39 ].

Molecular diagnostics

The primers were designed for the viral Fusion gene to confirm LSDV in the collected samples, forward primer- 5’ ATGGACAGAGCTTTATCA, reverse primer- 5’ TCATAGTGTTGTACTTCG (10pM/µl) with Tm 55 °C. The conventional PCR was carried out in 25 µl reaction with the following conditions used: 94 °C for 5 min for initial denaturation, 30 cycles for 94 °C for 1 min denaturation, 55 °C for 30 s annealing, 72 °C for 45 s extension and a final extension step of 72 °C for 5 min. The PCR result was visualized on 1% agarose (Himedia) gel having Ethidium Bromide (Amresco, Cole-Parmer) and analyzed against 100 bp DNA ladder (G-Biosciences) [ 46 ].

Individual PCR for sequencing

A total of 55 overlapping primer sets for 3 kbp amplicons with Tm 66 °C were designed by using PrimalScheme [ 47 ], a tool which is used to design primers for multiplex PCR, especially for viral outbreak strains. The complete genome sequence of LSDV from previous outbreak with GenBank ID: OK422493.1 was used to design primers. The DNA extracted from the skin scab sample collected from India was used to amplify the LSDV DNA. The 2-step PCR was performed using Phusion polymerase (Thermofisher Scientific) with the following conditions: 98 °C for initial denaturation, 98 °C for denaturation, 66 °C for annealing and extension, and 72 °C for 3 min for the final extension step. The PCR results were analyzed on 1% agarose (Himedia) gel having Ethidium Bromide (Amresco, Cole-Parmer) and analyzed against 1 kb plus DNA ladder (G-Biosciences).

Sequencing using Oxford-nanopore technology

The amplified LSDV DNA amplicons of 3 kbp size were pooled together, and the DNA concentration was determined by using a Qubit 3 fluorometer (Invitrogen) after calibrating with the standards, provided by manufacturer. The samples were processed for genome sequencing on MinION (Oxford Nanopore Technologies, Oxford, United Kingdom) with the ligation sequencing kit SQK-LSK109 Oxford Nanopore Technologies (ONT). Samples were sequenced on flow cell R10.3 version (FLO-MIN111) with 1570 active pores. For the MinION, the Barcode 01–09 from the native barcoding kit (ONT) were used. The DNA was cleaned up using KAPA Hyperpure beads (Roche) after every step during sample preparation, and DNA concentration was recorded. The DNA library was prepared as per the ONT protocol by adopting the following steps: end prep of the LSDV amplicons followed by Barcoding and adapter ligation. The flow cell was primed by using the EXP-FLP002 Flow cell priming kit (ONT). Around 300 fmol of DNA library was loaded along with the sequencing buffer and loading beads on the Spot ON MinION sequencer and was run for 11 h on R10.3 flow cell. Base calling and demultiplexing was performed by using Guppy v5.0 command-line software. The reads were aligned to the LSDV whole genome sequence GenBank ID: OK422493 (Ranchi, India, 2019). The alignment-based consensus was generated using a protocol adapted using Artic pipeline v1.2.1 developed for SARS-COV-2, and the gaps were filled by sequencing missing regions using same protocol.

Sequencing using Illumina NGS technology

NGS library preparation for LSDV whole genome sequencing was done using an in-solution tagmentation-based Illumina Nextera XT DNA Library Prep kit (Illumina) with starting DNA concentration of 1 ng/µl. Prepared libraries were quantified using Qubit 1X ds DNA HS kit (Invitrogen) and the quality check was done on Agilent 2100 bioanalyzer using High sensitivity DNA reagent kit (Agilent). All the libraries were normalized up to 4 nM and pooled. The equimolarized pool of library was processed for the denaturation with 1% PhiX library spike-in as the control library (Illumina). Sequencing was performed on Illumina NovaSeq 6000 SP Reagent Kit v1.5 (300 cycles) paired-end chemistry. Isolated Viral DNA library was sequenced on Illumina MiSeq using MiSeq reagent kit v2 (500 Cycles). The data generated was subject to reference-based assembly using CLC Genomics Workbench v22.0.2 against LSDV reference sequence (NC003027.1). The consensus of each sample were obtained using an inbuilt consensus sequence caller in CLC Genomics Workbench. The variants were visualized using IGV v 2.4.14.

Phylogenetic analysis and SNP analysis

The newly generated consensus sequence was used along with other LSDV complete genome sequences retrieved from the public domain database NCBI virus, accessed on Dec 25, 2023. The sequences used for phylogeny were aligned using Mafft v7.49, and a maximum likelihood phylogenetic tree was constructed using Iqtree2 version 2.0.7 with 1000 bootstrap value. The tree was visualized using interactive Tree Of Life (iTOL) v6.8.1. Nucmer v3.1 was used to generate SNPs in comparison to the Reference sequence of LSDV (NC003027.1). Nucmer does pairwise alignment to identify SNPs in query sequences compared to reference genome sequence. The table with SNPs was further analyzed using R script in R v4.3.1. The script [ 48 ] is available on GitHub (R script link Github).

Data availability

The data that supports the findings of this study is available in the NCBI (BioProject ID PRJNA998208 and PRJNA1004082).

Alexander R, Plowright W, Haig D. Cytopathogenic agents associated with lumpy skin disease of cattle. Bull Epizootic Dis Afr. 1957;5:489–92.

Google Scholar  

Diallo A, Viljoen GJ. Genus Capripoxvirus. In: Mercer AA, Schmidt A, Weber O, editors. Poxviruses. Basel: Birkhäuser Basel; 2007. pp. 167–81.

Chapter   Google Scholar  

Sanz-Bernardo B, Haga IR, Wijesiriwardana N, Hawes PC, Simpson J, Morrison LR, et al. Lumpy skin disease is characterized by severe multifocal Dermatitis with Necrotizing Fibrinoid Vasculitis following experimental infection. Vet Pathol. 2020;57(3):388–96.

Article   PubMed   PubMed Central   Google Scholar  

Davies F. Lumpy skin disease, an African capripox virus disease of cattle. Br Vet J. 1991;147(6):489–503.

Article   MathSciNet   CAS   PubMed   Google Scholar  

Babiuk S, Bowden TR, Boyle DB, Wallace DB, Kitching RP. Capripoxviruses: an emerging worldwide threat to sheep, goats and cattle. Transbound Emerg Dis. 2008;55(7):263–72.

Article   CAS   PubMed   Google Scholar  

Davies FG. Lumpy skin disease of cattle: a growing problem in Africa and the Near East. World Anim Rev. 1991;68(3):37–42.

Young E, Basson PA, Weiss KE. Experimental infection of game animals with lumpy skin disease virus (prototype strain Neethling). Onderstepoort J Vet Res. 1970;37(2):79–87.

CAS   PubMed   Google Scholar  

Dao TD, Tran LH, Nguyen HD, Hoang TT, Nguyen GH, Tran KVD, et al. Characterization of lumpy skin disease virus isolated from a giraffe in Vietnam. Transbound Emerg Dis. 2022;69(5):e3268–e72.

Kumar R, Godara B, Chander Y, Kachhawa JP, Dedar RK, Verma A, et al. Evidence of lumpy skin disease virus infection in camels. Acta Trop. 2023;242(106922):7.

Chihota CM, Rennie LF, Kitching RP, Mellor PS. Mechanical transmission of lumpy skin disease virus by Aedes aegypti (Diptera: Culicidae). Epidemiol Infect. 2001;126(2):317–21.

Article   CAS   PubMed   PubMed Central   Google Scholar  

Carn VM, Kitching RP. An investigation of possible routes of transmission of lumpy skin disease virus (Neethling). Epidemiol Infect. 1995;114(1):219–26.

Chihota CM, Rennie LF, Kitching RP, Mellor PS. Attempted mechanical transmission of lumpy skin disease virus by biting insects. Med Vet Entomol. 2003;17(3):294–300.

Tuppurainen ES, Venter EH, Coetzer JA, Bell-Sakyi L. Lumpy skin disease: attempted propagation in tick cell lines and presence of viral DNA in field ticks collected from naturally-infected cattle. Ticks Tick Borne Dis. 2015;6(2):134–40.

Gubbins S. Using the basic reproduction number to assess the risk of transmission of lumpy skin disease virus by biting insects. Transbound Emerg Dis. 2019;66(5):1873–83.

Kumar N, Tripathi BN. A serious skin virus epidemic sweeping through the Indian subcontinent is a threat to the livelihood of farmers. Virulence. 2022;13(1):1943–4. https://doi.org/10.1080/21505594.2022.2141971 .

Department SR. Cattle population in India 2016–2023 Dec 1, 2022 [Available from: https://www.statista.com/statistics/1181408/india-cattle-population/ .

Ali AA, Neamat-Allah ANF, Sheire HAE, Mohamed RI. Prevalence, intensity, and impacts of non-cutaneous lesions of lumpy skin disease among some infected cattle flocks in Nile Delta governorates, Egypt. Comp Clin Path. 2021;30(4):693–700.

Diseases OO-L. May. Version adopted by the World Assembly of Delegates of the OIE in 2010, OIE,Paris.: Terrestrial Manual of Lumpy Skin Disease; 2010 [Available from: https://www.woah.org/app/uploads/2021/03/a-reso-2016-public.pdf .

MacOwan K. Observations on the epizootiology of lumpy skin disease during the first year of its occurrence in Kenya. Bull Epizootic Dis of Africa. 1959;7:7–20.

Tuppurainen ES, Oura CA. Review: lumpy skin disease: an emerging threat to Europe, the Middle East and Asia. Transbound Emerg Dis. 2012;59(1):40–8.

House JA, Wilson TM, Nakashly SE, Karim IA, Ismail I, Danaf NE, et al. The isolation of lumpy skin disease virus and bovine herpesvirus-from cattle in Egypt. J Vet Diagn Invest. 1990;2(2):111–5.

Lojkić I, Šimić I, Krešić N, Bedeković T. Complete genome sequence of a lumpy skin disease virus strain isolated from the skin of a vaccinated animal. Genome Announcements. 2018;6(22):e00482–18.

Kumar N, Chander Y, Kumar R, Khandelwal N, Riyesh T, Chaudhary K, et al. Isolation and characterization of lumpy skin disease virus from cattle in India. PLoS ONE. 2021;16(1):e0241022.

Salib FA, Osman AH. Incidence of lumpy skin disease among Egyptian cattle in Giza Governorate. Egypt Veterinary World. 2011;4(4).

Babiuk S, Bowden T, Parkyn G, Dalman B, Manning L, Neufeld J, et al. Quantification of lumpy skin disease virus following experimental infection in cattle. Transbound Emerg Dis. 2008;55(7):299–307.

Irons P, Tuppurainen E, Venter E. Excretion of lumpy skin disease virus in bull semen. Theriogenology. 2005;63(5):1290–7.

Mathivanan EME, Raju K, Murugan R. Outbreak of lumpy skin disease in India 2022- an emerging threat to livestock & livelihoods. Global Biosecur. 2023;5.

Sudhakar SB, Mishra N, Kalaiyarasu S, Jhade SK, Hemadri D, Sood R, et al. Lumpy skin disease (LSD) outbreaks in cattle in Odisha state, India in August 2019: epidemiological features and molecular studies. Transbound Emerg Dis. 2020;67(6):2408–22.

Minor PD, John A, Ferguson M, Icenogle JP. Antigenic and molecular evolution of the vaccine strain of type 3 poliovirus during the period of excretion by a primary vaccinee. J Gen Virol. 1986;67(4):693–706.

Datten B, Chaudhary AA, Sharma S, Singh L, Rawat KD, Ashraf MS et al. An extensive examination of the warning signs, symptoms, diagnosis, available therapies, and prognosis for lumpy skin disease. Viruses. 2023;15(3).

Sprygin A, Pestova Y, Bjadovskaya O, Prutnikov P, Zinyakov N, Kononova S et al. Evidence of recombination of vaccine strains of lumpy skin disease virus with field strains, causing disease. PLoS ONE. 2020;15(5).

Chatterjee M, Osborne J, Bestetti G, Chang Y, Moore PS. Viral IL-6-induced cell proliferation and immune evasion of interferon activity. Science. 2002;298(5597):1432–5.

Article   ADS   CAS   PubMed   Google Scholar  

Tulman ER, Afonso CL, Lu Z, Zsak L, Kutish GF, Rock DL. Genome of lumpy skin disease virus. J Virol. 2001;75(15):7122–30.

Van Vliet K, Mohamed MR, Zhang L, Villa NY, Werden SJ, Liu J, et al. Poxvirus proteomics and virus-host protein interactions. Microbiol Mol Biol Rev. 2009;73(4):730–49.

Moon KB, Turner PC, Moyer RW. SPI-1-dependent host range of rabbitpox virus and complex formation with cathepsin G is associated with serpin motifs. J Virol. 1999;73(11):8999–9010.

Chibssa TR, Sombo M, Lichoti JK, Adam TIB, Liu Y, Elraouf YA, et al. Molecular Analysis of East African Lumpy skin Disease Viruses reveals a mixed isolate with features of both vaccine and field isolates. Microorganisms. 2021;9(6):1142.

Babkin I, Shchelkunov S. Time scale of poxvirus evolution. Mol Biol. 2006;40:16–9.

Article   CAS   Google Scholar  

Van Schalkwyk A, Byadovskaya O, Shumilova I, Wallace DB, Sprygin A. Estimating evolutionary changes between highly passaged and original parental lumpy skin disease virus strains. Transbound Emerg Dis. 2022;69(4):e486–e96.

PubMed   Google Scholar  

Manjunatha Reddy GB, Pabbineedi SM, Nagaraj S, Bijalwan S, Tadakod S, Bhutia Z, et al. Lumpy skin disease (LSD) in yak (Bos grunniens): an evidence of species spillover from cattle in India. Microorganisms. 2023;11(12):2823.

Sudhakar SB, Mishra N, Kalaiyarasu S, Ahirwar K, Chatterji S, Parihar O, et al. Lumpy skin disease virus infection in free-ranging Indian gazelles (Gazella bennettii), Rajasthan, India. Emerg Infect Dis. 2023;29(7):1407–10.

Das M, Chowdhury MSR, Akter S, Mondal AK, Uddin M, Rahman M. An updated review on lumpy skin disease: perspective of southeast Asian countries. J Adv Biotechnol Exp Ther. 2021;4(3):322–33.

Article   Google Scholar  

Strassburg MA. The global eradication of smallpox. Am J Infect Control. 1982;10(2):53–9.

Ben-Gera J, Klement E, Khinich E, Stram Y, Shpigel NY. Comparison of the efficacy of neethling lumpy skin disease virus and x10RM65 sheep-pox live attenuated vaccines for the prevention of lumpy skin disease - the results of a randomized controlled field study. Vaccine. 2015;33(38):4837–42.

Vandenbussche F, Mathijs E, Philips W, Saduakassova M, De Leeuw I, Sultanov A et al. Recombinant LSDV strains in Asia: vaccine spillover or natural emergence? Viruses. 2022;14(7).

Kumar N, Barua S, Kumar R, Khandelwal N, Kumar A, Verma A, et al. Evaluation of the safety, immunogenicity and efficacy of a new live-attenuated lumpy skin disease vaccine in India. Virulence. 2023;14(1):2190647.

K LK BS. Clinico-molecular diagnosis and characterization of bovine lumpy skin disease virus in Andhra Pradesh, India. Trop Anim Health Prod. 2021;53(4):424.

Article   PubMed   Google Scholar  

Quick J, Grubaugh ND, Pullan ST, Claro IM, Smith AD, Gangavarapu K, et al. Multiplex PCR method for MinION and Illumina sequencing of Zika and other virus genomes directly from clinical samples. Nat Protoc. 2017;12(6):1261–76.

Kumar A, Kaustubh M, Nath SS, Tatu U. Identification of clade-specific single nucleotide polymorphisms for improved rabies virus surveillance in Canis familiaris host. 2023.

Download references

UT acknowledges funding from IISc-DBT partnership and IOE grant. AK acknowledges financial support from CSIR. PY acknowledges financial support from IISc. GR, BG acknowledges funding from ICAR-NIVEDI, Bangalore. CJ acknowledges funding from Government of Gujarat State and NK acknowledges funding from ICAR-NRC on Equines, Hisar.

Author information

Authors and affiliations.

Department of Biochemistry, Indian Institute of Science, Bangalore, 560012, India

Priya Yadav, Ankeet Kumar, Sujith S Nath, Yashas Devasurmutt & Utpal Tatu

Central Research Laboratory, KIMS, Bangalore, 560070, India

Geetha Shashidhar

ICAR-National Institute of Veterinary Epidemiology and Disease Informatics, Bangalore, 560064, India

Sudeep Nagaraj, Yogisharadhya Revanaiah, Gundallahalli Bayyappa Manjunatha Reddy & Baldev Raj Gulati

Kamdhenu University, Gandhinagar, Gujarat, India

Deepak Patil, S K Raval & Jigar Raval

AH Department, GoG, Gujarat, India

Amit Kanani & Falguni Thakar

National Centre for Veterinary Type Cultures, ICAR-NRC on Equines, Sirsa Road, Hisar, Haryana, 125001, India

Naveen Kumar

Gujarat Biotechnology Research Centre, Gandhinagar, 382011, India

Madhvi Joshi, Apurvasinh Puvar, Sonal Sharma, Janvi Raval, Rameshchandra Pandit, Priyank Chavda & Chaitanya Joshi

You can also search for this author in PubMed   Google Scholar

Contributions

UT, GR, BG and CJ conceived and designed the study. PY, AK performed the ONT-based sequencing, analysed the data, and wrote the manuscript. YD helped in analysing sequencing data. AP, SS, JR, RP, PC prepared samples and performed NGS Illumina sequencing. MJ, DP, FT, NK helped in coordinating with team. SN, SN, YR and RSK diagnosed the cattle. SN, JR, AK, NK collected samples from field. UT, GR, BG, CJ reviewed the manuscript.

Corresponding author

Correspondence to Utpal Tatu .

Ethics declarations

Ethical statement.

Institute Bio-Safety certificate (IBSC) with Reference number IBSC/IISc/UT/16/2023 was obtained from the ethics committee to work on LSDV. The utilization of animal sample was undertaken as per ARRIVE guidelines. The LSDV samples were collected by the veterinarians as per the approved protocols and guidelines.

Consent for publication

Informed consent was obtained from the cow owners for the material and images used in the publication.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary File 1.

SNPs in different proteins of 22 genomes of LSDV reported in the study

Supplementary Fig. 1.

Complete agarose-gel image for PCR confirmation of LSDV in clinical samples (Figure 1B)

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/ . The Creative Commons Public Domain Dedication waiver ( http://creativecommons.org/publicdomain/zero/1.0/ ) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Cite this article.

Yadav, P., Kumar, A., Nath, S.S. et al. Unravelling the genomic origins of lumpy skin disease virus in recent outbreaks. BMC Genomics 25 , 196 (2024). https://doi.org/10.1186/s12864-024-10061-3

Download citation

Received : 17 September 2023

Accepted : 29 January 2024

Published : 19 February 2024

DOI : https://doi.org/10.1186/s12864-024-10061-3

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Lumpy skin disease virus
  • Whole genome sequencing
  • Mutation analysis

BMC Genomics

ISSN: 1471-2164

essay on lumpy virus in english

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

viruses-logo

Article Menu

essay on lumpy virus in english

  • Subscribe SciFeed
  • PubMed/Medline
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Lumpy skin disease: a systematic review of mode of transmission, risk of emergence and risk entry pathway.

essay on lumpy virus in english

1. Introduction

2. materials and methods, 3.1. selection process, 3.2. description of the retrieved articles, 3.3. host of lumpy skin disease virus, 3.4. modes of transmission, 3.4.1. direct transmission, 3.4.2. indirect transmission via vectors, 3.5. emergence of vaccine-like recombinant strains, 3.6. risk factors of lumpy skin disease outbreaks and spread, 3.7. risk analysis of introduction of lumpy skin disease to a free-area, 4. discussion, 5. conclusions, author contributions, institutional review board statement, informed consent statement, data availability statement, conflicts of interest.

Section and TopicItem #Checklist Item Location Where Item Is Reported
TITLE
Title 1Identify the report as a systematic review.1
ABSTRACT
Abstract 2See the PRISMA 2020 for Abstracts checklist.1
INTRODUCTION
Rationale 3Describe the rationale for the review in the context of existing knowledge.1–3
Objectives 4Provide an explicit statement of the objective(s) or question(s) the review addresses.3
METHODS
Eligibility criteria 5Specify the inclusion and exclusion criteria for the review and how studies were grouped for the syntheses.3–4
Information sources 6Specify all databases, registers, websites, organizations, reference lists and other sources searched or consulted to identify studies. Specify the date when each source was last searched or consulted.3–4
Search strategy7Present the full search strategies for all databases, registers and websites, including any filters and limits used.4
Selection process8Specify the methods used to decide whether a study met the inclusion criteria of the review, including how many reviewers screened each record and each report retrieved, whether they worked independently, and if applicable, details of automation tools used in the process.3 ( )
Data collection process 9Specify the methods used to collect data from reports, including how many reviewers collected data from each report, whether they worked independently, any processes for obtaining or confirming data from study investigators, and if applicable, details of automation tools used in the process.3
Data items 10aList and define all outcomes for which data were sought. Specify whether all results that were compatible with each outcome domain in each study were sought (e.g., for all measures, time points, analyses), and if not, the methods used to decide which results to collect.-
10bList and define all other variables for which data were sought (e.g., participant and intervention characteristics, funding sources). Describe any assumptions made about any missing or unclear information.-
Study risk of bias assessment11Specify the methods used to assess risk of bias in the included studies, including details of the tool(s) used, how many reviewers assessed each study and whether they worked independently, and if applicable, details of automation tools used in the process.3
Effect measures 12Specify for each outcome the effect measure(s) (e.g., risk ratio, mean difference) used in the synthesis or presentation of results.Not appropriate
Synthesis methods13aDescribe the processes used to decide which studies were eligible for each synthesis (e.g., tabulating the study intervention characteristics and comparing against the planned groups for each synthesis (item #5)).-
13bDescribe any methods required to prepare the data for presentation or synthesis, such as handling of missing summary statistics, or data conversions.-
13cDescribe any methods used to tabulate or visually display results of individual studies and syntheses.-
13dDescribe any methods used to synthesize results and provide a rationale for the choice(s). If meta-analysis was performed, describe the model(s), method(s) to identify the presence and extent of statistical heterogeneity, and software package(s) used.-
13eDescribe any methods used to explore possible causes of heterogeneity among study results (e.g., subgroup analysis, meta-regression).-
13fDescribe any sensitivity analyses conducted to assess robustness of the synthesized results.Not appropriate
Reporting bias assessment14Describe any methods used to assess risk of bias due to missing results in a synthesis (arising from reporting biases).Not appropriate
Certainty assessment15Describe any methods used to assess certainty (or confidence) in the body of evidence for an outcome.Not appropriate
RESULTS
Study selection 16aDescribe the results of the search and selection process, from the number of records identified in the search to the number of studies included in the review, ideally using a flow diagram.4.5
16bCite studies that might appear to meet the inclusion criteria, but which were excluded, and explain why they were excluded.4
Study characteristics 17Cite each included study and present its characteristics.6
Risk of bias in studies 18Present assessments of risk of bias for each included study.
Results of individual studies 19For all outcomes, present, for each study: (a) summary statistics for each group (where appropriate) and (b) an effect estimate and its precision (e.g., confidence/credible interval), ideally using structured tables or plots.6–20
Results of syntheses20aFor each synthesis, briefly summarize the characteristics and risk of bias among contributing studies.6–20
20bPresent results of all statistical syntheses conducted. If meta-analysis was performed, present for each the summary estimate and its precision (e.g., confidence/credible interval) and measures of statistical heterogeneity. If comparing groups, describe the direction of the effect.6–20
No meta-analysis
20cPresent results of all investigations of possible causes of heterogeneity among study results.6–20
20dPresent results of all sensitivity analyses conducted to assess the robustness of the synthesized results.Not appropriate
Reporting biases21Present assessments of risk of bias due to missing results (arising from reporting biases) for each synthesis assessed.Not appropriate
Certainty of evidence 22Present assessments of certainty (or confidence) in the body of evidence for each outcome assessed.6–20
DISCUSSION
Discussion 23aProvide a general interpretation of the results in the context of other evidence.6–20
23bDiscuss any limitations of the evidence included in the review.6–20
23cDiscuss any limitations of the review processes used.6–20
23dDiscuss implications of the results for practice, policy, and future research.21–26
OTHER INFORMATION
Registration and protocol24aProvide registration information for the review, including register name and registration number, or state that the review was not registered.Not registered
24bIndicate where the review protocol can be accessed, or state that a protocol was not prepared.-
24cDescribe and explain any amendments to information provided at registration or in the protocol.-
Support25Describe sources of financial or non-financial support for the review, and the role of the funders or sponsors in the review.26
Competing interests26Declare any competing interests of review authors.26
Availability of data, code and other materials27Report which of the following are publicly available and where they can be found: template data collection forms; data extracted from included studies; data used for all analyses; analytic code; any other materials used in the review.
Ref.Author/YearType of StudyStudy PurposeMethodologyMain Findings/ConclusionsLimitations of the StudyGeographical Area of Study
[ ]Yeruham et al. (1995) ObD
Vec-Ins
To describe the conditions and dairy herds affected by LSD outbreaks. Description of the area and herds in which LSD outbreaks were reported. Haematology, biochemistry and serology were performed on blood samples collected from affected animals, along with histopathology of skin lesions. Local wild ruminants, i.e., gazelles (Gazella gazella), and sheep and goats were examined in search of LSD clinical signs.It concluded that although the origin of the LSD outbreak in the dairy herds could not be traced with certainty, the circumstantial evidence (no cattle newly introduced in the village herds, thus, other means of introduction were therefore suggested) indicated that the LSDV was brought from Egypt by wind-carried Stomoxys calcitrans. Results are not very detailed. The study only mentions the number of herds affected but not the number of cattle heads.
The study only describes the epidemiology of the first LSD outbreak in Israel. Thus, all inferences on the modes of transmission and spread of the disease in the dairy herds were conducted using circumstantial evidence.
Israel
[ ]Davies
(1982)
ObD
Host
Attempts to define the maintenance of LSD in hosts living in high altitude indigenous forests by searching for antibodies to LSD virus in the sera from wild and domestic ruminants. Blood samples of cattle and wild ruminants were collected from different sources for LSDV isolation and serology through microserum neutralization tests and indirect fluorescent antibody test.The African buffalo (Syncerus caffer) had Abs to capripox virus: out of 254 buffaloes, 150 animals were seropositive to IFAT, along with a small number of domestic cattle. An LSD endemic area was proposed and authors suggested that the maintenance cycle involves the buffalo. No Ab was detected in the other wild ruminant species investigated. It concluded that, while an epidemic of LSD has occurred in Kenya, most cases were sporadic and probably the result of accidental contacts with a component of the maintenance cycle.Serology testing cannot distinguish the three viruses in the Capripoxvirus genus (sheep pox virus, goat pox virus and LSDV). The period of study and geographical environment were described. In the results, authors indicated that from the sera from positive to the IFAT test (150 out of 254) three groups of buffalo sera contained a significant number which neutralized the LSD/2490 strain of virus. There was no neutralization of cowpox virus by any of these positive sera, which increases the likelihood that the neutralization of LSD is due to specific antibody and not due to non-specific neutralizing properties of the sera. This is not a confirmation and the number of buffaloes was not specified.Kenya
[ ]Fagbo et al.
(2014)
ObD
Host
To expand the understanding of the role of buffalo in the maintenance of LSDV and Rift Valley Fever (RVF) by determining their seroprevalence during an inter-epidemic period.Between 2003 and 2004, blood samples were collected from African buffaloes in the Kruger National Park and Hluhluwe-iMfolozi Park, South Africa. They were tested for IgG Abs for LSD with ELISA and positive or suspected positive samples were further tested by SNT The I-ELISA for LSDV and RVFV detected IgG antibodies in 70 out of 248 (28.2%) and 15 out of 248 (6.1%) buffaloes, respectively. Using the SNT, LSDV and RVFV neutralizing Abs were found in 5 out of 66 (7.6%) and 12 out of 57 (21.1%) samples tested, respectively. Authors suggested that African buffaloes play a role in the epidemiology of these diseases during inter-epidemic periods.Limitations with serological tests as it is not possible to distinguish the three viruses in the Capripox virus (sheep and goat pox viruses and LSD). The SNT, only gave 5 positive out of 66 samples, i.e., the gold standard did not compare correctly with results obtained by the I-ELISA used in the study, as the I-ELISA is not validated for wildlife sera.
Authors mention that the African buffalo plays a role in the inter-epidemic period but the specific sampling period of the year (e.g., during the rainy or dry season, during an outbreak in the country) was never specified, thus it was not possible to draw that conclusion.
South Africa
[ ]Ahmed et al.
(2021)
ObD
Host
To identify and characterize the LSD virus outbreaks in Egypt, between 2016 and 2019, and determine the role of Egyptian buffaloes in the epidemiology of LSD.Forty-one and three skin biopsies were performed on clinically-affected cattle and buffaloes, respectively; 31 blood samples were collected from asymptomatic buffaloes in contact with clinically-infected cattle and tested by RT-PCR .
Samples were collected from 102 bovines showing clinical signs of LSD and 96 Egyptian buffaloes, with no vaccination history, and in contact with LSD clinically-affected cattle.
Positive samples were isolated and sequenced; phylogenetic trees were constructed.
Among the skin biopsies that underwent RT-PCR to detect LSDV, 31 cattle heads were positive and all buffaloes were negative. LSDV was isolated on CAM and MDBK cell culture in 19 positive samples.
ELISA results: 84/102 cattle were positive and 17/96 buffalo were positive.
The phylogenetic analysis was identical for all isolates, and presented a 99–100% identity with LSDV isolates from different countries in Africa, Asia, and Europe. ELISA analyses detected sero-reactivity to LSDV in Egyptian cattle and buffaloes. Conclusion: the Egyptian water buffalo is an accidental, non-adapted, host of the virus and the current vaccine strategy for LSD control should be re-evaluated to improve coverage and effectiveness.
Although it is proposed that Egyptian buffaloes are less susceptible to LSDV infection, only 3 samples of skin biopsies were used to confirm the presence of LSDV by RT-PCR. Antibodies were also detected by ELISA, but a low percentage were positive. These differences in results could be explained by a number of factors (e.g., sensitivity, specificity of the ELISA, Ab’s were produced due to another Capripoxvirus, low number of skin biopsies tested) which are not elaborated in the article.Egypt
[ ]Pandey et al.
(2022)
ObD
Host
To highlight the speed at which the disease can spread in animal populations, previously presumed to be naïve, and to quantify its impact with reference to subsistence agriculture in rural communities.Clinical signs were described and recorded after full clinical examination of affected animals (oxen, cows, Bos indicus calves and Asian water buffalo), in small village holdings around a tiger reserve. Questionnaires allowed gathering information on the clinical disease history and animal husbandry practices relevant to the spread of LSDV.The signs of LSD were recorded and described in 154 oxen, 34 cows, 13 calves (Bos indicus) and two Asian water buffaloes (Bubalus bubalis). The description of an LSD outbreak in naïve populations of cattle and buffaloes illustrated the need for increased awareness on the associated clinical signs and the maintenance of high biosecurity levels in hitherto disease-free countries.Diagnosis of LSD only relied on clinical signs, which could lead to false positives or negatives and thus to an over- or underestimation of the prevalence. India
[ ]Faris et al.
(2021)
ObC
RiskF.
To assess the prevalence of LSD in five selected localities in an Egyptian governorate and to detect the potential risk factors associated with LSD.Blood samples were collected from 599 cattle heads and 66 buffaloes, with and without clinical signs of LSD. Temperature humidity index (THI), resulting from the combination of air temperature and humidity, associated with the level of thermal stress was calculated. A multivariate logistic regression assessed the risk factors related to LSD prevalence. The risk factors identified in the study were: animal species (cattle and buffaloes), age, season (winter, spring and summer), THI, locality and immune status of animals (vaccinated vs. unvaccinated).The prevalence was 36.7% in cattle and 15.2% in buffaloes. Regarding the influence of age, the prevalence was 26.3% in animals <1 year, 42.2% in animals aged 1–2 years and 34.9% in the >2 years group. When considering the season, the prevalence reached 29.3% in the winter, 34.1% in the spring and 37.7% in the summer. A prevalence of 29.7%, 31.6% and 37.6% were calculated for a low, moderate and high THI, respectively. The prevalence in vaccinated vs. unvaccinated animals was 34.3% vs. 50%. The authors concluded that LSD had become endemic in Egypt and was responsible for sporadic outbreaks over the year, mainly in adult animals and during the summer; cattle was more likely to be infected than buffalo.The study assessed LSD prevalence in five localities using blood samples but no information on the diagnostic test used to confirm an LSDV infection was provided. Additionally, the authors did not specify if farmed cattle was randomly. Furthermore, no sample size was calculated. The season explanatory variables group did not include autumn and no explanation was given for its exclusion.Egypt
[ ]Aboud et al.
(2022)
ObD
Host
To confirm infection of Iraqi buffaloes with LSDV.Blood samples, clinical examination to detect skin lesions and collection of ticks from 150 buffaloes of different ages and sexes.
Tests used: PCR and histopathology.
Eight out of 150 buffaloes were positive by PCR. The histopathology performed on skin lesions revealed that one out of 13 samples were positive to LSDV. Among 29 ticks (species not specified) collected, none was positive. This is the first study to investigate LSD in buffaloes, to identify positive animals and to describe rare clinical signs. It concluded that an effective control of LSD requires an accurate and rapid laboratory diagnostic method such as PCR; histopathology could be a method to identify and confirm the disease along with clinical examination. Sampling was not random. Animals were selected based on information provided by veterinarians and buffalo owners who observed the clinical signs. Only 13 LSD-suspect skin were sampled and analyzed via histopathology.
The tick sample was very small (N = 29) and authors did not explain why.
Iraq
[ ]Greth et al. (1992)ObD
Host
Sampling of captive-bred Arabian Oryx (Oryx leucoryx) from a national wildlife research center after an animal showed clinical signs of LSD.Serology survey; virus was identified by electron microscopy. Virus neutralization was performed by antibody titer on paired sera. It was the first case of LSD infection described in the Arabian oryx, and also the first case reported in Saudi Arabia. The serologic survey of the herd (90 oryx) showed a low prevalence (2%) of infection and only one out of the two positive animals developed lesions.Sampling was performed in captive animals (i.e., not living freely), so the role of wildlife cannot be ascertained. The presence of LSDV was not confirmed by the tests used. The only certainty was that a Capripox virus was involved.Saudi Arabia
[ ]Molini et al. (2021)ObD
Host
To assess the presence of LSD in Namibian wildlife, the disease being endemic in cattle in the area.Nasal swabs and DNA samples, tested by PCR and RT-PCR, were collected from wild ruminants shot during the hunting season on a private farm in Namibia.
Only one sample from an asymptomatic eland (Taurotragus oryx) tested positive, out of 12 different wild animals. This is the first evidence of the presence of LSDV DNA in an eland. Forty swabs were analyzed, two were from eland.Although there was a limitation on the number of sampled animals, this confirmed a case of LSDV in a wild animal. No clinical signs were observed so the status of wild animals as reservoirs of LSD remains to be further investigated. Namibia
[ ]Barnard (1997)ObD
Host
To investigate the possibility that game animals (i.e., animals raised for hunting) are involved in the epidemiology of some of the most common viral diseases of livestock
In South Africa.
Authors tested 24 species of South African wild animals for the presence of Abs against 16 common viruses of domestic animals, including LSDV. Standard serological tests were used. The average annual rainfall of the sampling area was calculated, over a 20 year-period.The results of LSD prevalence, based on ELISA testing, were the following: 10% in black wildebeests (3/31 positive), 27% in blue wildebeests (4/15 positive), 23% in springboks (12/53 positive), 20% in impalas (5/25) and 7% in elands (1/15). The prevalence in the different zones varied from 17% in the grassland to 33% in the forest transition area.There is a limitation in using an LSDV serological test, i.e., it cannot confirm if Abs are synthetized vs. LSDV or vs. goat poxvirus. The test results are shown as positive or negative, but the cut-off was not properly defined (authors refer to their many years of experience with the test used for domestic animals). This could generate true or false positive or negative samples.
Only 15 buffalo samples were tested and they were all negative. This sample size was not representative of the real population of buffaloes potentially infected in the national park.
South Africa
[ ]Dao et al. (2022)ObD
Host
To investigate the cause of death of a giraffe in a zoo. Swab samples were collected from skin nodule biopsies and ruptured nodule wound for LSDV isolation. It is the firstly reported detection and isolation of LSDV genome in a sick giraffe. The phylogenetic analysis of the isolate showed its close relationship with previous Vietnamese and Chinese LSDV cattle strains. The source of infection of the giraffe was unknown; the authors presumed contacts with infected cattle but never confirmed such hypothesis.Vietnam
[ ]Carn and Kitching (1995)Exp.
R.T.
To attempt a transmission of LSDV from infected to susceptible cattle housed in close contact, in order to establish the potential for LSDV to spread in the absence of arthropods.Cattle was inoculated by three routes, consistent with a mechanical arthropod-borne transmission: on the conjunctival sac, intra-dermally and intravenously. Seven non-infected animals were housed in contact with infected animals for one month, in an insect-proof facility. Virus neutralization tests were performed to confirm the infection. Different contact experiments were carried through.No susceptible animal became positive. The conclusion was that the transmission of LSDV between animals by direct contact is extremely inefficient, and that a parenteral inoculation of the virus is required. The high proportion of animals who developed a generalized disease after intravenous inoculation implied that field cases of generalized LSD may follow a spread by blood-feeding arthropods.The study relied on an experimental infection, thus cattle are inoculated with a virulent strain and at high titers.
The number of animals used in the experiment was low and the length of the contact period may not have been sufficient.
Not applicable
[ ]Magori-Cohen et al. (2012)ObC
R.T
To evaluate LSD transmission via direct and indirect contact in field conditions.Using mathematical tools, transmissions via direct and indirect contact in field conditions were compared. A transmission model assessed outbreak dynamics and risk factors for LSD.
Data were collected during the 2006-LSD outbreak reported in a large Israeli dairy herd, which included ten separated cattle groups. Transmission by three contact modes was modelled, i.e., indirect contacts between the groups within a same herd, direct contacts or contacts via common drinking water within the groups, and transmission by contact during milking.
Indirect transmission was the only parameter that could solely explain the entire outbreak dynamics; its estimated overall effect was >5 times larger than all other combined routes of transmission. A 15.7-R0 (basic reproduction number) was induced by the indirect transmission from an infected cow remaining for one day in the herd, while the R0 induced by direct transmission was 0.36. These results indicated that LSDV spread within the herd could hardly be attributed to direct contacts between cattle or contacts during milking. The authors therefore concluded that transmission mostly occurs by indirect contact, probably by flying blood-feeding insects. This conclusion has important implications for the control of LSD.The epidemic in Israel was swiftly controlled. Hence, clinically affected animals were removed promptly and the herd was vaccinated, which may have affected the transmission parameters. Israel
[ ]Aleksandr et al. (2020)Exp.
R.T.
To assess the transmission by direct contact among infected and non-infected cows, in an insect-proof facility.This 60-day experiment involved five inoculated bulls (‘IN’ group) and two groups of in-contact animals (five cows per group, named C1 and C2). Cows belonging to C1 were in contact with the inoculated animals at the onset of the trial while C2 cows were introduced at day 33 of the experiment. The bulls were aged 6–8 months and were inoculated with the virulent vaccine-derived recombinant LSDV strain (Saratov/2017).The infection in both groups of contact animals was confirmed clinically, serologically and virologically. Viremia was demonstrated in blood, nasal and ocular excretions, using molecular tools. This is the first evidence of an indirect transmission for a naturally occurring recombinant LSDV isolated from the field. Further studies on LSDV biology are a priority: it is important to gain insights on whether the hypothesized indirect contact evidenced in this study is a de novo-created feature, absent from both parental strains of the novel (recombinant) LSDV isolate, or whether it was dormant but unlocked by genetic recombination.The virulent vaccine-derived recombinant LSDV strain (Saratov/2017) was directly inoculated to the experimental animals. As in other experimental studies, it is hard to establish conditions similar to the field. The virulent character of the strain may have helped the direct transmission.Not applicable
[ ]Osuagwuh et al. (2007)Exp
S.T.
To determine whether the LSD vaccine strain is excreted in semen after vaccination with modified live vaccines, and to determine the efficacy of vaccination in preventing LSDV excretion in semen of experimentally infected vaccinated bulls. Six unvaccinated and six vaccinated bulls were infected 27 days after the second vaccination with an LSD modified live vaccine. Furthermore, six unvaccinated bulls were infected experimentally with a virulent LSDV field strain. Blood and semen samples from the bulls were tested by serum neutralization test, virus isolation and PCR.Vaccinated bulls infected in laboratory conditions tested negative, while unvaccinated bulls were infected. Viral nucleic acid was detected in the severely affected bulls from day 10 post-infection (p.i.) until 28 p.i., end of the trial. LSDV was detected in semen of unvaccinated infected bulls, thus, the vaccine protect against the spread of LSDV via semen.A virulent strain was used and semen was tested when clinical signs of LSD were present. Although it is interesting to give insights on the seminal transmission, the field situation is unknown, i.e., the amount of LSDV recovered in semen of naturally infected bulls and the excretion dynamics are unknown.
[ ]Irons et al. (2005)Exp
S.T.
To establish the incidence and duration of LSDV excretion in the semen of naive bulls infected experimentally. Semen samples from six bulls experimentally infected with a virulent field isolate were collected intermittently over a 90-day period. Semen was collected for testing until three consecutive samples were found to be negative for LSDV by PCR or until the end of the testing period.
Authors conducted virus isolation and tested the infectivity of semen titration in tissue cultures.
All semen samples were LSDV-positive by PCR. The virus was only isolated in two severely affected bulls. This study confirmed the excretion of LSDV in bovine semen for prolonged periods (up to 159 days p.i.) even when obvious clinical signs of the disease were no longer apparent.The experimental infection used a virulent field isolate. Only six bulls were used. Although all samples were PCR-positive, the virus was only isolated from two severely affected bulls.
Although it was isolated by PCR over an extended period, it is unknown how infective the virus is in semen. Indeed, titration to determine the infectivity of viral particles was performed in tissue cultures of a single positive sample, i.e., a bull with obvious clinical signs.
[ ] Sudhakar et al. (2020) ObD
S.T.
Authors reported the first occurrence of LSD in cattle in India; they analyzed the epidemiological and genetic characterization data from LSD outbreaks in the districts of an Indian state. Clinical data were collected in the field.
Sampling (blood, scab), was performed on 60 cattle showing clinical signs of LSD. Seventeen samples of frozen bull semen were obtained from a semen bank farm. DNA extraction, conventional and real-time PCR and phylogenetic analysis were performed.
The study established the presence of LSDV in India the and involvement of LSDV field strains in the outbreaks. It provided evidence of LSDV shedding in semen of naturally infected bulls; 20.45% of frozen bull semen samples were positive. This is a descriptive study, thus only circumstantial inferences can be established. The provenance of the frozen field samples was not explained (i.e., small holdings, type of insemination, natural vs. artificial or mixed, dairy or beef herds), thus the effectiveness of seminal transmission under natural conditions has yet to be established.India
[ ]Annandale et al. (2010)Exp
S.T.
To determine the site of persistence of LSDV in bulls shedding the virus in semen for more than 28 days; to determine if the virus is present in all semen fractions and to study the lesions that develop in the genital tract.Six bulls were infected. Bulls that were PCR-positive on the whole semen sample collected on day 28 p.i. were slaughtered; tissue samples from their genital tracts were submitted to histopathology, electron microscopy, immune-peroxidase staining, virus isolation and PCR. Viral DNA was identified in all semen fractions from all bulls, but mostly from the cell-rich fraction and from the severely affected bulls. The PCR assay was positive on post-mortem samples of testes and epididymides 28 days p.i., from the two severely affected bulls. The authors isolated the virus from the testes of both bulls and from the epididymis of one of them. This study suggests that the testis and epididymis are sites of viral persistence in bulls shedding LSDV in semen for prolonged periods and revealed that viral DNA is present in all fractions of the ejaculate.The time of animal slaughtering conditioned the experimental infection. How long the virus remains in testes and epididymides still needs to be determined, as well as the way it would affect seminal transmission to a heifer.
[ ] Annandale et al. (2014)Exp
S.T.
Whether LSDV transmitted through semen can infect cows and their embryos.The authors performed two controlled trials simultaneously. Eleven beef heifers were synchronized and inseminated with fresh semen spiked with LSDV strain on day 0. Six animals were super-ovulated on day 1, then embryos were flushed from these heifers on day 6. Blood and serum samples were collected from day 4 until day 27 to determine the presence of LSDV and Abs. LSDV was detected by PCR, virus isolation or electron microscopy in blood, embryos and in the organs of experimentally infected animals. LSD was detected in blood, embryos and organs of experimentally infected heifers. This is the first report of experimental seminal transmission of LSDV in heifers and embryos through artificial insemination, thereby confirming the biological risk posed by LSDV-infected semen.The first positive SNT samples was detected 9 days p.i., whilst Irons et al., 2005 detected by 12 days p.i. This illustrates the variability in experimental studies, such as using a higher viral load in this study, or intra-uterine route of infection (previous was intravenously), which allows different exposures to the immune system.
[ ] Annandale et al. (2019) Exp
S.T.
To examine the effects of LSDV in frozen-thawed semen on in vitro embryo production parameters, including viral status of media and resulting embryos.Bovine oocytes were harvested from abattoir-collected ovaries and split into three experimental groups. After maturation, the oocytes were fertilized in vitro with frozen-thawed semen spiked with a high (HD) or a low (LD) dose of LSDV, or with LSDV-free semen (control). Eight day-blastocysts were examined for LSDV by PCR and virus isolation.The presence of LSDV in frozen-thawed semen reduced embryo yield significantly. Moreover, the presence of the virus in 8-day blastocysts confirmed that embryo transfer is a potential risk of virus transmission in cattle.Semen was infected in the laboratory, thus frozen immediately post infection after different dosages of LSDV. Although it clearly shows that frozen bull semen could be a mode of transmission, the risk of generating an LSD outbreak should be assessed. Embryos tested positive only by day 8, thus what happens after implantation and how viable this route of transmission is are still unknown. The laboratory conditions could confound a lower yield, and the optimal conditions to obtain viable embryos are not specified. Not applicable
[ ]Annandale et al. (2018) Exp
S.T.
To investigate the ability of common semen processing techniques to remove LSDV from cryopreserved bull semen, and to investigate the way the virus associates with the sperm cell.A semen sample was collected from an LSDV-negative bull and divided in three parts, two of which were spiked with different LSDV concentrations, i.e., large and small dose, and third one used as control. Samples were cryopreserved and later unfrozen using different processing methods (swim-up, single-layer centrifugation, Percoll gradient and Percoll gradient with trypsin). Semen evaluation methods for motility, PCR analysis, isolation and electron microscopy were performed on the unfrozen sperm.None of the common semen processing methods tested were able to clear (i.e., not effective) spiked frozen-thawed bull semen with LSDV, except for the Percoll gradient with added trypsin, but the semen quality was significantly deteriorated. That poses a biosecurity issue in the semen trade. It is unknown whether the concentrations of LSDV used in the study are comparable to those found in bulls naturally infected and shedding the virus in their semen.
Authors used laboratory infected semen, thus frozen immediately after infection by a virulent strain at different concentrations. Although it clearly shows that frozen bull semen could be a mode of transmission, it should further be tested for cow insemination, in order to determine if there is a risk of introducing LSD into a free area via highly contaminated semen.Not applicable
[ ] Rouby and Aboulsoud (2016) ObD
I.U.
To describe the clinical, histopathological, molecular and serological diagnostic of LSD in a premature one day old-calf, delivered from a cow with clinical signs of LSD.Description of the clinical, histopathological, molecular and serological diagnosis of LSD in the calf. PCR and gene sequencing confirmed the ELISA and serum neutralization tests.SNT confirmed that the one day old-calf had developed pre-colostrum serum Abs to LSDV, which indicated virus transmission in utero. All sera collected from animals located in the same area were serologically positive, which confirmed an exposure to LSDV.Description of a single case study. Although the authors inferred that LSD transmission occurred in utero, they did not explain why only one single calf exhibited clinical signs in the whole herd. It is unknown at what stage of pregnancy the cow was infected. This transmission route is viable but may be affected by other conditions.
[ ]Kononov et al. (2019)Exp
Meat and offal
To determine the potential presence of infectious virus and genetic material in meat and offal products, including testicles, from sub-clinically and clinically ill cattle inoculated with a virulent LSDV strain.Fourteen 6 to 7 month-old bulls were infected with LSDV. Infected animals were culled at 21 days p.i. and samples were collected from muscles, skin nodules, lymph nodes, tongue, trachea, lungs, heart, parenchymal organs, rumen, reticulum, omasum, small and large intestine and testicles. Real time-PCR was performed on the samples to detect LSDV.Findings demonstrated that lymph nodes and testicles of clinically and sub-clinically infected animals are reservoirs of live LSDV, whereas deep skeletal meat in both types of infection does not harbor live virus; the risk of transmission through this product is thus probably very low. The detection of LSDV in testicular tissues in sub-clinically ill animals is a concern, because of the potential spread of the virus through contaminated semen.Experimental infections used a virulent strain. The bulls were culled at an age different from the usual culling age for meat. Nevertheless, the study showed that the risk of virus transmission via sub-products was low.
[ ] Chihota et al. (2003)Exp.
Vec.I.
To investigate the transmission of LSDV from infected to susceptible animals by two species of mosquitoes, the stable fly and a species of biting midge.The mosquitoes Anopheles stephensi and Culex quinquefasciatus, the stable fly Stomoxys calcitrans and the biting midge Culicoides nubeculosus were allowed to feed on either LSD-infected animals or through a membrane on a blood meal containing LSDV. These arthropods were then allowed to feed on susceptible cattle at various intervals after the infective meal. Virus was searched for in the insects by PCR.The LSDV was not transmitted from infected to susceptible animals by An. Stephensi, S. calcitrans, C. nubeculosus and Cx. Quinquefasciatus. The transmission was attempted 24 h post-feeding. Inferences were that S. calcitrans may act as a mechanical vector of LSDV through interrupted feeding over 1–12 h periods, and not over longer periods. In C. nubeculosus midges, LSDV was not detected beyond day 0 post-feeding; the latter was not able to act as a biological vector as there was no evidence of virus replication. Mosquitoes may need to feed on a viraemic lesion to allow transmission. Authors suggested a far more elegant mode of transmission than a mere “dirty-pin” type of virus transfer. Overall, the insect species assessed in the study may be able to transmit LSDV to susceptible animals if their meal on an infected host is interrupted and they have to complete it on another susceptible animal, which is consistent with a mechanical transmission.Vectorial capacity and competency can be overestimated in experimental studies, i.e., animals are experimentally infected with a virulent strain, hence, have a higher viral load. Furthermore, the experimental hosts are shaved and put into adequate dispositions. Vectors feed when the animals show clinical signs, and at determined points of viremia, they are fed directly with infected blood or directly in a shaved portion of the animal skin or lesion. All these factors artificially increase the capacity and competence of vectors. Not applicable
[ ]Sanz-Bernardo et al. (2021)Exp.
Vec.I.
Authors used a highly relevant experimental LSD infection model, in the natural cattle host, and four representative blood-feeding insect species previously reported to have the capacity of acquiring LSDV. The study aimed at assessing their acquisition and retention of LSDV, and determining the LSDV R0 in cattle for each model insect species.Eight cattle were infected by intravenous and intradermal inoculation and all were exposed to: two mosquito species, i.e., Ae. Aegypti and Cx. Quinquefasciatus, C. nubeculosus biting midge and to the stable fly S. calcitrans on different days. Based on these quantitative data, and by combination with data from other studies, the authors used mathematical models to determine the R0 of LSDV in cattle, as mediated by each of these insect species. The probability of vectors acquiring LSDV from a sub-clinically-infected animal was very low (0.006) compared with an animal showing clinical signs (0.23). It means an insect feeding on a sub-clinically-infected animal was 97% less likely to acquire LSDV than one feeding on an animal showing clinical signs. These four potential vector species acquired LSDV from the host at a similar rate, but Ae. Aegypti and S. calcitrans retained the virus for a longer time, i.e., up to 8 days. There was no evidence of virus replication in the vectors, which is consistent with a mechanical rather than a biological transmission. The R0 was highest for Stomoxys calcitrans (19.1), followed by C. nubeculosus (7.1) and Ae. Aegypti (2.4), indicating that these three species are potentially efficient vectors of LSDV.Same limitations as experimental study [ ]. Not applicable
[ ]Sanz-Bernardo et al. (2022)Exp.
Vec.I.
To add results from a previous study on the role of hematophagous insects in the transmission of LSDV. The authors investigated the vector-borne transmission of LSDV in more details, by quantifying the acquisition and retention of LSDV in different anatomical parts of four vector species.Four vector species were focused on: S. calcitrans, Ae. Aegypti, Cx. Quinquefasciatus and C. nubeculosus. They were fed on either a lesion, normal skin of experimentally infected- cows or on an artificial membrane system containing viraemic blood. After feeding, insects were incubated for 0, 2, 4, or 8 days and then dissected into proboscis, head-thorax (including the upper digestive tract and salivary glands), and abdomen or proboscis and head-thorax-abdomen. The DNA of LSDV was searched for by PCR; LSDV titration was performed in skin biopsy. Mathematical models were generated to establish the parameters that influence the acquisition and retention of LSDV, by insects.For the four insect species, the probability of acquiring LSDV was substantially greater when feeding on a lesion compared with feeding on normal skin or blood from an animal showing clinical signs. After feeding on a skin lesion, LSDV was retained on the proboscis for a similar length of time (around 9 days) for the four species and for a shorter time in the rest of the body, ranging from 2.2 to 6.4 days. The insect body, rather than the proboscis, was more likely to be positive immediately after feeding. Acquisition and retention of LSDV by Ae. Aegypti after feeding on an artificial membrane feeding system that contained a high titre of LSDV was comparable to feeding on a skin lesion on an animal showing clinical signs, supporting the use of this laboratory model as a replacement in some animal studies. The probability of acquiring LSDV was highest for S. calcitrans, followed by Ae. Aegypti, Cx. Quinquefasciatus and C. nubeculosus. Same limitations as experimental study [ ].Not applicable
[ ]Gubbins et al. (2019) Exp.
Vec.I.
To estimate the risk of LSDV transmission by five different species of biting insects, based on the R0.The R0’s related to the mechanical transmission of LSDV were estimated based on previously published data of transmission experiments. Vector life history parameters were derived from published literature. The five species of biting insects were: the stable fly S. calcitrans, the biting midge C. nubeculosus, and three mosquito species, i.e., Ae. Aegypti, An. Stephensi, and Cx. Quinquefasciatus.With regard to R0 median (95% confidence interval), the results of skin lesions were the following: S. calcitrans 15.5 (1.4–81.9), Ae. Aegypti 7.4 (1.3–17.6), C. nubeculosus 1.8 (0.06–13.5), An. Stephensi, 1.6 (0.2–6.0) and Cx. Quinquefasciatus 0.8 (0.09–3.5). The results suggest that S. calcitrans is likely to be the most efficient in transmitting LSDV, but Ae. Aegypti would also be an efficient vector. By contrast, C. nubeculosus, An. Stephensi and Cx. Quinquefasciatus are likely inefficient vectors of LSDV.These parameters were estimated based on literature data, in particular, from experiments focusing on LSDV transmission by the five putative vector species. Parameters from the literature could vary as vector competence studies provided variable results. Not applicable
[ ]Kahana-Sutin et al. (2017)ObC
Vec.I.
To assess the possible vector(s) of LSDV under field conditions.A year-round trapping of dipterans was implemented in 12 Israeli dairy farms, one year after LSD outbreaks. Their abundance was compared with their abundance at the onset of 2012- and 2013-outbreaks, under the assumption that vector seasonality remains approximately the same over the years. Vector and environmental data were added to a weather-based model to explain the trapping results. The relative abundance of S. calcitrans during the outbreak period (December and April) was significantly higher compared to other dipterans. This model, based on weather parameters during the epidemic years, showed that S. calcitrans populations peaked in the months of LSD onset, in the studied farms. These observations and model predictions revealed a lower abundance of stable flies during October and November, when LSD affected adjacent grazing beef herds. Therefore, these findings suggest that S. calcitrans is a potential vector of LSD in Israeli dairy farms and that another vector is probably involved in LSDV transmission in grazing herds.The vectorial capacity of S. calcitrans was determined solely by its abundance, the detection of LSDV in the captured vectors was not performed. The study relied on the assumption that vector seasonality remains approximately the same over the years. Data were based on the occurrence of LSD in each farm affected during the 2012- and 2013-outbreaks in Israel (i.e., retrospective data). The vector availability for those years was inferred under the assumption that vector seasonality remains approximately the same over the years. Nonetheless, the study had a good design with a long time period of dipteran trapping; models were appropriate, which gave sound conclusions that S. calcitrans was the potential vector of LSD in Israeli non-grazing dairy farms. However, it also implied that another vector could be the culprit for the outbreaks in beef grazing herds, but no vector was suggested.Israel
[ ] Orynbayev et al. (2021)ObD
Vec.I. Ticks.
To describe the first cases of LSD in July 2016, in the Republic of Kazakhstan.Blood and samples of internal organs (lymph nodes, spleen, lungs, skin with nodular lesions) were taken from sick and dead animals. Ticks, horse flies and biting flies from affected areas or dead animals were submitted to LSDV testing. PCR and gene sequencing were applied.LSDV DNA was detected by PCR in all samples from dead animals and all ticks collected. Four Dermacentor marginatus and nine Hyalomma asiaticum ticks tested positive. LSDV DNA was also detected in three out of 21 horseflies (Tabanus bromius), and in one sample out of two S. calcitrans flies. The study concluded that the emergence of the disease coincided with a peak of vector activity; the introduction of LSDV in Kazakhstan was likely consecutive to the movements of infected livestock, with a subsequent transmission of the virus by blood-feeding insects.The number of vectors sampled for the detection of the virus was very small, i.e., 13 ticks, 21 horse flies and 2 Stomoxys flies. The vectors potentially involved in the outbreaks could not be determined.Kazakhstan
[ ] Makhahlela et al. (2022) ObD
Vec.I.
To increase the morphological and genetic information on the stable fly in South African feedlots, and to determine whether they may harbor LSDV and other pathogens of veterinary and economic importance.This field study consisted in the sampling of stable flies from different feedlots across three South African provinces. Flies were identified according to the standard key morphological characters. PCR were performed to detect the presence of LSDV DNA.LSDV DNA was detected in 8/53 samples, i.e., 15.08%. In South African feedlots, S. calcitrans harbours A. marginale and LSDV, which suggests that they may be involved in their mechanical transmission to livestock.The study only shows that some stable flies were positive to LSDV in several south African feedlots. No other conclusion can be drawn from that study. No information is provided on how the sampling size was determined. Pool samples varied in terms of number of flies per pool. The authors did not specify in the results section if they were dealing with the number of pools or the number of insects positive to LSD. However, the study did show that flies positive to LSDV are present in South African feedlots.South Africa
[ ]Issimov et al. (2020) Exp.
Vec.I.
To determine the vector competence of three Stomoxys spp. for the transmission of LSDV.S. calcitrans, S. sitiens and S. indica were allowed to feed to repletion in experimentally infected-cows, after which they were tested for LSDV. Another batch was allowed to feed incompletely and then was moved to a healthy animal to complete feeding. PCR, serum neutralization test and virus isolation were performed to detect LSDV.Recipient animals were all positive. St. calcitrans, S. sitiens and S. indica were negative 24 to 48 h post-feeding. All three species of flies demonstrated the capacity to ingest and harbor viral particles. They were able to transmit the virus within a 1 h time-interval between the meals. Moreover, LSDV was recovered from fly mouth parts within the same period and LSDV can survive in Stomoxys spp. at least 6 h following a meal on an infected animal. The mechanical transmission from infected to susceptible animals was demonstrated under laboratory conditions.The study only determined the competence of the Stomoxys fly under laboratory conditions. See also the Same limitations as experimental study [ ]. Not applicable
[ ]Issimov et al. (2021) Exp.
Vec.I.
The authors attempted to define the duration of LSDV retention in three Stomoxys spp., after intrathoracic inoculation, as well as virus potential to replicate after bypassing the midgut barrier.A virulent LSDV strain was inoculated directly in the thorax (to bypass the midgut barrier) of adult flies of S. calcitrans, S. sitiens and S. indica. The flies were tested for the presence of LSDV DNA by gel-based PCR and virus isolation, at different times and days post- inoculation. The virus was retained by the three Stomoxys spp., under laboratory conditions. LSDV was isolated from all three Stomoxys spp. up to 24 h post-inoculation while virus DNA was detectable up to 7 days post-inoculation. The outcomes illustrated the incompetence of Stomoxys spp. to serve as a biological vector of LSDV. Although it demonstrated the incompetence of Stomoxys spp. as a biological vector and the virus was retained in the Stomoxys, the virus was directly inoculated into the thorax, which would this increases the probability of the fly to be positive to LSDV. Not applicable
[ ]Paslaru et al. (2021)Exp.
Vec.I.
To investigate the role of S. calcitrans in the transmission of LSDV and its presence in four different farms in Switzerland. Laboratory-reared S. calcitrans flies were exposed to LSDV-spiked blood. Engorged flies were incubated and body parts, i.e., heads thorax and abdomens, were tested for the presence of LSDV DNA for up to 72 h post-feeding. LSDV DNA was tested with a DNA mini commercial kit. Correspondingly, virus isolation in cell culture from regurgitated blood and in fecal samples of the flies was carried through. The presence of the fly in different farms and at high altitudes was assessed by trapping.LSDV DNA was detected in heads, bodies, and regurgitated blood, up to 3 days post-feeding and up to 2 days post-feeding in the feces. Infectious virus was isolated from bodies and feces up to 2 days and up to 12 h post-feeding in the regurgitated blood. The viral load increased, which consolidates the role of S. calcitrans as a mechanical vector of LSDV. The fly was present in all farms investigated, including a farm located at 2128 m above sea level, showing that it is abundant and widespread. Feeding of the stable fly was performed by placing them in cotton pads soaked with blood spiked with LSDV and not by placing them onto LSDV infected animals, which could increase the competence of the fly. Despite such fact, the experimental study showed that S. calcitrans was a competent mechanical vector of LSDV; its abundance in the farms showed that it would be a capable vector for spreading the virus between the animals. Not applicable
[ ]Sohier et al. (2019) Exp
Vec.I.
To focus on the potential mechanical transmission of LSDV and to assess whether stable flies and horse flies could transmit LSDV when a shorter period between interrupted feeding on LSDV viraemic cattle followed by further feeding on naïve cattle would apply. Bulls were experimentally infected. Three independent experiments were performed wherein biting flies, i.e., S. calcitrans and tabanids Haematopota spp., were allowed to feed for 10 min on LSDV infected-bulls (when animals were viremic or upon emergence of nodules). Potentially infected-insects were then allowed to feed for 10 min on susceptible cattle, one hour after the infective meal. In the other two experiments, insects were placed on the animals for two to three consecutive days. Blood was collected and biopsies of nodules were performed for RT- PCR analysis and virus neutralization test.LSDV transmission by S. calcitrans was evidenced in the three independent experiments; LSDV transmission by Haematopota spp. was shown in one experiment. Results supported the mechanical transmission of the virus by these vectors. The study provided the first evidence of LSDV transmission by S. calcitrans and Haematopota spp. It is the first formal demonstration, under experimental conditions, that S. calcitrans is a vector of LSDV. LSDV was transferred from a donor to a receptive animal by flies exposed to the virus for maximum 3 days (and even 1 day for another animal) provides strong evidence that the transmission was mechanical and not biological. Horse flies also transmit LSDV, possibly more efficiently than stable flies. Indeed, one of the two horseflies put in contact with the receptive animal became positive. The large mouthparts of tabanids are helpful for mechanical transmission, as they can retain high blood volumes, and thus inoculate higher viral doses. The competence of both stable and horse flies was determined. The capacity of both species was inferred by their vector characteristics and not by modelling. See also the same limitations as experimental study [ ].Not applicable
[ ]Sprygin et al. (2020) LitRev
Vec.I.
That literature review gained insight on the relationship between climatic conditions, ecological characteristics of the stable fly (S. calcitrans L.) and the observed spread of LSD across the Russian Federation, in 2015–2019.Information on the entomology of S. calcitrans was compiled. Authors described the spread of LSD in cattle, in the Russian Federation, between 2015 and 2019; climatic conditions in the regions where the outbreaks occurred were recorded. The authors relied on data from domestic and foreign authors, on reports of Russian authorities on the spread of LSD in cattle and on meteorological data.Data analysis showed that the activity of the stable fly mainly fits during the seasonal pattern of LSD outbreaks. However, some outbreaks occurred outside the activity period of the stable fly, pointing to other routes of transmission.Vector capacity was based on previous studies.Not applicable
[ ] Chihota et al. (2001)Exp
Vec.I.
Given that Ae. Aegypti was identified as an important vector of poxviruses, e.g., the myxoma virus, the study was undertaken to determine whether that mosquito species can act as an efficient mechanical vector of LSDV.Fifty one week-old adult females of Ae. Aegypti fed on a lesion of experimentally infected steers. Transmission of the virus was then attempted by allowing these mosquitoes to feed on six susceptible cattle, at various times post- feeding. Transmission was confirmed by recording LSD clinical signs or recovering live virus from lesion material or blood of susceptible animals. DNA was extracted from infected mosquitoes and essayed by PCR. Cows were tested by PCR, virus isolation, virus neutralization index and their clinical score was recorded. The duration of virus transmission was also recorded.Results showed that LSDV could be transmitted by Ae.aegypti for at least 6 days after infection. LSDV was able to survive in infected mosquitoes for at least 6 days, at a quite similar titer, and was then transmitted. The virus could be localized within the mosquito in a site protected from inactivation. The authors suggested a far more complex mode of transmission than a mere ‘dirty pin’. In conclusion, Ae. Aegypti female mosquitoes have the capacity to transmit LSDV mechanically, from infected to susceptible cattle. The clinical signs recorded in animals exposed to infected mosquitoes were generally mild, only one case being moderate. LSDV was long-suspected to be transmitted by insects, but these findings are the first to demonstrate that theory unequivocally; authors suggested that Ae. Aegypti was a competent vector.Competence of the mosquito was determined by experimental infection. The main limitation is that mosquitoes were allowed to feed on a lesion, which is not necessarily the case in the field if one consider its anthropophilic behavior (i.e., preference to bite humans rather than animals).Not applicable
[ ]Paslaru et al. (2022) Exp
Vec.I.
To expand on the findings of the insect ‘model vector species’. The LSDV suitability of mosquitoes and biting midges was investigated. The mosquito species Ae. Aegypti, Cx. Pipiens and Ae. Japonicus were allowed infectious blood meals for 45 min. Field collected-Culicoides spp. and 2–3 day old laboratory reared-C. nubeculosus were exposed to an infectious blood meal for 30–45 min. The insects were tested for the presence of LSDV. DNA was extracted and isolated; bodies and head or wings were proxy for the virus dissemination at different time points after feeding. Post-feeding viral retention lasted for 10 days for Ae. Japonicas and 7 days for Cx. Pipiens. In the three mosquito species investigated, more body samples where PCR-positive compared to head samples, indicating that the virus was not efficiently retained in the mouthparts and that there was no virus dissemination. Thus, mechanical transmission of LSDV by these species seems feasible in case of interrupted feeding. Viral DNA could be detected in feces of Ae. Aegypti until day 4 after feeding, although the significance of that finding is unclear. Thus, mosquitoes might serve as mechanical vectors of LSDV in case of interrupted blood meals. In C. nubeculosus, the virus was isolated from homogenized bodies up to the end of the experiment (10 days p.i.). Interestingly, Cq values decreased over time, and a disseminated infection at day 10 p.i. was identified in one insect. Considering the postulated absence of salivary gland barriers in Culicoides spp., these findings indicated that the laboratory-reared C. nubeculosus might behave as a biological vector of LSDV under laboratory conditions. LSDV did not persist in field-collected biting midges. All insects were fed with LSDV-spiked blood meals and not directly on infected animals. Thus, competence may be inferred but the vectorial capacity of the mosquitoes cannot be implied. The virus was detected on homogenates of heads and body parts, rather than on the whole insects.
Viable virus was isolated from homogenized bodies until day 10 post-infection. Culicoides nubeculosus was assumed as a biological vector, under experimental conditions, but based on a single insect with disseminated infection at day 10 post feeding, and the absence of salivary gland barriers in the Culicoides spp. The field-collected C. nubeculosus showed no persistence of LSDV, which suggests its most likely low competence.
Not applicable
[ ]Şevik and Doğan (2017) ObC
RiskF.
To determine the epidemiological status of observed LSD in several regions of Turkey; to evaluate the risk factors associated with LSDV infection; to determine the phylogenetic relatedness of the LSDVs circulating in Turkey; to assess the economic cost of LSD in surveyed regions; to investigate the potential role of Culicoides spp. in the transmission of LSDV.Multiple samples were collected on dead animals: skin nodules, vesicle swabs, whole blood on EDTA tubes, lymph nodes, spleen, lungs, liver and heart; internal organs of aborted bovine foetuses were also sampled. Culicoides spp. were trapped in regions were the highest number of LSD cases was recorded. DNA was extracted and RT-PCR performed, along with sequence alignment and phylogenetic analysis.
A questionnaire was submitted to livestock owners to collect information on LSD occurrence and other farm characteristics (location, type of herd, dairy of beef, total number of cattle on farm, number of cattle affected and dead from LSD, animal age, breeds affected and history of vaccination). Generalized linear mixed models investigated the risk factors influencing LSD prevalence.
The generalized linear mixed model provided the following results: European cattle breeds, small-sized family farms and farms located near a lake were identified as risk factors influencing LSD prevalence.
The species of Culicoides in LSDV-positive pools was C. punctatus. The finding of LSDV in C. punctatus suggests that it may play a role in the transmission of LSDV. Furthermore, movements of infected animals to disease-free areas increase the risk of LSD introduction. Strategies of LSDV control should consider the risk factors identified in this study.
The model chosen to establish the factors influencing LSD prevalence was a linear model (not logistic), so it is hard to interpret the effect of a factor on LSD prevalence. Only LSD-suspected animals were sampled, and no sample size was calculated. The risk factors were not well established. Some factors assumed that the cattle died of LSD. The ‘near any lake’ factor is subjective as no distance from the farms affected by LSD was provided.
Culicoides spp. were positive but the study inferences on their role in LSDV transmission were subjective.
Turkey
[ ]Sprygin et al. (2018) ObD
Vec.I.
To report the epidemiological investigation of an LSDV case caused by a vaccine-like strain in Russia, including attempts to detect the vaccine-like strain in several insect species trapped at outbreak location.Samples of blood and scabs from cows of three affected farm (cows presenting clinical signs consistent with LSD) were collected and tested for field-LSDV DNA using a RT-PCR and vaccine-LSDV DNA using an assay developed for this specific work.
An entomological surveillance based on insect trapping was implemented during 2 weeks after confirmation of the outbreaks. Trapped houseflies were divided into two batches for pooled and individual testing. The other captured insects, stable flies and lesser flies were tested individually. The testing was for the presence of LSDV DNA and vaccine-like LSDV DNA.
There was no evidence of field-LSDV strain circulation. The DNA of vaccine-LSDV was present in cattle. Stable flies tested individually, and to a lesser extent houseflies, were negative. The pool tested included three to five houseflies sampled randomly; 14 out of the 25 pools tested positive to vaccine-like LSDV DNA, but not to field-LSDV DNA. Flies were washed four times and tested. In Musca domestica, LSDV DNA was mainly detected in the first wash fluid, suggesting genome or even viral contamination on the insect cadaver. Internal contamination of insect bodies, without any differentiation between body locations, was also revealed; however, the clinical relevance for mechanical transmission is unknown. In this study, we discovered that M. domestica flies carried vaccine-like LSDV DNA whereas stable flies trapped at the same time were negative for both field- and vaccine-like LSDV DNA.Although the first isolation of LSDV DNA from internal parts of non-biting insects is a very important finding, their role in LSDV transmission and spread still needs to be investigated.Russia
[ ]Wang et al. (2022) ObD
Vec.I.
To investigate the first LSDV case caused by a vaccine-like strain at the western border of China; search for LSDV DNA in several insects captured around the region during the outbreak.The authors implemented a surveillance of insects around the infected premises and the neighboring bordering areas. Insects were trapped; DNA was extracted and screened by RT-PCR and sequencing. A phylogenetic analysis was carried through.The most abundant species captured during the campaign was C. pipiens, but all were negative to LSDV. It suggests that species was not involved in the LSDV epidemic. The overwhelming majority of captured insects were non-biting. Two kinds of non-biting flies, i.e., Musca domestica L and Muscina stabulans, were positive for vaccine-like LSDV. Despite such finding, there was no direct evidence to support cross-border transmission of the vaccine-like LSDV. The positivity of surface and negativity of internal contents indicated that non-biting flies could only acquire the virus by physical contamination. The non-biting flies were the only insects to be positive to vaccine-like LSDV strain, and only on the surface of the body. Thus, their vectorial competence still needs to be determined. China
[ ] El-Ansary et al. (2022) ObD
Ticks
To investigate and assess LSDV isolated from ticks collected in various outbreaks in Egyptian governorates and to characterize the virus at the molecular level.Adult ticks were collected from cows in different Egyptian regions. Laboratory detection of LSDV was performed by PCR and sequencing. Further identification was carried on by non-serological methods.Rhipicephalus (Boophilus) annulatus was the most prevalent tick species on cattle in the investigated regions; 15% of them were positive to LSDV. The majority of recent LSD outbreaks occurred in a period with mild and wet weather, i.e., from May to September, which favors tick activity. The tick sample size was large, i.e., 4000 adult ticks. The number of positive samples was obtained by extrapolating the numbers of ticks from the positive pool samples wbich gave a total of 600 positive ticks out of 4000. Which extrapolated to 600 out 4000. Although it was a large sample size, the study only infers that ticks were positive to LSDV, which could determine their vectorial competence but not their capacity to transmit LSDV.Egypt
[ ] Rouby et al. (2017) Exp.
Tick
To investigate the role of R. annulatus ticks collected from naturally infected animals in the transmission of LSDV.Naturally infected cattle with LSD acute clinical signs underwent clinical examination. Samples of skin nodules and R. annulatus stages were collected from the sick cattle and examined by PCR; positive samples were confirmed by direct gene sequencing. Female engorged ticks were incubated for egg deposition; eggs and larvae that hatched were then screened for virus isolation and confirmed to be infected by PCR.Detection of LSDV in tick larvae proved the possibility for these to be a potential source of infection for susceptible animals.
The present study showed that females of naturally infected R. annulatus were able to transmit the virus vertically, via eggs to larvae. These findings suggest a high possibility for ticks to be a risk for the virus transmission and a field reservoir host of LSDV.
The competence of naturally infected ticks was established. Their role as a reservoir was not established, but only speculated. Not applicable
[ ]Tuppurainen et al. (2015)Exp.
Tick
To investigate in vitro replication and/or survival of LSDV in cell lines derived from the tick species R. appendiculatus, R. evertsi and R. (B.) decoloratus and investigate the presence of the virus in live ticks collected from naturally infected cattle during LSD outbreaks in Egypt and South Africa. LSDV was inoculated in tick cell lines: four semi-engorged female Rhiphicephalus spp. were collected in Egypt from three cows recovering from LSD but still showing some skin lesions and cabs. Tick samples were obtained from Egypt and South Africa. Detection of LSDV was carried out by real time PCR and virus titration. There was no evidence of LSDV replication in tick cell lines, although the virus was remarkably stable, i.e., remaining viable for 35 days at 28 °C in tick cell cultures. Viral DNA was detected in two-thirds of the 56 field ticks. This is the first report to highlight the presence of potentially virulent LSDV in ticks sampled on naturally infected animals. All four ticks collected from Egypt were positive to LSDV. Out of the 52 samples collected from South Africa, 11 were R. appendiculatus, four R. Boophilus, seven A. hebraeum, four H. truncatum, two Amblyomma sp., six Rhipicephalus Boophilus sp. The inability of LSDV to replicate in tick cell lines shed some information on the ability of the tick to act as a biological vector of LSDV. Not applicable
[ ]Lubinga et al. (2014) Exp.
Tick
To further understand the role of ixodid ticks in the transmission of LSDV. The study aimed at determining the specific organs of adult R. appendiculatus and A. hebraeum infected by LSDV following an interrupted meal (intrastadial), and the transstadial persistence.Nymphs and adult of R. appendiculatus and A. hebraeum ticks were orally infected by feeding on cattle infected experimentally by LSDV. For intrastadial infection, ticks were placed on infected animal for 4 days (on day 12 p.i.) after which they were collected for testing. LSDV was detected by immunohistochemistry, electron microscopy and RT-PCR. For transstadial persistence, nymphs fed on infected animals and once engorged, they were incubated for molting. Two months after emergence, they were put on LSD-free receptive animals and collected after for LSDV detection.Intrastadial and transstadial transmissions were demonstrated for R. appendiculatus. The same observation had been performed for A. hebraeum in a previous study. The virus was able to cross the midgut wall and infect various organs, indicating a potential for biological development and transmission of LSDV by ticks. The salivary glands were the most affected organs, strengthening the previous report of LSDV occurrence in tick saliva.Experimental infection affects the competence as it depends on the strain used and on direct feeding on an infected animal. A controlled environment facilitates infection, thus tick competence can be estimated. However, its vectorial capacity is still to be determined as these tick species do not spend their entire life cycle on the same host. Not applicable
[ ] Lubinga et al. (2014)Exp.
Tick
To investigate the passage of LSDV from engorged A. hebraeum nymphs to adults, and from engorged female R. decoloratus to larvae, under cold temperatures, in order to determine their possible role in the overwintering of LSDV.A. hebraeum and R. decoloratus female ticks were fed to repletion on LSD-free cattle. Thereafter, they were experimentally infected with LSDV on the day they dropped from the host. Nymphs were also infected and incubated at room temperature (25 °C), and at maximal and minimal winter temperatures, i.e., approximately 20 °C during the day and 5 °C at night. Virus isolation, RT-PCR and immunoperoxidase staining were performed to detect LSDV in the corresponding samples. Transmission electron microscopy was used in tick organs. Transstadial and transovarial persistence of LSDV were observed in experimentally infected A. hebraeum nymphs and R. decoloratus females, after a 2 month-exposure to cold temperatures, i.e., 5 °C at night and 20 °C during the day. This finding suggests a possible overwintering of the virus in these tick species.Same limitations as for study [ ].Not applicable
[ ] Lubinga et al. (2014)Exp.
Tick
To study the egg-transmission of LSDV from infected female ticks to the larvae in A. hebraeum, R. appendiculatus and R. decoloratus. Laboratory infected cattle hosted adult A. hebraeum, R. appendiculatus and R. decoloratus during the viraemic stage. Two other animals were used as receptive hosts to assess the transmission of LSDV by A. hebraeum and R. appendiculatus larvae, respectively. Subsequently, these ticks fed on LSD-free animals to observe if mechanical transmission occurs.The detection of LSDV in larvae of A. hebraeum, R. decoloratus and R appendiculatus indicates a transovarial passage of LSDV in these species. Authors showed LSDV transmission to receptive animals by A. hebraeum, R. appendiculatus larvae. These findings, in accordance with other studies, suggest a high possibility that ticks act as reservoir hosts of LSDV in the field. The overwintering in some tick species such as R. decoloratus may play a significant role in the overwintering of LSDV.Same limitations as for study [ ].Not applicable
[ ]Lubinga et al. (2015)Exp.
Tick
To investigate the potential role of Amblyoma hebraeum ticks in mechanical/intrastadial and transstadial transmission of LSDV.Adults and nymphs of A. hebraeum ticks were placed to feed on animals artificially infected with LSDV and subsequently transferred (nymphs after incubation up to 35 days to molt to adults) to naïve recipient cattle. Successful transmission of LSDV to recipient animals was determined through monitoring of clinical signs and laboratory detection of LSDV by RT-PCR, SNT and virus isolation. This report provides further evidence of mechanical intrastadial and, for the first time transstadial, transmission of LSDV by A. hebraeum. These findings implicate A. hebraeum as a possible reservoir host in the epidemiology of the disease.Same limitations as for study [ ].Not applicable
[ ]Tuppurainen et al. (2013)Exp.
Tick
To examine the potential for transovarial transmission of LSDV in R. decoloratus ticks.Tick larvae were put on infected cows up to completion of life cycle and were allowed to lay eggs. After hatching, larvae were transferred to non-infected receptive cattle. Blood samples were collected from these cattle hosts at different days p.i. Laboratory detection of LSDV was performed by RT-PCR, SNT and virus isolation. Receptive animals showed mild clinical signs with characteristic lesions. Thus, R. decoloratus ticks were able to transmit LSDV transovarially; this is the first report of such type of transmission for a poxvirus. Same limitations as for study [ ].Not applicable
[ ]Tuppurainen et al. (2013)Exp.
Tick
To investigate if LSDV can be transmitted mechanically by African brown ear ticks Rhipicephalus appendiculatus.Laboratory-bred R. appendiculatus males fed on experimentally infected viraemic cattle. Partially fed male ticks were then transferred on non-infected cows. The receptive animal did not develop any visible skin lesion post-infection.The receptive animal became viraemic, showed mild clinical signs of LSD and seroconverted. Thus, R. appendiculatus ticks are able to act as mechanical vectors of LSDV. Additionally, R. appendiculatus males transmitted LSDV though feeding on visibly intact skin, which demonstrated that viraemic animals with no lesion at the tick-feeding site may be a source of infection. This is the first demonstration of poxvirus transmission by a tick species.Same limitations as for study [ ].Not applicable
[ ] Lubinga et al. (2013) Exp.
Tick
To detect LSDV in saliva of A. hebraeum and R. appendiculatus adult ticks fed, as nymphs or adults, on LSDV-infected animals; thereby, the authors also aim at demonstrating transstadial or mechanical/intrastadial passage of the virus in these tick species.Cattle were experimentally infected with LSDV and used to host nymphs and adult ticks of A. hebraeum and R. appendiculatus. The presence of LSDV in the saliva of these adult ticks was investigated by RT-PCR and virus isolation.For the first time, LSDV was detected in the saliva of both A. hebraeum and R. appendiculatus ticks. At the same time, the authors demonstrated the persistence of LSDV in ticks between developmental stages (transstadial) and within the same stage (intrastadial) in both tick species. Same limitations as for study [ ].Not applicable
[ ] Tuppurainen et al. (2011)Exp.
Tick
To investigate the potential role of ixodid (hard) ticks in the transmission of LSD.Three common African tick species, i.e., R. appendiculatus, A. hebraeum and R. (B.) decoloratus, at different life stages, were fed on the skin lesion of infected animals during the viraemic stage. After feeding, the partially fed male ticks were transferred to the skin of non-infected “receptive” animals, while females were allowed to lay eggs; these eggs were tested by PCR and virus isolation. Nymphs were allowed to develop for 2–3 weeks before testing. The receptive cattle were tested for LSDV.This is the first molecular evidence of potential LSDV transmission by ixodid ticks. The study evidenced transstadial and transovarial transmissions of LSDV by R. (B.) decoloratus ticks and mechanical or intrastadial transmission by R. appendiculatus and A. hebraeum ticks.Same limitations as for study [ ].Not applicable
[ ] Kononov et al. (2019)ObD
Vac.
The present study follows up the epidemiological situation since 2016, and further examines samples containing vaccine-like LSDV strains, in the Privolzhsky Federal District, in 2017. That area is geospatially outside the zone affected in 2016 and where live vaccines against LSDV had never been authorized or knowingly used.That field study investigated 13 out of 42 outbreaks. Whole blood, nasal swabs, and scabs were sampled and tested by PCR.
Sequence analysis by amplifying the nucleotide sequences of RPO30 and GPCR gene to determine the type of strain of the LSDV.
Four outbreaks, i.e., two in backyard cattle and two in commercial farms were caused by vaccine-like LSDV strains, whereas the nine other outbreaks were attributed to field strains.
Vaccine-like LSDV strains were isolated in two out of 21 backyard cattle and in 96 out of 2112 animals sampled in two commercial farms.
Although live attenuated LSDV vaccines are prohibited in Russia, several vaccine-like LSDV strains were identified in the 2017 outbreaks, including commercial farms and backyard animals exhibiting clinical signs consistent with field LSDV strains. Sequence alignments of three vaccine-like LSDV strains showed a clear similarity to the corresponding RPO30 and GPCR gene sequences of vaccine attenuated viruses. How vaccine-like strains spread into Russian cattle remains to be clarified.
Not all outbreaks were sampled. The study managed to show that vaccine-like strains of LSD were the culprits of some outbreaks occurring in the region.Russia
[ ]Aleksandr et al. (2020) ObD
Vac.
To report the emergence of a novel vaccine-like LSDV variant in Kurgan Oblast (Russia), along the southern Kazakh border, in 2018.Samples of blood, serum and skin were collected from cows. DNA was extracted and RT-PCR performed to isolate the virus. Sequence and melt curve analysis were carried out as well.Phylogenetic analysis of these additional loci placed the Kurgan/2018 strain in either vaccine or field groups, strongly suggesting a novel recombinant profile. This is another piece of evidence exposing the potential for recombination in capripoxviruses and the ignored danger of using live homologous vaccines against LSD. Authors discussed the need to revise the PCR-based strategy to differentiate infected from vaccinated animals and the potential scenarios of incursion. The contribution of KSGP/NI-2490-like strain to the emergence of the recently identified vaccine-like recombinant is discussed.A new variant is described. That descriptive study accurately detected the vaccine-like strain. Russia
[ ]Sprygin et al. (2020)ObD
Vac.
To perform full genome sequencing of the strain in order to characterize the genetic background of the strain responsible of an LSD outbreak that occurred during the winter 2019. Field samples were collected, then the virus was isolated and cultured on lamb testis cells before purification. Genomic DNA was extracted and sequenced.The proteins encoded by the ORFs are of high importance, since the findings show they mutated repeatedly from attenuated vaccine profiles to virulent wild-type profiles. Further work is needed to assess the extent to which recombinant vaccine-like strains spread in the country. Experimental work aimed at correlating the genetics of recombinant progeny with the virulence observed in infected hosts would also be interesting.The study describes the importance of the proteins encoded by ORFs. This can provide indications on how recombinant vaccine-like LSDV strains spread in Russia.Russia
[ ]Shumilova et al. (2022) ObD
Vac.
To report the detection and analysis of another recombinant strain from Saratov in 2019; that strain seems to be a clonal progeny of Russia/Saratov/2017, that overwintered in the region since 2017.Viral samples were collected in the Saratov region, Russian Federation, in 2019. The samples were seeded on propagated and purified goat ovarian culture. DNA was extracted and sequenced. The findings demonstrated the persistence of LSDV during winter and successful overwintering in a cold climate, which encourages additional research on LSDV biology.No inferences were made on the origin of the vaccine strain. The reported outbreaks occurred in cold climates (i.e., outside the normal range of vector activity), which shows the overwintering of the virus. This conclusion is very important and the authors should have explained it more in details. Russia
[ ] Sprygin et al. (2020)ObD
Vac.
To provide an overview of LSDV evolution in the Russian Federation since its first occurrence in North Caucasus in 2015 and further spread eastward, along the Kazakh border.Blood samples were collected between 2015 and 2018 from cows presenting clinical signs of LSDV. DNA extraction was performed on 21 LSDV isolates from different regions and the presence of LSDV DNA was initially confirmed by PCR. Phylogenetic analysis was performed.The findings showed that, between 2015 and 2018, the molecular epidemiology of LSDV in Russia split into two independent waves. The 2015–2016-epidemic was attributable to a field isolate, whereas the 2017-epidemic and even more the 2018-epidemic, were caused by novel importations of the virus, not genetically linked to the 2015–2016 field-strain. Such observations demonstrated a new emergence rather than the continuation of a field-type epidemic. Since recombinant vaccine-like LSDV isolates seem to have entrenched across the country border, the policy of using certain live vaccines requires revision as it is a clear biosecurity threat.The study design describing the Russian LSD epidemic was well conducted. Inferences showed that new disease importations occurred in 2018. The authors could have provided hypothesis on the origin of emergence. Biosafety of the vaccine was questioned, but not the fact that it may have been poorly produced. Russia
[ ] Byadovskaya et al. (2022)ObC
Vac.
To summarize the LSD outbreaks occurring between 2015 and 2020 across the Russian Federation and discuss the epidemiological features and possible risk factors in the current epidemiological situation. Location data (i.e., geographical coordinates) of LSD outbreaks were collected, along with the date of the disease onset, the number of susceptible, infected and dead animals, average monthly temperatures and cattle density (2010 national statistics).
A spatiotemporal analysis was performed, i.e., spatiotemporal clusters, a permutation model, a Poisson model and a directionality test.
The outbreaks of LSD occurred primarily in small holdings (backyard) rather than in commercial farms, mainly during the warm months, with the majority of outbreak peaks occurring in mid-summer.
A highlight was made that in 2018 LSD cases continued until November and in snowy March 2019, i.e., winter conditions (snow and freezing temperatures) that preclude vector activity.
Disease tended to form annual spatiotemporal clusters in 2016–2018, whereas in 2019 and 2020, such segregation was not evident.
The spatial-temporal analysis was well-conducted and gave a general picture of the clusters that occurred in Russia. Cold weather conditions, precluding vector activity, were highlighted. Although there were evident clusters, the effect of vaccination during the outbreaks was not mentioned neither included in the analysis. Russia
[ ]Ma et al. (2021)ObD
Vac.
To characterize the genomic and phylogenetic features of an LSDV strain detected from cattle with typical LSD clinical signs in farms of southeast China.Skin nodules, wounds, ocular, nasal, oral and rectal swabs were sampled from six affected cattle. The authors performed viral DNA detection, genomic sequencing and recombination analysis. At least 25 putative recombination events between a vaccine strain and a field strain were identified in the genome of GD01/2020, which could affect the virulence and transmissibility of the virus. These results suggest that a virulent vaccine-recombinant LSDV, from an unknown origin, was introduced China, in Xinjiang, in 2019, and spread to Guangdong in 2020.The study focused on the characterization of the LSDV strains detected in cattle clinically affected by LSDV. The question on how LSD was introduced in China remains unanswered. China
[ ] Vandenbussche et al. (2022)ObD
Vac.
The aim of the study was twofold: (1) to analyze the composition of two batches of the Lumpivax vaccine and (2) to investigate a possible link between the vaccine and the recent vaccine-like recombinant LSDV strains.The following processes were carried out: virus sequencing, reconstruction of vaccine strains, genome-wide analysis,
recombination and breakpoint analysis.
The great divergence of recombinant strains in the batches (Neethling-like LSDV vaccine strain, KSGP-like LSDV vaccine strain and Sudan-like GTPV strain) suggests that they arose during seed production. The recent emergence of vaccine-like LSDV strains in large parts of Asia is, therefore, most likely the result of a spill over from animals vaccinated with the Lumpivax vaccine.The study is well conducted and provides reasonable evidence that the vaccine-like strains causing the latest outbreaks in Russia and Asia are due to poorly manufactured Lumpivax vaccine.Not applicable
[ ] Selim et al. (2021)ObC
RiskF.
To investigate LSDV seroprevalence in cattle with no history of vaccination, in some governorates of northern Egypt, and to assess the risk factors of infection.Samples were collected randomly and classified according to the type of herd (dairy and beef), the breed (Baladi, mixed, and Holstein), the season (autumn, winter, spring, and summer), the age (range between <1 and >3 years old) and sex (male/female), if the sample came from animals with contact with other animals, water sources and feeding. Serum samples were analyzed by ELISA testing.
The authors performed a multivariate logistic regression model and a chi-square analysis.
The multivariate logistic regression gave the following results. The risk of infection by LSDV was higher in Holstein breed, adult cattle and in the summer. Furthermore, communal grazing (i.e., sharing pastures) communal water points (i.e., shared water sources), introduction of new animal in a herd, and contact with other animals were identified as significant risk factors for the occurrence of LSDV infection in cattle.The study was well designed and the risk factors well established. The only limitation of the study is that it relied on serology testing, and thus, the authors can only assume that the samples tested positive because of LSDV. Only unvaccinated cattle were assessed. Egypt
[ ] Ince and Türk (2019) ObC
RiskF.
To analyze potential risk factors of LSD by a GIS and provide information to control its spread.GIS systems and user interface programs were developed. The following data on LSD outbreaks were used: farms, cattle movements as well as temperature by the time of the outbreak. The authors assessed by combining an active disease follow-up, a questionnaire and retrospective data that focused on 70 pastoral and agro-pastoral farms, from August 2013 to December 2014.
A multivariate logistic regression computed the strength of contribution of these risk factors to LSD occurrence.
The most significant risk factor affecting LSD prevalence was the proximity with the southern border of Turkey; the transmission of the disease to Turkey may have occurred from Syria and Iraq, since movements of live animals across the Syria–Iraq border exist and the first outbreak was recorded near the border. Analyses of morbidity risk factors of animal movements and animal markets showed that cattle purchased from other farms were at risk. For the transmission of LSD among farms, the most significant factor was cattle movements. LSD prevalence was significantly associated with purchasing infected animals that had not been tested or quarantined. The number of registered LSD outbreaks was higher in the summer, which suggests a seasonal distribution of LSD outbreaks during dry seasons. A seasonal trend of LSD outbreaks was observed in 2014. The number of reported outbreaks increased from June to October 2014, with a peak in August.
The multivariate logistic regression concluded that cattle < 24 months old were more likely to be infected; females were more at risk than males and vaccinated animals were less at risk.
The risk factors were not well defined. The results of the final model were badly displayed and hard to interpret. Large confidence intervals show that there may be an issue in sample size or in the number of cattle tested. Conclusions were based on circumstantial evidence of movements across the Turkey-Syria border.Turkey
[ ] Ochwo et al. (2019) ObC
RiskF.
To provide additional epidemiological information on LSD by estimating the herd and animal-level seroprevalence, and risk factors for seropositivity in herds with no history of vaccination, in the four major geographical regions of Uganda.Blood was collected between July 2016 and August 2017, in Uganda districts; samples were screened by indirect ELISA for the presence of Abs against LSDV. The following herd characteristics were considered: cattle’s sex, age and breed, type of management, mean annual rainfall, region, contact with buffaloes, communal water source, newly introduced cattle, contact with wildlife and herd size.
The authors applied multivariate logistic regression models.
The multivariate logistic regression model showed that pastoral and shared pastures, as well as fenced farms, were significantly associated with LSDV seropositivity. Other risk factors were: mean annual rainfalls of 1001–1200 mm and 1201–1400 mm, female cattle, age > 25 months and 13–24 months, and drinking from communal water sources.Specific regions of Uganda were focused on in the study. This study relied on seroprevalence, by using an ELISA test to detect Abs against Capripoxviruses. It would have been useful to include the presence/absence of goats or sheep on and near the farms. Regarding the ‘communal’ water sources, the authors did not specify what they meant by ‘communal’. They did not explain either why they considered only herds with ≥ 20 cattle. The history of vaccination against LSD was included, which could give false positives to the serology test and over-estimate the prevalence, and finally the affect the results of the final model. Uganda
[ ] Hailu et al. (2014) ObC
RiskF.
To estimate herd-level prevalence of LSD, and to assess the risk factors associated with the disease in Ethiopia; LSD is one of the major livestock disease problems in that country. Questionnaires were carried out on affected Ethiopian farms between October 2012 and February 2013. The questionnaire was designed to ascertain the presence of LSD based on the farmer’s ability to recognize LSD clinical signs; it also gathered information on herd size, cattle age structure and management practices. The approach aimed at assessing the epidemiological factors associated with LSD in the previous two years. A multivariate logistic regression was carried out; the odds ratios of the potential risk factors of LSD occurrence were estimated. The risk factors of LSD occurrence were: herd size (>22 animals), use of shared pastures and watering points, introduction of a new animal in the herd. Given that the characteristics of local management practices cannot be readily changed, disease control should rely on a greater use of effective LSD vaccines.No sample size was determined. Herds were randomly selected but included in the study based on herd owner’s willingness to complete the questionnaire. The LSD status was determined by the farmer’s ability to recognize clinical signs associated with the disease. Although, the authors tried to account for it by recording commonly occurring skin diseases of cattle in the study areas: they were recorded from the district veterinary clinic for the differential diagnoses and by crosschecking whether the herd owner correctly related the disease event with the clinical signs of LSD. The possibility of error in detecting LSD signs or not would have affected the number of positive animals. Vaccination status was included.Ethiopia
[ ] Issimov et al. (2022)ObC
RiskF.
To determine the prevalence of LSD, at individual and herd levels, and risk factors of LSD in West Kazakhstan.The authors developed a questionnaire to assess the magnitude of LSD occurrence (based on the observation of clinical signs by the farmer) and associated risk factors. They considered herd size, breed, contact with other domestic animals, year and month of LSD occurrence and herd management (feeding and watering management, animal movement, vaccination, treatment).
Multivariate logistic regression models were used to investigate the potential risk factors.
At animal level, the factors associated with LSD outbreaks included: medium and large herd size, purchase of animals and the sale of animals during an LSD outbreak. Herd management system had not altered after the outbreak. Therefore, the implementation of nationwide training programs is essential to improve the preparedness and awareness of farmers and veterinary personnel to control future emerging diseases.The authors only considered farms located in west Kazakhstan. The categorization of farms, i.e., LSD-affected or not, relied on the presence or absence of LSD-affected animals in the farm. A farm was considered as affected if clinical signs characteristic of LSD were observed in at least one animal of the herd. This could have affected the number of true positives to LSD, as reporting the farm as positive or negative relied on the cattle holder’s observation only.Kazakhstan
[ ] Gari et al. (2010)ObC
RiskF.
To address important knowledge gaps regarding the magnitude of LSD occurrence in different agro-climatic conditions and to identify associated risk factors.The authors developed a questionnaire to gather the following information: year and month of LSD occurrence (LSD identified by the farmer), number, sex and age of affected animals that subsequently died, herd management (i.e., sedentary/transhumant farming system), herd size, vaccination against LSD, management of grazing/watering points, contacts with sheep and goats and introduction of new animals. The peak of biting fly activity (months) was observed and recorded. Data related to LSD occurrence in the study area and countrywide, as well as annual rainfall for the period 2000–2007 were registered as well.
A multivariate logistic regression model was used, based on LSD occurrence at herd level.
The odds ratios of LSD occurrence in midland vs. highland, and in lowland vs. highland, were 3.86 (95% CI = 2.61–5.11) and 4.85 (95% CI = 2.59–7.1), respectively. A significantly higher risk of LSD occurrence was associated with communal grazing and watering management, as well as with the introduction of new cattle. No sample size was calculated. The classify animals as LSD positive authors used farmers’ reports, reports from the district agricultural office documentation, and the national disease outbreak report database. All reporting systems were based on observations of clinical signs, The study used a crosschecking validation on clinical signs and described the disease to account for such bias. However, there is still the issue of confirming the true LSD status as it relied on the farmers’ ability to recognize clinical signs of LSD; the signs could be confounded with other co-morbidities. Vaccination status was not taken into account, which could have affected the risk factors.Ethiopia
[ ]Odonchimeg et al. (2022) ObC
RiskF.
To investigate the current LSD outbreak in Mongolia to determine the prevalence and identify potentially associated risk factors.The authors developed a questionnaire to gather the following information: general knowledge of LSD, herd’s proximity to water sources, vector activity, and water source, among others.
Samples of suspected clinical cases were obtained. Cattle skin nodules were collected and submitted to PCR, virus isolation, DNA sequencing and histopathology. A phylogenetic analysis was also performed.
Data were submitted to a multivariate logistic regression analysis.
In the multivariate model, females showed a significantly higher risk of LSD occurrence compared to males. On the contrary, adult animals, young cattle and locations near a tube well and pond (vs. near a river) were protecting factors. The authors did not describe the study design. They did not calculate the sample size nor explained the sampling methodology (e.g., random, selected herd). The questionnaire was not described and the risk factor that were taken into account were not listed. Only factors which were significant in the univariate model. Only suspected clinical cases of LSD were sampled, thus could be an underestimation of cases. Locations of the farm near the tube well, pond or river, were used as risk factors but the distance classified as ‘near’ was never specified.Mongolia
[ ]Hasib et al. (2021)ObC
RiskF.
To confirm LSD occurrence based on clinical, molecular and pathological identification and to unveil the plausible risk factors of LSDV infection in a region of Bangladesh.The authors developed a questionnaire to collect demographic data on farms with suspected cases of LSD, i.e., breed, age, sex, and management practices such as source of water supply). A case was considered as LSD positive when an animal showed two or more defined clinical signs. Biopsy of nodular lesions was performed on sick or suspicious cattle, for confirmation; PCR, nucleotide sequencing and phylogenetic analysis were conducted on positive samples. Prevalence maps and multivariate logistic models were obtained. A total of 19 farms, accounting a total of 3327 animals, were considered. Out of those, 120 were deemed as sick or suspected, and skin biopsies were collected from nodular lesions. The final multivariate model revealed that only foreign breeds and females were at higher risk. Sampling was performed on suspect animals. Although no sampling size nor methodology were described, a large number of farms and animals were included in the study. Cattle were physically examined and farmers interviewed. Biopsy was taken only on suspect or clinically affected animals. No animal was considered as vaccinated, as it was a new outbreak in Bangladesh.Bangladesh
[ ]Molla et al. (2018) ObC
RiskF.
To estimate the seroprevalence, to identify and quantify the risk factors contributing to the occurrence of LSD.Sampling was performed in different regions. Antibody neutralization test detected Abs against LSDV. Herd level sensitivity and specificity were calculated.
The variables included in the multivariate logistic regression model were: altitude (<2000/2000–2400/>2400 m above sea level), contact with other animals (yes/no), free animal movements (yes/no), presence of water bodies (river/pond/lake/damp swampy/irrigated lands) (yes/no), animal trade route in the study area (yes/no) and animal characteristics (breed, age and sex). Animals were categorized as calf (0.5–1 year), young (1–4 years old) and adult (≥4 years old); breeds were Holstein-Frisian cross and local Zebu.
A total of 2386 serum samples were collected. Generally, cattle population accounting many adults and that live in wet areas were at higher risk, whereas cattle in frequent contact with other cattle and other animal species had a lower risk, potentially due to a dilution effect of vectors.
The final multivariate model identified the age as a risk factor, with animals aged 1–4 years old and ≥4 years were more at risk, compared to cows aged 6 months to 1 year old.
Contacts with other animal species were protective. The presence of water bodies was a risk factor also.
The study focused on the central and north western parts of Ethiopia. The limitation of seroprevalence is that it cannot determine which Capripoxvirus causes the immune response. The authors did not consider the vaccination status in the analysis. The high number of serum samples ensured a robust estimation of the prevalence. Ethiopia
[ ] Machado et al. (2019)ObC
RiskF.
To identify factors associated with 2014–2016 LSDV outbreaks and explore geographic areas at-risk, based on potential ecologically favorable conditions and the spatiotemporal dynamics of the disease.Ecological niche modelling and fine spatiotemporally explicit Bayesian hierarchical model were applied to 2014–2016 LSDV outbreak data, from Middle Eastern , Central Asian and Eastern European countries. The outbreak database contained information on the geographical coordinates, date of occurrence, and numbers of susceptible and infected animals per herd. Several independent variables influenced the spatiotemporal variability of LSDV. A risk was positively associated with precipitation and temperature, and negatively affected by wind. A contradiction and unresolved debate is the role of wind in the spread of the virus or via potential vectors, such as S. calcitrans. Authors found a negative effect of wind speed, i.e., the risk of LSDV would be reduced when winds are stronger. They also identified temperature as a factor increasing the relative risk of LSDV. Land cover may play a role in determining the risk.The study covered a large geographic area, ignoring administrative boundaries, and instead, used a grid cell construction based on previous studies that estimated the distances over which LSDV could spread.Middle East
Central Europe and Asia
[ ] Alkhamis and VanderWaal (2016)ObC
RiskF.
To characterize the spatial-temporal dynamics of LSDV in Middle Eastern countries and to assess whether environmental and demographic variables could predict the geographic distribution of LSDV outbreaks reported in these countries between 2012 and 2015.The authors used a maximum entropy ecological niche modelling method. They assessed multiple effective reproductive numbers to assess the transmission potential and efficacy of control and prevention measures during the epidemic that occurred in Middle Eastern countries.
Outbreak data from July 2012 to May 2015. The following environmental variables were included in the ecological niche model: climate variables, cattle/buffalo/sheep and goat density, global land cover and type ofl livestock production system. The following climatic variables were added to the model: monthly average, minimum and maximum temperatures, monthly rainfalls and altitude.
The most important environmental predictors that contributed to the ecological niche of LSDV included: annual rainfalls, land cover, average diurnal range temperature, type of livestock production system, and global livestock densities. Average monthly effective reproductive number (R-TD) was 2.2 (95% CI: 1.2–3.5), whereas the largest R-TD was estimated in Israel (R-TD = 22.2 (95% CI: 15.2–31.5) in September 2013, which indicated that the demographic and environmental conditions during this period were suitable to LSDV super-spreading events.When using such approach to infer spatial patterns of infection risk, it is important to remember that there is no single ‘true’ model that predicts the risk across all contexts. Indeed, environmental factors contributing to the risk may differ across space and time. Authors did acknowledge that results might differ according to the input dataset. However, it also allowed the identification of spatial and environmental patterns that are consistent, regardless of the input dataset. The identified environmental predictors matched those identified in the literature, but it is important to consider that the resulting risk maps for LSDV occurrence are not definitive and need to be updated periodically as new data emerge. Thus, in the event of future epidemics, these analyses need to be repeated and refined in order to be subsequently used in surveillance, control, and prevention strategies. Middle East
[ ]Ardestani et al. (2020) ObC
RiskF.
To assess the relationship between 2012 and 2016 LSDV outbreaks and environmental variables, in order to identify the most important environmental variables; to produce a distribution map of LSDV outbreaks in certain Iran areas, in order to determine at risk-areas based on potential ecologically-desirable conditions.The authors used data on 2012–2016 LSDV outbreaks in Iranian provinces. For each LSDV outbreak, the database included information on its geographical coordinates (latitude and longitude), time data (month, season and year), social and political divisions of locations, type of herd, total number of farms, number of examined and affected animals and number of dead animals recorded. Ecological niche models were applied to data.Rainfalls of the wettest period and the coldest season, as well as isothermality, were the bioclimatic variables explaining LSD prevalence. Coexistence of specific weather conditions, including defined humidity and temperature, is necessary for an LSD outbreak.Although the authors present a fast and accurate approach to model the probability of LSDV, it is only worth for a specific area of Iran. Thus, inferences derived from this model need to be interpreted with caution.Iran
[ ]Allepuz et al. (2019) ObC
RiskF.
To analyze and identify the association between the LSD outbreaks reported in Turkey, Russia, the Balkans and Israel, with climatic variables, land cover, and cattle density in order to predict the risk of LSD spread in neighboring free-countries of Europe and Central Asia. The following data were added to the model: LSD outbreak locations, date of occurrence, geographical coordinates, animals at risk and animals clinically sick and dead. These data were gathered between July 2012 and December 2018 in the Balkans , Caucasus and Middle East . The following variables, i.e., density of cattle, land cover and climate, were included in spatial regression models.The results showed a significant effect of land cover on the occurrence of an LSD outbreak: areas at risk were mostly croplands, grassland, or shrub land. Cattle density, as well as areas with higher annual average temperature and higher diurnal range of temperatures, were also identified as risk factors.Data used for this study relied mostly on passive reports of the veterinary services from the countries included in the analysis. The use of passive surveillance data has its limitations as cases or outbreaks could be underreported. This should be considered when interpreting the results.Balkans
Caucasus
Middle East
[ ]Molla et al. (2017) ObC
RiskF.
To evaluate the spatial and temporal distribution of LSD outbreaks and to forecast future patterns of outbreaks in Ethiopia, based on data reported over the 2000–2015 period.The authors used data of Ethiopian LSD outbreaks that occurred between the years 2000 and 2015. The records contained monthly information on place, time, and number of cases, deaths and animals at risk. The geographical distribution of LSD outbreaks over the 16 years was mapped, per administrative zone, using a geographic information system (GIS) software. The spread of the epidemic was also shown using SPMAP programs. Monthly average rainfalls for the period 1999–2013 were considered as well. Three seasons exist in Ethiopia (a) February to May, (b) June to September and (c) October to January, which registers the highest rainfall. Time series analysis and spectral analysis were conducted to detect seasonality and cyclical patterns in the LSD outbreak time series.The highest LSD incidences were registered in warm and humid highlands, while the lowest occurred in hot and dry lowland areas. The regions receiving relatively high rainfalls for a reasonable period are conducive to the replication and survival of blood-feeding arthropods and thus, to the spread of the disease. The occurrence of LSD outbreaks was seasonal, with a peak registered in October and the lowest number in May and at the end of the long rainy season.
Additionally, LSD outbreaks do not occur at random over time: authors demonstrated the seasonality by spectral analysis. The seasonal variation of LSD outbreaks might be related to the variation in temperatures and rainfalls between seasons, leading to variable arthropod densities in the environment.
The presence of a long-term trend or season effect was determined only by simple examination of the graph, no statistical analysis was conducted to assess statistical significance.
The existence of al long-term trend in LSD outbreaks was modelled by linear regression and using the number of LSD outbreaks (or trend component of the outbreak).
Authors establish the limitations of the model that does not consider the correlation between successive values of the time series. This means one can only gain advantage of using short-term forecasts. Additionally, the wide confidence interval indicates the need of frequent updating of the model by incorporating the latest outbreak reports.
Ethiopia
[ ] Molla et al. (2017)ObC
RiskF.
To better understand the dynamics of LSDV outbreaks and to quantify transmission rate and reproductive ratio (R0) between animals.The transmission parameters relied on a susceptible-infectious recovered (SIR) epidemic model with environmental transmission, and estimated using generalized linear models. The survival rate of infectious virus in the environment equaled 0.325 per day, based on the best-fitting statistical model. The daily transmission rate between animals reached 0.071 (95% CI = 0.068–0.076) in the crop-livestock production system and 0.076 in the intensive production system (95% CI = 0.068–0.085). The R0 of LSD between animals was 1.07 in the crop-livestock production system and 1.09 in the intensive production system. These R0’s provides a baseline to assess the efficacy of various control options.The daily transmission rates of crop livestock systems and intensive systems did not differ significantly. That suggests that the knowledge of these parameters alone is not sufficient to predict the risk of LSD in the different production systems.Ethiopia
[ ] Mercier et al. (2018) ObC
RiskF.
To estimate the LSDV spread rate for a further use in risk analysis of LSDV introduction in other European countries.LSD outbreaks were mapped according to their geographical coordinates. Study time period ranged from the date of the first occurrence, in May 2015 (western Turkey), to August 2016.
Outbreak mapping and thin plate spline regression models were used.
The frequency of outbreaks was highly seasonal, with little or no transmission in the winter period. The skewed distribution of spread rates suggested two distinct underlying epidemiological processes, i.e., (i) local and distant spread possibly related to vectors and (ii) cattle trade movements. Low spread rates were probably related to local LSDV transmission by infected arthropods and contacts between infected and naïve cattle, covering small daily distances. On the other hand, high spread rates might be related to the movements of infected animals between farms trade, to/from cattle markets or to slaughterhouses.This analysis considered only the outbreaks reported up to the end of August 2016, and did not include all Albanian outbreaks; 2323 out of 3585 outbreaks occurred after this date. In addition, the analysis implicitly includes the impact of stamping out infected herds on the rate of spread, which was implemented in all affected countries except Albania. Although unavoidable, the maximum spread rate due to possible under- or delayed reporting is probably unstable. Vaccination campaigns must have strongly influenced the spread of the disease and vaccination data were not incorporated in the model.Balkans
[ ]Gubbins et al. (2020)ObC
RiskF.
To explore how the force of infection depends on the distance between non-infected and infected herds, to assess evidence for seasonality in the force of infection and to estimate the impact of vaccination on the spread of LSDV.The authors used LSD outbreak data from Albania collected in 2016. A kernel-based approach described the transmission of LSDV between herds. In this approach, all transmission routes were combined in a single generic mechanism with the probability of transmission from an infected to a non-infected herd assumed to depend on the distance between them (i.e., the transmission kernel).It was shown that most of the transmission occurred over short distances (<5 km), but with an appreciable probability of transmission over longer distances. The authors evidenced a seasonal variation in the force of infection associated with temperature, possibly through its influence on the relative abundance of the stable fly S. calcitrans. Both results are consistent with a transmission of LSDV by the bites of blood-feeding insects, though further work is required to incriminate the vector species.The approach of combining all transmission routes into a single generic mechanism, and the assumption of susceptibility of an uninfected herd and the infectiousness of an infected herd to be both proportional to the number of cattle in the herd, could affect the kernel shape. Albania
[ ]Punyapornwithaya et al. (2022) ObC
RiskF.
To determine the spatio-temporal patterns of LSD outbreaks in dairy farms, in northeastern Thailand, in order to better understand the epidemiology of LSD outbreaks affecting dairy farms.An LSD case was defined as a dairy cow displaying LSD clinical signs. Blood was sampled to confirm an infection by LSDV. The following epidemiological data were collected: number of dairy cattle with LSD clinical signs, deaths with clinical signs and the number of all dairy cattle on the farms. The geographical coordinates of each farm were recorded. A spatio-temporal analysis using space-time permutation models, Poisson and Bernoulli models was performed.The authors concluded that, because there are few cattle movements between dairy farms, the spread of LSD was less likely due to close contacts between cattle from different farms. Furthermore, the spread of LSD was likely caused by insect vectors, which are abundant in most dairy farms in Thailand. Indeed, the finding that LSD outbreaks were located in a large number of farms and over a short period, and that several farms were concentrated in the area, suggests that LSDV was probably transmitted by insect vectors.The authors did not draw any direct conclusion from the model regarding the vector transmission. Spatial temporal patterns showed that several farms concentrated on the same were affected over a short period of time. This with the fact that there were few cattle movements among farms made authors reach the conclusion that it the spread was attributed to insect vectors. Thailand
[ ]Klausner et al. (2017) ObC
RiskF.
To examine the possibility of LSDV introduction in Israel, in 1989 and 2006, by long-distance wind-associated movements of infected vectors from Egypt. Israeli outbreaks were reported in August 1989 and on 7 June 2006. Backwards Lagrangian trajectories (BLTs) analysis was conducted. It consists in reconstructing the travelling path of an air parcel from its source to a given receptor. These trajectories are calculated using the re-analysis of available meteorological fields as inputs. Synoptic systems climatologically associated with the period preceding the outbreaks were identified, along with typical atmospheric transport routes during the synoptic systems. Three-dimensional backwards Lagrangian trajectories (BLTs) were calculated using the hybrid single-particle Lagrangian integrated trajectory model.At the first stage, the relevant synoptic systems that allowed wind transport from Egypt to Israel during the 3 months preceding each outbreak were identified. The analysis revealed several events in which atmospheric connection routes between the affected locations in Egypt and Israel were established. Specifically, in 1989, Damietta and Port Said stand out as likely sources for the outbreak in Israel. In 2006, different locations acted simultaneously as potential sources of Israeli outbreak. The analysis pointed out Sharav low and Shallow Cyprus, low to the North, to be the most likely systems to enable windborne transport from Egypt to Israel. These findings are of high importance to analyze the risk of transmission of vector-borne viruses in the eastern Mediterranean region.The study only considered Israel. The difficulty in conducting such type of analysis stems from the uncertainty regarding the exact arrival of the virus on the receptor site (in this case, Israel). Although authors concluded that winds could have carried the infected vector from Egypt to Israel, the vectors (in this case Stomoxys) competency was not mentioned, and thus atmospheric travel under dry conditions is possible but not ideal for the survival of the flies. Hence, a doubt remains on the viability of such route of spread. Israel
[ ]Horigan et al. (2018) Risk A.
QL
The qualitative assessment focused on the probability of LSDV introduction in the UK, between June 2017 and June 2018, and the probability of onward transmission in the country.A qualitative risk assessment was conducted. The approach was based on the framework set out by the OIE. The risk questions to be addressed were: (a) what is the probability of introduction of LSDV in the UK within the next year? (b) what is the probability of onward transmission of LSDV in the UK; could it be introduced within the next year? The following risk pathways of introduction were considered: infected live animals legally/illegally imported in the UK, contaminated animal products legally/illegally imported and infected vector imported to the UK.The overall risk of potential introduction and further onward transmission of LSDV was “very low” through livestock, but with a “high” probability of onward transmission. The risk of introduction was considered ‘very low’ via vectors, but the probability of onward transmission was ‘high’. Exotic animals, germplasm, hides/skins, meat and milk products were negligible for both probabilities.The study conducted the risk assessment of entry using and describing the correct guidelines. As any other qualitative risk assessments, it depends on the knowledge of the experts who conducted the categorization. United Kingdom
[ ]Gale et al. (2016) Risk A.
QL
A qualitative assessment of the risk of importation of one infected product (i.e., skin/hide or bale of wool) through legal trade into the UK.A qualitative risk assessment was conducted. The approach relied on the framework set out by the OIE. The specific risk question was: what is the probability that a whole skin/hide or bale of wool legally imported from a European Union Member State (MS) experiencing an ongoing outbreak is infected with capripoxvirus at the point of entry into the UK?The predicted risk of importation of LSD virus per cattle hide/skin was also low (assuming LSD was to emerge in a EU MS with similar herd prevalence to sheep and goat pox in 2013/14 in Greece). The amount of LSDV on an infected cow hide, if imported, may be very low. It is recommended to recalculate the risks of entry for capripoxviruses if outbreaks occur elsewhere within the EU.The risk assessment used the correct guidelines. Given that only EU Member States were considered in the analysis, the risk assessment is most likely to give a low risk.United Kingdom
[ ]Farra et al. (2021)Risk A.
QL
A qualitative risk assessment was conducted, with the aim (i) to investigate the probability of LSDV introduction in Ukraine and, (ii) if introduced, the probability of onward transmission in the country within the next year.A qualitative risk assessment was conducted. The approach relied on the framework set out by the OIE Handbook on Import risk analysis. The overall questions of the risk assessment were: (a) probability for LSDV to be introduced in Ukraine within the next year; (b) if LSDV is introduced in Ukraine, probability of onward transmission in the country within the next year; (c) risk pathways, i.e., cattle, wild ruminants, semen, embryos, biomaterials, skin, hides, trophies, meat, milk and vectors.The illegal trade of cattle was considered the highest risk of LSD introduction. However, the probability was estimated to be low. When assessing the probability of an animal to be exposed to the virus and responsible for the further transmission in Ukraine, a high probability was estimated for flying vectors. The risk assessment was very complete in using all the risk pathways with the right guideline. The study was described very well. The limitations are similar to any qualitative risk assessment, i.e., it relies on the knowledge of experts. Ukraine
[ ]Saegerman et al. (2018)Risk A.
QT
In order to estimate, for France, the threat of introduction of vectors through animal trucks (cattle or horses) coming from at-risk countries (Balkans and neighboring countries), a quantitative import risk analysis (QIRA) model was developed according to the international standard.The authors used a stochastic model to assess the probability of importing cattle from an at-risk area, that can be infected with LSDV before its detection.
They also estimated the probability that trucks come from an infected farm located in the at-risk area and the probability of an animal to be infected already in the farm but without clinical signs. The authors also considered the probability of the virus surviving in Stomoxys spp. and the probability that Stomoxys spp. would survive during transport (survival of the fly was estimated at 2–3 days).
The authors used stochastic QIRA modelling and combined experimental/field data and expert opinion. The yearly risk of LSDV being introduced by stable flies (S. calcitrans) travelling in animal trucks was between 6 × 10 and 5.93 × 10 with a median value of 89.9 × 10 ; it was mainly due to the risk related to insects entering farms in France from vehicles transporting cattle from the at-risk area. The risk related to the transport of cattle going to slaughterhouses or the transport of horses was much lower (between 2 × 10 and 3.73 × 10 and between 5 × 10 and 3.95 × 10 for cattle and horses, respectively). The disinsection of trucks transporting live animals is important to reduce this risk.Authors mentioned the limitation of the QIRA modelling which were related to the choice of assumptions and worst case scenarios (proportion of infected Stomoxys equivalent to the proportion of contagious cattle, absence of cleaning, disinfection and disinsectisation of the truck used for the transport of animals, absence of unloading of animals during transport, only Stomoxys calcitrans considered as mechanical vector of LSDV, proportion of mixed cattle and equine activities in countries of origin unknown and consequently estimated at the same as in France, and probability of infecting cattle on the destination farm of 100%.France
[ ]ANSES, 2017Risk A.
QL
To assess the risk of LSD introduction in France.The authors assessed the risk of LSD introduction in France taking into account the different risk factors of introduction; The probability was ‘only’ for the probability of a first LSD outbreak on the French territory for the specific year of when the study was conducted and it was based on the epidemiological situation of LSD in January 2017, according to the exiting European regulations at that date and using trade data of the year 2016. An assessment of the risk of a first LSD outbreak in France was performed, depending on the different virus sources and their possible ways of introduction (live animals and their products—semen and embryos, vectors, inert media, etc.). The risk assessment was carried out according to a quantitative approach for the introduction pathways considered by the experts as most likely (movements of animals, movements of arthropod vectors,); in the other cases, the approach was qualitative. Only animals from EU at-risk areas (MS that reported outbreaks) were taken into account in the analysis. The probability of LSD introduction by live animals was limited to the risk of introduction by live cattle.
The quantitative probability of a first LSD outbreak in France following the introduction of infected live cattle was estimated between 0.004% and 0.32% (95% CI), which corresponds to an ‘extremely low to low’ qualitative probability (3 to 5 on AFSSA 2008 scale, which ranges from 0 to 9). The probability of a first LSD outbreak in France following the introduction of infected live cattle for the slaughterhouse is therefore estimated to be null.
The risk of LSD introduction by long-distance road transports of vectors is limited to the risk of introduction by Stomoxys spp. The quantitative probability of a first LSD outbreak in France following the introduction of infective vectors transported with live cattle was therefore estimated between 0.002% and 0.44% (95% CI), which corresponds to an ‘extremely low to low’ qualitative probability (3 to 5 on AFSSA scale). The probability of introduction via other modes was considered as null.
The qualitative risk assessment was very thorough. Not only experts’ opinion was used but also quantitative data regarding cattle and horse entering France, which gave a more certain assessment. France
[ ]Saegerman et al. (2019)Risk A.
QT
To assess the risk of LSD introduction through cattle imports.In order to estimate the threat for France, a QIRA model was developed to assess the risk of LSD introduction in France through cattle imports.Based on available information, and using a stochastic model, the probability of a first LSD outbreak in France, following the import of batches of infected live cattle for breeding or fattening, was estimated at 5.4 × 10 (95% probability interval [PI]: 0.4 × 10 ; 28.7 × 10 ) in summer months (during high vector activity) and 1.8 × 10 (95% PI: 0.14 × 10 ; 15 × 10 ) in the winter.The QIRA model depends on the available data and information on live-animal trade between European countries other than France (in particular between infected countries and countries bordering France) was not available for the French experts.France
[ ]Ince et al. (2016)Risk A.
QL
To assess the epidemiology of LSD, its transmission mechanisms, and the potential role of risk factors. Qualitative estimates of the risk, spatial variation in risk, and the factors associated with the risk of LSD introduction and spread in animal markets are a prerequisite for developing specific policies to prevent or control epidemics.The authors performed a qualitative risk assessment. The approach relied on the framework set out by the OIE Handbook on Import risk analysis. The risk question was the probability of cattle with LSD being introduced to the animal market? The farms with reported outbreaks were observed by a veterinarian, who examined any suspect animal. The risk estimation and management were carried out. Two risk pathways were identified, i.e., (1) probability of cattle to be exposed to LSDV from seasonal migration, and (2) probability of exposing cattle to LSD through veterinary equipment.The risk (probability) of a farm being infected was estimated as ‘medium to high’, such as the risk (probability) of an animal being infected on a farm. The risk (probability) of not detecting LSD in non-certified and infected cattle was ‘high, such as the risk (probability) of LSD introduction to non-infected provinces through animal movements and the risk (probability) of cattle to be exposed to LSDV from seasonal migration. Finally, the risk (probability) of exposing cattle to LSD through veterinary equipment was estimated as ‘medium’.The release assessment categories were not clearly detailed in the results. The same limitation as above, regarding qualitative risk assessments, apply. Turkey
[ ]Taylor et al. (2019)Risk A.
QT
To provide a generic framework for quantitative risk assessment of disease introduction using LSD as a case study.The authors created a generic framework, i.e., they defined the risk of infection as the probability of one or more initial infections in the native susceptible population in a specific area. Then the framework was applied to a single pathway using LSD as a case study (2016-outbreak in the Balkans). The risk assessment was performed on three spatial scales, i.e., countries, regions and individual farms.Croatia (assuming no vaccination occurred) had the highest mean probability of infection, beating out Italy, Hungary and Spain. The detection of infected cattle at importation does reduces the risk, but proportionally lower for countries with the highest risk. The results were consistent across the spatial scales, while in addition, at the finer spatial scales, specific areas or individual locations on which to focus surveillance were identified.Only a single pathway of introduction was used, i.e., the number of cattle traded within the EU, and on the basis of LSD prevalence in the country of origin of cattle. Thus, results are conditioned by the prevalence of LSD in the EU.Europe
  • House, J.A.; Wilson, T.M.; El Nakashly, S.; Karim, I.A.; Ismail, I.; El Danaf, N.; Moussa, A.M.; Ayoub, N.N. The isolation of lumpy skin disease virus and bovine herpesvirus-4 from cattle in Egypt. J. Vet. Diagn. Investig. 1990 , 2 , 111–115. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Yeruham, I.; Nir, O.; Braverman, Y.; Davidson, M.; Grinstein, H.; Haymovitch, M.; Zamir, O. Spread of lumpy skin disease in Israeli dairy herds. Vet. Rec. 1995 , 137 , 91–93. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Davies, F.G. Lumpy skin disease of cattle: A growing problem in Africa and the Near East. World Anim. Rev. 1991 , 68 , 37–42. [ Google Scholar ]
  • Kumar, S.M. An Outbreak of Lumpy Skin Disease in a Holstein Dairy Herd in Oman: A Clinical Report. Asian J. Anim. Vet. Adv. 2011 , 6 , 851–859. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Tageldin, M.H.; Wallace, D.B.; Gerdes, G.H.; Putterill, J.F.; Greyling, R.R.; Phosiwa, M.N.; Al Busaidy, R.M.; Al Ismaaily, S.I. Lumpy skin disease of cattle: An emerging problem in the Sultanate of Oman. Trop. Anim. Health Prod. 2014 , 46 , 241–246. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • WAHIS. World Animal Health Information System. 2022. Available online: https://wahis.woah.org/#/home (accessed on 10 January 2023).
  • Page, M.J.; McKenzie, J.E.; Bossuyt, P.M.; Boutron, I.; Hoffmann, T.C.; Mulrow, C.D.; Shamseer, L.; Tetzlaff, J.M.; Akl, E.A.; Brennan, S.E.; et al. The PRISMA 2020 statement: An updated guideline for reporting systematic reviews. BMJ 2021 , 372 , n71. [ Google Scholar ] [ CrossRef ]
  • Chihota, C.M.; Rennie, L.F.; Kitching, R.P.; Mellor, P.S. Attempted mechanical transmission of lumpy skin disease virus by biting insects. Med. Vet. Entomol. 2003 , 17 , 294–300. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Sanz-Bernardo, B.; Haga, I.R.; Wijesiriwardana, N.; Basu, S.; Larner, W.; Diaz, A.V.; Langlands, Z.; Denison, E.; Stoner, J.; White, M.; et al. Quantifying and Modeling the Acquisition and Retention of Lumpy Skin Disease Virus by Hematophagus Insects Reveals Clinically but Not Subclinically Affected Cattle Are Promoters of Viral Transmission and Key Targets for Control of Disease Outbreaks. J. Virol. 2021 , 95 , e02239-20. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Sanz-Bernardo, B.; Suckoo, R.; Haga, I.R.; Wijesiriwardana, N.; Harvey, A.; Basu, S.; Larner, W.; Rooney, S.; Sy, V.; Langlands, Z.; et al. The Acquisition and Retention of Lumpy Skin Disease Virus by Blood-Feeding Insects Is Influenced by the Source of Virus, the Insect Body Part, and the Time since Feeding. J. Virol. 2022 , 96 , e0075122. [ Google Scholar ] [ CrossRef ]
  • Gubbins, S. Using the basic reproduction number to assess the risk of transmission of lumpy skin disease virus by biting insects. Transbound. Emerg. Dis. 2019 , 66 , 1873–1883. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Issimov, A.; Kutumbetov, L.; Orynbayev, M.B.; Khairullin, B.; Myrzakhmetova, B.; Sultankulova, K.; White, P.J. Mechanical Transmission of Lumpy Skin Disease Virus by Stomoxys Spp ( Stomoxys calsitrans , Stomoxys sitiens , Stomoxys indica ), Diptera: Muscidae. Animals 2020 , 10 , 477. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Issimov, A.; Taylor, D.B.; Shalmenov, M.; Nurgaliyev, B.; Zhubantayev, I.; Abekeshev, N.; Kushaliyev, K.; Kereyev, A.; Kutumbetov, L.; Zhanabayev, A.; et al. Retention of lumpy skin disease virus in Stomoxys spp ( Stomoxys calcitrans , Stomoxys sitiens , Stomoxys indica ) following intrathoracic inoculation, Diptera: Muscidae. PLoS ONE 2021 , 16 , e0238210. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Paslaru, A.I.; Verhulst, N.O.; Maurer, L.M.; Brendle, A.; Pauli, N.; Vögtlin, A.; Renzullo, S.; Ruedin, Y.; Hoffmann, B.; Torgerson, P.R.; et al. Potential mechanical transmission of Lumpy skin disease virus (LSDV) by the stable fly ( Stomoxys calcitrans ) through regurgitation and defecation. Curr. Res. Insect Sci. 2021 , 1 , 100007. [ Google Scholar ] [ CrossRef ]
  • Sohier, C.; Haegeman, A.; Mostin, L.; De Leeuw, I.; Campe, W.V.; De Vleeschauwer, A.; Tuppurainen, E.S.M.; van den Berg, T.; De Regge, N.; De Clercq, K. Experimental evidence of mechanical lumpy skin disease virus transmission by Stomoxys calcitrans biting flies and Haematopota spp. horseflies. Sci. Rep. 2019 , 9 , 20076. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Chihota, C.M.; Rennie, L.F.; Kitching, R.P.; Mellor, P.S. Mechanical transmission of lumpy skin disease virus by Aedes aegypti (Diptera: Culicidae). Epidemiol. Infect. 2001 , 126 , 317–321. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Paslaru, A.I.; Maurer, L.M.; Vögtlin, A.; Hoffmann, B.; Torgerson, P.R.; Mathis, A.; Veronesi, E. Putative roles of mosquitoes (Culicidae) and biting midges ( Culicoides spp.) as mechanical or biological vectors of lumpy skin disease virus. Med. Vet. Entomol. 2022 , 36 , 381–389. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Rouby, S.R.; Hussein, K.H.; Aboelhadid, S.M.; El-Sherif, A.M. Role of rhipicephalus annulatus tick in transmission of lumpy skin disease virus in naturally infected cattle in Egypt. Adv. Anim. Vet. Sci. 2017 , 5 , 185–191. [ Google Scholar ] [ CrossRef ]
  • Tuppurainen, E.S.; Venter, E.H.; Coetzer, J.A.; Bell-Sakyi, L. Lumpy skin disease: Attempted propagation in tick cell lines and presence of viral DNA in field ticks collected from naturally-infected cattle. Ticks Tick-Borne Dis. 2015 , 6 , 134–140. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Lubinga, J.C.; Clift, S.J.; Tuppurainen, E.S.; Stoltsz, W.H.; Babiuk, S.; Coetzer, J.A.; Venter, E.H. Demonstration of lumpy skin disease virus infection in Amblyomma hebraeum and Rhipicephalus appendiculatus ticks using immunohistochemistry. Ticks Tick-Borne Dis. 2014 , 5 , 113–120. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Lubinga, J.C.; Tuppurainen, E.S.; Coetzer, J.A.; Stoltsz, W.H.; Venter, E.H. Evidence of lumpy skin disease virus over-wintering by transstadial persistence in Amblyomma hebraeum and transovarial persistence in Rhipicephalus decoloratus ticks. Exp. Appl. Acarol. 2014 , 62 , 77–90. [ Google Scholar ] [ CrossRef ]
  • Lubinga, J.C.; Tuppurainen, E.S.; Coetzer, J.A.; Stoltsz, W.H.; Venter, E.H. Transovarial passage and transmission of LSDV by Amblyomma hebraeum , Rhipicephalus appendiculatus and Rhipicephalus decoloratus . Exp. Appl. Acarol. 2014 , 62 , 67–75. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Lubinga, J.C.; Tuppurainen, E.S.; Mahlare, R.; Coetzer, J.A.; Stoltsz, W.H.; Venter, E.H. Evidence of transstadial and mechanical transmission of lumpy skin disease virus by Amblyomma hebraeum ticks. Transbound. Emerg. Dis. 2015 , 62 , 174–182. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Tuppurainen, E.S.; Lubinga, J.C.; Stoltsz, W.H.; Troskie, M.; Carpenter, S.T.; Coetzer, J.A.; Venter, E.H.; Oura, C.A. Evidence of vertical transmission of lumpy skin disease virus in Rhipicephalus decoloratus ticks. Ticks Tick-Borne Dis. 2013 , 4 , 329–333. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Tuppurainen, E.S.; Lubinga, J.C.; Stoltsz, W.H.; Troskie, M.; Carpenter, S.T.; Coetzer, J.A.; Venter, E.H.; Oura, C.A. Mechanical transmission of lumpy skin disease virus by Rhipicephalus appendiculatus male ticks. Epidemiol. Infect. 2013 , 141 , 425–430. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Lubinga, J.C.; Tuppurainen, E.S.; Stoltsz, W.H.; Ebersohn, K.; Coetzer, J.A.; Venter, E.H. Detection of lumpy skin disease virus in saliva of ticks fed on lumpy skin disease virus-infected cattle. Exp. Appl. Acarol. 2013 , 61 , 129–138. [ Google Scholar ] [ CrossRef ]
  • Tuppurainen, E.S.; Stoltsz, W.H.; Troskie, M.; Wallace, D.B.; Oura, C.A.; Mellor, P.S.; Coetzer, J.A.; Venter, E.H. A potential role for ixodid (hard) tick vectors in the transmission of lumpy skin disease virus in cattle. Transbound. Emerg. Dis. 2011 , 58 , 93–104. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Osuagwuh, U.I.; Bagla, V.; Venter, E.H.; Annandale, C.H.; Irons, P.C. Absence of lumpy skin disease virus in semen of vaccinated bulls following vaccination and subsequent experimental infection. Vaccine 2007 , 25 , 2238–2243. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Irons, P.C.; Tuppurainen, E.S.; Venter, E.H. Excretion of lumpy skin disease virus in bull semen. Theriogenology 2005 , 63 , 1290–1297. [ Google Scholar ] [ CrossRef ]
  • Annandale, C.H.; Irons, P.C.; Bagla, V.P.; Osuagwuh, U.I.; Venter, E.H. Sites of persistence of lumpy skin disease virus in the genital tract of experimentally infected bulls. Reprod. Domest. Anim. 2010 , 45 , 250–255. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Annandale, C.H.; Holm, D.E.; Ebersohn, K.; Venter, E.H. Seminal transmission of lumpy skin disease virus in heifers. Transbound. Emerg. Dis. 2014 , 61 , 443–448. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Annandale, C.H.; Smuts, M.P.; Ebersohn, K.; du Plessis, L.; Thompson, P.N.; Venter, E.H.; Stout, T.A.E. Effect of using frozen-thawed bovine semen contaminated with lumpy skin disease virus on in vitro embryo production. Transbound. Emerg. Dis. 2019 , 66 , 1539–1547. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Annandale, C.H.; Smuts, M.P.; Ebersohn, K.; du Plessis, L.; Venter, E.H.; Stout, T.A.E. Effect of semen processing methods on lumpy skin disease virus status in cryopreserved bull semen. Anim. Reprod. Sci. 2018 , 195 , 24–29. [ Google Scholar ] [ CrossRef ]
  • Carn, V.M.; Kitching, R.P. An investigation of possible routes of transmission of lumpy skin disease virus (Neethling). Epidemiol. Infect. 1995 , 114 , 219–226. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Aleksandr, K.; Olga, B.; David, W.B.; Pavel, P.; Yana, P.; Svetlana, K.; Alexander, N.; Vladimir, R.; Dmitriy, L.; Alexander, S. Non-vector-borne transmission of lumpy skin disease virus. Sci. Rep. 2020 , 10 , 7436. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kononov, A.; Prutnikov, P.; Shumilova, I.; Kononova, S.; Nesterov, A.; Byadovskaya, O.; Pestova, Y.; Diev, V.; Sprygin, A. Determination of lumpy skin disease virus in bovine meat and offal products following experimental infection. Transbound. Emerg. Dis. 2019 , 66 , 1332–1340. [ Google Scholar ] [ CrossRef ]
  • Ma, J.; Yuan, Y.; Shao, J.; Sun, M.; He, W.; Chen, J.; Liu, Q. Genomic characterization of lumpy skin disease virus in southern China. Transbound. Emerg. Dis. 2021 , 69 , 2788–2799. [ Google Scholar ] [ CrossRef ]
  • Faris, D.N.; El-Bayoumi, K.; El-TaraBany, M.; Kamel, E.R. Prevalence and Risk Factors for Lumpy Skin Disease in Cattle and Buffalo under Subtropical Environmental Conditions. Adv. Anim. Vet. Sci. 2021 , 9 , 1311–1316. [ Google Scholar ] [ CrossRef ]
  • Şevik, M.; Doğan, M. Epidemiological and Molecular Studies on Lumpy Skin Disease Outbreaks in Turkey during 2014–2015. Transbound. Emerg. Dis. 2017 , 64 , 1268–1279. [ Google Scholar ] [ CrossRef ]
  • Selim, A.; Manaa, E.; Khater, H. Seroprevalence and risk factors for lumpy skin disease in cattle in Northern Egypt. Trop. Anim. Health Prod. 2021 , 53 , 350. [ Google Scholar ] [ CrossRef ]
  • Ince, O.B.; Türk, T. Analyzing risk factors for lumpy skin disease by a geographic information system (GIS) in Turkey. J. Hell. Vet. Med. Soc. 2019 , 70 , 1797–1804. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Ochwo, S.; VanderWaal, K.; Munsey, A.; Nkamwesiga, J.; Ndekezi, C.; Auma, E.; Mwiine, F.N. Seroprevalence and risk factors for lumpy skin disease virus seropositivity in cattle in Uganda. BMC Vet. Res. 2019 , 15 , 236. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Hailu, B.; Tolosa, T.; Gari, G.; Teklue, T.; Beyene, B. Estimated prevalence and risk factors associated with clinical Lumpy skin disease in north-eastern Ethiopia. Prev. Vet. Med. 2014 , 115 , 64–68. [ Google Scholar ] [ CrossRef ]
  • Issimov, A.; Kushaliyev, K.; Abekeshev, N.; Molla, W.; Rametov, N.; Bayantassova, S.; Zhanabayev, A.; Paritova, A.; Shalmenov, M.; Ussenbayev, A.; et al. Risk factors associated with lumpy skin disease in cattle in West Kazakhstan. Prev. Vet. Med. 2022 , 207 , 105660. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Gari, G.; Waret-Szkuta, A.; Grosbois, V.; Jacquiet, P.; Roger, F. Risk factors associated with observed clinical lumpy skin disease in Ethiopia. Epidemiol. Infect. 2010 , 138 , 1657–1666. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Odonchimeg, M.; Erdenechimeg, D.; Tuvshinbayar, A.; Tsogtgerel, M.; Bazarragchaa, E.; Ulaankhuu, A.; Selenge, T.; Munkhgerel, D.; Munkhtsetseg, A.; Altanchimeg, A.; et al. Molecular identification and risk factor analysis of the first Lumpy skin disease outbreak in cattle in Mongolia. J. Vet. Med. Sci. 2022 , 84 , 1244–1252. [ Google Scholar ] [ CrossRef ]
  • Hasib, F.M.Y.; Islam, M.S.; Das, T.; Rana, E.A.; Uddin, M.H.; Bayzid, M.; Nath, C.; Hossain, M.A.; Masuduzzaman, M.; Das, S.; et al. Lumpy skin disease outbreak in cattle population of Chattogram, Bangladesh. Vet. Med. Sci. 2021 , 7 , 1616–1624. [ Google Scholar ] [ CrossRef ]
  • Molla, W.; Frankena, K.; Gari, G.; Kidane, M.; Shegu, D.; de Jong, M.C.M. Seroprevalence and risk factors of lumpy skin disease in Ethiopia. Prev. Vet. Med. 2018 , 160 , 99–104. [ Google Scholar ] [ CrossRef ]
  • Machado, G.; Korennoy, F.; Alvarez, J.; Picasso-Risso, C.; Perez, A.; VanderWaal, K. Mapping changes in the spatiotemporal distribution of lumpy skin disease virus. Transbound. Emerg. Dis. 2019 , 66 , 2045–2057. [ Google Scholar ] [ CrossRef ]
  • Alkhamis, M.A.; VanderWaal, K. Spatial and Temporal Epidemiology of Lumpy Skin Disease in the Middle East, 2012–2015. Front. Vet. Sci. 2016 , 3 , 19. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Ardestani, E.G.; Mokhtari, A. Modeling the lumpy skin disease risk probability in central Zagros Mountains of Iran. Prev. Vet. Med. 2020 , 176 , 104887. [ Google Scholar ] [ CrossRef ]
  • Magori-Cohen, R.; Louzoun, Y.; Herziger, Y.; Oron, E.; Arazi, A.; Tuppurainen, E.; Shpigel, N.Y.; Klement, E. Mathematical modelling and evaluation of the different routes of transmission of lumpy skin disease virus. Vet. Res. 2012 , 43 , 1. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Molla, W.; Frankena, K.; MCM, D.E.J. Transmission dynamics of lumpy skin disease in Ethiopia. Epidemiol. Infect. 2017 , 145 , 2856–2863. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Mercier, A.; Arsevska, E.; Bournez, L.; Bronner, A.; Calavas, D.; Cauchard, J.; Falala, S.; Caufour, P.; Tisseuil, C.; Lefrançois, T.; et al. Spread rate of lumpy skin disease in the Balkans, 2015–2016. Transbound. Emerg. Dis. 2018 , 65 , 240–243. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Molla, W.; de Jong, M.C.M.; Frankena, K. Temporal and spatial distribution of lumpy skin disease outbreaks in Ethiopia in the period 2000 to 2015. BMC Vet. Res. 2017 , 13 , 310. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Byadovskaya, O.; Prutnikov, P.; Shalina, K.; Babiuk, S.; Perevozchikova, N.; Korennoy, F.; Chvala, I.; Kononov, A.; Sprygin, A. The changing epidemiology of lumpy skin disease in Russia since the first introduction from 2015 to 2020. Transbound. Emerg. Dis. 2022 , 69 , e2551–e2562. [ Google Scholar ] [ CrossRef ]
  • Allepuz, A.; Casal, J.; Beltrán-Alcrudo, D. Spatial analysis of lumpy skin disease in Eurasia-Predicting areas at risk for further spread within the region. Transbound. Emerg. Dis. 2019 , 66 , 813–822. [ Google Scholar ] [ CrossRef ]
  • Punyapornwithaya, V.; Seesupa, S.; Phuykhamsingha, S.; Arjkumpa, O.; Sansamur, C.; Jarassaeng, C. Spatio-temporal patterns of lumpy skin disease outbreaks in dairy farms in northeastern Thailand. Front. Vet. Sci. 2022 , 9 , 957306. [ Google Scholar ] [ CrossRef ]
  • Kahana-Sutin, E.; Klement, E.; Lensky, I.; Gottlieb, Y. High relative abundance of the stable fly Stomoxys calcitrans is associated with lumpy skin disease outbreaks in Israeli dairy farms. Med. Vet. Entomol. 2017 , 31 , 150–160. [ Google Scholar ] [ CrossRef ]
  • Gubbins, S.; Stegeman, A.; Klement, E.; Pite, L.; Broglia, A.; Cortiñas Abrahantes, J. Inferences about the transmission of lumpy skin disease virus between herds from outbreaks in Albania in 2016. Prev. Vet. Med. 2020 , 181 , 104602. [ Google Scholar ] [ CrossRef ]
  • Klausner, Z.; Fattal, E.; Klement, E. Using Synoptic Systems’ Typical Wind Trajectories for the Analysis of Potential Atmospheric Long-Distance Dispersal of Lumpy Skin Disease Virus. Transbound. Emerg. Dis. 2017 , 64 , 398–410. [ Google Scholar ] [ CrossRef ]
  • Davies, F.G. Observations on the epidemiology of lumpy skin disease in Kenya. J. Hyg. (Lond.) 1982 , 88 , 95–102. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Fagbo, S.; Coetzer, J.A.; Venter, E.H. Seroprevalence of Rift Valley fever and lumpy skin disease in African buffalo ( Syncerus caffer ) in the Kruger National Park and Hluhluwe-iMfolozi Park, South Africa. J. S. Afr. Vet. Assoc. 2014 , 85 , e1–e7. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Ahmed, E.M.; Eltarabilli, M.M.A.; Shahein, M.A.; Fawzy, M. Lumpy skin disease outbreaks investigation in Egyptian cattle and buffaloes: Serological evidence and molecular characterization of genome termini. Comp. Immunol. Microbiol. Infect. Dis. 2021 , 76 , 101639. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Pandey, N.; Hopker, A.; Prajapati, G.; Rahangdale, N.; Gore, K.; Sargison, N. Observations on presumptive lumpy skin disease in native cattle and Asian water buffaloes around the tiger reserves of the central Indian highlands. N. Z. Vet. J. 2022 , 70 , 101–108. [ Google Scholar ] [ CrossRef ]
  • Aboud, A.M.; Luaibi, O.K. Molecular and histopathological detection of lumpy skin disease in Buffaloes, Iraq. Int. J. Health Sci. 2022 , 6 , 4127–4137. [ Google Scholar ] [ CrossRef ]
  • Greth, A.; Gourreau, J.M.; Vassart, M.; Nguyen Ba, V.; Wyers, M.; Lefevre, P.C. Capripoxvirus disease in an Arabian oryx ( Oryx leucoryx ) from Saudi Arabia. J. Wildl. Dis. 1992 , 28 , 295–300. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Molini, U.; Boshoff, E.; Niel, A.P.; Phillips, J.; Khaiseb, S.; Settypalli, T.B.K.; Dundon, W.G.; Cattoli, G.; Lamien, C.E. Detection of Lumpy Skin Disease Virus in an Asymptomatic Eland ( Taurotragus oryx ) in Namibia. J. Wildl. Dis. 2021 , 57 , 708–711. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Barnard, B.J. Antibodies against some viruses of domestic animals in southern African wild animals. Onderstepoort J. Vet. Res. 1997 , 64 , 95–110. [ Google Scholar ]
  • Dao, T.D.; Tran, L.H.; Nguyen, H.D.; Hoang, T.T.; Nguyen, G.H.; Tran, K.V.D.; Nguyen, H.X.; Van Dong, H.; Bui, A.N.; Bui, V.N. Characterization of Lumpy skin disease virus isolated from a giraffe in Vietnam. Transbound. Emerg. Dis. 2022 , 69 , e3268–e3272. [ Google Scholar ] [ CrossRef ]
  • Rouby, S.; Aboulsoud, E. Evidence of intrauterine transmission of lumpy skin disease virus. Vet. J. 2016 , 209 , 193–195. [ Google Scholar ] [ CrossRef ]
  • Sudhakar, S.B.; Mishra, N.; Kalaiyarasu, S.; Jhade, S.K.; Hemadri, D.; Sood, R.; Bal, G.C.; Nayak, M.K.; Pradhan, S.K.; Singh, V.P. Lumpy skin disease (LSD) outbreaks in cattle in Odisha state, India in August 2019: Epidemiological features and molecular studies. Transbound. Emerg. Dis. 2020 , 67 , 2408–2422. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kononov, A.; Byadovskaya, O.; Kononova, S.; Yashin, R.; Zinyakov, N.; Mischenko, V.; Perevozchikova, N.; Sprygin, A. Detection of vaccine-like strains of lumpy skin disease virus in outbreaks in Russia in 2017. Arch. Virol. 2019 , 164 , 1575–1585. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Aleksandr, K.; Pavel, P.; Olga, B.; Svetlana, K.; Vladimir, R.; Yana, P.; Alexander, S. Emergence of a new lumpy skin disease virus variant in Kurgan Oblast, Russia, in 2018. Arch. Virol. 2020 , 165 , 1343–1356. [ Google Scholar ] [ CrossRef ]
  • Sprygin, A.; Van Schalkwyk, A.; Shumilova, I.; Nesterov, A.; Kononova, S.; Prutnikov, P.; Byadovskaya, O.; Kononov, A. Full-length genome characterization of a novel recombinant vaccine-like lumpy skin disease virus strain detected during the climatic winter in Russia, 2019. Arch. Virol. 2020 , 165 , 2675–2677. [ Google Scholar ] [ CrossRef ]
  • Shumilova, I.; Krotova, A.; Nesterov, A.; Byadovskaya, O.; van Schalkwyk, A.; Sprygin, A. Overwintering of recombinant lumpy skin disease virus in northern latitudes, Russia. Transbound. Emerg. Dis. 2022 , 69 , e3239–e3243. [ Google Scholar ] [ CrossRef ]
  • Sprygin, A.; Pestova, Y.; Bjadovskaya, O.; Prutnikov, P.; Zinyakov, N.; Kononova, S.; Ruchnova, O.; Lozovoy, D.; Chvala, I.; Kononov, A. Evidence of recombination of vaccine strains of lumpy skin disease virus with field strains, causing disease. PLoS ONE 2020 , 15 , e0232584. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Vandenbussche, F.; Mathijs, E.; Philips, W.; Saduakassova, M.; De Leeuw, I.; Sultanov, A.; Haegeman, A.; De Clercq, K. Recombinant LSDV Strains in Asia: Vaccine Spillover or Natural Emergence? Viruses 2022 , 14 , 1429. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Orynbayev, M.B.; Nissanova, R.K.; Khairullin, B.M.; Issimov, A.; Zakarya, K.D.; Sultankulova, K.T.; Kutumbetov, L.B.; Tulendibayev, A.B.; Myrzakhmetova, B.S.; Burashev, E.D.; et al. Lumpy skin disease in Kazakhstan. Trop. Anim. Health Prod. 2021 , 53 , 166. [ Google Scholar ] [ CrossRef ]
  • Makhahlela, N.B.; Liebenberg, D.; Van Hamburg, H.; Taioe, M.O.; Onyiche, T.; Ramatla, T.; Thekisoe, O.M.M. Detection of pathogens of veterinary importance harboured by Stomoxys calcitrans in South African feedlots. Sci. Afr. 2022 , 15 , e01112. [ Google Scholar ] [ CrossRef ]
  • Sprygin, A.; Pestova, Y.; Prutnikov, P.; Kononov, A. Detection of vaccine-like lumpy skin disease virus in cattle and Musca domestica L. flies in an outbreak of lumpy skin disease in Russia in 2017. Transbound. Emerg. Dis. 2018 , 65 , 1137–1144. [ Google Scholar ] [ CrossRef ]
  • Wang, Y.; Zhao, L.; Yang, J.; Shi, M.; Nie, F.; Liu, S.; Wang, Z.; Huang, D.; Wu, H.; Li, D.; et al. Analysis of vaccine-like lumpy skin disease virus from flies near the western border of China. Transbound. Emerg. Dis. 2022 , 69 , 1813–1823. [ Google Scholar ] [ CrossRef ]
  • El-Ansary, R.E.; El-Dabae, W.H.; Bream, A.S.; El Wakil, A. Isolation and molecular characterization of lumpy skin disease virus from hard ticks, Rhipicephalus (Boophilus) annulatus in Egypt. BMC Vet. Res. 2022 , 18 , 302. [ Google Scholar ] [ CrossRef ]
  • Horigan, V.; Beard, P.M.; Roberts, H.; Adkin, A.; Gale, P.; Batten, C.A.; Kelly, L. Assessing the probability of introduction and transmission of Lumpy skin disease virus within the United Kingdom. Microb. Risk Anal. 2018 , 9 , 1–10. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Gale, P.; Kelly, L.; Snary, E.L. Qualitative assessment of the entry of capripoxviruses into Great Britain from the European Union through importation of ruminant hides, skins and wool. Microb. Risk Anal. 2016 , 1 , 13–18. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Farra, D.; De Nardi, M.; Lets, V.; Holopura, S.; Klymenok, O.; Stephan, R.; Boreiko, O. Qualitative assessment of the probability of introduction and onward transmission of lumpy skin disease in Ukraine. Microb. Risk Anal. 2021 , 20 , 100200. [ Google Scholar ] [ CrossRef ]
  • Ince, Ö.B.; Çakir, S.; Dereli, M.A. Risk analysis of lumpy skin disease in Turkey. Indian J. Anim. Res. 2016 , 50 , 1013–1017. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • ANSES. Risk of Introduction of Lumpy Skin Disease into France ; ANSES Opinion Request No 2016-SA-0120; French Agency for Food, Environmental and Occupational Health & Safety: Maisons-Alfort, France, 2017; p. 134. Available online: https://www.anses.fr/en/system/files/SABA2016SA0120RaEN.pdf (accessed on 1 October 2022).
  • Saegerman, C.; Bertagnoli, S.; Meyer, G.; Ganière, J.P.; Caufour, P.; De Clercq, K.; Jacquiet, P.; Fournié, G.; Hautefeuille, C.; Etore, F.; et al. Risk of introduction of lumpy skin disease in France by the import of vectors in animal trucks. PLoS ONE 2018 , 13 , e0198506. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Saegerman, C.; Bertagnoli, S.; Meyer, G.; Ganière, J.P.; Caufour, P.; De Clercq, K.; Jacquiet, P.; Hautefeuille, C.; Etore, F.; Casal, J. Risk of introduction of Lumpy Skin Disease into France through imports of cattle. Transbound. Emerg. Dis. 2019 , 66 , 957–967. [ Google Scholar ] [ CrossRef ]
  • Taylor, R.A.; Berriman, A.D.C.; Gale, P.; Kelly, L.A.; Snary, E.L. A generic framework for s patial quantitative risk assessments of infectious diseases: Lumpy skin disease case study. Transbound. Emerg. Dis. 2019 , 66 , 131–143. [ Google Scholar ] [ CrossRef ] [ Green Version ]
  • Sprygin, A.V.; Fedorova, O.A.; Nesterov, A.A.; Shumilova, I.N.; Byadovskaya, O.P. The Stable Fly Stomoxys calcitrans L as a Potential Vector in the Spread of Lumpy Skin Disease Virus in Russia: Short Review. In E3S Web of Conferences ; EDP Sciences: Les Ulis, France, 2020. [ Google Scholar ]
  • Gubler, D.J. Vector-borne diseases. Rev. Sci. Tech. 2009 , 28 , 583–588. [ Google Scholar ] [ CrossRef ]
  • Christofferson, R.C.; Mores, C.N. Estimating the magnitude and direction of altered arbovirus transmission due to viral phenotype. PLoS ONE 2011 , 6 , e16298. [ Google Scholar ] [ CrossRef ]
  • Pitkin, A.; Deen, J.; Otake, S.; Moon, R.; Dee, S. Further assessment of houseflies ( Musca domestica ) as vectors for the mechanical transport and transmission of porcine reproductive and respiratory syndrome virus under field conditions. Can. J. Vet. Res. 2009 , 73 , 91–96. [ Google Scholar ] [ PubMed ]
  • Saegerman, C.; Hubaux, M.; Urbain, B.; Lengelé, L.; Berkvens, D. Regulatory issues surrounding the temporary authorisation of animal vaccination in emergency situations: The example of bluetongue in Europe. Rev. Sci. Tech. 2007 , 26 , 395–413. [ Google Scholar ] [ CrossRef ] [ PubMed ] [ Green Version ]
  • Tuppurainen, E.; Alexandrov, T.; Beltran-Alcrudo, D. Lumpy skin disease field manual-a manual for veterinarians. FAO Anim. Prod. Health Man. 2017 , 20 , 1–60. [ Google Scholar ]
  • Ritchie, S.A.; Rochester, W. Wind-blown mosquitoes and introduction of Japanese encephalitis into Australia. Emerg. Infect. Dis. 2001 , 7 , 900–903. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Hendrickx, G.; Gilbert, M.; Staubach, C.; Elbers, A.; Mintiens, K.; Gerbier, G.; Ducheyne, E. A wind density model to quantify the airborne spread of Culicoides species during north-western Europe bluetongue epidemic, 2006. Prev. Vet. Med. 2008 , 87 , 162–181. [ Google Scholar ] [ CrossRef ]

Click here to enlarge figure

Inclusion CriteriaArticles published from 1980 to September 2022
Studies focused on epidemiological characteristics of LSDV (i.e., hosts, animal reservoirs, vectors)
Studies reporting LSD modes of transmission
Studies analyzing historical or new outbreaks data with the purpose to highlight LSD risk factors
Studies describing quantitative and/or qualitative risk modelling of LSD
Studies reporting LSDV in ruminants other than cattle
Exclusion CriteriaFirst exclusion criteria
Editorials, letters to the editor
Studies related to a pathogen other than LSDV
Studies concerning the investigation of LSDV molecular characteristic
Studies on surveillance of LSDV
Second exclusion criteria
Articles describing modelling of economic impacts of LSD
Studies reporting vaccine efficiency, molecular interaction of LSD, or LSDV characteristics
Studies to evaluate test performance or surveillance systems
Studies on outbreak control
Reports on clinical signs
Studies focusing on the prevalence of LSD and excluding its transmission and the risk factors of outbreaks
General literature reviews of LSD
Type of StudyMethodologyObjective of the StudyCountReferences
Experimental studiesExperimental infections
Molecular techniques to detected LSD. PCR, neutralization, gene sequencing
Vector competence of blood-sucking insects/ticks.20[ , , , , , , , , , , , , , , , , , , , ]
Semen/oocytes: determine if there is LSDV in reproductive organs of cattle and bulls, semen, oocytes after experimental infection6[ , , , , , ]
Direct transmission: detect if there is a direct transmission between experimentally infected animals and healthy animals in a vector-proof environment2[ , ]
Establish the presence of LSDV in meat and offal products1[ ]
Establish the spill over from a vaccine1[ ]
Observational studies
Cross-sectional studiesMultivariable logistic or regression modellingRisk factors for LSD outbreaks, i.e., herd size, movement of animals, weather conditions11[ , , , , , , , , , , ]
Ecological niche models Bayesian hierarchical modelsIdentification of geographic locations and weather conditions which are suitable for the occurrence/spread of LSDV3[ , , ]
Mathematical modellingEvaluation of modes of transmission; establish transmission parameters and the R0 between animals 2[ , ]
Thin-plate spline regressionDetermine the spread rate1[ ]
Time series and spectral analysisTemporal trends and seasonal effects1[ ]
Spatial temporal analysisEvaluate the epidemic between different geographical areas3[ , , ]
Weather based modelEstimation of population dynamics of potential vectors1[ ]
Kernel-based modellingDetermine the force of infection based on distance and seasonality1[ ]
Hybrid single particle. Lagrangian-integrated trajectory modelIdentify wind events that condition vector transport1[ ]
Descriptive studiesField sampling of animals/suspected vectorsDetecting LSDV in animals other than cattle9[ , , , , , , , , ]
Intrauterine transmission of LSDV in natural conditions1[ ]
Semen from naturally infected bulls1[ ]
Detection, isolation of vaccine strains6[ , , , , , ]
Isolation of LSDV in field-collected vector 6[ , , , , , ]
Risk Assessment
QualitativeWOAH Risk analysis guidelinesProbability of introduction and/or spread into a country considering different pathways4[ , , , ]
WOAH Risk analysis guidelines and trade dataProbability of introduction and/or spread into a country considering different pathways1[ ]
QuantitativeQuantitative import risk analysisStochastic model for the probability of LSD introduction in a free country via a specific pathway2[ , ]
Created a generic frameworkA single pathway of introduction, i.e., live cattle trade1[ ]
Literature ReviewLiterature reviewLiterature review of the Stomoxys fly with additional information of outbreak data1[ ]
Animal (Species)Type of Samples, Test and LocationCountry
Year of Sampling
Reference
African buffalo (Syncerus caffer) Kenya 1981[ ]
African buffalo (Syncerus caffer) South Africa 2014[ ]
Egyptian buffalo (*) Egypt
2016 to 2019
[ ]
Asian buffalo
(Bubalus bubalis)
India
2020
[ ]
Buffalo (*) Egypt
2018
[ ]
Buffalo (*) Iraq
2021 to 2022
[ ]
Arabian oryx
(Oryx leucoryx)
Saudi Arabia 1989 [ ]
Southern eland (Taurotragus oryx) Namibia
2019
[ ]
Southern eland (Taurotragus oryx); Springbok (Antidorcas marsupialis); Impala (Aepyceros melampus); Wildebeest (Connachaetes gnou, C. taurinus) , 23% of C taurinus, 7% of southern eland, 23% of springboks and 20% of impalasSouth Africa 1993–1995[ ]
Giraffe (Giraffa Camelopardalis) Vietnam
2021
[ ]
Vector InvestigatedDetection of LSDV on the VectorDetection of LSDV in a Specific Body Part of the VectorLSD Viral Retention on the InsectEvidence of LSDV Replication in the InsectTransmission Attempts of LSDVBasic Reproduction Number (R0)Detection of LSDV in Field-Collected Samples
Stable fly
Stomoxys calcitrans[ , , , , , , ][ , , , ] [ , , , , , , ][ , , , , , , ][ , , ][ , ][ , , ]
Stomoxys sitiens[ , ][ ][ , ][ , ][ ]
Stomoxys indica[ , ][ ][ , ][ , ][ ]
Mosquitoes
Aedes aegypti[ , , , ][ , ][ , , , ][ , , , ][ ][ , ]
Anopheles stephensi[ ] [ ] [ ][ ][ ]
Culex quinquefasciatus[ , , ][ ][ , , ][ , ][ ][ , ]
Culex pipiens[ ][ ][ ][ ]
Aedes japonicus[ ][ ][ ][ ]
Biting midges
Culicoides nubeculosus[ , , , ][ , ][ , , ][ , , ][ ][ , ][ ]
Culicoidess spp.,
C. punctatus
[ , ][ ][ ][ ] [ , ]
Horseflies
Haematopota spp.[ ] [ ] [ ]
Tabanus bromiums [ ]
Non biting flies
Musca domestica L.[ , ] [ , ]
Muscina stabulans[ ] [ ]
Type of Infection/TransmissionTick Species
Amblyoma hebraeumRhipicephalus appendiculatusRhipicephalus decoloratusRhipicephalus annulatus
Intrastadial infection. Either nymphs or adult ticks without LSDV were allowed to feed on LSD infected cattle and then tested for the presence of the virus (body, or specific organs, e.g., salivary gland, gut)[ , , ][ , , ][ ][ ] *
Intrastadial/mechanical transmission. Adult ticks are interrupted in their feeding from a cow experimentally infected with LSDV and placed onto susceptible cows which are later tested for LSDV infection (i.e., transmission occurred)[ ][ ]
Transstadial persistence. Ticks at the larvae or nymphal stage are fed to repletion on cattle experimentally infected with LSDV. Nymphs then are incubated for molting into adults which are later tested for LSDV presence[ , , , ] **[ , , ]
Transstadial/mechanical transmission. Ticks at the larvae or nymphal stage are fed to repletion in cattle experimentally infected with LSDV. Emerging adult ticks are transferred onto healthy cattle to check if they were infected (i.e., transmission occurred)[ ][ ]
Transovarial passage. Female ticks were allowed to feed on LSDV experimentally infected cattle and later incubated to oviposit and for eggs to hatch. Eggs and/or mature larvae were tested for LSDV infection[ ][ ][ , , ][ ]
Transovarial transmission. Female adult ticks or larvae were allowed to feed on LSDV infected cattle and later incubated to oviposit and for eggs to hatch. Hatched larvae were place into healthy cows which are later tested to check if they were infected (i.e., transmission occurred)[ ][ ][ , ]
Identified Main Risk FactorsCountry/Region of StudyReference
Seasonality
   Risk of outbreaks increases with higher temperature and/or rainfallEgypt, Middle East, Balkans, Iran, Ethiopia, Albania, Eurasia, Uganda, Eastern and central Asia, Turkey, Russia[ , , , , , , , , , , , ]
Animal movements or tradeEgypt, Balkans, Ethiopia, Turkey, Kazakhstan[ , , , , , , ]
Herd characteristics
   Type of holdings, i.e., backyard, commercial farms Turkey, Middle East, Russia[ , , ]
   Herd size Ethiopia, Kazakhstan[ , ]
Cattle characteristics
   AgeMongolia, Egypt, Uganda, Ethiopia, Turkey[ , , , , , ]
   BreedTurkey, Egypt, Bangladesh[ , , , ]
   SexTurkey, Uganda, Mongolia, Bangladesh[ , , , ]
Farm location/landscape
   Urban and mixed rain-fed arid livestock systemMiddle East[ ]
   Areas mostly covered with croplands, grassland or shrub land Eurasia[ ]
   Presence of a water body near the farm (e.g., lake, river, pond, well)Turkey, Ethiopia, Mongolia[ , , ]
   Type of agro-climateEthiopia[ , ]
Type of herd management
   Water sources: communal or located in farmEgypt, Uganda, Ethiopia, Mongolia[ , , , ]
   Grazing: private or communal/pastoralUganda, Egypt, Ethiopia[ , , , ]
   Contact of cattle with other animals (e.g., buffaloes, sheep) Egypt, Uganda, Ethiopia[ , , ]
Cattle densityEurasia, Middle East[ , ]
Category FactorRisk Factor Odds Ratio
(95% C.I.)
Reference
Herd characteristicsGenus/breedBuffalo
Cattle
Reference
4.08 (1.98–8.4)
[ ]
Baladi
Mixed
Holstein
Reference
4.59 (1.83–11.48)
4.58 (1.73–12.12)
[ ]
Local
Cross breed
Reference
3.58 (1.40–9.17)
[ ]
SexMaleReference
Female19.29 (2. 46–151.32)[ ]
1.72 (1.02–2.92)[ ]
2.40 (1.11–5.16)[ ]
3.96 (2.16–7.27)[ ]
Age<1 year
1–2 years
>2 years
Reference
2.35 (1.48–3.7)
1.33 (0.88–2.01)
[ ]
<1 year
1–3 years
>3 years
Reference
1.41 (0.63–3.11)
2.49 (1.17–5.32)
[ ]
>24 months
<24 months
Reference
21.1 (8.83–50.43)
[ ]
0–12 months
13–24 months
>25 months
Reference
1.24 (0.63–2.44)
1.96 (1.15–3.34)
[ ]
0.5–1 year
1–4 years
≥4 years
Reference
1.38 (0.90- 2.09)
2.44 (1.67- 3.55)
[ ]
Calf
Young
Adult
Reference
0.21 (0.02–1.71)
0.05 (0.01–0.37)
[ ]
Herd sizeSmall (2–11 animals)
Medium and large
(>12 animals)
Reference
19.3 (1.4–50)
[ ]
Small
Medium
Large
Reference
0.68 (0.54–0.84)
0.63 (0.49–0.81)
[ ]
ManagementGrazing systemCommunal/pastoral
Fenced farm
Zero grazing
Reference
5.26 (2.64–10.48)
0.28 (0.06–1.44)
[ ]
Separate
Communal
Both
Reference
1.55 (0.91–2.60)
0.75 (0.39–1.42)
[ ]
Communal water sourcesNo
Yes
Reference
3.28 (2.11–5.09)
[ ]
3.31 (1.42–7.71)[ ]
Grazing and water sourcesSeparate/Private
Communal
Reference
4.1 (2.02–6.18)
[ ]
14.44 (2.23–94.0)[ ]
Water sourceRiver
Pond
Tube well
Reference
0.18 (0.06–0.53)
0.16 (0.05–0.47)
[ ]
ManagementFree animal movement No
Yes
Reference
0.36 (0.24–0.52)
[ ]
Contact with other animalsNo
Yes
Reference
3.40 (1.62–7.10)
[ ]
0.41 (0.23- 0.74)[ ]
Contact with buffaloNever
Daily
Weekly/monthly
Reference
1.78 (0.50–6.31)
0.49 (0.29–0.85)
[ ]
New introduction of cattle in the herdNo
Yes
Reference
8.5 (6.0–11)
[ ]
2.22 (1.32–3.71)[ ]
4.43 (2.6–7.5)[ ]
Purchase of animalsNo
Yes
Reference
11.67 (8.87–15.35)
[ ]
Sale(s) of animals during LSD outbreaksNo
Yes
Reference
1.24 (1.06–1.45)
[ ]
VaccinationNo
Yes
Reference
0.13 (0.05–0.34)
[ ]
EnvironmentSeasonWinter
Autumn
Spring
Summer
Reference
0.19 (0.02–1.50)
0.87 (0.29–2.51)
7.30 (3.97–13.42)
[ ]
Mean annual rainfall800–1000 mm
1001–1200 mm
1201–1400 mm
Reference
5.60 (2.35–13.34)
4.58 (2.23–9.40)
[ ]
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

Bianchini, J.; Simons, X.; Humblet, M.-F.; Saegerman, C. Lumpy Skin Disease: A Systematic Review of Mode of Transmission, Risk of Emergence and Risk Entry Pathway. Viruses 2023 , 15 , 1622. https://doi.org/10.3390/v15081622

Bianchini J, Simons X, Humblet M-F, Saegerman C. Lumpy Skin Disease: A Systematic Review of Mode of Transmission, Risk of Emergence and Risk Entry Pathway. Viruses . 2023; 15(8):1622. https://doi.org/10.3390/v15081622

Bianchini, Juana, Xavier Simons, Marie-France Humblet, and Claude Saegerman. 2023. "Lumpy Skin Disease: A Systematic Review of Mode of Transmission, Risk of Emergence and Risk Entry Pathway" Viruses 15, no. 8: 1622. https://doi.org/10.3390/v15081622

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

  • DOI: 10.1016/j.virusres.2019.05.015
  • Corpus ID: 173187984

Transmission of lumpy skin disease virus: A short review.

  • A. Sprygin , Y. Pestova , +2 authors A. Kononov
  • Published in Virus Research 1 August 2019
  • Environmental Science, Medicine

112 Citations

Lumpy skin disease, an emerging transboundary viral disease: a review, lumpy skin disease: insights into current status and geographical expansion of a transboundary viral disease., the role of modeling in the epidemiology and control of lumpy skin disease: a systematic review, global burden of lumpy skin disease, outbreaks, and future challenges.

  • Highly Influenced

Molecular Characterization of the 2020 Outbreak of Lumpy Skin Disease in Nepal

Lumpy skin disease: history, current understanding and research gaps in the context of recent geographic expansion, emergence and transboundary spread of lumpy skin disease in south asia, epidemiology, diagnosis and control of lumpy skin disease in egyptian ruminants.

  • 17 Excerpts

A Recombinant Vaccine-like Strain of Lumpy Skin Disease Virus Causes Low-Level Infection of Cattle through Virus-Inoculated Feed

Molecular characterization of lumpy skin disease virus from recent outbreaks in pakistan, 65 references, lumpy skin disease, an african capripox virus disease of cattle., evidence of vertical transmission of lumpy skin disease virus in rhipicephalus decoloratus ticks., epidemiological characterization of lumpy skin disease outbreaks in russia in 2016, detection of vaccine‐like lumpy skin disease virus in cattle and musca domestica l. flies in an outbreak of lumpy skin disease in russia in 2017, mechanical transmission of lumpy skin disease virus by rhipicephalus appendiculatus male ticks, a potential role for ixodid (hard) tick vectors in the transmission of lumpy skin disease virus in cattle., the detection of lumpy skin disease virus in samples of experimentally infected cattle using different diagnostic techniques., molecular detection and seasonal distribution of lumpy skin disease virus in cattle breeds in turkey, an investigation of possible routes of transmission of lumpy skin disease virus (neethling), epidemiological and molecular studies on lumpy skin disease outbreaks in turkey during 2014–2015, related papers.

Showing 1 through 3 of 0 Related Papers

Molecular characterization of lumpy skin disease virus from recent outbreaks in Pakistan

  • Original Article
  • Published: 25 November 2023
  • Volume 168 , article number  297 , ( 2023 )

Cite this article

essay on lumpy virus in english

  • Shumaila Manzoor   ORCID: orcid.org/0000-0002-3623-9086 1 ,
  • Muhammad Abubakar 1 ,
  • Aziz Ul-Rahman 2 ,
  • Zainab Syed 3 ,
  • Khurshid Ahmad 1 &
  • Muhammad Afzal 3  

488 Accesses

2 Altmetric

Explore all metrics

Lumpy skin disease (LSD) is a contagious viral transboundary disease listed as a notifiable disease by the World Organization of Animal Health (WOAH). The first case of this disease was reported in Pakistan in late 2021. Since then, numerous outbreaks have been documented in various regions and provinces across the country. The current study primarily aimed to analyze samples collected during LSD outbreaks in cattle populations in the Sindh and Punjab provinces of Pakistan. Phylogenetic analysis was conducted using partial sequences of the GPCR, p32, and RP030 genes. Collectively, the LSDV strains originating from outbreaks in Pakistan exhibited a noticeable clustering pattern with LSDV strains reported in African, Middle Eastern, and Asian countries, including Egypt, the Kingdom of Saudi Arabia, India, China, and Thailand. The precise reasons behind the origin of the virus strain and its subsequent spread to Pakistan remain unknown. This underscores the need for further investigations into outbreaks across the country. The findings of the current study can contribute to the establishment of effective disease control strategies, including the implementation of a mass vaccination campaign in disease-endemic countries such as Pakistan.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save.

  • Get 10 units per month
  • Download Article/Chapter or eBook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime

Price excludes VAT (USA) Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Rent this article via DeepDyve

Institutional subscriptions

essay on lumpy virus in english

Similar content being viewed by others

essay on lumpy virus in english

Molecular characterization of lumpy skin disease virus (LSDV) emerged in Bangladesh reveals unique genetic features compared to contemporary field strains

essay on lumpy virus in english

Phylogenetic analysis of the lumpy skin disease viruses in northwest of Iran

Incidence and molecular characterisation of lumpy skin disease virus in zimbabwe using the p32 gene, data availability.

The assembled sequences have been submitted to the GenBank database (accession no. OP807837-57).

McInnes CJ, Damon IK, Smith GL et al (2023) ICTV Virus Taxonomy Profile: Poxviridae 2023. J Gen Virol 104(5):001849. https://doi.org/10.1099/jgv.0.001849

Article   CAS   Google Scholar  

Annandale CH, Irons PC, Bagla VP et al (2010) Sites of persistence of lumpy skin disease virus in the genital tract of experimentally infected bulls. Reprod Domest Anim 45(2):250–255. https://doi.org/10.1111/j.1439-0531.2008.01274.x

Article   CAS   PubMed   Google Scholar  

Bianchini J, Simons X, Humblet MF et al (2023) Lumpy skin disease: a systematic review of mode of transmission, risk of emergence and risk entry pathway. Viruses 15(8):1622. https://doi.org/10.3390/v15081622

Article   PubMed   PubMed Central   Google Scholar  

Tuppurainen ES, Lubinga JC, Stoltsz WH et al (2013) Mechanical transmission of lumpy skin disease virus by Rhipicephalus appendiculatus male ticks. Epidemiol Infect 141(2):425–430. https://doi.org/10.1017/S0950268812000805

Gelaye E, Lamien CE (2019) Lumpy skin disease and vectors of LSDV. In: Kardjadj M, Diallo A, Lancelot R (eds) Transboundary animal diseases in Sahelian Africa and connected regions. Springer, Cham, pp 267–288. https://doi.org/10.1007/978-3-030-25385-1_13

Chapter   Google Scholar  

Paslaru AI, Maurer LM, Vögtlin A et al (2022) Putative roles of mosquitoes ( Culicidae ) and biting midges ( Culicoides spp.) as mechanical or biological vectors of lumpy skin disease virus. Med Vet Entomol 36(3):381–389. https://doi.org/10.1111/mve.12576

Article   CAS   PubMed   PubMed Central   Google Scholar  

Tuppurainen ES, Oura CA (2012) Lumpy skin disease: an emerging threat to Europe, the Middle East and Asia. Transbound Emerg Dis 59(1):40–48. https://doi.org/10.1111/j.1865-1682.2011.01242.x

Pestova Y, Byadovskaya O, Kononov A, Sprygin A (2018) A real time high-resolution melting PCR assay for detection and differentiation among sheep pox virus, goat pox virus, field and vaccine strains of lumpy skin disease virus. Mol Cell Probes 41:57–60. https://doi.org/10.1016/j.mcp.2018.08.003

Sudhakar SB, Mishra N, Kalaiyarasu S et al (2020) Lumpy skin disease (LSD) outbreaks in cattle in Odisha state, India in August 2019: Epidemiological features and molecular studies. Transbound Emerg Dis 67(6):2408–2422. https://doi.org/10.1111/tbed.13579

Ul-Rahman A, Niaz N, Raza MA et al (2022) First emergence of lumpy skin disease in cattle in Pakistan. Transbound Emerg Dis 69(6):3150–3152. https://doi.org/10.1111/tbed.14742

Article   PubMed   Google Scholar  

Ul-Rahman A, Shahid MF, Iqbal MZ et al (2023) Evaluation of haematological, serum biochemical and oxidative stress parameters in cattle naturally infected with lumpy skin disease virus. Trop Anim Health Prod 55(3):184. https://doi.org/10.1007/s11250-023-03608-1

Khatri G, Rai A, Aashish S et al (2023) Epidemic of lumpy skin disease in Pakistan. Vet Med Sci 9(2):982–984. https://doi.org/10.1002/vms3.1037

Rouby SR (2018) RPO30 gene based PCR for detection and differentiation of lumpy skin disease virus and sheep poxvirus field and vaccinal strains. Vet Sci Res Rev 4(1):1–8. https://doi.org/10.17582/journal.vsrr/2018.4.1.1.8

Article   Google Scholar  

Timurkan MÖ, Özkaraca M, Aydın H, Sağlam YS (2016) The detection and molecular characterization of lumpy skin disease virus, northeast turkey. Int J Vet Sci 5(1):44–47

Google Scholar  

Hall TA (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp Ser 41:95–98

CAS   Google Scholar  

Tamura K, Stecher G, Peterson D et al (2013) MEGA6: molecular evolutionary genetics analysis version 6.0. Mol Biol Evol 30:2725–2729. https://doi.org/10.1093/molbev/mst197

Babiuk S, Bowden TR, Parkyn G et al (2008) Quantification of lumpy skin disease virus following experimental infection in cattle. Transbound Emerg Dis 55(7):299–307. https://doi.org/10.1111/j.1865-1682.2008.01024.x

Aleksandr K, Pavel P, Olga B et al (2020) Emergence of a new lumpy skin disease virus variant in Kurgan Oblast, Russia, in 2018. Arch Virol 165(6):1343–1356. https://doi.org/10.1007/s00705-020-04607-5

Ochwo S, VanderWaal K, Ndekezi C et al (2020) Molecular detection and phylogenetic analysis of lumpy skin disease virus from outbreaks in Uganda 2017–2018. BMC Vet Res 16(1):66. https://doi.org/10.1186/s12917-020-02288-5

Wang J, Xu Z, Wang Z et al (2022) Isolation, identification and phylogenetic analysis of lumpy skin disease virus strain of outbreak in Guangdong, China. Transbound Emerg Dis 69(5):e2291–e2301. https://doi.org/10.1111/tbed.14570

Ma J, Yuan Y, Shao J et al (2021) Genomic characterization of lumpy skin disease virus in southern China. Transbound Emerg Dis 69(5):2788–2799. https://doi.org/10.1111/tbed.14432

Selim A, Manaa E, Khater H (2021) Molecular characterization and phylogenetic analysis of lumpy skin disease in Egypt. Comp Immunol Microbiol Infect Dis 79:101699. https://doi.org/10.1016/j.cimid.2021.101699

Şevik M, Doğan M (2017) Epidemiological and molecular studies on lumpy skin disease outbreaks in Turkey during 2014–2015. Transbound Emerg Dis 64(4):1268–1279. https://doi.org/10.1111/tbed.12501

Kumar N, Chander Y, Kumar R et al (2021) Isolation and characterization of lumpy skin disease virus from cattle in India. PLoS One 16(1):e0241022. https://doi.org/10.1371/journal.pone.0241022

Maw MT, Khin MM, Hadrill D et al (2022) First report of lumpy skin disease in Myanmar and molecular analysis of the field virus isolates. Microorganisms 10(5):897. https://doi.org/10.3390/microorganisms10050897

Tulman ER, Afonso CL, Lu Z et al (2001) Genome of lumpy skin disease virus. J Virol 75(15):7122–7130. https://doi.org/10.1128/jvi.75.15.7122-7130.2

Klement E, Broglia A, Antoniou SE et al (2020) Neethling vaccine proved highly effective in controlling lumpy skin disease epidemics in the Balkans. Prev Vet Med 181:104595. https://doi.org/10.1016/j.prevetmed.2018.12.001

Wolff J, Moritz T, Schlottau K et al (2020) Development of a safe and highly efficient inactivated vaccine candidate against lumpy skin disease virus. Vaccines 9(1):4. https://doi.org/10.3390/vaccines9010004

Hasib FM, Islam MS, Das T et al (2021) Lumpy skin disease outbreak in cattle population of Chattogram, Bangladesh. Vet Med Sci 7(5):1616–1624. https://doi.org/10.1002/vms3.524

Sudhakar SB, Mishra N, Kalaiyarasu S et al (2022) Genetic and phylogenetic analysis of lumpy skin disease viruses (LSDV) isolated from the first and subsequent field outbreaks in India during 2019 reveals close proximity with unique signatures of historical Kenyan NI-2490/Kenya/KSGP-like field strains. Transbound Emerg Dis 69(4):e451–e462. https://doi.org/10.1111/tbed.14322

Sprygin A, Pestova Y, Wallace DB et al (2019) Transmission of lumpy skin disease virus: a short review. Virus Res 269:197637. https://doi.org/10.1016/j.virusres.2019.05.015

Tuppurainen ES, Stoltsz WH, Troskie M et al (2011) A potential role for ixodid (hard) tick vectors in the transmission of lumpy skin disease virus in cattle. Transbound Emerg Dis 58(2):93–104. https://doi.org/10.1111/j.1865-1682.2010.01184.x

Sohier C, Haegeman A, Mostin L et al (2019) Experimental evidence of mechanical lumpy skin disease virus transmission by Stomoxys calcitrans biting flies and Haematopota spp. horseflies. Sci Rep 9(1):20076. https://doi.org/10.1038/s41598-019-56605-6

Mafirakureva P, Saidi B, Mbanga J (2017) Incidence and molecular characterisation of lumpy skin disease virus in Zimbabwe using the P32 gene. Trop Anim Health Prod 49(1):47–54. https://doi.org/10.1007/s11250-016-1156-9

Mansour KA (2017) Phylogenetic tree analysis study of Lumpy skin disease virus based envelope protein P32 gene in Al-Qadisiyah Province, Iraq. Al-Qadisiyah J Vet Med Sci 16(1):106–111

Rashid PM, Baba Sheikh MO, Raheem ZH, Marouf AS (2017) Molecular characterisation of lumpy skin disease virus and sheeppox virus based on p32 gene. Bulg J Vet Med 20(2):131–140. https://doi.org/10.15547/bjvm.984

Hedayati Z, Varshovi HR, Mohammadi A, Tabatabaei M (2021) Molecular characterization of lumpy skin disease virus in Iran (2014–2018). Arch Virol 166(8):2279–2283. https://doi.org/10.1007/s00705-021-05119-6

Lu G, Xie J, Luo J et al (2021) Lumpy skin disease outbreaks in China, since 3 August 2019. Transbound Emerg Dis 68(2):216–219. https://doi.org/10.1111/tbed.13898

Download references

Acknowledgments

The authors would like to thank the veterinary field staff and Livestock & Dairy Development Department Pakistan, for their support during sample collection.

Author information

Authors and affiliations.

National Veterinary Laboratory, Ministry of National Food Security and Research, Park Road, Islamabad, Pakistan

Shumaila Manzoor, Muhammad Abubakar & Khurshid Ahmad

Department of Pathobiology & Biomedical Sciences, Faculty of Veterinary and Animal Sciences, MNS University of Agriculture, Multan, Pakistan

Aziz Ul-Rahman

FAO Project, National Agriculture Research Centre premises, Islamabad, Pakistan

Zainab Syed & Muhammad Afzal

You can also search for this author in PubMed   Google Scholar

Contributions

Manzoor S, Abubakar M, and Afzal M conceived the study; Abubakar M, Syed Z, and Ahmad K collected data from field outbreaks; Ul-Rahman A did data analysis; and Manzoor S, Abubakar M, and Ul-Rahman A wrote the manuscript and edited the final draft.

Corresponding authors

Correspondence to Shumaila Manzoor or Aziz Ul-Rahman .

Ethics declarations

Conflict of interest.

The authors declare that they have no conflict of interest.

Ethical approval

This article does not contain any studies with human participants or animals performed by any of the authors.

Additional information

Handling Editor: William G. Dundon.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary file1 (DOCX 15 KB)

Rights and permissions.

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Manzoor, S., Abubakar, M., Ul-Rahman, A. et al. Molecular characterization of lumpy skin disease virus from recent outbreaks in Pakistan. Arch Virol 168 , 297 (2023). https://doi.org/10.1007/s00705-023-05925-0

Download citation

Received : 04 January 2023

Accepted : 20 October 2023

Published : 25 November 2023

DOI : https://doi.org/10.1007/s00705-023-05925-0

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Find a journal
  • Publish with us
  • Track your research

preview

Lumpy Skin Disease Essay

INTRODUCTION: Lumpy skin disease (LSD) is viral in origin called "Neethling virus" belongs to a family of Poxviridae which includes pox virus for both sheep and goat (sheep pox virus and goat pox virus), and it is infectious and eruptive disease the insects play important role as a vector by biting specially within wet summer and autumn months. All cattle breed susceptible to infected and there is no evidence or study present that a specific age or breed of cattle are more than others to exposed to infection, generally; younger are more severely affected. Spreading of the infection may occur less commonly by direct contact through saliva, skin lesions, milk, semen, or nasal discharge of infected animals. Typically, nodules or lumps on the

Cellulitis Essay

Main purpose of cellulitis treatment process is reduction of severity cellulitis infection, fast recovery, pain relieve, cure affected skin and prevention of recurrence - definition of treatment for cellulitis. In most cases healing process contains treatment with antibiotics drugs. Antibiotics are used orally or intravenous depending of severity of affected skin area. Period for using oral antibiotics (by mouth) is 10 to 14 days. In this period is crucial to take every single pill that doctor has prescribed, even later when you begin feel better .Treatment for cellulitis with intravenously antibiotics (IV) is recommended when we have more serious infection,in most cases lasts 3 to 5 days.

Dermatofibromas Research Paper

Dermatofibromas (also known as Fibrous histiocytoma and Fibroma simplex, Nodular subepidermal fibrosis, Sclerosing hemangioma), are common noncancerous (benign) small skin growths. Dermatofibromas are typically detected on the arms and legs. Other common areas are located on breast, face and hands. This growth is usually a benign, single structure that resembles a nipple. Its appearance can be discolored and contains hard, scar-like tissue. A minor injury such as an insect bite, puncture or most recently discovered arsenic, can result in the formation of a dermatofibroma. These growths or nodes only can be found on humans and have not been reported or found other animals. It is estimated that only 1:10 women

Lyme Disease Research Paper

Lyme disease is an infection produced by bacteria called Borrelia Burgdorferi. This bacteria or germ is ordinarily found in shrews, deer, mice, and squirrels. Ixodes bugs, normally called deer ticks, often feed on the blood from an infected animal. When this happens, the tick then becomes a carrier of the bacteria that causes Lyme disease and can infect you with this germ through your skin. Infected deer ticks are normally found in the northeast and upper Midwest United States because of the climate and humidity levels, and become more active in the late spring and early summer months after the birth of new larvae. Lyme disease is known to cause a skin rash called erythema migrans and can leave you problems with your joints, brain, heart, and nerves. The

Asheboro Dermatology Research Paper

When it comes to treating acne, Asheboro Dermatology & Skin Surgery in West End, NC, has successfully treated patients with Accutane (Isotretinoin), a drug taken orally to help eliminate the redness, scarring and scabbing that’s often associated with the skin condition.

Skin Cancer Research Paper

Discuss the different types of skin cancer, risk factors, diagnosis, treatment options and preventative measures.

Essay On Skin Cancer

Skin cancer is a very real and common health problem for Australians, with 2 in 3 developing this deadly disease by the age of 70 (Lynch, 2006). “Due to the Australian climate and lifestyle, as well as a predominantly fair skinned population, skin cancer is the most prevalent form of cancer in Australia,” (Australian Institute of Health and Welfare, 2000). For the most effective prevention against skin cancer, it is recommended that a combination of sun safety measures are met; slip, slop, slap, seek, slide (Slip on sun protective clothing, slop on water resistant SPF30+ sunscreen, slap on a hat, seek shade, slide on sunglasses), (Preventing Skin Cancer, 2017). By simply reducing recreational sun exposure, the risk of developing melanoma –

treatments: Medicines (creams), Medicines that you take by mouth, A treatment using ultraviolet A (UVA) light (PUVA) and extracting the infected area surgically.

Cancer is usually a terrifying word. Those who have never received a diagnosis of cancer, likely cannot fathom the anxiety such an event produces. Even forms of cancer that are highly treatable often bring about significant and immediate life changes for patients. Hopefully, the ideas from this article can help to demystify cancer, and make it somewhat easier to cope with the disease and its effects.

Skin Cancer Essay

  • 2 Works Cited

Skin cells that lose the ability to grow and divide are called skin cancer. Another name for skin cancer is neoplasia. Skin cancer begins on the outer layer of the skin called the epidermis and is the most common form of cancer in humans. All skin cancers are important but the most commonly seem is the basal cell carcinoma, squamous cell carcinoma, melanoma. These skin cancers happens when the skin starts to grows abnormal skin cells and form a mass called a skin tumor. Basal cell carcinoma and squamous are the most common skin cancer and they are referred to as nonmelanoma cancer. The least dangerous skin cancer of the three is the basal cell carcinoma. The most dangerous skin cancer of the three is melanoma because it spreads quickly to

Skin Pathology

In the article “Skin Pathology in the Cretaceous: Evidence for Probable Failed Predation in a Dinosaur” written by Brice M. Rothschild and Robert Depalma, they report their findings on the observation of the skin of a duck billed dinosaur regarding its’ scaled pattern and forming of scar tissue (Rothschild and Depalma 2013). Rothschild and Depalma introduce the issue of dinosaur skin pathology through the finding of damaged dinosaur skin through Lingham-Soliar ‘s (2008) report on the damaged skin and bite impression of the Psittacosaurus skin, which did not heal from the bite; indicating that the Psittacosaurus had died from either a scavenger or a predator (Rothschild and Depalma 2013). However Rothschild and Depalma (2013) focus their research

Pinkeye In Cattle

One of our bucket calves this year caught pink eye. Pinkeye is a very painful virus any type of animal can catch, including humans. Pinkeye is caused by a bacteria that infects the surface of the eye. A problem with M. Bovis is that it can spread easily and rapidly. It is difficult to control the pink eye virus because of face flies. Face Flies are flies that spread M. Bovis. The flies are attracted to the tearing that is produced by the infected eye. The fly then picks up the virus and can transfer it to other animals. The flies can infect a whole herd of cattle in a day, so it is best to treat the virus as quickly as possible. Because we have two calves in one barn, we had to treat the virus immediately.

What Are The Physiological Origins Of The Following Skin Lesions

Skin tags are made up of loose collagen fibres and blood vessels surrounded by the skin, collagen is found throughout the body.

North Dallas Dermatology Research Paper

North Dallas Dermatology Associates is a dermatology clinic that is located in Dallas, Texas. North Dallas Dermatology Associates was established in 2000. Their services include cosmetic dermatology, medical dermatology, facial treatments, body contouring and skin tightening, and laser treatments. Their cosmetic dermatology treatments include Botox, Dysport, Cosmetic Fillers, Juvederm Voluma, Sculptra Aesthetic, and more. Their facial treatments include chemical peels, facials, OxyGeneo, microdermabrasion, and MicroNeedling with PRP. Their dermatologists are board-certified.

Skin cancer is an abnormal growth of cells, and there are three main types of skin cancer. Squamous cell carcinoma, basal cell carcinoma and lastly malignant melanoma, the one that we are going to focus on. While it is not the most common type of skin cancer, that is basal cell carcinoma, it is the most dangerous/deadly.

Non Melanoma Research Paper

Basal Cell Carcinoma – Most common type of skin cancer it accounts for 80% of non-melanoma cancers. If left untreated it can affect skin, bone tissue, and cause death.

Related Topics

  • Infectious disease
  • Immune system
  • Animal rights
  • Bovine spongiform encephalopathy
  • Transmissible spongiform encephalopathy
  • Diseases A-Z
  • Diet & Fitness
  • Naturopathy
  • Healthy Relationships
  • Web Stories
  • Women's Health
  • Home remedies

Select Language

  • English इंग्लिश
  • Hindi हिंदी
  • Diseases Conditions

What Is Lumpy Skin Disease, Can It Infect Humans? Symptoms And Causes of LSD

What Is Lumpy Skin Disease, Can It Infect Humans? Symptoms And Causes of LSD

The Lumpy Skin Disease (LSD) is caused by a virus called the Capripoxvirus and is one of the biggest threats to livestock worldwide. This virus genetically belongs to the goatpox and sheeppox virus family.

Lumpy Skin Disease (LSD) is spreading rapidly in India. In the latest data, the health officials have confirmed that the viral disease has infected over 5,000 cattle in Uttar Pradesh, which has disrupted milk production as well. Not just in UP, the viral disease has entered other states too. Over the last few weeks, nearly 3,000 cattle have died in Rajasthan and Gujarat. Focusing on the seriousness of the disease and its high transmission rate, many states have also put a ban on the movement of cattle. Now, to know more about this disease which is killing cattle across India, let's understand this virus. How does it transmit from one person to another, what are the symptoms that it can cause and everything else related to it.

What Is The Lumpy Skin Disease?

Who is the carrier of this virus.

  • MoU Signed For Commercial Production Of Lumpi-Provac, Indigenous Vaccine For Lumpy Skin Disease
  • Lumpy Virus Continues To Wreak Havoc In Agra: 61 Cows Infected In 36 Villages
  • Lumpy Skin Disease Reaches Delhi: 173 Cases Reported In One Day

Some of the other symptoms of LSD include excessive nasal and salivary secretion. One of the severe effects of LSD also includes miscarriage. Experts say that pregnant cows and buffaloes often suffer miscarriage and in some cases, infected cattle/animals can die due to the infection as well.

Can Lumpy Skin Disease Infect Humans?

Lumpy Skin Disease (LSD) is not a new disease, according to the available data, this disease was already endemic in most African countries, and since 2012it is spreading across the Middle East, Southeast Europe and West and Central Asia. Also, when it comes to Asia, LSD has been infecting cattle since 2019. In May 2022, Pakistan's Punjab reported the deaths of over 300 cows due to LSD. Currently, the concerning part is that the virus is not only spreading across the country but also killing cattle.

You may like to read

Can we drink milk, trending now, how can one stop the viral disease from spreading further.

This viral disease spread among animals, however, it is important to save the cattle also. In order to do so, the first thing that is required is proper vigilance of the virus, which include - early detection, followed by a rapid and widespread vaccination campaign. The next few steps can include sanitising cattle sheds, isolating the infected animals, and contacting the nearest veterinarian for treatment of the infected animal.

  • Lumpy Skin Disease
  • Lumpy Skin Disease Cases
  • lumpy skin disease symptoms
  • Lumpy Skin Disease Treatment
  • Rajasthan Lumpy Skin Disease

Don’t Miss Out on the Latest Updates. Subscribe to Our Newsletter Today!

thehealthsite subscribe now

Subscribe Now

Enroll for our free updates

thehealthsite subscribe now

Thank You for Subscribing

Thanks for Updating Your Information

By clicking “Accept All Cookies”, you agree to the storing of cookies on your device to enhance site navigation, analyze site usage, and assist in our marketing efforts. Cookie Policy .

TALK WITH SHIVI

10 lines on Lumpy Virus in English / What is Lampi Virus?

10 lines on Lumpy Virus in English / What is Lampi Virus?

10 lines on lumpy virus.

Hello friends in this video we are presenting a 10 line essay on “ Lumpy Virus ” & “ Lampy Virus “. We hope this is helpful for you.

What is lumpy Virus?/ Short essay on Lampy Virus

  • The lumpy virus is spreading in cows in many states of the country.
  • According to the Global Alliance for Vaccines and Immunization (GAVI), the lumpy virus is a disease transmitted by cows and buffaloes.
  • This is a type of skin disease which is caused by a virus called Capripoxvirus . It spreads from one animal to another.
  • Due to this disease, lumps start appearing in the skin of infected cows and buffaloes.
  • This virus is spreading so fast that thousands of animals have died in Rajasthan, UP, Bihar and MP.
  • The main symptoms of lumpy disease are fever, loss of weight, watery eyes, drooling, rash on the body, loss of appetite and giving less milk.
  • It is a disease that spreads through direct contact with cattle and also through contaminated food and water.
  • To prevent lumpy disease, keep the infected animal isolated and clean the living space properly. Also regularly spray mosquitoes and get the infected animal vaccinated.
  • When the animal dies, do not leave it uncovered and spray the entire area with disinfectants.
  • This virus especially affects cows with weak immunity. Only a vaccine can control and prevent this disease.
  • While milking a cow-buffalo infected with lumpy virus, take these measures to protect yourself –  do milking by wearing a mask, sanitize your hands after milking, cleaning your hands before and after milking.

10 Lines on Lumpy Virus

Leave a Comment Cancel reply

Save my name, email, and website in this browser for the next time I comment.

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • v.15(12); 2022 Dec

Lumpy skin disease: A newly emerging disease in Southeast Asia

Kanokwan ratyotha.

1 Faculty of Veterinary Sciences, Mahasarakham University, Maha Sarakham 44000, Thailand

Suksanti Prakobwong

2 Department of Biology, The Parasitology, Geoinformatics, Environment and Health Science Research Group, Faculty of Science, Udon Thani Rajabhat University, Udon Thani 41000, Thailand

Supawadee Piratae

3 One Health Research Unit, Faculty of Veterinary Sciences, Mahasarakham University, Maha Sarakham 44000, Thailand

Lumpy skin disease (LSD) is caused by LSD virus (LSDV). This virus has been classified in the genus Capripoxvirus , family Poxviridae which generally affects large ruminants, especially cattle and domestic water buffalo. The first outbreak of LSD was found in 1929 in Zambia, then spreading throughout Africa and with an ongoing expanding distribution to Asia and Europe. In 2020, LSD was found from Southeast Asia in Vietnam and Myanmar before reaching Thailand and Laos in 2021. Therefore, LSD is a newly emerging disease that occurs in Southeast Asia and needs more research about pathology, transmission, diagnosis, distribution, prevention, and control. The results from this review show the nature of LSD, distribution, and epidemic maps which are helpful for further information on the control and prevention of LSD.

Introduction

Lumpy skin disease (LSD) is caused by LSD virus (LSDV), a virus in the family Poxvirus, genus Capripoxvirus as well as sheep pox virus (SPPV) and goat pox virus (GTPV) [ 1 ]. This virus can cause infection mainly in cattle ( Bos spp.) and buffaloes ( Bubalus spp.); there are also reports in other wild ruminant species, such as giraffes, bulls, and springboks [ 2 ]. This arbovirus is probably transmitted by mechanical transmission through blood–sucking arthropods, including mosquitoes, ticks, and flies [ 3 ]. This virus can also be transmitted to susceptible animals through direct contact with the secretions of other infected animals and indirect contact from contaminants of the owner of the animals and objects (vehicle, equipment, etc.) [ 1 , 4 ]. Infected animals may present variations of clinical signs ranging from subclinical to high morbidity and mortality. The clinical signs are fever (40.0°C–41.5°C), lacrimation, nasal discharge, hypersalivation, lethargy, anorexia, and weakness, followed by the development of nodular lesions in the skin and mucous membranes of the whole body. In addition, these lesions may develop into the muscular layer [ 3 , 4 ]. The resulting wound lesions can develop necrotic tissue and scarring, which may occur with secondary infection with other types of complications such as bacteria, viruses, or myiasis and cause severe clinical symptoms [ 5 ]. In general, prevalence of LSD ranges from 1%–2% to 80%–90% in different situations in the endemic region [ 2 ]. Mostly LSD can cause low mortality and the differences in the mortality rate may be explained by differing susceptibility of hosts (strain, age, and host immune response).

The first reported case of LSD occurred in 1929 in Zambia and after that, it spread throughout South Africa with sporadic outbreaks in some areas [ 5 ]. At present, this disease is an endemic disease in Africa. Lumpy skin disease has recently spread in Asia during 1988–1989 following outbreaks in Europe and the Middle East in 1990 [ 6 ]. The disease emerged in South Asia in 2019 and then rapidly spread throughout Southeast Asia in 2020 [ 1 ]. Infection with LSDV could affect not only the health status of ruminants, but furthermore, might impact the economic activity of ruminants, especially cattle.

Here, we review the general information about LSD and show where research is needed for a better understanding of the biology, pathology, transmission, diagnosis, distribution, prevention, and control of this newly emerging disease in Southeast Asia.

Virus and Classification

Lumpy skin disease virus (LSDV) is a virus in the family Poxviridae, subfamily Chordopoxviridae, genus Capripoxvirus . The genus Capripoxvirus comprises three viruses; SPPV, GTPV, and LSDV. Lumpy skin disease virus is large–sized (230–260 nm) enclosed in a lipid enveloped with a genome of approximately 150 kilobase pairs (kbp) and shared 97% identity in the nucleotide sequences with SPPV and GTPV genome ( Figure-1 ). The LSDV genome included at least 146 putative genes, which displayed proteins that play roles in virion structure, DNA replication, transcription and metabolism, protein processing and assembly, virus stability, and evading host immune response [ 7 ]. Lumpy skin disease virus can cause infection mainly in large ruminants; specifically in cattle, buffaloes, and other wild ruminant species. However, the disease is not contagious from animals to humans. Signs and symptoms of LSDV vary widely and depend on many factors, such as status of the host’s immune response, strain of the virus, and the environment. Severe clinical signs of disease may occur in young animals, lactating animals, or animals with lower immunity than healthy animals. European beef cattle strains ( Bos taurus ) are more susceptible to LSD than tropical or Indian beef cattle strains ( Bos indicus ). Moreover, domestic buffalo ( Bubalus bubalis ) have a lower incidence rate than cattle [ 1 , 4 , 8 ]. The LSDV genome can be detected in nodules, ulceration, blood, secretions, and semen in both vertebrate (ruminant) ( Table-1 ) [ 1 , 2 , 4 ] and invertebrate animals (arthropods) ( Table-2 ) [ 9 – 19 ].

An external file that holds a picture, illustration, etc.
Object name is Vetworld-15-2764-g001.jpg

Schematic diagram of the poxvirus structure. (a) cross-section; (b) longitudinal section. (Figure prepared by Kanokwan Ratyotha).

Vertebrate hosts susceptible to LSDV infection.

Vertebrate hostsCountries/RegionsReferences
Giraffe
South Africa[ , ]
Impala
South Africa[ , ]
Eland
South Africa[ ]
Wildebeest
South Africa[ ]
Thomson’s Gazelle
South Africa[ ]
Oryx
South Africa, Saudi Arabia[ , , ]
South Africa, Saudi Arabia[ , ]
Springbox
Namibia[ , , ]
African wild buffalo
Kenya[ , ]

LSDV=Lumpy skin disease virus

LSDV genome detected in vertebrate animals.

Invertebrate hostsCountries/RegionsReferences
Ticks
South Africa[ – ]
South Africa[ – , – ]
South Africa[ , ]
Glimpses
  spp.South Africa[ ]
Flies
Belgium, Egypt[ , ]
Mosquitoes
Egypt[ ]
Egypt[ ]
Egypt[ ]
Egypt[ ]

transmission, Spread, and Virus Stability

Lumpy skin disease is a vector-borne disease transmitted by mosquitoes ( Aedes aegypti, Anopheles stephensi, Culex quinquefasciatus , and Culicoides nubeculosus ), ticks ( Rhipicephalus appendiculatus, Rhipicephalus decoloratus , and Amblyomma hebraeum ), and Diptera ( Haematopota spp. and Stomoxys calcitrans ). The LSDV can survive in skin nodules for 1 month and at least 3 weeks in air–dried hides. The virus is excreted in the blood, nasal secretions, saliva, ear notches, semen, and milk and can be transmitted to suckling calves [ 20 – 22 ]. In general, vectors enhance the distribution of LSDV by mechanical and biological transmissions. Many studies have reported that after blood-sucking vectors (mosquitoes, ticks, glimpse, and flies) take a blood meal from infected cattle (viremia stage), the virus can propagate and shed in the salivary glands, head, body, and feces of insects. This allows the infected insects to become a reservoir for further transmission [ 15 ]. The infectivity of LSDV has been studied in both egg and juvenile of ticks ( R. decoloratus ) in which it was found that the disease can be transmitted by transovarial transmission [ 14 ]. In addition, the mechanical contact of flies is also a way they can be carriers of disease. Direct contact between cattle in a cage has been commonly found in endemic areas. However, the viral transmission was also accomplished by the contamination of veterinary equipment and vehicles as well as stockholders from a farm translocating to distant areas [ 1 , 23 ]. Most LSD infections have been found in the summer when vectors are active; it may designate the blood–feeding insects and the virus spread. The enhancement of risk factors was associated with a warm and humid climate that supported the reproduction of vector populations. The introduction of new animals to a herd is one of the risk factors.

Lumpy skin disease virus is well tolerated in the environment in the pH range of 6.3–8.3 and it can survive in dry scabs on the skin for up to 3 months [ 1 ]. Lumpy skin disease virus can grow in cell cultures at 4°C for up to 6 months, in phosphate buffer saline at 28°C for up to 35 days, in skin nodule lesions collected from frozen at –80°C for up to 10 years. However, it can be destroyed by ultraviolet heat for an exact time, for example, 55°C for 2 h, 60°C for 1 h, and 65°C for 30 min. Moreover, this virus is sensitive to excess acid or base and therefore, it can be destroyed by common disinfectants [ 9 ].

Pathogenesis and Clinical Signs

In some outbreaks that occurred in Africa and middle Asia countries, the mortality rate was generally low (1%–3%) but may reach 40% in some regions. After infection, the incubation period ranges from 4 to 7 days, as determined experimentally, but for naturally occurring infections, the incubation period is 28 days or is prolonged to 35 days [ 1 , 4 ]. Lumpy skin disease virus infection by intradermal replication in fibroblasts, macrophage, pericytes, and endothelial cells leads to viremia causing vacuities and lymphangitis in affected areas [ 2 ]. After cattle recover from infection, they acquire antibodies for about 6 months [ 6 ]. Lumpy skin disease is an economically impact disease in an outbreak because the severity in cows peaks during lactation and causes a decreased milk harvest during the high fever caused by the viral infection and bacterial mastitis. Clinical signs after the incubation period can be classified into 4 phases.

Phase 1 (acute phase) after the incubation period, animals have fever as high as 41°C for about 7 days sometimes prolonged to 10 days with anoxia, depression, lacrimation, increased nasal discharge, saliva secretions, lack of milk, found multinodular lesions around skin, and mucous membrane. Some cases are non–febrile.

Phase 2 subscapular and precrural lymph nodes develop noticeable enlargement 3–5 times of their normal, and there are increased multi nodules, mostly on the head, neck, limbs, genitalia, udder, mucous membrane, nasal and oral cavities, or plaques at the site of inoculation. The diameter of the nodule lesion is 0.5–5 cm, apparent in varying numbers and sizes, from only a few to multiple lesions covering the entire animal ( Figure-2a ). After 1–2 days, nodules rupture, also shedding virus depending on the concentration of virus. Sometimes found edema of limbs is caused by lymphangitis and vasculitis.

An external file that holds a picture, illustration, etc.
Object name is Vetworld-15-2764-g002.jpg

The clinical sign of lumpy skin disease in cattle in Thailand. (a) The lymph nodes develop noticeably enlarged 3–5 times of its normal; (b) the nodule lesions’ transformation into ulceration and necrosis.

Phase 3, after 2–3 weeks, nodules lesions into ulceration and become necrotic. Also, beaded serum exudes, especially from limbs and causes lameness, pain, and lack of movement. In severe cases, an ulcerative lesion appears in the mucous membranes at various points, such as the eye and nasal cavities; there is excessive salivation, lachrymation, and nasal discharge. The secretions from animals may contain LSDV [ 19 ].

Phase 4, after at least 1 month, there is complete healing of ulcerations and also thickening of the skin and hyperpigmentation of the lesion ( Figure-2b ).

Bacterial and virus infections can occur in lesions, and these pathogens are able to be inhaled by the host, causing a sequence of complications. These complications include keratitis, mastitis, pneumonia, and myiasis and can increase mortality. Other clinical signs of complications are abortion, decreased lactation, anestrus, infertility, and subclinical sign may be present. Post-mortal LSDV infected cattle have been found with nodule formation and ulceration in the trachea, lung, and gallbladder [ 6 , 19 , 24 ]. Differential diagnosis include pseudo LSD/bovine Herpes mammillitis by bovine Herpes virus type 2, pseudocowpox and bovine papular stomatitis by Parapoxvirus , insect biting, urticaria, and demodicosis [ 4 ]. In outbreaks of the disease, the morbidity rate varies widely depending on the immune status of the hosts and the abundance of mechanical arthropod vectors.

Lumpy skin disease virus affects large ruminants, especially cattle and domestic water buffalo; however, this virus has also been reported in wildlife due to these animals belonging to the suborder Ruminantia, as well as cattle and buffaloes. Clinical signs of LSDV infection in wildlife are difficult to monitor and may range from asymptomatic to severe clinical signs [ 24 ]. Although reports of disease in wildlife are low incidence, it is still a concern whether the disease spreads from pets to wildlife [ 25 ]. It is difficult to prevent and control the disease due to the inability to control livestock and make preventive vaccines for wildlife as well as livestock. There are also many species of ruminant wildlife that are not known how to transmit disease, which can affect their natural population.

Humoral and Cellular Immunity

In cases of natural infection, the immunity caused by infection can be detected approximately 2 weeks after infection. From experimental infection, antibodies are detectable from 6 to 8 day post-infection. The highest immunity can be detected during 3–4 weeks after infection and remain detectable for up to 5 months [ 26 ]. Although antibodies are able to limit the spread of extracellular organisms, most LSDV are predominantly intracellular. Humoral immunity cannot be enough to eliminate the proliferation of viruses inside cells. Therefore, cell-mediated immunity is essential for the effective control of infection in animals. Once an animal has been vaccinated, the humoral immune response can last longer than 7 months which is capable of preventing disease [ 27 ]. However, animals in endemic regions are recommended to receive an annual vaccination booster due to the duration of humoral and cellular immunity is still unknown.

Diagnosis of LSDV is based on characteristic clinical signs combined with various laboratory approaches. Virus isolation and electron microscopy can be done but are rather expensive, labor and time–consuming, and cannot differentiate between poxvirus virions. The immunological–based techniques such as enzyme-linked immunosorbent assay have been developed to detect antibodies for LSDV infection; however, false detection caused by non-specific binding can occur between Parapoxvirus and Capripoxvirus [ 28 ]. In the laboratory, DNA–based detection methods by polymerase chain reaction (PCR) or real-time-PCR (RT-PCR) is used to detect viral DNA in specimens, including nodules, secretion, semen, and blood of suspected animals [ 29 ]. Target genes for PCR or RT-PCR detection are often used for the gene–specific viral attachment proteins such as P32, RPO30, and GPCR [ 30 , 31 ]. For diagnosis, LSDV, sheep and goat poxviruses can be distinguished by real-time PCR technique [ 32 ]. The differentiation among natural genotypic targets of either vaccine or field strain genomes was developed by using a universal TaqMan probe to cover the field, vaccine, and recombinant strains of LSD [ 33 ]. The virulent LSDV from the vaccine strain was established by the restriction fragment length polymorphism [ 34 ].

Necropsy and Histopathological Finding

Infection with LSDV classically causes an acute disease with fever, depression, and appearance of nodules and lesions in the skin. The clinical signs of LSD are lymph nodes to form a skin blister about 2–5 cm in diameter on the skin, such as head, neck, legs, breast, and genitals. The necrotic nodules were ulcerated and formed deep scabs [ 35 ]. Larger vesicles may become necrotic and scarred for several months, while smaller vesicles heal faster. Blisters or ruptures of the cyst can be found. The mucus in the mouth, gastrointestinal tract, trachea, and lungs may be seen with edema. For asymptomatic infection, lesions are found in the subcutaneous or muscular layer. After the postmortem, lesions are often found in the respiratory organs, gastrointestinal tract, breast, lungs, bladder, kidneys, uterus, or testicles [ 36 ]. Histopathological examination of nodular skin biopsies showed edema, hyperemia, acanthosis, hydropic degeneration, and hyperkeratosis in epidermis [ 37 ].

Epidemiology (Susceptibility Hosts, Prevalence in Other Countries)

Lumpy skin disease virus was first investigated in Zambia in 1929 and was endemic in Africa during 1988–1989. However, the disease had been transmitted to middle Asian countries, including Saudi Arabia, Iran, Israel, and Iraq, by 1990 [ 6 ]. The incidence of LSD worldwide in 2016–2020 included African and Asian countries, and the epidemiology dynamically changed between the years. The prevalence and incidence of LSD detected by RT-PCR were reviewed during 2016–2020 ( Figure-3 ). In 2016, the prevalence of LSD in the countries was 4.66% in Iraq [ 38 ], 4.77% in Uganda [ 39 ], 5.00% in Nigeria [ 40 ], 6.00% in Saudi Arabia [ 23 ], 7.22%–18.0% in Ethiopia [ 41 , 42 ], and 12.9%–22.00% in Kazakhstan [ 43 , 44 ]. In 2017, the prevalence of LSD in countries was 24.00% in Egypt [ 45 ] and 5.67% in Ethiopia [ 46 ]. In 2018, the prevalence of LSD was 29.00% in Russia [ 47 ] and 31.20%–88.80% in Egypt [ 48 – 50 ]. In 2019, the prevalence of LSD was 10.00% in Bangladesh [ 51 ], 19.50% in China [ 52 ], 22.28%–27.50% in Egypt [ 36 , 53 ], and 37.66% in India [ 54 ]. In 2020, the prevalence of LSD was 3.00%–6.00% in Myanmar [ 55 ], 4.85% and 53.20% in Nepal [ 56 , 57 ], 13.93% in India [ 58 ], and 78.00% in Bangladesh [ 59 ]. In 2021–2022, the prevalence of LSD was 70% in Egypt [ 60 ], 4.17% in Thailand [ 61 ], 5.9% in Mongolia [ 62 ], and 36.2% in Ethiopia [ 63 ].

An external file that holds a picture, illustration, etc.
Object name is Vetworld-15-2764-g003.jpg

Prevalence of lumpy skin disease positively detected by polymerase chain reaction assay during 2016–2022. [Source: base map from the public Geo-Informatics and Space Technology Development Agency (GISTDA) using ArcGIS software (ESRI Inc., Redlands, CA, USA)].

Prevalence of LSD in Southeast Asia

Lumpy skin disease was shown to be distributed to South Asia through Southeast Asia (SEA) in 2019–2020 [ 1 ]. The first detection was found in the upper part of Vietnam in 2020 by the World Organization for Animal Health or Office International des Epizooties; OIE. The virus was isolated, identified, and investigated. The virus was similar to that endemic in Russia in 2017 and in China in 2019, indicating the disease was introduced from China–Vietnam border, and became distributed through 27 provinces in the country [ 64 , 65 ]. Thereafter, LSD was transmitted to other countries in SEA, such as Laos, Cambodia, Thailand, and Myanmar [ 55 , 66 , 67 ]. In Malaysia, the disease has been reported but some cases have not been confirmed as LSD [ 68 ]. In Indonesia and the Philippines, LSD has not been found and critical notification in the prevention and control of LSD were concerned [ 69 , 70 ]. In Thailand, at least 65 of 76 Provinces had reported infected animals to OIE by April, 2021. The Thai government produced strategies for prevention and controls, such as non–transfer of the animals from endemic areas, and vaccination in cattle and buffalo [ 71 ], which resulted in a sharp drop in morbidity rate in cattle in Thailand in 2022. Our experiences are agreeable with the encounter of LSD in Turkey in 2013 where uncontrolled animal movement is an important risk factor for spreading LSD to the neighboring countries, including Balkan, Caucasus, Iran, and Asia. Therefore, when any case occurs, isolation, quarantine, and vector control are necessary and have to be applied immediately [ 72 ].

Prevention Control and Eradication (Risk Factors, Regulations, Actions, and Vaccinations)

The stability of viruses in ambient conditions for a long period has certainly been established. It can persist in desiccated lesions on skin for 25–50 days and persist for many months in the dark environment in animal sheds. Veterinary education is needed for livestock workers to enable the performance of timely diagnoses of the disease to diminish the spread of the disease. Effective treatment against LSD has not been recognized. Symptomatic treatment for anti–inflammatory symptoms and antibiotics for preventing secondary microbial infection was used. Supportive therapy such as the Vitamin B–complex, Vitamin AD3E to retain the feeding capacity, and reproductive maintenance was frequently used ( Table-3 ) [ 73 – 75 ].

Therapeutic agents for LSD treatment.

Therapeutic agentsPharmacological effectsReferences
EnrofloxacinAntibiotic[ , ]
Oxytetracycline[ , ]
Penicillin[ ]
Cephalosporin[ ]
Tetracycline[ ]
Fluoroquinolone[ ]
Chlorpheniramine maleateAntihistamine[ , ]
MeloxicamNonsteroidal anti–inflammatory[ , ]
Dexamethasone suspensionSteroidal anti–inflammatory[ ]

LSD=Lumpy skin disease

Disease spreading can occur by infected animals or contaminated equipment or vectors. Early outbreaks can be controlled if the animal population is quarantined, sanitation on equipment or locality and biosafety. Controlling the movement or quarantine of newly imported animals for at least 3–4 weeks before being imported to the farm is one of the practices during the epidemic period [ 26 ]. Blood–sucking insects are the main vectors that cause the rapid spread of the disease; therefore, destroying breeding grounds, removing manure and cleaning with pesticides to disinfect and eliminate vectors regularly are recommended. Moreover, vaccination is an important preventive measure to reduce the spread and severity of the disease.

To prevent and control strategies, including the assortment of risks and affected animals, movement restrictions, and compulsory and consistent vaccination are recommended, including the following:

Restriction of animal movement

The movement of animals infected with LSDV and/or effects of LSD must be exactingly prohibited to prevent the distribution of the disease to other areas and crossover to non-infected farms. The prevention of transboundary movements and the restriction of animal movement should be strict. Animals with skin lesions should be investigated and should be isolated for assessment.

Control the distribution of vectors

The distribution of arthropod vectors’ movement and spread is risk factors for LSD transmission. Insecticides and traps are frequently used in livestock to prevent the disease.

Vaccination

The immunization of LSD by live attenuated vaccines has been effectively used in endemic areas. Antigenic homology of Carpipoxvirus , including SPPV, GTPV, and LSDV, cross-protection of immune response was beneficial [ 1 , 4 , 24 , 76 ]. A live attenuated vaccine is commercially available for LSD eradication. Sheep pox vaccine from SPPV and GTPV is used for control in countries with high LSD outbreaks. Types of LSD vaccines are as follows: (1) Attenuated LSDV vaccines (Neethling vaccines) are the currently effective vaccine to prevent LSD in cattle. The effective control success possibility is 80% in the livestock, (2) Attenuated SPPV vaccines are suitable for the areas that SPPV and LSDV outbreak and (3) Attenuated Gorgan GTPV vaccine is suitable for the areas where outbreaks are a combination of SPPV and LSDV [ 1 ]. The live attenuated LSDV vaccine (Neethling vaccines), which is used in livestock worldwide, is the only ubiquitous LSDV vaccine. After vaccination, the immunity is raised within 10–30 days. This vaccine is recommended at any age unless contemporaries showing signs of infection have already occurred.

Lumpy skin disease is an infectious disease in large ruminants, cattle, and domestic water buffalo. This disease regularly occurs in Africa, Europe, and some regions of Asia and spread to Southeast Asia in 2020. The clinical signs range from subclinical to high fever, lymph node enlargement, and apparent nodules over the entire body, followed by developing necrotic tissue and scarring. Although LSD has a low mortality rate, the development of lesions can cause complications and has no specific treatment. Successful prevention of LSD is vaccination together with vector control, controlling animals, especially from endemic regions, and monitoring the situations of LSD outbreaks continuously by disease surveillance.

Authors’ Contributions

KR: Performed literature search and drafted the manuscript. SukP: Reviewed and edited the manuscript. SP: Conceived the study, reviewed and edited the manuscript, and performed final manuscript revision. All authors have read and approved the final manuscript.

Acknowledgments

The authors are thankful to the One Health Research Unit and the Veterinary Infectious Disease Research Unit, Mahasarakham University, Thailand, for supporting the study. The authors did not receive any funds for this study.

Competing Interests

The authors declare that they have no competing interests.

Publisher’s Note

Veterinary World remains neutral with regard to jurisdictional claims in published map and institutional affiliation.

Advertisement

Supported by

What We Know About the Global Microsoft Outage

Airlines to banks to retailers were affected in many countries. Businesses are struggling to recover.

  • Share full article

Video player loading

By Eshe Nelson and Danielle Kaye

Eshe Nelson reported from London and Danielle Kaye from New York.

Across the world, critical businesses and services including airlines, hospitals, train networks and TV stations, were disrupted on Friday by a global tech outage affecting Microsoft users.

In many countries, flights were grounded, workers could not get access to their systems and, in some cases, customers could not make card payments in stores. While some of the problems were resolved within hours, many businesses, websites and airlines continued to struggle to recover.

What happened?

A series of outages rippled across the globe as information displays, login systems and broadcasting networks went dark.

The problem affecting the majority of services was caused by a flawed update by CrowdStrike , an American cybersecurity firm, whose systems are intended to protect users from hackers. Microsoft said on Friday that it was aware of an issue affecting machines running “CrowdStrike Falcon.”

But Microsoft had also said there was an earlier outage affecting U.S. users of Azure, its cloud service system. Some users may have been affected by both. Even as CrowdStrike sent out a fix, some systems were still affected by midday in the United States as businesses needed to make manual updates to their systems to resolve the issue.

George Kurtz, the president and chief executive of CrowdStrike, said on Friday morning that it could take some time for some systems to recover.

essay on lumpy virus in english

How a Software Update Crashed Computers Around the World

Here’s a visual explanation for how a faulty software update crippled machines.

How the airline cancellations rippled around the world (and across time zones)

Share of canceled flights at 25 airports on Friday

essay on lumpy virus in english

50% of flights

Ai r po r t

Bengalu r u K empeg o wda

Dhaka Shahjalal

Minneapolis-Saint P aul

Stuttga r t

Melbou r ne

Be r lin B r anden b urg

London City

Amsterdam Schiphol

Chicago O'Hare

Raleigh−Durham

B r adl e y

Cha r lotte

Reagan National

Philadelphia

1:20 a.m. ET

essay on lumpy virus in english

We are having trouble retrieving the article content.

Please enable JavaScript in your browser settings.

Thank you for your patience while we verify access. If you are in Reader mode please exit and  log into  your Times account, or  subscribe  for all of The Times.

Thank you for your patience while we verify access.

Already a subscriber?  Log in .

Want all of The Times?  Subscribe .

IMAGES

  1. (PDF) Mechanical transmission of lumpy skin disease virus by

    essay on lumpy virus in english

  2. Influenza virus

    essay on lumpy virus in english

  3. (PDF) Isolation and Molecular Characterization of Lumpy Skin Disease

    essay on lumpy virus in english

  4. 10 lines on Lumpy Virus in English

    essay on lumpy virus in english

  5. Excretion of Lumpy Skin Disease Virus followingVaccination, 978-3-639

    essay on lumpy virus in english

  6. Essay writing on Coronavirus in english || Coronavirus essay with heading

    essay on lumpy virus in english

VIDEO

  1. Grading Sophia test 🤦😯🧐#grading #school #test #teacher #correction #shorts #professor #asmr #viral

  2. Lumpy Virus Ki Vaccine Laga Di

  3. Essay writing on Corona virus #english #essaywritinginenglish #coronavirus

  4. Essay on Coronavirus in English

  5. Lumpy virus disease #goumata #shorts #cowvideos #cowdisease #cure

  6. virus English or Spanish

COMMENTS

  1. Lumpy skin disease, an emerging transboundary viral disease: A review

    Lumpy skin disease is an emerging bovine viral disease, which is endemic in most African countries and some Middle East ones, and the elevated risk of the spread of disease into the rest of Asia and Europe should be considered. The recent rapid spread of disease in currently disease‐free countries indicates the importance of understanding the ...

  2. Lumpy Skin Disease: A Systematic Review of Mode of Transmission, Risk

    Ticks transmit several viruses, e.g., Flaviviridae, that cause encephalitis-like diseases (e.g., tick-borne encephalitis virus, Kumlinge virus and louping ill virus), and Bunyaviridae, responsible of hemorrhagic fevers (e.g., Nairobi sheep disease virus and Crimean-Congo hemorrhagic fever virus). Thus, the role of ticks as biological vectors of ...

  3. Lumpy skin disease: history, current understanding and research gaps in

    1. Introduction. The previous decade has seen an unprecedented increase in scientific reports and publications on lumpy skin disease virus (LSDV), mainly due to its spread into new geographic regions and its adverse economic impacts on a global scale leading to greater interest in the disease and the causative agent (Bianchini et al., 2023).LSDV is a double-stranded DNA (dsDNA) poxvirus virus ...

  4. Transmission of lumpy skin disease virus: A short review

    Lumpy skin disease has moved beyond its native range in Africa via poorly understood mechanisms or pathways. •. Transmission implicates movement of live animals, naturally, or "human-assisted", coupled with either vector-borne, or contact modes. •. Although transmission via contact appears to be of low efficacy, arthropods could play a ...

  5. Lumpy skin disease, an emerging transboundary viral disease: A review

    1 INTRODUCTION. Lumpy skin disease (LSD), a major threat to stockbreeding, can cause acute or subacute disease in cattle and water buffalo (Givens, 2018; Tuppurainen, Venter, et al., 2017).All ages and breeds of cattle are affected, but especially the young and cattle in the peak of lactation (Tuppurainen et al., 2011).The reason why the World Organization for Animal Health (OIE) has placed ...

  6. Lumpy Skin Disease: Epidemiology, Phylogeny and Transmission

    This article presents the results of a study on the susceptibility of laboratory animals to the lumpy skin disease virus (LSDV). Mice weighing 15-20 g, hamsters weighing 40-60 g, guinea pigs weighing 600-1200 g, and rabbits weighing 2.5-3 kg were used in this study.

  7. Non-vector-borne transmission of lumpy skin disease virus

    The transmission of "lumpy skin disease virus" (LSDV) has prompted intensive research efforts due to the rapid spread and high impact of the disease in recent years, especially in Eastern ...

  8. (PDF) A review on current epidemiology and molecular studies of lumpy

    and ide ntification of lumpy skin disease virus from naturally infected buf faloes at Ka luobia, Egy pt. Global Veterinar ia, 7, Pages-234-237, DOI : 10.1186/s12917-022-03398- y.

  9. Unravelling the genomic origins of lumpy skin disease virus in recent

    Lumpy skin disease virus (LSDV) belongs to the genus Capripoxvirus and family Poxviridae. LSDV was endemic in most of Africa, the Middle East and Turkey, but since 2015, several outbreaks have been reported in other countries. In this study, we used whole genome sequencing approach to investigate the origin of the outbreak and understand the genomic landscape of the virus. Our study showed ...

  10. Global Burden of Lumpy Skin Disease, Outbreaks, and Future Challenges

    2.1. Lumpy Skin Disease Virus. Lumpy skin disease is a WOAH-marked highly contagious vector-borne emerging transboundary pox-viral infection of bovine species [23,24].The disease is caused by lumpy skin disease virus (LSDV), which belongs to the genus Capripoxvirus (CaPV) under the subfamily of Chordopoxvirinae within the family of Poxviridae [].The genus capripoxvirus is comprised of sheep ...

  11. Lumpy skin disease: A comprehensive review on virus biology

    Lumpy skin disease: A comprehensive review on virus biology, pathogenesis, and sudden global emergence Gaurav Moudgil †1, Jatin Chadha †1, Lavanya Khullar 1, Sanjay Chhibber 1, and Kusum Harjai1,* 1 Department of Microbiology, Panjab University, Chandigarh, India † Both the authors contributed equally to this work. * To whom correspondence should be addressed: Prof. Kusum Harjai ...

  12. Viruses

    The spread of lumpy skin disease (LSD) to free countries over the last 10 years, particularly countries in Europe, Central and South East Asia, has highlighted the threat of emergence in new areas or re-emergence in countries that achieved eradication. This review aimed to identify studies on LSD epidemiology. A focus was made on hosts, modes of transmission and spread, risks of outbreaks and ...

  13. Transmission of lumpy skin disease virus: A short review

    The role of modeling in the epidemiology and control of lumpy skin disease: a systematic review. Modeling has made a significant contribution in addressing challenges associated with the epidemiology and control of LSD, especially in the areas of risk factors, disease transmission, diagnosis and forecasting as well as intervention strategies.

  14. Lumpy Skin Disease Virus

    Lumpy skin disease virus, sheeppox virus, and goatpox virus are isolated from African cattle, sheep, and goats of Asia and Africa, respectively. Lumpy skin disease is found in most African countries south of 10°N latitude and is thought to be vector borne (Losos, 1986 ). All capripoxviruses can be spread by direct contact within a herd.

  15. Lumpy skin disease

    Lumpy skin disease (LSD) is an infectious disease in cattle caused by a virus of the family Poxviridae, also known as Neethling virus.The disease is characterized by fever, enlarged superficial lymph nodes, and multiple nodules (measuring 2-5 centimetres (1-2 in) in diameter) on the skin and mucous membranes, including those of the respiratory and gastrointestinal tracts.

  16. (PDF) Lumpy Skin disease: Review of literature

    Abstract. Lumpy skin disease (LSD) causes hu ge economic losses in the livestock industry. It is. caused by Lumpy skin di sease virus ( LSDV), which belon gs to th e family Poxviridae, with the ...

  17. Molecular characterization of lumpy skin disease virus from recent

    Lumpy skin disease (LSD) is a contagious viral transboundary disease listed as a notifiable disease by the World Organization of Animal Health (WOAH). The first case of this disease was reported in Pakistan in late 2021. Since then, numerous outbreaks have been documented in various regions and provinces across the country. The current study primarily aimed to analyze samples collected during ...

  18. Understanding the research advances on lumpy skin disease: A

    Etiology. LSD is a viral contagious cattle disease caused by Lumpy skin disease virus (LSDV; Murphy et al., 2008).The virus is a large linear double-stranded DNA genomes of 151 kb and belongs to one of the Capripoxvirus genus, subfamily Chordopoxvirniae, family Poxviridae (Tulman et al., 2001; Bhanuprakash et al., 2006; Moss, 2007; K, 2014).Viruses of the Poxviridae family are very similar in ...

  19. Lumpy Skin Disease Essay

    514 Words. 3 Pages. Open Document. INTRODUCTION: Lumpy skin disease (LSD) is viral in origin called "Neethling virus" belongs to a family of Poxviridae which includes pox virus for both sheep and goat (sheep pox virus and goat pox virus), and it is infectious and eruptive disease the insects play important role as a vector by biting specially ...

  20. What Is Lumpy Skin Disease, Can It Infect Humans?

    Whether humans can get infected by the lumpy skin disease-causing virus or not is the biggest question right now. In a statement, the World Organisation for Animal Health (WOAH) said that LSD has ...

  21. 10 lines on Lumpy Virus in English

    What is lumpy Virus?/ Short essay on Lampy Virus. The lumpy virus is spreading in cows in many states of the country. According to the Global Alliance for Vaccines and Immunization (GAVI), the lumpy virus is a disease transmitted by cows and buffaloes. This is a type of skin disease which is caused by a virus called Capripoxvirus. It spreads ...

  22. PDF Lumpy Skin Disease: A Systematic Review of Mode of Transmission, Risk

    (www.scopus.com(accessed on 1 September 2022)), with the search term "Lumpy Skin Disease". Only English-written articles, with an available abstract, and published between January 1980 and September 2022, were extracted. Editorials and books were excluded. These articles investigated LSD hosts, transmission modes, risk factors of an outbreak

  23. Lumpy skin disease: A newly emerging disease in Southeast Asia

    Virus and Classification. Lumpy skin disease virus (LSDV) is a virus in the family Poxviridae, subfamily Chordopoxviridae, genus Capripoxvirus.The genus Capripoxvirus comprises three viruses; SPPV, GTPV, and LSDV. Lumpy skin disease virus is large-sized (230-260 nm) enclosed in a lipid enveloped with a genome of approximately 150 kilobase pairs (kbp) and shared 97% identity in the ...

  24. What We Know About the Global Microsoft Outage

    Across the world, critical businesses and services including airlines, hospitals, train networks and TV stations, were disrupted on Friday by a global tech outage affecting Microsoft users.