Hostname: page-component-68c7f8b79f-qcl88 Total loading time: 0 Render date: 2025-12-18T13:26:14.657Z Has data issue: false hasContentIssue false

Current state of knowledge about African swine fever: a review

Published online by Cambridge University Press:  10 December 2025

Zi-bin Li
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Bao-bao Wang
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Yong-yu Gao
Affiliation:
College of Life and Health, Dalian University, Dalian, China College of Animal Medicine, Jilin Agricultural University, Changchun, China
Yu-han Xian
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Hong-sheng Feng
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Hang Jin
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Hai-yang Li
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Si-yu Yang
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Chen-jun Sang
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Yu-die Cao
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Yue Tang
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Yong-xin Cui
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Zhi-qiang Ding
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Hui He
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
Feng-shan Gao*
Affiliation:
College of Life and Health, Dalian University, Dalian, China The Dalian Animal Virus Antigen Epitope Screening and Protein Engineering Drug Developing Key Laboratory, Dalian, China
*
Corresponding author: Feng-shan Gao; Email: gfsh0626@126.com
Rights & Permissions [Opens in a new window]

Abstract

African swine fever (ASF) is a highly contagious animal disease caused by African swine fever virus (ASFV). It is listed by the World Organization for Animal Health (WOAH) as an animal disease subject to statutory reporting. ASFV, a large, enveloped double-stranded DNA virus with high genomic complexity, exhibits a case fatality rate of up to 100%, posing a significant threat to the global pig industry and food safety. To date, the absence of a safe commercial ASFV vaccine primarily stems from challenges in identifying immunogenic viral antigens, insufficient characterization of ASFV pathogenesis, and limited understanding of the virus’s immune evasion mechanisms. Here, we review the pathogenic characteristics (morphological structure, clinical symptoms, and epidemiological characteristics), molecular biological characteristics, and infection mechanism of ASFV, as well as the immune response mechanism, vaccine research, and the latest information on ASFV in other areas. This review will be in favour of understanding the current state of knowledge of ASF and developing effective vaccines to control this disease.

Information

Type
Review Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press.

Introduction

African swine fever (ASF) is a viral disease that causes high mortality in pigs. The disease is caused by African swine fever virus (ASFV; Cowan, Reference Cowan1961; Montgomery, Reference Montgomery1921). ASFV, the sole member of the genus Asfivirus, possesses a double-stranded DNA genome with a complex structure. Notably, it is the only known tick-borne DNA virus transmitted biologically by soft ticks of the genus Ornithodoros (Liu et al., Reference Liu, He, Zhu, Li, Wang, Li, Pan, Wang, Chu, Yang, Zhang, He, Li, Li and Zhang2024; Ramirez-Medina et al., Reference Ramirez-Medina, Vuono, Silva, Rai, Valladares, Pruitt, Espinoza, Velazquez-Salinas, Borca and Gladue2022). ASFV can infect domestic pigs, wild boars, and Ornithodoros, which is known as a vector and reservoir of ASFV (Chenais et al., Reference Chenais, Depner, Ebata, Penrith, Pfeiffer, Price, Stahl and Fischer2022; Pereira de Oliveira et al., Reference Pereira de Oliveira, Hutet, Paboeuf, Duhayon, Boinas, Perez de Leon, Filatov, Vial and Le Potier2019). ASF was first reported in Kenya in 1921, and subsequently, ASFV spread throughout the African continent (Montgomery, Reference Montgomery1921; Mur et al., Reference Mur, Atzeni, Martinez-Lopez, Feliziani, Rolesu and Sanchez-Vizcaino2016). ASF first appeared in Portugal in 1957 and again in 1960, at which point it quickly spread into Spain, France, Malta, Belgium, Italy, and the Netherlands (Gaudreault et al., Reference Gaudreault, Madden, Wilson, Trujillo and Richt2020). In 1971, Cuba was the first country in North America to report ASF (1971; Tury et al., Reference Tury, Ramos and Urquiaga1973), and the virus was believed to have been introduced from Spain (Costard et al., Reference Costard, Wieland, de Glanville, Jori, Rowlands, Vosloo, Roger, Pfeiffer and Dixon2009). ASFV was also reported in Ukraine in 1977 (Korennoy et al., Reference Korennoy, Gulenkin, Gogin, Vergne and Karaulov2017). In 1978, a new ASF incursion occurred in Sardinia (Italy). ASF was further reported in the late 1970s in several Caribbean island countries, including Cuba (1978, with the last occurrence in 1980), Dominican Republic (1978, with the last occurrence in 1981), and Haiti (1979, with the last occurrence in 1984) (Costard et al., Reference Costard, Wieland, de Glanville, Jori, Rowlands, Vosloo, Roger, Pfeiffer and Dixon2009). ASF was reported in Brazil in 1978 and eradicated in the 1980s (Lyra, Reference Lyra2006). By the mid-1990s, ASFV was eradicated outside Africa, except for an isolated outbreak in Portugal in 1999, caused by its introduction into a shelter for pigs infested by Ornithodoros erraticus ticks, and the island of Sardinia (Italy) (p72 genotype I), where it has remained endemic since 1978 (Boinas et al., Reference Boinas, Wilson, Hutchings, Martins and Dixon2011; Laddomada et al., Reference Laddomada, Rolesu, Loi, Cappai, Oggiano, Madrau, Sanna, Pilo, Bandino, Brundu, Cherchi, Masala, Marongiu, Bitti, Desini, Floris, Mundula, Carboni, Pittau, Feliziani, Sanchez-Vizcaino, Jurado, Guberti, Chessa, Muzzeddu, Sardo, Borrello, Mulas, Salis, Zinzula, Piredda, De Martini and Sgarangella2019; Mur et al., Reference Mur, Atzeni, Martinez-Lopez, Feliziani, Rolesu and Sanchez-Vizcaino2016; Pavone et al., Reference Pavone, Iscaro, Dettori and Feliziani2023). After a new introduction from Africa to Georgia in 2007 (Gogin et al., Reference Gogin, Gerasimov, Malogolovkin and Kolbasov2013; Rowlands et al., Reference Rowlands, Michaud, Heath, Hutchings, Oura, Vosloo, Dwarka, Onashvili, Albina and Dixon2008), ASFV spread from the Black Sea port of Poti across the Caucasus region; after reaching the Russian Federation in 2007, it spread to the eastern territories of the European Union in 2014, to Poland and the Baltic countries in 2014, and to Moldova in 2016. In August 2018, the virus spread to China, the world’s largest pig-producing country, and it is currently spreading in Asian countries including Republic of Korea, India, Malaysia, Vietnam, Japan, and Philippines (Hsu et al., Reference Hsu, Chang, Otake, Molitor and Perez2024; Kim et al., Reference Kim, Cho, Lee, Kim, Nah, Kim, Kim, Hwang, Sohn, Choi, Kang and Kim2020; Ligue-Sabio et al., Reference Ligue-Sabio, Lacaba, Mijares, Murao and Alviola2025; Liu et al., Reference Liu, Liu, Shan, Wei, An, Shen and Chen2020; Oh et al., Reference Oh, NA, Vn, Dao, Cho, Lee, Jung, Do, Kim, Bok, Hur and Lee2023; Rajukumar et al., Reference Rajukumar, Senthilkumar, Venkatesh, Singh, Patil, Kombiah, Tosh, Dubey, Sen, Barman, Chakravarty, Dutta, Pegu, Bharali and Singh2021) and some Eastern European countries including Russia, Ukraine, Poland, Romania, and Lithuania (Anastasia et al., Reference Anastasia, Elena, Daria, Tatiana, Olga, Ruslan, Timofey and Elena2025; Bezymennyi et al., Reference Bezymennyi, Tarasov, Kyivska, Mezhenska, Mandyhra, Kovalenko, Sushko, Hudz, Skorokhod, Datsenko, Muzykina, Milton, Sapachova, Nychyk, Halka, Frant, Huettmann, Drown, Gerilovych, Mezhenskyi, Bortz and Lange2023; Chernyshev et al., Reference Chernyshev, Igolkin, Zinyakov and Chvala2024; Dhollander et al., Reference Dhollander, Cattaneo, Cortinas Abrahantes, Boklund, Szczotka-Bochniarz, Mihalca, Papanikolaou, Mur, Balmos, Frant, Gal-Cison, Kwasnik, Rozek, Malakauskas, Masiulis, Turcinaviciene, Rusina, Aminalragia-Giamini, Chesnoiu, Jazdzewski, Rola, Barbuceanu and Stegeman2025; Igolkin et al., Reference Igolkin, Mazloum, Zinyakov, Chernyshev, Schalkwyk, Shotin, Lavrentiev, Gruzdev and Chvala2024). The current virus isolate is referred to as the Georgia strain and is grouped within genotype II (Rowlands et al., Reference Rowlands, Michaud, Heath, Hutchings, Oura, Vosloo, Dwarka, Onashvili, Albina and Dixon2008). In the absence of an effective vaccine, ASFV has caused severe economic losses worldwide. The most effective control methods are isolating the affected area and slaughtering infected animals. Due to the complex structure, high variability, and immune evasion of ASFV, the development of various vaccines has failed, and there are still few approved commercial vaccines (Bosch-Camós et al., Reference Bosch-Camós, López and Rodriguez2020). This article introduces the latest research based on the status of ASFV.

Pathogenic characteristics

Morphological and structural characteristics of ASFV

ASFV is a large, complex double-stranded DNA arbovirus and a member of the Asfarviridae family (Blome et al., Reference Blome, Franzke and Beer2020). ASFV has a complex multilayered structure with an overall icosahedral morphology (Fig. 1), and the average diameter of the particles is 260–300 nm. ASFV can be divided into five layers from the inside to the outside (Fig. 2). The innermost layer is the central genomic nucleolus, and the second layer is the thick protein core shell covering the nucleolus. The third layer is the inner lipid membrane. The fourth layer is an icosahedral protein capsid, and ASFV acquires its outermost envelope, the fifth layer, when budding through the host plasma membrane. The core-shell diameter is 180 nm, and it is covered by a 70 Å-thick lipid bilayer membrane (Wang et al., Reference Wang, Zhao, Wang, Zhang, Wang, Gao, Li, Wang, Bu, Rao and Wang2019). The maximum diameter of the main structural capsid of ASFV is 250 nm, and its structure has been solved at a resolution of 4.1 Å. The capsid structure consists of multiple proteins, including one major capsid protein (p72) and four minor proteins (M1249L, p17, p49, and H240R). These capsid proteins consist of 12 five-symmetrical and 20 three-symmetrical structures. Small capsid proteins (p17, p49, and M1249L) form a complex network beneath the outer shell. The capsid is fixed by adjacent proteins that assemble together to maintain its stability. Three 100 nm-long M1249L proteins run along each edge, connecting two adjacent five- and three-symmetrical regions and forming an extensive intermolecular network with other capsid proteins to enable the formation of the capsid skeleton (Alejo et al., Reference Alejo, Matamoros, Guerra and Andres2018; Liu et al., Reference Liu, Ma, Qian, Zhang, Tan, Lei and Xiang2019b).

Figure 1. Morphological structure of ASFV (Liu et al., Reference Liu, Ma, Qian, Zhang, Tan, Lei and Xiang2019b). (Reproduced directly from Liu et al., Reference Liu, Ma, Qian, Zhang, Tan, Lei and Xiang2019b Cell Host Microbe with permission from the Cell Press).

Figure 2. Schematic diagram of ASFV’s structure.

Clinical symptoms and pathological changes upon ASFV infection

According to the virulence of the virus, infection is divided into the following categories based on clinical symptoms: severe acute, acute, subacute, chronic, and atypical course (Avagyan et al., Reference Avagyan, Hakobyan, Baghdasaryan, Arzumanyan, Poghosyan, Bayramyan, Semerjyan, Sargsyan, Voskanyan, Vardanyan, Karalyan, Hakobyan, Abroyan, Avetisyan, Karalova, Semerjyan and Karalyan2024; Sanchez-Cordon et al., Reference Sanchez-Cordon, Floyd, Hicks, Crooke, McCleary, McCarthy, Strong, Dixon, Neimanis, Wikstrom-Lassa, Gavier-Widen and Nunez2021; Wang et al., Reference Wang, Zhang, Hou, Yang and Wen2021b). Severe acute infection is caused by a highly virulent strain, which leads to the sudden death of pigs without clinical symptoms, resulting in a 100% mortality rate. Within 3–4 days after infection, pigs with acute disease typically exhibit a high fever, haemorrhaging of the reticuloendothelial system, loss of appetite and lethargy, and dyskinesia. The lymph nodes are usually swollen, fragile, and bleeding in post-mortem examinations (Sanchez-Cordon et al., Reference Sanchez-Cordon, Floyd, Hicks, Crooke, McCleary, McCarthy, Strong, Dixon, Neimanis, Wikstrom-Lassa, Gavier-Widen and Nunez2021). Some pigs with acute disease exhibit lymph node congestion, digestive problems, including gastrointestinal tract haemorrhage and oedema; central nervous system oedema; and perivascular haemorrhage. Within 4–15 days after infection, the case fatality rate is nearly 100% (Sanchez-Cordon et al., Reference Sanchez-Cordon, Floyd, Hicks, Crooke, McCleary, McCarthy, Strong, Dixon, Neimanis, Wikstrom-Lassa, Gavier-Widen and Nunez2021; Wang et al., Reference Wang, Zhang, Hou, Yang and Wen2021b). Medium-virulence strain infection causes a subacute infection, characterized by mild symptoms, primarily manifesting as respiratory distress, joint pain, and swelling, with low mortality (30–70%). Abortion may result from ASFV infection in pregnant sows. Low-virulence isolates can cause chronic symptoms, such as faeces containing mucus, miscarriage, vomiting, diarrhoea, bleeding, breathing changes, and other symptoms, but the mortality rate is low (Sánchez-Cordón et al., Reference Sánchez-Cordón, Montoya, Reis and Dixon2018). The incubation period of ASFV generally ranges from 5 to 15 days, with some individual cases extending to 28 days. The atypical course has emerged in recent years, caused by lower virulent strains of ASFV, characterized by a significantly prolonged disease duration, sometimes reduced mortality, prolonged viremia, and no apparent clinical symptoms (Avagyan et al., Reference Avagyan, Hakobyan, Baghdasaryan, Arzumanyan, Poghosyan, Bayramyan, Semerjyan, Sargsyan, Voskanyan, Vardanyan, Karalyan, Hakobyan, Abroyan, Avetisyan, Karalova, Semerjyan and Karalyan2024; Sun et al. Reference Sun, Zhang, Wang, He, Zhang, Wang, Wang, Huang, Xi, Huangfu, Tsegay, Huo, Sun, Tian, Xia, Yu, Li, Liu, Guan, Zhao and Bu2021).

ASFV is highly stable in the environment; therefore, it can remain infectious for several days in faeces or urine under suitable conditions and can survive even longer in organic matter (Mazur-Panasiuk et al., Reference Mazur-Panasiuk, Zmudzki and Wozniakowski2019). In the secretions, excrement, blood, and tissues of infected pigs, its activity can be maintained for several weeks, and it can also hold its activity for a long time in smoked and air-dried foods, such as cured dry ham, in which it can survive for 150 days. ASFV is also highly resistant to acidic and alkaline conditions, and the virus can remain stable at pH values ranging from 3.9 to 11.5. The virus is resistant to low temperatures but is sensitive to high temperatures. It can survive for 30 min at 55 ℃, 10 min at 60 ℃, several weeks at 25–37 ℃, and more than 1 year at 4 ℃. Additionally, it can survive in frozen pork up to 2 years (Mazur-Panasiuk and Woźniakowski, Reference Mazur-Panasiuk and Woźniakowski2020). Numerous studies have demonstrated the efficacy of various lipid solvents, detergents, and disinfectants containing aldehyde-, phenol-, or iodide-based compounds in inactivating ASFV (Juszkiewicz et al., Reference Juszkiewicz, Walczak, Wozniakowski, Pejsak and Podgorska2025; Ni et al., Reference Ni, Chen, Yun, Xie, Ye, Hua, Zhu and Zhang2023). In particular, commercially available disinfectants such as compound peroxymonosulfate, sodium dichloroisocyanurate, and glutaraldehyde have been widely implemented in biosecurity protocols, especially in China, where ASF outbreaks have been most prevalent (Juszkiewicz et al., Reference Juszkiewicz, Walczak, Mazur-Panasiuk and Woźniakowski2020). These compounds have shown significant virucidal activity against ASFV in both laboratory and field conditions.

Epidemiological characteristics of ASFV

ASF was first observed in settlers’ pigs in Kenya, Africa, in 1909 and was first reported in 1921 as a distinct disease from classical swine fever (Montgomery, Reference Montgomery1921). Reports of ASF in South Africa and Angola (Steyn, 1932) followed. ASF broke out in the Lisbon area of Portugal In 1957 and again in 1960 (Wilkinson, Reference Wilkinson and Pensaert1989); then, it spread to the Iberian Peninsula and spread further in Europe as well as in the Caribbean and Brazil in South America (Costard et al., Reference Costard, Wieland, de Glanville, Jori, Rowlands, Vosloo, Roger, Pfeiffer and Dixon2009). During the same time, ASF was also found in Malawi and Mozambique (Abreu et al., Reference Abreu, de Valadão, Limpo Serra, Ornelas Mário and Sousa Montenegro1962; Matson, Reference Matson1960; Mendes, Reference Mendes1971). Researchers have shown that some intermediate animals, such as warthogs (Phacochoerus africanus) and Ornithodoros moubata complex ticks, were associated with virus ASFV infection and transmission at that time in sub-Saharan Africa (Montgomery, Reference Montgomery1921; Penrith et al., Reference Penrith, Vosloo, Jori and Bastos2013). ASFV strains are currently grouped into 24 genotypes based on the reference B646 L gene (p72), and all the gene variants are associated with the disease (Quembo et al., Reference Quembo, Jori, Vosloo and Heath2018). The 24 genotypes are clustered into three distinct evolutionary lineages, each confined to a broad geographical area. Lineage I comprised 14 genotypes (including I, XVII, II, XXIV, XXI, V, VI, XVIII, VII, XXII, IV, III, XX, and XIX) associated with viruses from West and southern Africa. In contrast, lineage II (including XIV, XVI, XV, XIII, XII, VIII, and XI) consisted of viruses from East Africa and lineage III (including X, IX, and XIII) consisted of viruses from the Great Lakes Region of East and Central Africa (Bastos et al., Reference Bastos, Penrith, Cruciere, Edrich, Hutchings, Roger, Couacy-Hymann and RT2003; Boshoff et al., Reference Boshoff, Bastos, Dube and Heath2014; Quembo et al., Reference Quembo, Jori, Vosloo and Heath2018). Most of these genotypes have been associated with ASF outbreaks in various parts of sub-Saharan Africa (Mulumba-Mfumu et al., Reference Mulumba-Mfumu, Saegerman, Dixon, Madimba, Kazadi, Mukalakata, Oura, Chenais, Masembe, Stahl, Thiry and Penrith2019). Genotype I has been shown to predominate in Central Africa (Kouakou et al., Reference Kouakou, Michaud, Biego, Gnabro, Kouakou, Mossoun, Awuni, Minoungou, Aplogan, Awoume, Albina, Lancelot and Couacy-Hymann2017) and West Africa (Hakizimana et al., Reference Hakizimana, Nyabongo, Ntirandekura, Yona, Ntakirutimana, Kamana, Nauwynck and Misinzo2020). Most Tanzanian ASF outbreaks have been linked to genotypes II, IX, X, XV, and XVI of the virus (Misinzo et al., Reference Misinzo, Magambo, Masambu, Yongolo, Van Doorsselaere and Nauwynck2011; Njau et al., Reference Njau, Machuka, Cleaveland, Shirima, Kusiluka, Okoth and Pelle2021), while in other areas in Mauritius, Madagascar, Tanzania, and Zimbabwe, the prevalent highly virulent strain of ASFV has been identified as genotype II (Wambura et al., Reference Wambura, Masambu and Msami2006). In addition, genotypes IX and X were identified in Uganda and Kenya (Atuhaire et al., Reference Atuhaire, Afayoa, Ochwo, Mwesigwa, Okuni, Olaho-Mukani and Ojok2013; Gallardo et al., Reference Gallardo, Okoth, Pelayo, Anchuelo, Martin, Simon, Llorente, Nieto, Soler, Martin, Arias and Bishop2011). From 1968 to 1995, ASFV of the p72 genotype I was present in European countries, including Malta, Sardinia, Italy, France, Belgium, and the Netherlands. The prevalent genotype in the Gulu district is genotype IX (Barongo et al., Reference Barongo, Stahl, Bett, Bishop, Fevre, Aliro, Okoth, Masembe, Knobel and Ssematimba2015). ASF was first reported in Cuba in North America in 1971, after which ASF epidemics caused by ASFV of genotype I also occurred in Brazil, the Dominican Republic, and other countries (Ankhanbaatar et al., Reference Ankhanbaatar, Auer, Ulziibat, Settypalli, Gombo-Ochir, Basan, Takemura, Tseren-Ochir, Ouled Ahmed, Meki, Datta, Soumare, Metlin, Cattoli and Lamien2023). By the mid-1990s, ASFV had been eradicated outside Africa, except for an isolated outbreak in Portugal in 1999 and an infection on the island of Sardinia (Italy) (p72 genotype I) (Boinas et al., Reference Boinas, Wilson, Hutchings, Martins and Dixon2011; Laddomada et al., Reference Laddomada, Rolesu, Loi, Cappai, Oggiano, Madrau, Sanna, Pilo, Bandino, Brundu, Cherchi, Masala, Marongiu, Bitti, Desini, Floris, Mundula, Carboni, Pittau, Feliziani, Sanchez-Vizcaino, Jurado, Guberti, Chessa, Muzzeddu, Sardo, Borrello, Mulas, Salis, Zinzula, Piredda, De Martini and Sgarangella2019; Pavone et al., Reference Pavone, Iscaro, Dettori and Feliziani2023). After a new introduction of the virus from Africa into Georgia in 2007, ASFV genotype II spread to the Caucasus and reached the Russian Federation (Beltrán-Alcrudo et al., Reference Beltrán-Alcrudo, Lubroth, Depner and De La Rocque2008). It then spread further from Russia to the European Union in 2014 (Dixon et al., Reference Dixon, Stahl, Jori, Vial and Pfeiffer2020). In August 2018, the virus reached China, the world’s largest pig-producing country, and is currently spreading in most of Southeast Asia and Oceania (Hakizimana et al., Reference Hakizimana, Nyabongo, Ntirandekura, Yona, Ntakirutimana, Kamana, Nauwynck and Misinzo2020; Mulumba-Mfumu et al., Reference Mulumba-Mfumu, Saegerman, Dixon, Madimba, Kazadi, Mukalakata, Oura, Chenais, Masembe, Stahl, Thiry and Penrith2019; Njau et al., Reference Njau, Machuka, Cleaveland, Shirima, Kusiluka, Okoth and Pelle2021; Quembo et al., Reference Quembo, Jori, Vosloo and Heath2018; Urbano and Ferreira, Reference Urbano and Ferreira2022). Since then, the epidemiological situation of ASF has continued to deteriorate. In July 2021, after nearly 40 years of absence, ASF was reintroduced in the Dominican Republic and later in Haiti. In January 2022, it reappeared on the Italian mainland (Urbano and Ferreira, Reference Urbano and Ferreira2022). Although ASFV genotype II is currently the dominant strain, genotype I has also been identified in China (Cao et al., Reference Cao, Lu, Wu and Zhu2022). Furthermore, some highly lethal genotype I and II recombinant ASFV strains have been detected in pigs (Zhao et al., Reference Zhao, Sun, Huang, Ding, Zhu, Zhang, Shen, Zhang, Zhang, Ren, Wang, Li, He and Bu2023). The new defined genotype XXIV was identified in soft ticks from Gorongosa National Park in Africa (Quembo et al., Reference Quembo, Jori, Vosloo and Heath2018). The World Animal Health Information System of the World Organization for Animal Health (WOAH) has released a map of the real-time global distribution of ASF (Fig. 3) (https://wahis.woah.org/#/event-management), from which it is evident that ASF is spreading rapidly in Africa, Asia, Europe, and other regions. Research on ASF prevention, control, and treatment is therefore urgent. In addition to wild pigs and soft ticks, the natural hosts of ASFV, the virus can also spread through direct or indirect contact with infected pigs and their products. The ASFV transmission cycle mainly includes the sylvatic cycle (wild boar–soft tick–wild boar cycle), wild boar–domestic pig cycle, and domestic pig–domestic pig cycle (Karger et al., Reference Karger, Pérez-Núñez, Urquiza, Hinojar, Alonso, Freitas, Revilla, Le Potier and Montoya2019). Given the above, vigilance and continuous surveillance will be crucial while the quest for an ASF vaccine continues.

Note: The number on the map represents the number of outbreaks of ASF that occurred in the area from 3 October 2018 to 29 June 2025.

Figure 3. Global distribution of ASFV as of 29 June 2025.

In the sylvatic cycle (wild boar–soft tick–wild boar), soft ticks play an important role in transmission between wild boars, and ASFV can be transmitted between different soft ticks. After a soft tick bites an infected wild boar, the virus can survive in the soft tick for a long time and then enter the next cycle when the tick bites another wild boar. The blood, excreta, and secretions of infected wild pigs contain high levels of the virus. Therefore, through direct or indirect contact with infected pigs or contaminated surfaces, feed, and water, ASFV spreads rapidly between wild pigs, leading to long-term stability of the sylvatic cycle (Mazur-Panasiuk et al., Reference Mazur-Panasiuk, Zmudzki and Wozniakowski2019). Wild boar–domestic pig cycle: ASFV spreads rapidly in wild boar herds; this virus can cross borders through the wildlife corridor and become a source of infection in domestic pigs. In regions with high wild boar densities, direct or indirect contact with domestic pigs, particularly in low-biosecurity settings where pigs are allowed to forage outdoors, facilitates the transmission of viral pathogens to domestic populations. ASFV spreads among domestic pigs primarily through direct contact with infected individuals, with transmission occurring via the oronasal route or skin abrasions upon exposure to blood, excreta, or secretions. Notably, pigs recovering from infection with moderately or low-virulent strains may become chronic carriers, perpetuating viral circulation within herds. Human-mediated transmission further exacerbates spread through fomites, including contaminated clothing, footwear, vehicles, equipment, and slaughter tools. Viral dissemination via contaminated pork products, swill, feed, faeces, and bedding can result in both localized and long-distance transmission, including cross-border spread (Golnar et al., Reference Golnar, Martin, Wormington, Kading, Teel, Hamer and Hamer2019; Niederwerder et al., Reference Niederwerder, Stoian, Rowland, Dritz, Petrovan, Constance, Gebhardt, Olcha, Jones, Woodworth, Fang, Liang and Hefley2019).

Molecular biological characteristics of ASFV

ASFV has a 170–194 kb linear dsDNA genome (Fig. 4) consisting of three main regions: a left variable region (LVR) of approximately 38–47 kb, a right variable region (RVR) of approximately 13–16 kb, and a conserved central region of 125 kb (Dixon et al., Reference Dixon, Chapman, Netherton and Upton2013; Hakizimana et al., Reference Hakizimana, Ntirandekura, Yona, Nyabongo, Kamwendo, Chulu, Ntakirutimana, Kamana, Nauwynck and Misinzo2021). The end of the LVR contains a hairpin loop structure formed by 37nt partial base pairing or an inversion. Both the LVR and the RVR contain multigene families (MGFs) that can cause variations in the genome lengths of different ASFV strains (Hakizimana et al., Reference Hakizimana, Ntirandekura, Yona, Nyabongo, Kamwendo, Chulu, Ntakirutimana, Kamana, Nauwynck and Misinzo2021). The ASFV genome includes 150–167 open reading frames (ORFs) encoding 68 structural proteins and over 100 nonstructural proteins (Salas and Andrés, Reference Salas and Andrés2013). These proteins include enzymes involved in evading host defences, DNA replication and repair, and gene expression regulation, as well as structural proteins for virus assembly (Cackett et al., Reference Cackett, Matelska, Sýkora, Portugal, Malecki, Bähler, Dixon and Werner2020; Jia et al., Reference Jia, Ou, Pejsak, Zhang and Zhang2017). The functional proteins of ASFV include envelope proteins, capsid proteins, nucleocapsid proteins, and several binding proteins involved in viral replication (Andres et al., Reference Andres, Charro, Matamoros, Dillard and Abrescia2020). Notably, epidemiological studies have provided evidence that mutations in functional proteins enable the virus to evade the host immune system, which is an essential barrier to antiviral therapy and vaccine development.

Figure 4. Annotated genome structure of ASFV (Dixon et al., Reference Dixon, Chapman, Netherton and Upton2013; Hakizimana et al., Reference Hakizimana, Ntirandekura, Yona, Nyabongo, Kamwendo, Chulu, Ntakirutimana, Kamana, Nauwynck and Misinzo2021). (Reproduced directly from Dixon et al. Reference Dixon, Chapman, Netherton and Upton2013 Virus Research with permission from the Elsevier and slightly modified according to Hakizimana et al., Reference Hakizimana, Ntirandekura, Yona, Nyabongo, Kamwendo, Chulu, Ntakirutimana, Kamana, Nauwynck and Misinzo2021.) A, The overall genomic structure of open reading frames (ORFs) in ASFV, represented by Georgia 2007/1. B, The arrangement of genes (open reading frames [ORFs]) in the genome of ASFV, represented by Georgia 2007/1. The direction and size of the arrows indicate the transcriptional direction and length of the ORFs, respectively. The different colours indicate the various functions of the ORFs. The black arrow indicates ORFs encoding enzymes and factors involved in genome replication, repair, or transcription. The grey arrow indicates ORFs that can be expressed as proteins. The pink arrow suggests ORFs associated with escape from the host immune system. The yellow arrow indicates ORFs with other predicted functions. The white arrow indicates ORFs with unknown functions. The turquoise, blue, green, brown, and mauve arrows indicate members of multigene families. Deletion of the genes shown in red can reduce virulence.

Comparison of the ASFV genome with that of other dsDNA viruses in the NCLDV group

ASFVs, as members of the Asfarviridae family, also belong to the group of nucleocytoplasmic large DNA viruses (NCLDVs), which infect a wide range of eukaryotic species, including amoebae, algae, fish, amphibians, arthropods, birds, and mammals (Campillo-Balderas et al., Reference Campillo-Balderas, Lazcano, Cottom-Salas, Jacome and Becerra2023). This group of viruses has linear or circular double-stranded DNA genomes whose sizes span approximately one order of magnitude, from 100 to 2500 kbp. The International Committee on Taxonomy of Viruses currently recognizes seven taxonomic families as members of the NCLDVs (ICTV 2020): Ascoviridae, Asfarviridae, Iridoviridae, Marseilleviridae, Mimiviridae, Phycodnaviridae, and Poxviridae (Campillo-Balderas et al., Reference Campillo-Balderas, Lazcano, Cottom-Salas, Jacome and Becerra2023; Tidona C, Reference Tidona C2011). These viral families possess special metabolic characteristics because they either synthesize their DNA exclusively in the cytoplasm or undergo first-stage replication and early transcription in the host nucleus, and late transcription in the cytoplasm. Therefore, they have also been classified within the phylum Nucleocytoviricota (Asgari et al., Reference Asgari, Bideshi, Bigot, Federici, Cheng and Ictv Report2017; Campillo-Balderas et al., Reference Campillo-Balderas, Lazcano, Cottom-Salas, Jacome and Becerra2023). The genome sizes of Asfarviridae, Marseilleviridae, and Mimiviridae are 170, 370, and 1180 kb, respectively, while other members, including Iridoviridae, Ascoviridae, Phycodnaviridae, and Poxviridae, have variable ranges of 100–220, 150–190, 150–400, and 130–380 kb, respectively (Yutin et al., Reference Yutin, Wolf, Raoult and Koonin2009). ASFV, as a representative of Asfarviridae, seems to be a relatively petite virus compared to other NCLDVs.

Nucleocapsid protein

There are 110 highly conserved ORFs in the ASFV genome, two of which encode the ASFV polyprotein precursors pp220 and pp62 (or pp60). The mature particles form after these two precursor proteins are hydrolysed and processed to assemble the core-shell of the virus (Liu et al., Reference Liu, Ma, Qian, Zhang, Tan, Lei and Xiang2019b).

The pp220 protein encoded by the CP2475 L gene has a molecular weight of 281.5 kDa and is participates in late-stage viral infection. As a protein scaffold, polyprotein pp220 can bind the nucleocapsid to the inner membrane of the virus and promote the assembly of the empty nucleocapsid. After polyprotein pp220 is hydrolysed and processed by the virus-encoded SUMO-like protease S273 R, it is converted into the mature virus particle proteins p150, p37, p34, p14, and the recently identified p5 (Andres et al., Reference Andres, Charro, Matamoros, Dillard and Abrescia2020). The relative molecular weight of the polyprotein pp62 encoded by the CP530R gene is 60.5 kDa, and this protein plays an important role in the replication, assembly, and maturation of ASFV. Upon hydrolysis by the S273R protease, pp62 is converted into the mature virus particle proteins p35 and p15 (Fu et al., Reference Fu, Zhao, Zhang, Zhang, Li, Zhang, Wang, Sun, Jiao, Chen, Guo and Rao2020), and the recently identified p8 (Alejo et al., Reference Alejo, Matamoros, Guerra and Andres2018). The mature virus particle proteins p150, p37, p14, p34, and p35 (Li et al., Reference Li, Fu, Zhang, Zhao, Li, Geng, Sun, Wang, Chen, Jiao, Cao, Guo and Rao2020) and p15 account for approximately 30% of the total viral protein and constitute the main components of the virus nucleocapsid. The p37 transport protein plays an important role in the ASFV replication cycle. The p34 protein has a protective effect on trypsin. The p150 protein is a component of the membrane. p5 and p8 are structural components of the ASFV virus particles, and with other polymer proteins, they are packed into the core of a virus particle (Alejo et al., Reference Alejo, Matamoros, Guerra and Andres2018).

Capsid proteins

The p72 protein is the main structural protein of the capsid, with a relative molecular mass of 73.2 kDa, and is the key antigen protein encoded by the B646L (VP72) gene (Liu et al., Reference Liu, Ma, Qian, Zhang, Tan, Lei and Xiang2019a). Newly synthesized p72 is evenly distributed among the soluble cytoplasm, the membrane, and the endoplasmic reticulum, where it accumulates to form a large capsid or matrix precursor. p72 has a highly conserved sequence with good antigenicity and immunogenicity; however, the anti-p72 antibody does not play a decisive role in antibody-mediated immune protection (Gallagher and Harris, Reference Gallagher and Harris2020).

The p49 protein is encoded by the B438L gene and is an essential structural protein for the formation of infectious virus particles (Dixon et al., Reference Dixon, Chapman, Netherton and Upton2013). The gene ORF consists of 1317 nucleotides, encodes 438 amino acids, and is expressed in viral factories. Upon p49 binding to the membrane, other proteins are recruited to the inner viral membrane to initiate assembly into a complete capsid. In the absence of the p49 protein, the virus particle has an abnormal tubular structure instead of a complete icosahedral symmetric structure.

The p14.5 protein, encoded by the E120R gene, and also known as the pE120R protein, has a molecular weight of 13.6 kDa. As a DNA-binding protein, p14.5 plays an important role in the formation of the core and capsid of the virus. It participates in the intracellular transport of virus particles. It is essential for the transfer of virus particles from the viral factories, where virus particles assemble in discrete cytoplasmic areas near the nucleus, to the plasma membrane of the cells (Andres et al., Reference Andres, Alejo, Salas and Salas2002; Jia et al., Reference Jia, Ou, Pejsak, Zhang and Zhang2017). pE120R is reportedly involved in regulating ASFV immune evasion, which inhibits the cGAS-STING-mediated immune response (He et al., Reference He, Yuan, Ma, Zhao, Yang, Zhang, Han, Wan and Zhang2022).

Inner envelope proteins

p54, encoded by the E183L gene, is also a critical ASFV antigenic structural protein with a relative molecular mass of 25 kDa. The protein is located in the outer lipid membrane of the virus particle. This protein can recruit the envelope precursor to the assembly site and plays a crucial role in virus adsorption and invasion, as well as in the inoculation of attenuated strains into pigs to induce specific antibodies (Rodríguez et al., Reference Rodríguez, García-Escudero, Salas and Andrés2004). Studies have shown that the p54 protein can stimulate the body to produce antibodies against this protein (Neilan et al., Reference Neilan, Zsak, Lu, Burrage, Kutish and Rock2004). In the early stage of the virus infection cycle, p54 protein inhibitors can suppress the activity of proteins related to virus attachment. Transient p54 expression in Vero cells demonstrated that this protein can activate caspase-3 and caspase-9, mediating cell apoptosis (Wang et al., Reference Wang, Zhang, Hou, Yang and Wen2021a).

p30 is encoded by the CP204 L gene and has a molecular mass of 30 kDa. The expression of the p30 protein indicates that the ASFV virus has entered the cell and is in the early stages of infection. The yeast two-hybrid system was used to screen for cellular proteins in the porcine macrophage cDNA library that may interact with the p30 protein, and heterogeneous ribonucleoprotein K (hnRNP-K) was identified as the most likely cellular ligand for the p30 protein (Hernaez et al., Reference Hernaez, Escribano and Alonso2008). It is now generally accepted that the p30 protein is one of the primary antigenic structural proteins in ASFV, and p30 has been identified as a subunit vaccine candidate (Petrovan et al., Reference Petrovan, Yuan, Li, Shang, Murgia, Misra, Rowland and Fang2019).

The pE248R protein, encoded by the E248R gene, is a crucial internal envelope protein involved in forming disulfide bonds (Alejo et al., Reference Alejo, Matamoros, Guerra and Andres2018). Studies have shown that the adsorption and internalization of the virus are not affected by the lack of the pE248R protein. Deletion of the pE248R protein inhibits the replication of virus particles in host cells, as well as the early and late expression of viral genes.

p17 is encoded by the ORF in the D117L gene, has a molecular weight of 13.1 kDa, and is a transmembrane protein (Xia et al., Reference Xia, Wang, Liu, Shao, Ao, Xu, Jiang, Luo, Zhang, Chen, Meurens, Zheng and Zhu2020). This protein facilitates the further assembly of the viral membrane precursor into an icosahedral intermediate, thereby increasing the viability of the virus. Blocking p17 expression inhibits the hydrolysis of the pp220 and pp62 proteins.

p12 is a membrane protein expressed in the late stage of ASFV infection, is encoded by an ORF in the O61 R gene, is 6.7 kDa in length, and is located in the inner envelope of the virus (Salas and Andrés, Reference Salas and Andrés2013). p12 plays an important role in ASFV attachment to susceptible cells. In addition to the abovementioned main proteins, p22, encoded by KP177L, and the pE199L, pH108R, and pH108R proteins play known roles, among which pE248R and pE199L are related to virus invasion.

Outer envelope protein

CD2v is the only characterized protein in the outer envelope and is encoded by an ORF in the EP402 R gene. It is also known as pEP402R and is similar to the T lymphocyte surface adhesion receptor CD2, with a relative molecular weight of 105 kDa. CD2v is a glycoprotein composed of a signal peptide, a transmembrane region, and two immunoglobulin-like domains, and it is located near the viral factories. Due to the presence of ligands on the surface of pig red blood cells, ASFV is easily internalized by red blood cells. Studies have shown that CD2v-mediated red blood cell adsorption facilitates the spread of the virus in the host (Burmakina et al., Reference Burmakina, Malogolovkin, Tulman, Xu, Delhon, Kolbasov and Rock2019). Proline repeats in the cytoplasmic region of CD2v can interact with the actin adaptor protein SH3P7, participating in vesicle transport and signal transduction (Rodríguez et al., Reference Rodríguez, Yáñez, Almazán, Viñuela and Rodriguez1993). The differences in the repeated sequences of different strains can be used to analyse the serotype of a strain. The CD2v protein exhibits good immunogenicity and is a choice for the development of cross-protective vaccines. Additionally, the CD2v protein can impair lymphocyte function and is involved in cell adhesion, enhancing virulence and immune regulation. CD2v also plays essential roles in ASFV tissue tropism, immune evasion, and the promotion of virus replication in the host (Malogolovkin et al., Reference Malogolovkin, Burmakina, Titov, Sereda, Gogin, Baryshnikova and Kolbasov2015).

Other proteins

The p10 protein is encoded by the K78R gene and has a relative molecular mass of 8.4 kDa. This protein participates in the adsorption of ASFV and can bind single-stranded or double-stranded DNA. Amino acids 71–73 of p10 play an important role in the introduction of virus particles into nuclei. The pA104 R protein may participate in the transcription, DNA replication, and genome packaging of ASFV (Alejo et al., Reference Alejo, Matamoros, Guerra and Andres2018).

Replication of ASFV and the host immune response

The replication mechanism of ASFV

ASFV enters host cells through clathrin-mediated and dynein-dependent endocytosis, as well as macropinocytosis, including actin-dependent endocytosis and endocytosis involving microtubule activity (Andrés, Reference Andrés2017; Hernaez et al., Reference Hernaez, Guerra, Salas and Andres2016). ASFV can enter cells through a process known as classic phagocytosis, also referred to as receptor-mediated endocytosis. However, the cell receptors and viral ligands involved are currently unknown (Galindo et al., Reference Galindo, Cuesta-Geijo, Hlavova, Muñoz-Moreno, Barrado-Gil, Dominguez and Alonso2015).

The endocytosed virions are partially disrupted in late endosomes, where they lose their outer membrane and outer capsid. The capsid decomposes in the acidic endosomal lumen (Matamoros et al., Reference Matamoros, Alejo, Rodríguez, Hernáez, Guerra, Fraile-Ramos and Andrés2020). After capsid degradation and host membrane fusion mediated by the pE248R protein, the ASFV core is released into the cytoplasm (Hernaez et al., Reference Hernaez, Guerra, Salas and Andres2016). The core is transported through microtubules to the viral factories, where virus-encoded RNA polymerase and transcription factors replicate the virus. Approximately, 20% of the genes in ASFV are involved in transcription and mRNA modification. The newly generated virus particles are released from infected cells through a process known as budding. ASFV disrupts the structure of the subnuclear domain and chromatin, inducing changes in nuclear structure, facilitating the inhibitory nuclear environment, and enabling the virus to replicate efficiently in host cells (Sánchez et al., Reference Sánchez, Quintas, Nogal, Castelló and Revilla2013; Simões et al., Reference Simões, Freitas, Leitão, Martins and Ferreira2019); however, the full role of the nucleus in this process is still unclear. Therefore, further research is needed.

The mechanism of the host immune response to ASFV

Early experimental studies have demonstrated that immune serum and monoclonal antibodies from convalescent pigs can neutralize many ASFV isolates and that these antibodies can prevent the virus from attaching to cells in the early stage (Dixon et al., Reference Dixon, Islam, Nash and Reis2019). Recent experiments have demonstrated that antibodies specific to ASFV can provide immune protection, but cannot offer cross-protection (O’Donnell et al., Reference O’Donnell, Holinka, Krug, Gladue, Carlson, Sanford, Alfano, Kramer, Lu, Arzt, Reese, Carrillo, Risatti and Borca2015). Antibody neutralization is a sensitive and easily reversible process, and ASFV aggregates easily and cannot be recognized by specific antibodies. Therefore, antibody-mediated protection cannot completely neutralize the virus. Viruses that are not neutralized by antibodies can cause persistent infection in pigs (Neilan et al., Reference Neilan, Zsak, Lu, Burrage, Kutish and Rock2004). Antibodies may neutralize the virus in two ways: in one process, the virus binds to Vero cells or porcine alveolar macrophages, and through saturation binding, the antibody blocks the virus from attaching to the cell (Sánchez et al., Reference Sánchez, Pérez-Núñez and Revilla2017); in the other process, the ASFV neutralization-mediated mechanism is activated to inhibit virus internalization by Vero cells or porcine alveolar macrophages in the presence of immune serum. In summary, ASFV-specific antibodies partially inhibit the virus’s attachment but cannot provide complete protection (Escribano et al., Reference Escribano, Galindo and Alonso2013; Sunwoo et al., Reference Sunwoo, Pérez-Núñez, Morozov, Sánchez, Gaudreault, Trujillo, Mur, Nogal, Madden, Urbaniak, Kim, Ma, Revilla and Richt2019).

An increasing number of experiments have demonstrated that the cellular immune response mediated by cytotoxic T lymphocytes (CTLs) plays a crucial role in the process of ASFV infection (Argilaguet et al., Reference Argilaguet, Perez-Martin, Nofrarias, Gallardo, Accensi, Lacasta, Mora, Ballester, Galindo-Cardiel, Lopez-Soria, Escribano, Reche and Rodriguez2012; Bosch-Camós et al., Reference Bosch-Camós, López, Collado, Navas, Blanco-Fuertes, Pina-Pedrero, Accensi, Salas, Mundt, Nikolin and Rodríguez2021). The classical method of antigen presentation by major histocompatibility complex (MHC) class I molecules is related to the activation of cellular immunity (Rock and Shen, Reference Rock and Shen2005). The protein produced by ASFV replication and translation in the cell is degraded by the host cell’s proteasome to produce a series of short peptides. Some of these short peptides are transported to the endoplasmic reticulum by transport associated with antigen processing (TAP). The heavy chain and light chain β2 m microglobulin (β2 m) of class I molecules assemble to form a heterodimer, and the protein is then processed and modified by the Golgi apparatus to achieve correct spatial folding. Finally, the protein is transported to the cell membrane surface, where it binds to and activates the CD8+ T lymphocyte receptor (TCR), thereby inducing cellular immune function that can kill and eliminate target cells infected by ASFV (Gao et al., Reference Gao, He, Quan, Jiang, Lin, Chen and Qu2017).

Experiments have shown that the targets of this virus-specific CTL protein are usually limited to MHC class I antigens (Netherton et al., Reference Netherton, Goatley, Reis, Portugal, Nash, Morgan, Gault, Nieto, Norlin, Gallardo, Ho, Sánchez-Cordón, Taylor and Dixon2019). Blocking target cell antigens with monoclonal antibodies against MHC class I antigens (monoclonal antibodies, MAbs) significantly decreases CTL activity. The absence of CD8+ T lymphocytes results in a lack of specific cellular immune protection (Zhu et al., Reference Zhu, Ramanathan, Bishop, O’Donnell, Gladue and Borca2019). ASFV epitopes (p72, p32, p54, and CD2v) fused with ubiquitin serve as immune antigens; in the absence of specific antibodies, partial protection against lethal attacks is provided (Oura et al., Reference Oura, Denyer, Takamatsu and Parkhouse2005). The CD2v/p32/p54 fusion antigen, expressed by the baculovirus BacMam, also provides partial protection (Argilaguet et al., Reference Argilaguet, Pérez-Martín, López, Goethe, Escribano, Giesow, Keil and Rodríguez2013). An ASFV DNA expression library lacking CD2v, p32, and p54 has been used to develop vaccines that can partially protect animals from virulent strain attacks (Lacasta et al., Reference Lacasta, Ballester, Monteagudo, Rodriguez, Salas, Accensi, Pina-Pedrero, Bensaid, Argilaguet, Lopez-Soria, Hutet, Le Potier and Rodriguez2014). Pigs inoculated with attenuated ASFV strains are protected from the same or closely related virulent strains with a degree of cross-protection. The BA71ACD2 strain, with artificially deleted CD2v, can induce immune cross-protection against both homologous and heterologous viral strains, and can produce specific CD8+ T lymphocyte immunity (Burmakina et al., Reference Burmakina, Malogolovkin, Tulman, Xu, Delhon, Kolbasov and Rock2019).

Cytokines play a crucial role in activating cellular immunity, thereby influencing the treatment of ASFV infection (Salguero et al., Reference Salguero, Sánchez-Cordón, Núñez, Fernández de Marco and Gómez-Villamandos2005). For example, TNF-α is a pleiotropic cytokine that can inhibit viral infections, regulate the function of immune cells, and the production of molecules that mediate inflammatory responses (Barrado-Gil et al., Reference Barrado-Gil, Del Puerto, Galindo, Cuesta-Geijo, García-Dorival, de Motes and Alonso2021). The main source of TNF-α is macrophages, and TNF-α transcription is regulated by several transcription factors, such as NF-κB, NFAT, and c-Jun. Overexpression of these transcription coactivators can restore the activity of the TNF-α promoter. In addition to the effects of cytokines, different vectors expressing the same ASFV-specific antigen (e.g., poxvirus, adenovirus, pseudorabies virus) can also induce various degrees of immune response, and mixed immunizing cocktails can induce strong immune protection (Lokhandwala et al., Reference Lokhandwala, Waghela, Bray, Sangewar, Charendoff, Martin, Hassan, Koynarski, Gabbert, Burrage, Brake, Neilan and Mwangi2017; Maeda et al., Reference Maeda, West, Hayasaka, Ennis and Terajima2005). Therefore, the CTL-mediated cellular immune response plays an important protective role against ASFV infection. Future research on cellular immunity will also be beneficial for the production of effective subunit vaccines, as it eliminates the effect of ASFV infection on host cells and provides long-term protection. Therefore, CTLs play a crucial role in eliminating ASFV infection.

The immune escape mechanism of ASFV

Repair mechanism of ASFV

When the virus invades, the host cell response causes DNA damage, leading to potentially lethal gene mutations in the viral genes or inhibition of the activity of viral DNA and RNA polymerases. ASFV utilizes the encoded DNA base excision repair (EBR) pathway enzyme to repair mutations, ensuring that the sequence is correctly copied (Redrejo-Rodríguez and Salas, Reference Redrejo-Rodríguez and Salas2014). The enzymes in the EBR pathway include DNA polymerase X repair, AP endonuclease, and DNA ligase. Additionally, the nucleotide-metabolizing enzyme and thymidine kinase encoded by ASFV are important for the effective replication of ASFV in macrophages. It is speculated that their actions increase the size of the dNTP library needed for viral genome replication (Redrejo-Rodríguez et al., Reference Redrejo-Rodríguez, Rodríguez, Suárez, Salas and Salas2013).

ASFV inhibits the type I interferon response

A series of proteins encoded by ASFV can inhibit the expression of type I interferons, cytokines, chemokines, adhesion molecules, interferon-stimulated genes (ISGs), and other immunoregulatory genes (Reis et al., Reference Reis, Abrams, Goatley, Netherton, Chapman, Sanchez-Cordon and Dixon2016). MGF360-15 R (A276R) inhibits IRF3 expression through a mechanism independent of the transcription factors IRF7 and NF-κB. The IRF3 receptor is part of a necessary pathway for activating the type I interferon response; therefore, the pathways activated by TLR3 and cytoplasmic sensing ultimately inhibit the induction of type I IFN. MGF505-7 R (A528R) inhibits the induction of type I IFN by inhibiting the transcription factors IRF3 and NF-κB, as well as inhibiting the type I and type II IFN signalling pathways (O’Donnell et al., Reference O’Donnell, Holinka, Sanford, Krug, Carlson, Pacheco, Reese, Risatti, Gladue and Borca2016). The ASFV pI329 L protein is similar to the cellular TLR3, a highly glycosylated protein expressed in the cell membrane. It can inhibit the TLR3-mediated induction of IFN-β and activation of NF-κB and IRF3. The pI329 L protein targets TRIF, and overexpression of TRIF can reverse the inhibitory effect of NF-κB and IRF3 activation (Zhu et al., Reference Zhu, Ramanathan, Bishop, O’Donnell, Gladue and Borca2019).

ASFV inhibits cell apoptosis

During viral entry and replication, ASFV inhibits the apoptosis and necrosis of the host cell through encoded proteins, ensuring that the virus has enough time for replication (Dixon et al., Reference Dixon, Sánchez-Cordón, Galindo and Alonso2017). The first early antiapoptotic protein encoded by the A179L gene is a homolog of Bcl-2 (Galindo et al., Reference Galindo, Hernaez, Díaz-Gil, Escribano and Alonso2008). The proapoptotic proteins in the BH3-only family include Bim, Bid, Puma, Noxa, Bmf, Bik, Bad, and Hrk. Bax and Bak are activated downstream of the BH3-only protein and play important regulatory roles in cell apoptosis. The A179L-encoded protein acts on upstream proapoptotic BH3 domain proteins (Bid, Bim, and Puma) and downstream Bak and Bax proteins to inhibit host cell apoptosis. Additionally, the A179L protein inhibits autophagy mediated by the BH3 domain of the autophagy regulator Beclin-only (Banjara et al., Reference Banjara, Caria, Dixon, Hinds and Kvansakul2017). The second late apoptosis inhibitor protein, A224L, is encoded by the A224L gene and is similar to the IAP gene-encoded apoptosis-like protein (IAP). A224L contains a single BIR motif at the NH2 end, and a 4-cysteine-type zinc finger domain is formed in the C-terminal region (Banjara et al., Reference Banjara, Shimmon, Dixon, Netherton, Hinds and Kvansakul2019). The expression of the A224L protein inhibited the apoptotic pathway involving caspase 3 and apoptosis induced by stimuli such as TNF-α signalling. The expression of the A224L protein can activate the NF-κB signalling pathway, enabling escape from host immunity and activating the transcription of many antiapoptotic genes, including IAP and Bcl-2 family members, to inhibit cell apoptosis (Portugal et al., Reference Portugal, Leitão and Martins2009).

Cell infection leads to a cellular stress response. As activated protein kinases phosphorylate translation initiation factors, eIF2-α reduces overall protein synthesis and targets CHOP, thereby activating the transcription of proapoptotic proteins. The ASFV DP71L protein interacts with PP1 and eIF-2α, thereby inducing the dephosphorylation of eIF-2α and inhibiting the activation of the proapoptotic transcription factor CHOP, which in turn inhibits the apoptosis induced by this pathway. pE153R contains a C-type lectin domain, and p53 inhibits cell apoptosis by activating the transcription of many apoptosis inhibitors (Hurtado et al., Reference Hurtado, Bustos, Granja, de León, Sabina, López-Viñas, Gómez-Puertas, Revilla and Carrascosa2011). Apoptosis can be observed in the later stages of infection because it may promote the spread of the virus through the apoptotic body and its subsequent absorption by monocytes/macrophages, thus preventing the inflammatory response caused by cell death through necrosis. Infected cells may secrete or present the cell surface factor TNF-α, which can induce the apoptosis of uninfected lymphocytes through the bystander effect. The large-scale destruction of lymphocytes inhibits the immune response mechanism.

The mechanism by which ASFV infection inhibits inflammation

ASFV has an elegant mechanism for evading the host immune defence system. ASFV can inhibit the secretion of IFN-α and TNF-α from macrophages, as well as the transcription of inflammatory cytokines, thereby reducing the level of mRNA transcription and translation (Golding et al., Reference Golding, Goatley, Goodbourn, Dixon, Taylor and Netherton2016). The pA238L protein encoded by the A238L gene is similar to the I-kB inhibitor of the NF-κB transcription factor and can inhibit NF-κB activation (Salguero et al., Reference Salguero, Gil, Revilla, Gallardo, Arias and Martins2008). In addition to activating the type I interferon promoter, the NF-κB transcription factor also transcriptionally activates many inflammatory responses. In resting cells, after calcineurin dephosphorylation, the NFAT factor is transferred to the nucleus to activate NFAT-dependent gene transcription. Meanwhile, pA238L binds to the calcium/calmodulin-regulated phosphatase calmodulin (CaN), inhibiting the pathway activated by the phosphatase and thus ultimately inhibiting the inflammatory response (Correia et al., Reference Correia, Ventura and Parkhouse2013).

ASFV vaccines

Although the first attempts to develop an ASF vaccine date back to the 1960s, the lack of safe and effective commercial vaccines persists, primarily due to the virus’s large genomic complexity and sophisticated immune evasion mechanisms. Early work by Malmquist et al. (Malmquist and Hay, Reference Malmquist and Hay1960) revealed that FMDV can replicate efficiently only in primary swine macrophages, which implies a hurdle for the large-scale commercial production of live attenuated vaccines (LAVs). Later, Malmqusit (Malmquist, Reference Malmquist1962) established the foundation by adapting the Hinde isolate to grow in primary pig kidney cell cultures, followed by serial passage in a porcine kidney (PK2a) cell line. Then, a series of other passage cell lines were used to propagate ASFV strains, including Vero cells for ASFV-G (Krug et al., Reference Krug, Holinka, O’Donnell, Reese, Sanford, Fernandez-Sainz, Gladue, Arzt, Rodriguez, Risatti and Borca2015) and COS-1 cells for the BA71CD2 strain (Monteagudo et al., Reference Monteagudo, Lacasta, Lopez, Bosch, Collado, Pina-Pedrero, Correa-Fiz, Accensi, Navas, Vidal, Bustos, Rodriguez, Gallei, Nikolin, Salas and Rodriguez2017), which suggests that the large-scale cultivation of ASFV strains appears feasible. However, the adaptation of ASFV field isolates to replicate in established cell lines through successive serial passaging is usually accompanied by significant modifications to the virus genome, which can even result in the loss of virus genes and phenotypic changes such as loss of the ability to replicate in swine. Therefore, the long-standing lack of suitable cell lines for the production of commercial ASFV vaccine is a great limitation until Borca et al. (Reference Borca, Rai, Ramirez-Medina, Silva, Velazquez-Salinas, Vuono, Pruitt, Espinoza and Gladue2021) recently developed a stable porcine cell line, Plum Island porcine epithelial cells (PIPEC), thereby enabling suitable commercial vaccine production of ASFV-G-DI177L/DLVR that replicates efficiently in PIPEC and primary swine macrophages. In addition, animals that have recovered from acute ASFV infection can only resist infection by closely related strains, while antibodies and T cells play vital roles in virus control (Sang et al., Reference Sang, Miller, Lokhandwala, Sangewar, Waghela, Bishop and Mwangi2020). Therefore, although vaccine design is feasible, safety is a core issue of ASF vaccines. The safety concerns associated with conventional LAVs remain an unavoidable challenge. These include risks such as incomplete effectiveness against heterogeneous strains, reversion to virulence under animal passages, unpredictable strain-specific genetic modifications, altered phenotypic traits, and even host-environment interactions, all of which could potentially contribute to large-scale ASFV outbreaks or persistent infection (Blome et al., Reference Blome, Friedrichs, Schafer and Beer2025; Cadenas-Fernandez et al., Reference Cadenas-Fernandez, Barroso-Arevalo, Kosowska, Diaz-Frutos, Gallardo, Rodriguez-Bertos, Bosch, Sanchez-Vizcaino and Barasona2024; Sun et al., Reference Sun, Huang, Zhang, Zhang, Shen, Zhang, Wang, Huo, Wang, Huangfu, Wang, Li, Liu, Sun, Tian, Xia, Guan, He, Zhu, Zhao and Bu2021; van den Born et al., Reference van den Born, Olasz, Meszaros, Goltl, Olah, Joshi, van Kilsdonk, Segers and Zadori2025). Currently, the ASFV vaccine development landscape is primarily focused on inactivated vaccines, attenuated vaccines, subunit vaccines, and live vector vaccines.

Inactivated vaccines

Inactivated vaccines are a class of conventional vaccines that use physical or chemical methods to inactivate pathogenic microorganisms, causing them to lose their pathogenicity while retaining their antigenicity, which activates the body’s immune system upon immunization. Inactivated vaccines have the advantages of simple preparation and low cost. Research on inactivated ASF vaccines began in the 1960s; however, extensive studies have demonstrated that these vaccines primarily elicit humoral immunity, failing to induce robust cellular immune responses mediated by cytotoxic T cells. Consequently, they are unable to provide adequate protection against ASF (Cadenas-Fernandez et al., Reference Cadenas-Fernandez, Sanchez-Vizcaino, van den Born, Kosowska, van Kilsdonk, Fernandez-Pacheco, Gallardo, Arias and Barasona2021; Zhang et al., Reference Zhang, Zhao, Zhang, Qin, Shan and Cai2023). Although some adjuvants (e.g., Polygen, Emulsigen-D) or immune enhancers may enhance cellular immunity and increase vaccine efficacy in inactivated vaccines of other pathogens (Dar et al., Reference Dar, Kalaivanan, Sied, Mamo, Kishore, Suryanarayana and Kondabattula2013; Kaur et al., Reference Kaur, Saxena, Rai and Bhatnagar2010; Milian-Suazo et al., Reference Milian-Suazo, Gutierrez-Pabello, Bojorquez-Narvaez, Anaya-Escalera, Canto-Alarcon, Gonzalez-Enriquez and Campos-Guillen2011), inactivated ASF vaccines, when formulated with these adjuvants, continue to be insufficient to confer sterilizing immunity or durable immunoprotection (Blome et al., Reference Blome, Gabriel and Beer2014; Xie et al., Reference Xie, Liu, Di, Liu, Gong, Chen, Li, Yu, Lv, Zhong, Song, Liao, Song, Wang and Chen2022).

Attenuated vaccines

LAVs can be viruses attenuated through cell passage, naturally occurring ASFV strains with reduced virulence, or strains that are attenuated using gene-editing techniques to delete virulence factors (Krug et al., Reference Krug, Holinka, O’Donnell, Reese, Sanford, Fernandez-Sainz, Gladue, Arzt, Rodriguez, Risatti and Borca2015; Monteagudo et al., Reference Monteagudo, Lacasta, Lopez, Bosch, Collado, Pina-Pedrero, Correa-Fiz, Accensi, Navas, Vidal, Bustos, Rodriguez, Gallei, Nikolin, Salas and Rodriguez2017; Mulumba-Mfumu et al., Reference Mulumba-Mfumu, Goatley, Saegerman, Takamatsu and Dixon2016; Urbano and Ferreira, Reference Urbano and Ferreira2022).

Research on attenuated ASFV vaccines obtained by cell passage was first reported in 1957 (Detray, Reference Detray1957). However, previous research has shown that attenuated ASFV vaccines generated by cell passage can only provide homologous protection in animal challenge experiments (Manso-Ribeiro et al., Reference Manso-Ribeiro, Lopez-Frazao and Sobral1963). Spain and Portugal have isolated attenuated ASFV strains as vaccines; these strains provide homologous protection and partial heterologous protection, but vaccinated pigs exhibit various degrees of side-effects from the vaccine: a debilitating chronic disease showing viraemia and fever in a late phase of infection in many vaccinated pigs was caused because of these vaccines were insufficiently tested for their safety; as a result, the use of these vaccines was stopped (Gavier-Widen et al., Reference Gavier-Widen, Stahl and Dixon2020; Leitao et al., Reference Leitao, Cartaxeiro, Coelho, Cruz, Parkhouse, Portugal, Vigario and Martins2001). In 2015, an attenuated variant of the ASFV strain Stavropol 01/08 (Stavropol) isolated from the Russian Federation was obtained after 24 continuous passages in homologous pig kidney cells (A4C2/9k) and 20 passages in heterologous African green monkey CV-1 cells. The virus lost pathogenicity in pigs but did not protect them from death after challenge with a virulent virus, indicating that the attenuated Stavropol 01/08 could not provide enough protection for animals (Balysheva et al., Reference Balysheva, Galnbek and Balyshev2015). In the same year, Krug et al. (Krug et al., Reference Krug, Holinka, O’Donnell, Reese, Sanford, Fernandez-Sainz, Gladue, Arzt, Rodriguez, Risatti and Borca2015) reported the complete attenuation of a virulent ASFV field isolate from the Republic of Georgia (ASFV-G) by 110 successive passages in Vero cells. However, immunization of swine with the fully attenuated virus did not confer protection against challenge with the virulent parental strain ASFV-G. Further detection revealed that amino acid substitutions and frameshift mutations occurred in the genome of the attenuated ASFV-G strain. According to Krug’s research, developing effective attenuated ASFV vaccines by cell passage is very difficult. Contemporaneously, Lacasta et al. (Reference Lacasta, Monteagudo, Jimenez-Marin, Accensi, Ballester, Argilaguet, Galindo-Cardiel, Segales, Salas, Dominguez, Moreno, Garrido and Rodriguez2015) reported that the cell culture-adapted strain E75CV1, obtained by adapting E75 to grow in the CV1 cell line, can potentially protect against homologous E75 virus challenge but was shown to have poor cross-protection against BA71; this finding indicated that E75CV1 provided partial heterologous protection against BA71. Recently, a novel naturally attenuated ASFV strain, ASFV-989, isolated from a blood sample of pigs inoculated with the thermo-attenuated ASFV Georgia 2007/1 strain, was reported by Bourry’s group to provide full clinical protection against challenge with Georgia 2007/1 through oronasal or intramuscular immunization (Bourry et al., Reference Bourry, Hutet, Le Dimna, Lucas, Blanchard, Chastagner, Paboeuf and Le Potier2022). ASFV-989 is a promising vaccine because it has not been genetically manipulated, which facilitates its use through open distribution in the wild, although a natural deletion of 7458 nucleotides located in the 5′ end encoding region of MGF 505/360, compared to that in the Georgia population (2007/1), was identified.

Another approach to developing attenuated vaccines is the screening of naturally attenuated, non-haem-adsorbing (non-HAD) strains and/or strains with reduced virulence that occur naturally during ASF epidemics (Urbano and Ferreira, Reference Urbano and Ferreira2022). Leitao et al. (Reference Leitao, Cartaxeiro, Coelho, Cruz, Parkhouse, Portugal, Vigario and Martins2001) reported a case in which a non-HAD, nonfatal genotype I ASFV strain, NH/P68, was isolated from a chronically infected pig and used to infect oronasally or intramuscularly to evaluate its effectiveness. The vaccinated pigs exhibited good cellular and humoral immunity and were protected against highly virulent ASFV/L60 challenge. Sanchez-Cordon et al. (Reference Sanchez-Cordon, Chapman, Jabbar, Reis, Goatley, Netherton, Taylor, Montoya and Dixon2017) also reported the naturally attenuated ASFV genotype I isolate OURT88/3, showing that intranasal immunization with low or moderate doses (103 and 104 TCID50) of OURT88/3 provided complete protection (100%) against challenge with the virulent genotype I OURT88/1 isolate. Recently, Barasona et al. (Reference Barasona, Gallardo, Cadenas-Fernandez, Jurado, Rivera, Rodriguez-Bertos, Arias and Sanchez-Vizcaino2019) developed a non-HAD, attenuated ASF virus of genotype II isolated in Latvia in 2017 (Lv17/WB/Rie1), which conferred 92% protection against challenge with a virulent ASF virus isolate (Arm07). Gallardo et al. (Reference Gallardo, Soler, Rodze, Nieto, Cano-Gomez, Fernandez-Pinero and Arias2019) also confirmed that pigs intramuscularly infected with Lv/17/WB/Rie1 ASFV developed nonspecific clinical signs and, in some cases, remained asymptomatic, exhibiting intermittent and weak viremia, along with a high antibody response. Furthermore, 2 months following primary infection with Lv17/WB/Rie1, the two pigs were fully resistant to challenge with the virulent genotype II HAD Latvian ASFV. Although naturally attenuated strains have the potential to be developed as LAVs, their residual virulence and safety concerns regarding dosage remain significant (Urbano and Ferreira, Reference Urbano and Ferreira2022).

A third strategy for developing live attenuated ASFV vaccines involves the use of gene-editing techniques to change the virus genome and obtain deletion mutant strains. The quality of the gene-edited vaccine is controllable, and this approach can be used to distinguish between vaccine-immunized and wild-infected animals. Researchers from the European Union ASFV Reference Laboratory (Gallardo et al., Reference Gallardo, Sánchez, Pérez-Núñez, Nogal, de León, Carrascosa, Nieto, Soler, Arias and Revilla2018) in Spain used the ASFV wild-type strain NH/P68 as the backbone to construct multiple protein-deficient strains by using gene-editing techniques. After immunization, the vaccine candidates fully protected pigs against homologous challenge with L60, although they failed to protect against genotype II Arm07. The protection rate against this strain was thus unsatisfactory. The most promising candidate gene for an ASFV vaccine is a deletion mutant produced by homologous recombination or the CRISPR/Cas9 technique (Perez-Nunez et al., Reference Perez-Nunez, Sunwoo, Garcia-Belmonte, Kim, Vigara-Astillero, Riera, Kim, Jeong, Tark, Ko, You and Revilla2022; Rathakrishnan et al., Reference Rathakrishnan, Moffat, Reis and Dixon2020). Through these methods, reasonable deletions of virulence genes and interferon inhibitors have been attempted in several strains. The deleted virulence-related genes included 9GL (B119L), UK (DP96R), CD2v (EP402R), DP148R, and various members of the multigene family (MGF; Sang et al., Reference Sang, Miller, Lokhandwala, Sangewar, Waghela, Bishop and Mwangi2020). Along these lines, the deletion of different types of I interferon inhibitors can prevent the activation of the attenuated virus. The same deletion exhibits varying degrees of attenuation in different strains, and the duration of immunity remains a weakness of this potential ASF vaccine (Arias et al., Reference Arias, de la Torre, Dixon, Gallardo, Jori, Laddomada, Martins, Parkhouse, Revilla and Rodriguez2017). For example, the candidate vaccine HLJ/18-7GD, based on the highly virulent Chinese ASFV strain HLJ/18, was constructed by deleting gene fragments encoding 1–7 different proteins, including MGF505-1R, MGF505-2R, MGF505-3R, MGF360-12 L, MGF360-13 L, MGF360-14 L, and CD2v (EP402R) (Chen et al., Reference Chen, Zhao, He, Liu, Wang, Zhang, Li, Shan, Chen, Zhang, Wang, Wen, Wang, Guan, Liu and Bu2020). To examine whether these gene-deleted viruses were attenuated in pigs, specific pathogen-free (SPF) piglets were inoculated intramuscularly, and body temperature and survival rate were measured for 3 weeks. Except for HLJ/18-7GD, which was significantly weakened in pigs and exhibited a strong protective effect against the parents of the HLJ/18 strain and the genotype II Georgia07-like ASFV (Wang et al., Reference Wang, Zhang, Li, Zhang, Chen, Zhang, Sun, Zhu, Liu, He, Bu and Zhao2024), the other gene-deleted strains were lethal (Chen et al., Reference Chen, Zhao, He, Liu, Wang, Zhang, Li, Shan, Chen, Zhang, Wang, Wen, Wang, Guan, Liu and Bu2020). Recently, Borca and Tran reported a highly effective ASFV vaccine obtained by deleting the I177L gene via gene-editing techniques, named ASFV-G-DeltaI177L; this vaccine can induce effective protection against its parental virulent virus strain, Georgia 2007, as well as against a virulent Vietnamese field strain isolated in 2019. In addition, large-scale experiments to detect virus shedding and transmission have confirmed that even under varying conditions, ASFV-G-∆I177L remains a safe live-attenuated vaccine (Borca et al., Reference Borca, Ramirez-Medina, Silva, Vuono, Rai, Pruitt, Holinka, Velazquez-Salinas, Zhu and Gladue2020; Tran et al., Reference Tran, Phuong, Huy, Thuy, Nguyen, Quang, Ngon, Rai, Gay, Gladue and Borca2022). Based on the ASFV-G-∆I177L vaccine, Borca et al. (Reference Borca, Ramirez-Medina, Espinoza, Rai, Spinard, Velazquez-Salinas, Valladares, Silva, Burton, Meyers, Clark, Wu, Gay and Gladue2024) recently developed an ASFV-G-ΔI177L/ΔEP402R mutant by deleting the EP402R gene, which encodes the viral CD2v protein responsible for mediating haemadsorption of swine erythrocytes. This deletion not only attenuates the virus but also enables differentiation between infected and vaccinated animals (DIVA capability), a critical feature for vaccine-based control strategies. Currently, a new AVAC ASF LIVE vaccine, produced from an attenuated genotype II ASF virus (ASFV) strain with the deletion of six MGF genes and cultured in a Diep’s macrophage (DMAC) cell line, has been officially licensed for use and commercialization in Vietnam (Diep et al., Reference Diep, Ngoc, Duc, Dang, Tiep, Quy, Tham and Doanh2025). It is reported that this vaccine has demonstrated safety and high efficacy in protecting pigs from genotype II ASFV infection. However, to determine whether this vaccine can protect animals from heterologous ASFV infection, further investigation is needed. Certain naturally immunized pigs can resist attacks from homologous strains, but most of these pigs cannot resist attacks from heterologous strains (O’Donnell et al., Reference O’Donnell, Holinka, Sanford, Krug, Carlson, Pacheco, Reese, Risatti, Gladue and Borca2016). In fact, researchers have made progress in cross-protection using live attenuated ASFV vaccines. In 2017, Monteagudo reported a new recombinant live attenuated ASFV vaccine, named BA71∆CD2 (genotype I-based), in which deletion of the viral CD2v (EP402R) gene greatly attenuated the virulent BA71 strain in vivo (Monteagudo et al., Reference Monteagudo, Lacasta, Lopez, Bosch, Collado, Pina-Pedrero, Correa-Fiz, Accensi, Navas, Vidal, Bustos, Rodriguez, Gallei, Nikolin, Salas and Rodriguez2017). Inoculation of pigs with this kind of vaccine provided full protection not only against lethal challenge with the parental BA71 but also against the heterologous E75 (both genotype I strains). Interestingly, 100% of the pigs immunized with BA71ΔCD2 also survived lethal challenge with Georgia 2007/1, the genotype II strain of ASFV currently circulating mainly in continental Europe, Asia, and Oceania. In 2020, Lopez extended this research using BA71∆CD2 as a tool for understanding ASFV cross-protection via the use of phylogenetically distant ASFV strains. They first observed that five out of six (83.3%) of the pigs immunized once with 106 PFU of BA71∆CD2 survived the tick bite challenge using Ornithodoros sp. soft ticks naturally infected with the RSA/11/2017 strain (genotype XIX, clade D, while BA71, E75 and Georgia2007/01 belonged to the same clade C). Second, they found that homologous prime boosting with BA71∆CD2 improved the survival rate only to 50% after Ken06.Bus (genotype IX, clade A) challenge, and all the infected pigs exhibited mild clinical signs consistent with ASF. However, they also noted that cross-protection is a multifactorial phenomenon that does not depend solely on sequence similarity, because they found that 100% of the pigs immunized with BA71∆CD2 and boosted with the parental BA71 virulent strain survived lethal challenge with Ken06.Bus, with almost no clinical signs of this disease (Lopez et al., Reference Lopez, van Heerden, Bosch-Camos, Accensi, Navas, Lopez-Monteagudo, Argilaguet, Gallardo, Pina-Pedrero, Salas, Salt and Rodriguez2020). Later, researchers further identified that the onset of cross-protective immunity is triggered by the LAV candidate BA71∆CD2 after 12 days, accompanied by humoral immunity with the presence of virus-specific IgG and cellular immunity with a cytotoxic response before the challenge (Marin-Moraleda et al., Reference Marin-Moraleda, Munoz-Basagoiti, Tort-Miro, Navas, Munoz, Vidal, Cobos, Martin-Mur, Meas, Motuzova, Chang, Gut, Accensi, Pina-Pedrero, Nunez, Esteve-Codina, Gavrilov, Rodriguez, Liu and Argilaguet2024). Recently, Sereda developed another attenuated ASFV seroimmunotype I vaccine, named Katanga-350. The attenuated vaccine was shown to protect 80% of pigs from a virulent strain of the same genotype and seroimmunotype (Lisbon-57). In contrast, for heterologous ASFV strain attacks, at least 50% of the surviving pigs received protection from subsequent intramuscular infection with a heterologous (genotype II, seroimmunotype VIII) virulent strain (Stavropol 08/01) (Sereda et al., Reference Sereda, Vlasov, Koltsova, Morgunov, Kudrjashov, Sindryakova, Kolbasova, Lyska, Koltsov, Zhivoderov, Pivova, Baluishev, Gogin and Kolbasov2022). Although progress has been made in attenuated ASFV vaccines, potential problems, such as reversion of virulence and a shortage of a universal ASFV vaccine that can match most prevalent ASFV mutant strains, must be considered.

Genetically engineered vaccines

Compared to traditional inactivated vaccines and LAVs, genetically engineered vaccines may elicit more effective immune responses and greater safety (Chandana et al., Reference Chandana, Nair, Chaturvedi, Abhishek, Pal, Charan, Balaji, Saini, Vasavi and Deepa2024; Gaudreault and Richt, Reference Gaudreault and Richt2019; Goatley et al., Reference Goatley, Reis, Portugal, Goldswain, Shimmon, Hargreaves, Ho, Montoya, Sánchez-Cordón, Taylor, Dixon and Netherton2020). Genetically engineered vaccines for ASF include subunit vaccines, live virus vector vaccines, and nucleic acid (DNA) vaccines.

Subunit vaccines and live vector vaccines

A subunit vaccine expresses the specific protective antigen genes of ASFV in a system suitable for inducing protection against ASF. Barderas et al. (Reference Barderas, Rodríguez, Gómez-Puertas, Avilés, Beitia, Alonso and Escribano2001) used recombinant baculovirus to express the p54/30 chimeric protein. Pigs immunized with the chimeric protein produced neutralizing antibodies. This vaccine was safer than traditional vaccines, but its protective effect was sometimes unsatisfactory. The live vector vaccine is obtained by inserting antigen-encoding genes into the genome of avirulent or attenuated pathogens. Live virus vector vaccines mainly use poxvirus, adenovirus, or baculovirus as expression vectors. Oviedo et al. (Reference Oviedo, Rodríguez, Gómez-Puertas, Brun, Gómez, Alonso and Escribano1997) studied the main antigenic proteins p54 and p30, which are encoded by the E183L and CP204L genes of ASFV and are highly expressed in baculovirus. The results showed that the constructed antigen protein can be used for ASFV diagnosis or further immunological research. Viral vector vaccines represent a promising alternative strategy for ASFV control, though their efficacy heavily depends on the selection of appropriate target antigens (Lokhandwala et al., Reference Lokhandwala, Petrovan, Popescu, Sangewar, Elijah, Stoian, Olcha, Ennen, Bray, Bishop, Waghela, Sheahan, Rowland and Mwangi2019). For both subunit and viral vector vaccine platforms, protective immunity requires the expressed antigen to elicit robust ASFV-specific antibody responses and cellular immune activation. Notably, merely delaying mortality in vaccinated pigs without preventing infection or clinical disease does not constitute full protection (Lopera-Madrid et al., Reference Lopera-Madrid, Osorio, He, Xiang, Adams, Laughlin, Mwangi, Subramanya, Neilan, Brake, Burrage, Brown, Clavijo and Bounpheng2017). With the guidance of replication-deficient human adenovirus 5 (primary immunization) and modified vaccinia Ankara (enhanced), eight ASFV antigen pools were found to protect pigs against ASFV genotype I. The protective antigen mixture contained the gene products of 18 ASFV genes, including B646L (p72), CP204L (p30), and CP530R (pp62). A modified vaccinia Ankara (MVA) expressing codon-optimized ASFV genes constituted by 18 selected ORFs with immunological activity boost led to reduced clinical signs and reduced levels of viremia in a portion of pigs after challenge with the virulent OUR T88/1 isolate (Netherton et al., Reference Netherton, Goatley, Reis, Portugal, Nash, Morgan, Gault, Nieto, Norlin, Gallardo, Ho, Sánchez-Cordón, Taylor and Dixon2019). In another study, a combination of the ASFV genes B646L (p72), CP204L (p30), CP530R (pp62), E183L (p54), the mature p37 product, and two sections of the mature p150 protein of the pp220 polyprotein (CP2475 L gene) was evaluated using the same prime-boost strategy with two different adjuvants. Approximately half of the animals in one of the adjuvanted groups survived challenge with the Georgia 2007/1 isolate (Lokhandwala et al., Reference Lokhandwala, Petrovan, Popescu, Sangewar, Elijah, Stoian, Olcha, Ennen, Bray, Bishop, Waghela, Sheahan, Rowland and Mwangi2019). Recently, Goatley et al. (Reference Goatley, Reis, Portugal, Goldswain, Shimmon, Hargreaves, Ho, Montoya, Sánchez-Cordón, Taylor, Dixon and Netherton2020) combined two immunization protocols using a pool of eight virally vectored ASF antigens, namely, E199L, EP153R, EP364R, F317L, I329L, MGF360-11 L, MGF505-4R, and MGF505-5 R, derived from the attenuated genotype I OUR T88/3, for delivery to pigs using an adenovirus prime and MVA boost. The pooled antigens completely protected pigs against fatal disease caused by the genotype I ASFV strain, which indicated that an effective subunit vaccine against ASF might be successfully developed. Additionally, an appropriate virus carrier may be crucial for the development of an ASFV subunit vaccine. As expected, Lopera-Madrid et al. (Reference Lopera-Madrid, Medina-Magues, Gladue, Borca and Osorio2021) suggested that expression vectors need to be optimized to improve the immunogenicity of ASFV subunit vaccines due to the complex nature of ASFV. They constructed several recombinant MVA vectors to evaluate the efficiency of different promoters and secretory signal sequences in the expression and immunogenicity of the p30 protein from ASFV and found that the natural poxvirus PrMVA13.5L promoter induced high levels of both p30 mRNA and specific anti-p30 antibodies in mice, while the synthetic PrS5E promoter and the S E/L promoter linked to a secretory signal produced lower mRNA and antibody levels. Therefore, promoter selection may be as crucial as the antigen used to develop ASFV subunit vaccines with MVA as the delivery vector. In addition, research on ASFV antigen gene combinations, adjuvant optimization, and other aspects is crucial for the development of subunit vaccines and carrier vaccines.

DNA vaccines

The DNA vaccine directly enters the host cell to express the antigen protein through a plasmid vector containing the gene sequence of the encoded antigen protein, stimulating the cells to produce the specific antigen protein and directly activating the cellular and humoral immunity of the host. Research on these vaccines has primarily focused on antigen genes, such as p30, p54, p72, and CD2v, as well as their proteins. In accordance with the serological response of recovered animals, the antigenic structural proteins p30 (CP204L), p54 (E183L), p72 (B646L), pp62 (CP530R), and CD2v (EP402R) have become the main targets of this type of vaccine design. Argilaguet et al. (Reference Argilaguet, Perez-Martin, Gallardo, Salguero, Borrego, Lacasta, Accensi, Diaz, Nofrarias, Pujols, Blanco, Perez-Filgueira, Escribano and Rodriguez2011) developed DNA vaccines encoding a fusion of p30 and p54 (pCMV-PQ) that induced good antibody responses in mice but no detectable antibody response in pigs. These authors subsequently extended the studies by demonstrating that DNA immunization in pigs could be exponentially improved by adding the extracellular domain of ASFV Hemagglutinin (sHA) to vaccine-encoded antigens fused to ubiquitin. Pigs immunized with pCMV-UbsHAPQ exhibited a strong CD8+ T-cell response, and 33% of the pigs vaccinated with pCMV-UbsHAPQ were protected in the absence of specific antibodies (Argilaguet et al., Reference Argilaguet, Perez-Martin, Nofrarias, Gallardo, Accensi, Lacasta, Mora, Ballester, Galindo-Cardiel, Lopez-Soria, Escribano, Reche and Rodriguez2012). Following the above research, Lacasta et al. (Reference Lacasta, Ballester, Monteagudo, Rodriguez, Salas, Accensi, Pina-Pedrero, Bensaid, Argilaguet, Lopez-Soria, Hutet, Le Potier and Rodriguez2014) developed an expression library containing more than 4,000 individual plasmid clones, each (except for p54, p30, and hemagglutinin ORFs) fused to ubiquitin. Immunization of farm pigs with the expression library yielded 60% protection against lethal challenge with the virulent E75 strain. In addition, Sereda et al. (Reference Sereda, Kazakova, Namsrayn, Vlasov, Sindryakova and Kolbasov2023) evaluated the protective efficacy of a heterologous vaccination strategy in pigs, wherein an attenuated vaccine strain FK-32/135 (seroimmunotype IV) was administered in combination with a DNA vaccine encoding a chimeric nucleotide sequence derived from the CD2v protein (EP402R, nucleotides 49–651) of the MK-200 strain (seroimmunotype III). This approach aimed to confer protection against the virulent ASFV strain Mozambique-78 (seroimmunotype III). Vaccination with the ASFV vaccine strain FK-32/135 protects pigs from disease caused by the strain with homologous seroimmunotype-France-32 (seroimmunotype IV). Still, it cannot induce sufficient cellular and humoral immunity to protect animals from challenge with the virulent strain Mozambique-78 (seroimmunotype III). These findings indicate that the protective effects of DNA vaccines against challenge with a homologous virulent strain of ASFV remain unsatisfactory, even in combination with a heterologous attenuated ASFV vaccine.

Conclusions

The global spread of ASF shows that the prevention and control measures implemented by veterinary and agricultural departments are inadequate. Since the development of a safe and effective vaccine remains challenging due to ASFV’s large and complex genome, its sophisticated immune evasion strategies, and the ongoing characterization of protective antigens, traditional culling of infected pigs and enhanced disinfection protocols remain necessary control measures. In addition, reasonable isolation plans should be developed. Simultaneously, research on epidemiology and molecular epidemiology should be promoted, and the spread of the virus should be strictly controlled. The complex structure of ASFV has been solved at moderate resolution, and 68 protein structures have been determined. Some progress has also been made in characterizing ASFV–host interactions and virus-encoded proteins that regulate the host response to infection. However, the interactions between ASFV and its host and vector are still unknown. In the future, additional attention should be directed toward ASFV at the DNA, RNA, and protein levels. Fully elucidating the interaction between the virus and host cell, identifying the receptor of ASFV, determining the complete ASFV protein profile, and conducting further ASFV virology and functional genomics research will provide strong support for the development of effective ASF vaccines in the future.

Acknowledgements

This research was supported financially by the National Natural Science Foundation of China (grant number 32273022), the Innovative Development Project in the Department of Education of Liaoning Province (LJ242511258002), and also funded by the Dalian University Discipline Construction Special Project (grant number DLUXK-2025-FX-003).

Conflict of interest statement

The authors declare that they have no competing interests.

Footnotes

These authors contributed equally to this work.

References

(1971) Preliminary report on the African swine fever epizootic in Cuba. Methods of diagnosis and control. Bull Off Int Epizoot 75(7), 367437. https://pubmed.ncbi.nlm.nih.gov/5168939/.Google Scholar
Abreu, EF, de Valadão, FB, Limpo Serra, JJB, Ornelas Mário, R and Sousa Montenegro, A (1962) Peste suína africana em Moc¸ ambique. Anais Dos Servic¸ Os Veterinárias de Moc¸ Ambique 8, 105123.Google Scholar
Alejo, A, Matamoros, T, Guerra, M and Andres, G (2018) A proteomic atlas of the African swine fever virus particle. Journal of Virology 92(23). https://doi.org/10.1128/JVI.01293-18.CrossRefGoogle ScholarPubMed
Anastasia, G, Elena, K, Daria, L, Tatiana, B, Olga, K, Ruslan, G, Timofey, S and Elena, K (2025) Assessment of hunters’ awareness of African swine fever in samara oblast, the Russian Federation. BMC Veterinary Research 21(1), 219. https://doi.org/10.1186/s12917-025-04664-5.CrossRefGoogle Scholar
Andrés, G (2017) African swine fever virus gets undressed: new insights on the entry pathway. Journal of Virology 91(4). https://doi.org/10.1128/jvi.01906-16.CrossRefGoogle ScholarPubMed
Andres, G, Alejo, A, Salas, J and Salas, ML (2002) African swine fever virus polyproteins pp220 and pp62 assemble into the core shell. Journal of Virology 76(24), 1247312482. https://doi.org/10.1128/jvi.76.24.12473-12482.2002.CrossRefGoogle ScholarPubMed
Andres, G, Charro, D, Matamoros, T, Dillard, RS and Abrescia, NGA (2020) The cryo-EM structure of African swine fever virus unravels a unique architecture comprising two icosahedral protein capsids and two lipoprotein membranes. Journal of Biological Chemistry 295(1), 112. https://doi.org/10.1074/jbc.AC119.011196.CrossRefGoogle ScholarPubMed
Ankhanbaatar, U, Auer, A, Ulziibat, G, Settypalli, TBK, Gombo-Ochir, D, Basan, G, Takemura, T, Tseren-Ochir, EO, Ouled Ahmed, H, Meki, IK, Datta, S, Soumare, B, Metlin, A, Cattoli, G and Lamien, CE (2023) Comparison of the whole-genome sequence of the African swine fever virus from a Mongolian wild boar with genotype II viruses from Asia and Europe. Pathogens 12(9), 1143. https://doi.org/10.3390/pathogens12091143.CrossRefGoogle ScholarPubMed
Argilaguet, JM, Perez-Martin, E, Gallardo, C, Salguero, FJ, Borrego, B, Lacasta, A, Accensi, F, Diaz, I, Nofrarias, M, Pujols, J, Blanco, E, Perez-Filgueira, M, Escribano, JM and Rodriguez, F (2011) Enhancing DNA immunization by targeting ASFV antigens to SLA-II bearing cells. Vaccine 29(33), 53795385. https://doi.org/10.1016/j.vaccine.2011.05.084.CrossRefGoogle ScholarPubMed
Argilaguet, JM, Pérez-Martín, E, López, S, Goethe, M, Escribano, JM, Giesow, K, Keil, GM and Rodríguez, F (2013) BacMam immunization partially protects pigs against sublethal challenge with African swine fever virus. Antiviral Research 98(1), 6165. https://doi.org/10.1016/j.antiviral.2013.02.005.CrossRefGoogle ScholarPubMed
Argilaguet, JM, Perez-Martin, E, Nofrarias, M, Gallardo, C, Accensi, F, Lacasta, A, Mora, M, Ballester, M, Galindo-Cardiel, I, Lopez-Soria, S, Escribano, JM, Reche, PA and Rodriguez, F (2012) DNA vaccination partially protects against African swine fever virus lethal challenge in the absence of antibodies. PLoS One 7(9), e40942. https://doi.org/10.1371/journal.pone.0040942.CrossRefGoogle ScholarPubMed
Arias, M, de la Torre, A, Dixon, L, Gallardo, C, Jori, F, Laddomada, A, Martins, C, Parkhouse, RM, Revilla, Y and Rodriguez, FAJ (2017) Approaches and perspectives for development of African swine fever virus vaccines. Vaccines (Basel) 5(4). https://doi.org/10.3390/vaccines5040035.Google ScholarPubMed
Asgari, S, Bideshi, DK, Bigot, Y, Federici, BA, Cheng, XW and Ictv Report, C (2017) ICTV virus taxonomy profile: ascoviridae. Journal of General Virology 98(1), 45. https://doi.org/10.1099/jgv.0.000677.CrossRefGoogle ScholarPubMed
Atuhaire, DK, Afayoa, M, Ochwo, S, Mwesigwa, S, Okuni, JB, Olaho-Mukani, W and Ojok, L (2013) Molecular characterization and phylogenetic study of African swine fever virus isolates from recent outbreaks in Uganda (2010-2013). Virology Journal 10, 247. https://doi.org/10.1186/1743-422X-10-247.CrossRefGoogle ScholarPubMed
Avagyan, H, Hakobyan, S, Baghdasaryan, B, Arzumanyan, H, Poghosyan, A, Bayramyan, N, Semerjyan, A, Sargsyan, M, Voskanyan, H, Vardanyan, T, Karalyan, N, Hakobyan, L, Abroyan, L, Avetisyan, A, Karalova, E, Semerjyan, Z and Karalyan, Z (2024) Pathology and clinics of naturally occurring low-virulence variants of African swine fever emerged in domestic pigs in the south caucasus. Pathogens 13(2), 130. https://doi.org/10.3390/pathogens13020130.CrossRefGoogle ScholarPubMed
Balysheva, VIPE, Galnbek, T and Balyshev, VM (2015) Immunological properties of attenuated variants of African swine fever virus isolated in the Russian Federation. Russian Agricultural Sciences 41(2–3), 178182. https://link.springer.com/article/110.3103S1068367415020056.10.3103/S1068367415020056CrossRefGoogle Scholar
Banjara, S, Caria, S, Dixon, LK, Hinds, MG and Kvansakul, M (2017) Structural insight into African swine fever virus A179L-mediated inhibition of apoptosis. Journal of Virology 91(6). https://doi.org/10.1128/jvi.02228-16.CrossRefGoogle ScholarPubMed
Banjara, S, Shimmon, GL, Dixon, LK, Netherton, CL, Hinds, MG and Kvansakul, M (2019) Crystal structure of African swine fever virus A179L with the autophagy regulator beclin. Viruses 11(9), 789. https://doi.org/10.3390/v11090789.CrossRefGoogle ScholarPubMed
Barasona, JA, Gallardo, C, Cadenas-Fernandez, E, Jurado, C, Rivera, B, Rodriguez-Bertos, A, Arias, M and Sanchez-Vizcaino, JM (2019) First oral vaccination of Eurasian wild boar against African swine fever virus genotype II. Frontiers in Veterinary Science 6, 137. https://doi.org/10.3389/fvets.2019.00137.CrossRefGoogle ScholarPubMed
Barderas, MG, Rodríguez, F, Gómez-Puertas, P, Avilés, M, Beitia, F, Alonso, C and Escribano, JM (2001) Antigenic and immunogenic properties of a chimera of two immunodominant African swine fever virus proteins. Archives of Virology 146(9), 16811691. https://doi.org/10.1007/s007050170056.CrossRefGoogle ScholarPubMed
Barongo, MB, Stahl, K, Bett, B, Bishop, RP, Fevre, EM, Aliro, T, Okoth, E, Masembe, C, Knobel, D and Ssematimba, A (2015) Estimating the basic reproductive number (R0) for African swine fever virus (ASFV) transmission between pig herds in Uganda. PLoS One 10(5), e0125842. https://doi.org/10.1371/journal.pone.0125842.CrossRefGoogle ScholarPubMed
Barrado-Gil, L, Del Puerto, A, Galindo, I, Cuesta-Geijo, M, García-Dorival, I, de Motes, CM and Alonso, C (2021) African swine fever virus ubiquitin-conjugating enzyme is an immunomodulator targeting NF-κB activation. Viruses 13(6), 1160. https://doi.org/10.3390/v13061160.CrossRefGoogle Scholar
Bastos, AD, Penrith, ML, Cruciere, C, Edrich, JL, Hutchings, G, Roger, F, Couacy-Hymann, E and RT, G (2003) Genotyping field strains of African swine fever virus by partial p72 gene characterisation. Archives of Virology 148(4), 693706. https://doi.org/10.1007/s00705-002-0946-8.CrossRefGoogle ScholarPubMed
Beltrán-Alcrudo, D, Lubroth, J, Depner, K and De La Rocque, S (2008) African swine fever in the Caucasus. FAO EMPRES Watch 1, 18.Google Scholar
Bezymennyi, M, Tarasov, O, Kyivska, GV, Mezhenska, NA, Mandyhra, S, Kovalenko, G, Sushko, M, Hudz, N, Skorokhod, SV, Datsenko, R, Muzykina, L, Milton, E, Sapachova, MA, Nychyk, S, Halka, I, Frant, M, Huettmann, F, Drown, DM, Gerilovych, A, Mezhenskyi, AA, Bortz, E and Lange, CE (2023) Epidemiological characterization of African swine fever dynamics in Ukraine, 2012-2023. Vaccines (Basel) 11(7). https://doi.org/10.3390/vaccines11071145.Google ScholarPubMed
Blome, S, Franzke, K and Beer, M (2020) African swine fever - A review of current knowledge. Virus Research 287, 198099. https://doi.org/10.1016/j.virusres.2020.198099.CrossRefGoogle ScholarPubMed
Blome, S, Friedrichs, V, Schafer, A and Beer, M (2025) Benefit risk considerations for African swine fever virus live attenuated vaccines. NPJ Vaccines 10(1), 147. https://doi.org/10.1038/s41541-025-01208-8.CrossRefGoogle ScholarPubMed
Blome, S, Gabriel, C and Beer, M (2014) Modern adjuvants do not enhance the efficacy of an inactivated African swine fever virus vaccine preparation. Vaccine 32(31), 38793882. https://doi.org/10.1016/j.vaccine.2014.05.051.CrossRefGoogle Scholar
Boinas, FS, Wilson, AJ, Hutchings, GH, Martins, C and Dixon, LJ (2011) The persistence of African swine fever virus in field-infected Ornithodoros erraticus during the ASF endemic period in Portugal. PLoS One 6(5), e20383. https://doi.org/10.1371/journal.pone.0020383.CrossRefGoogle ScholarPubMed
Borca, MV, Rai, A, Ramirez-Medina, E, Silva, E, Velazquez-Salinas, L, Vuono, E, Pruitt, S, Espinoza, N and Gladue, DP (2021) A cell culture-adapted vaccine virus against the current African swine fever virus pandemic strain. Journal of Virology 95(14), e0012321. https://doi.org/10.1128/JVI.00123-21.CrossRefGoogle ScholarPubMed
Borca, MV, Ramirez-Medina, E, Espinoza, N, Rai, A, Spinard, E, Velazquez-Salinas, L, Valladares, A, Silva, E, Burton, L, Meyers, A, Clark, J, Wu, P, Gay, CG and Gladue, DP (2024) Deletion of the EP402R gene from the genome of African swine fever vaccine strain ASFV-G-∆I177L provides the potential capability of differentiating between infected and vaccinated animals. Viruses 16(3), 376. https://doi.org/10.3390/v16030376.CrossRefGoogle ScholarPubMed
Borca, MV, Ramirez-Medina, E, Silva, E, Vuono, E, Rai, A, Pruitt, S, Holinka, LG, Velazquez-Salinas, L, Zhu, J and Gladue, DP (2020) Development of a highly effective African swine fever virus vaccine by deletion of the I177L gene results in sterile immunity against the current epidemic Eurasia Strain. Journal of Virology 94(7). https://doi.org/10.1128/JVI.02017-19.CrossRefGoogle ScholarPubMed
Bosch-Camós, L, López, E, Collado, J, Navas, MJ, Blanco-Fuertes, M, Pina-Pedrero, S, Accensi, F, Salas, ML, Mundt, E, Nikolin, V and Rodríguez, F (2021) M448R and MGF505-7R: two African swine fever virus antigens commonly recognized by ASFV-specific T-cells and with protective potential. Vaccines (Basel) 9(5). https://doi.org/10.3390/vaccines9050508.Google ScholarPubMed
Bosch-Camós, L, López, E and Rodriguez, F (2020) African swine fever vaccines: A promising work still in progress. Porcine Health Management 6, 17. https://doi.org/10.1186/s40813-020-00154-2.CrossRefGoogle ScholarPubMed
Boshoff, CI, Bastos, AD, Dube, MM and Heath, L (2014) First molecular assessment of the African swine fever virus status of Ornithodoros ticks from Swaziland. Onderstepoort Journal of Veterinary Research 81(1), E15. https://doi.org/10.4102/ojvr.v81i1.846.CrossRefGoogle ScholarPubMed
Bourry, O, Hutet, E, Le Dimna, M, Lucas, P, Blanchard, Y, Chastagner, A, Paboeuf, F and Le Potier, MF (2022) Oronasal or intramuscular immunization with a thermo-attenuated ASFV strain provides full clinical protection against Georgia 2007/1 challenge. Viruses 14(12), 2777. https://doi.org/10.3390/v14122777.CrossRefGoogle ScholarPubMed
Burmakina, G, Malogolovkin, A, Tulman, ER, Xu, W, Delhon, G, Kolbasov, D and Rock, DL (2019) Identification of T-cell epitopes in African swine fever virus CD2v and C-type lectin proteins. Journal of General Virology 100(2), 259265. https://doi.org/10.1099/jgv.0.001195.CrossRefGoogle ScholarPubMed
Cackett, G, Matelska, D, Sýkora, M, Portugal, R, Malecki, M, Bähler, J, Dixon, L and Werner, F (2020) The African swine fever virus transcriptome. Journal of Virology 94(9). https://doi.org/10.1128/jvi.00119-20.CrossRefGoogle ScholarPubMed
Cadenas-Fernandez, E, Barroso-Arevalo, S, Kosowska, A, Diaz-Frutos, M, Gallardo, C, Rodriguez-Bertos, A, Bosch, J, Sanchez-Vizcaino, JM and Barasona, JA (2024) Challenging boundaries: Is cross-protection evaluation necessary for African swine fever vaccine development? A case of oral vaccination in wild boar. Frontiers in Immunology 15, 1388812. https://doi.org/10.3389/fimmu.2024.1388812.CrossRefGoogle ScholarPubMed
Cadenas-Fernandez, E, Sanchez-Vizcaino, JM, van den Born, E, Kosowska, A, van Kilsdonk, E, Fernandez-Pacheco, P, Gallardo, C, Arias, M and Barasona, JA (2021) High doses of inactivated African swine fever virus are safe, but do not confer protection against a virulent challenge. Vaccines (Basel) 9(3). https://doi.org/10.3390/vaccines9030242.Google Scholar
Campillo-Balderas, JA, Lazcano, A, Cottom-Salas, W, Jacome, R and Becerra, A (2023) Pangenomic analysis of nucleo-cytoplasmic large DNA viruses. i: the phylogenetic distribution of conserved oxygen-dependent enzymes reveals a capture-gene process. Journal of Molecular Evolution 91(5), 647668. https://doi.org/10.1007/s00239-023-10126-z.CrossRefGoogle ScholarPubMed
Cao, S, Lu, H, Wu, Z and Zhu, S (2022) A duplex fluorescent quantitative PCR assay to distinguish the genotype I and II strains of African swine fever virus in Chinese epidemic strains. Frontiers in Veterinary Science 9, 998874. https://doi.org/10.3389/fvets.2022.998874.CrossRefGoogle ScholarPubMed
Chandana, MS, Nair, SS, Chaturvedi, VK, Abhishek, , Pal, S, Charan, MSS, Balaji, S, Saini, S, Vasavi, K and Deepa, P (2024) Recent progress and major gaps in the vaccine development for African swine fever. Brazilian Journal of Microbiology 55(1), 9971010. https://doi.org/10.1007/s42770-024-01264-7.CrossRefGoogle ScholarPubMed
Chen, W, Zhao, D, He, X, Liu, R, Wang, Z, Zhang, X, Li, F, Shan, D, Chen, H, Zhang, J, Wang, L, Wen, Z, Wang, X, Guan, Y, Liu, J and Bu, Z (2020) A seven-gene-deleted African swine fever virus is safe and effective as a live attenuated vaccine in pigs. Science China Life Sciences 63(5), 623634. https://doi.org/10.1007/s11427-020-1657-9.CrossRefGoogle ScholarPubMed
Chenais, E, Depner, K, Ebata, A, Penrith, ML, Pfeiffer, DU, Price, C, Stahl, K and Fischer, K (2022) Exploring the hurdles that remain for control of African swine fever in smallholder farming settings. Transboundary and Emerging Diseases 69. https://doi.org/10.1111/tbed.14642.CrossRefGoogle ScholarPubMed
Chernyshev, RS, Igolkin, AS, Zinyakov, NG and Chvala, IA (2024) Comparative analysis of whole-genome sequences of African swine fever virus (Asfarviridae: Asfivirus) isolates small es, cyrillicollected on the territory of the left bank of the Dnieper River in 2023. Problems of Virology 69(5), 481494. https://doi.org/10.36233/0507-4088-263.CrossRefGoogle Scholar
Correia, S, Ventura, S and Parkhouse, RM (2013) Identification and utility of innate immune system evasion mechanisms of ASFV. Virus Research 173(1), 87100. https://doi.org/10.1016/j.virusres.2012.10.013.CrossRefGoogle ScholarPubMed
Costard, S, Wieland, B, de Glanville, W, Jori, F, Rowlands, R, Vosloo, W, Roger, F, Pfeiffer, DU and Dixon, LK (2009) African swine fever: How can global spread be prevented? Philosophical Transactions of the Royal Society of London B: Biological Sciences 364(1530), 26832696. https://doi.org/10.1098/rstb.2009.0098.CrossRefGoogle ScholarPubMed
Cowan, KM (1961) Immunological studies on African swine fever virus. I. Elimination of the procomplementary activity of swine serum with formalin. The Journal of Immunology 86, 465470.10.4049/jimmunol.86.4.465CrossRefGoogle ScholarPubMed
Dar, P, Kalaivanan, R, Sied, N, Mamo, B, Kishore, S, Suryanarayana, VV and Kondabattula, G (2013) Montanide ISA 201 adjuvanted FMD vaccine induces improved immune responses and protection in cattle. Vaccine 31(33), 33273332. https://doi.org/10.1016/j.vaccine.2013.05.078.CrossRefGoogle ScholarPubMed
Detray, DE (1957) African swine fever in wart hogs (Phacochoerus aethiopicus). Journal of the American Veterinary Medical Association 130(12), 537540.Google ScholarPubMed
Dhollander, S, Cattaneo, E, Cortinas Abrahantes, J, Boklund, AE, Szczotka-Bochniarz, A, Mihalca, AD, Papanikolaou, A, Mur, L, Balmos, OM, Frant, M, Gal-Cison, A, Kwasnik, M, Rozek, W, Malakauskas, A, Masiulis, M, Turcinaviciene, J, Rusina, A, Aminalragia-Giamini, R, Chesnoiu, T, Jazdzewski, K, Rola, J, Barbuceanu, F and Stegeman, JA (2025) Prospective case-control study of determinants for African swine fever introduction in commercial pig farms in Poland, Romania, and Lithuania. Transboundary and Emerging Diseases 2025, 5419764. https://doi.org/10.1155/tbed/5419764.CrossRefGoogle ScholarPubMed
Diep, NV, Ngoc, NT, Duc, NV, Dang, VX, Tiep, TN, Quy, CT, Tham, BT and Doanh, PN (2025) Safety and efficacy profiles of the live attenuated vaccine AVAC ASF LIVE for preventing African swine fever in pigs. Transboundary and Emerging Diseases 2025, 8623876. https://doi.org/10.1155/tbed/8623876.CrossRefGoogle ScholarPubMed
Dixon, LK, Chapman, DA, Netherton, CL and Upton, C (2013) African swine fever virus replication and genomics. Virus Research 173(1), 314. https://doi.org/10.1016/j.virusres.2012.10.020.CrossRefGoogle ScholarPubMed
Dixon, LK, Islam, M, Nash, R and Reis, AL (2019) African swine fever virus evasion of host defences. Virus Research 266, 2533. https://doi.org/10.1016/j.virusres.2019.04.002.CrossRefGoogle ScholarPubMed
Dixon, LK, Sánchez-Cordón, PJ, Galindo, I and Alonso, C (2017) Investigations of pro- and anti-apoptotic factors affecting African swine fever virus replication and pathogenesis. Viruses 9(9), 241. https://doi.org/10.3390/v9090241.CrossRefGoogle ScholarPubMed
Dixon, LK, Stahl, K, Jori, F, Vial, L and Pfeiffer, DU (2020) African swine fever epidemiology and control. Annual Review of Animal Biosciences 8, 221246. https://doi.org/10.1146/annurev-animal-021419-083741.CrossRefGoogle ScholarPubMed
Escribano, JM, Galindo, I and Alonso, C (2013) Antibody-mediated neutralization of African swine fever virus: Myths and facts. Virus Research 173(1), 101109. https://doi.org/10.1016/j.virusres.2012.10.012.CrossRefGoogle ScholarPubMed
Fu, D, Zhao, D, Zhang, W, Zhang, G, Li, M, Zhang, Z, Wang, Y, Sun, D, Jiao, P, Chen, C, Guo, Y and Rao, Z (2020) Structure of African swine fever virus p15 reveals its dual role for membrane-association and DNA binding. Protein and Cell 11(8), 606612. https://doi.org/10.1007/s13238-020-00731-9.CrossRefGoogle ScholarPubMed
Galindo, I, Cuesta-Geijo, MA, Hlavova, K, Muñoz-Moreno, R, Barrado-Gil, L, Dominguez, J and Alonso, C (2015) African swine fever virus infects macrophages, the natural host cells, via clathrin- and cholesterol-dependent endocytosis. Virus Research 200, 4555. https://doi.org/10.1016/j.virusres.2015.01.022.CrossRefGoogle ScholarPubMed
Galindo, I, Hernaez, B, Díaz-Gil, G, Escribano, JM and Alonso, C (2008) A179L, a viral Bcl-2 homologue, targets the core Bcl-2 apoptotic machinery and its upstream BH3 activators with selective binding restrictions for Bid and Noxa. Virology 375(2), 561572. https://doi.org/10.1016/j.virol.2008.01.050.CrossRefGoogle ScholarPubMed
Gallagher, JR and Harris, AK (2020) Cryo-EM cools down swine fever. Journal of Biological Chemistry 295(1), 1314. https://doi.org/10.1074/jbc.H119.012169.CrossRefGoogle ScholarPubMed
Gallardo, C, Okoth, E, Pelayo, V, Anchuelo, R, Martin, E, Simon, A, Llorente, A, Nieto, R, Soler, A, Martin, R, Arias, M and Bishop, RP (2011) African swine fever viruses with two different genotypes, both of which occur in domestic pigs, are associated with ticks and adult warthogs, respectively, at a single geographical site. Journal of General Virology 92(Pt 2), 432444. https://doi.org/10.1099/vir.0.025874-0.CrossRefGoogle Scholar
Gallardo, C, Sánchez, EG, Pérez-Núñez, D, Nogal, M, de León, P, Carrascosa, ÁL, Nieto, R, Soler, A, Arias, ML and Revilla, Y (2018) African swine fever virus (ASFV) protection mediated by NH/P68 and NH/P68 recombinant live-attenuated viruses. Vaccine 36(19), 26942704. https://doi.org/10.1016/j.vaccine.2018.03.040.CrossRefGoogle ScholarPubMed
Gallardo, C, Soler, A, Rodze, I, Nieto, R, Cano-Gomez, C, Fernandez-Pinero, J and Arias, M (2019) Attenuated and non-haemadsorbing (non- HAD) genotype II African swine fever virus (ASFV) isolated in Europe, Latvia 2017. Transboundary and Emerging Diseases 66(3), 13991404. https://doi.org/10.1111/tbed.13132.CrossRefGoogle ScholarPubMed
Gao, C, He, X, Quan, J, Jiang, Q, Lin, H, Chen, H and Qu, L (2017) Specificity Characterization of SLA class I molecules binding to swine-origin viral cytotoxic T lymphocyte epitope peptides in vitro. Frontiers in Microbiology 8, 2524. https://doi.org/10.3389/fmicb.2017.02524.CrossRefGoogle Scholar
Gaudreault, NN, Madden, DW, Wilson, WC, Trujillo, JD and Richt, JA (2020) African swine fever virus: an emerging DNA arbovirus. Frontiers in Veterinary Science 7, 215. https://doi.org/10.3389/fvets.2020.00215.CrossRefGoogle ScholarPubMed
Gaudreault, NN and Richt, JA (2019) Subunit vaccine approaches for African swine fever virus. Vaccines (Basel) 7(2). https://doi.org/10.3390/vaccines7020056.Google ScholarPubMed
Gavier-Widen, D, Stahl, K and Dixon, L (2020) No hasty solutions for African swine fever. Science 367(6478), 622624. https://doi.org/10.1126/science.aaz8590.CrossRefGoogle ScholarPubMed
Goatley, LC, Reis, AL, Portugal, R, Goldswain, H, Shimmon, GL, Hargreaves, Z, Ho, CS, Montoya, M, Sánchez-Cordón, PJ, Taylor, G, Dixon, LK and Netherton, CL (2020) A pool of eight virally vectored African swine fever antigens protect pigs against fatal disease. Vaccines (Basel) 8(2). https://doi.org/10.3390/vaccines8020234.Google ScholarPubMed
Gogin, A, Gerasimov, V, Malogolovkin, A and Kolbasov, D (2013) African swine fever in the north caucasus region and the Russian Federation in years 2007-2012. Virus Research 173(1), 198203. https://doi.org/10.1016/j.virusres.2012.12.007.CrossRefGoogle ScholarPubMed
Golding, JP, Goatley, L, Goodbourn, S, Dixon, LK, Taylor, G and Netherton, CL (2016) Sensitivity of African swine fever virus to type I interferon is linked to genes within multigene families 360 and 505. Virology 493, 154161. https://doi.org/10.1016/j.virol.2016.03.019.CrossRefGoogle ScholarPubMed
Golnar, AJ, Martin, E, Wormington, JD, Kading, RC, Teel, PD, Hamer, SA and Hamer, GL (2019) Reviewing the potential vectors and hosts of African swine fever virus transmission in the United States. Vector-Borne and Zoonotic Diseases 19(7), 512524. https://doi.org/10.1089/vbz.2018.2387.CrossRefGoogle ScholarPubMed
Hakizimana, JN, Ntirandekura, JB, Yona, C, Nyabongo, L, Kamwendo, G, Chulu, JLC, Ntakirutimana, D, Kamana, O, Nauwynck, H and Misinzo, G (2021) Complete genome analysis of African swine fever virus responsible for outbreaks in domestic pigs in 2018 in Burundi and 2019 in Malawi. Tropical Animal Health and Production 53(4), 438. https://doi.org/10.1007/s11250-021-02877-y.CrossRefGoogle ScholarPubMed
Hakizimana, JN, Nyabongo, L, Ntirandekura, JB, Yona, C, Ntakirutimana, D, Kamana, O, Nauwynck, H and Misinzo, G (2020) Genetic analysis of African swine fever virus from the 2018 outbreak in south-eastern burundi. Frontiers in Veterinary Science 7, 578474. https://doi.org/10.3389/fvets.2020.578474.CrossRefGoogle ScholarPubMed
He, WR, Yuan, J, Ma, YH, Zhao, CY, Yang, ZY, Zhang, Y, Han, S, Wan, B and Zhang, GP (2022) Modulation of host antiviral innate immunity by African swine fever virus: a review. Animals (Basel) 12(21). https://doi.org/10.3390/ani12212935.Google ScholarPubMed
Hernaez, B, Escribano, JM and Alonso, C (2008) African swine fever virus protein p30 interaction with heterogeneous nuclear ribonucleoprotein K (hnRNP-K) during infection. FEBS Letters 582(23–24), 32753280. https://doi.org/10.1016/j.febslet.2008.08.031.CrossRefGoogle ScholarPubMed
Hernaez, B, Guerra, M, Salas, ML and Andres, G (2016) African swine fever virus undergoes outer envelope disruption, capsid disassembly and inner envelope fusion before core release from multivesicular endosomes. PLOS Pathogens 12(4), e1005595. https://doi.org/10.1371/journal.ppat.1005595.CrossRefGoogle ScholarPubMed
Hsu, CH, Chang, CY, Otake, S, Molitor, TW and Perez, A (2024) Strategies for transboundary swine disease management in Asian Islands: foot and mouth disease, classical swine fever, and African swine fever in Taiwan, Japan, and the Philippines. Veterinary Sciences 11(3). https://doi.org/10.3390/vetsci11030130.CrossRefGoogle ScholarPubMed
Hurtado, C, Bustos, MJ, Granja, AG, de León, P, Sabina, P, López-Viñas, E, Gómez-Puertas, P, Revilla, Y and Carrascosa, AL (2011) The African swine fever virus lectin EP153R modulates the surface membrane expression of MHC class I antigens. Archives of Virology 156(2), 219234. https://doi.org/10.1007/s00705-010-0846-2.CrossRefGoogle ScholarPubMed
Igolkin, A, Mazloum, A, Zinyakov, N, Chernyshev, R, Schalkwyk, AV, Shotin, A, Lavrentiev, I, Gruzdev, K and Chvala, I (2024) Detection of the first recombinant African swine fever virus (genotypes I and II) in domestic pigs in Russia. Molecular Biology Reports 51(1), 1011. https://doi.org/10.1007/s11033-024-09961-0.CrossRefGoogle Scholar
Jia, N, Ou, Y, Pejsak, Z, Zhang, Y and Zhang, J (2017) Roles of African swine fever virus structural proteins in viral infection. Journal of Veterinary Research 61(2), 135143. https://doi.org/10.1515/jvetres-2017-0017.CrossRefGoogle ScholarPubMed
Juszkiewicz, M, Walczak, M, Mazur-Panasiuk, N and Woźniakowski, G (2020) Effectiveness of chemical compounds used against African swine fever virus in commercial available disinfectants. Pathogens 9(11), 878. https://doi.org/10.3390/pathogens9110878.CrossRefGoogle ScholarPubMed
Juszkiewicz, M, Walczak, M, Wozniakowski, G, Pejsak, Z and Podgorska, K (2025) The influence of the temperature on effectiveness of selected disinfectants against African swine fever virus (ASFV). Viruses 17(2), 156. https://doi.org/10.3390/v17020156.CrossRefGoogle ScholarPubMed
Karger, A, Pérez-Núñez, D, Urquiza, J, Hinojar, P, Alonso, C, Freitas, FB, Revilla, Y, Le Potier, MF and Montoya, M (2019) An update on African swine fever virology. Viruses 11(9), 864. https://doi.org/10.3390/v11090864.CrossRefGoogle ScholarPubMed
Kaur, M, Saxena, A, Rai, A and Bhatnagar, R (2010) Rabies DNA vaccine encoding lysosome-targeted glycoprotein supplemented with Emulsigen-D confers complete protection in preexposure and postexposure studies in BALB/c mice. The FASEB Journal 24(1), 173183. https://doi.org/10.1096/fj.09-138644.CrossRefGoogle ScholarPubMed
Kim, HJ, Cho, KH, Lee, SK, Kim, DY, Nah, JJ, Kim, HJ, Kim, HJ, Hwang, JY, Sohn, HJ, Choi, JG, Kang, HE and Kim, YJ (2020) Outbreak of African swine fever in South Korea, 2019. Transboundary and Emerging Diseases 67(2), 473475. https://doi.org/10.1111/tbed.13483.CrossRefGoogle ScholarPubMed
Korennoy, FI, Gulenkin, VM, Gogin, AE, Vergne, T and Karaulov, AK (2017) Estimating the basic reproductive number for African swine fever using the Ukrainian historical epidemic of 1977. Transboundary and Emerging Diseases 64(6), 18581866. https://doi.org/10.1111/tbed.12583.CrossRefGoogle ScholarPubMed
Kouakou, KV, Michaud, V, Biego, HG, Gnabro, HPG, Kouakou, AV, Mossoun, AM, Awuni, JA, Minoungou, GL, Aplogan, GL, Awoume, FK, Albina, E, Lancelot, R and Couacy-Hymann, E (2017) African and classical swine fever situation in ivory-coast and neighboring countries, 2008-2013. Acta Tropica 166, 241248. https://doi.org/10.1016/j.actatropica.2016.10.027.CrossRefGoogle ScholarPubMed
Krug, PW, Holinka, LG, O’Donnell, V, Reese, B, Sanford, B, Fernandez-Sainz, I, Gladue, DP, Arzt, J, Rodriguez, L, Risatti, GR and Borca, MV (2015) The progressive adaptation of a georgian isolate of African swine fever virus to vero cells leads to a gradual attenuation of virulence in swine corresponding to major modifications of the viral genome. Journal of Virology 89(4), 23242332. https://doi.org/10.1128/JVI.03250-14.CrossRefGoogle ScholarPubMed
Lacasta, A, Ballester, M, Monteagudo, PL, Rodriguez, JM, Salas, ML, Accensi, F, Pina-Pedrero, S, Bensaid, A, Argilaguet, J, Lopez-Soria, S, Hutet, E, Le Potier, MF and Rodriguez, F (2014) Expression library immunization can confer protection against lethal challenge with African swine fever virus. Journal of Virology 88(22), 1332213332. https://doi.org/10.1128/JVI.01893-14.CrossRefGoogle ScholarPubMed
Lacasta, A, Monteagudo, PL, Jimenez-Marin, A, Accensi, F, Ballester, M, Argilaguet, J, Galindo-Cardiel, I, Segales, J, Salas, ML, Dominguez, J, Moreno, A, Garrido, JJ and Rodriguez, F (2015) Live attenuated African swine fever viruses as ideal tools to dissect the mechanisms involved in viral pathogenesis and immune protection. Veterinary Research 46, 135. https://doi.org/10.1186/s13567-015-0275-z.CrossRefGoogle ScholarPubMed
Laddomada, A, Rolesu, S, Loi, F, Cappai, S, Oggiano, A, Madrau, MP, Sanna, ML, Pilo, G, Bandino, E, Brundu, D, Cherchi, S, Masala, S, Marongiu, D, Bitti, G, Desini, P, Floris, V, Mundula, L, Carboni, G, Pittau, M, Feliziani, F, Sanchez-Vizcaino, JM, Jurado, C, Guberti, V, Chessa, M, Muzzeddu, M, Sardo, D, Borrello, S, Mulas, D, Salis, G, Zinzula, P, Piredda, S, De Martini, A and Sgarangella, F (2019) Surveillance and control of African swine fever in free-ranging pigs in Sardinia. Transboundary and Emerging Diseases 66(3), 11141119. https://doi.org/10.1111/tbed.13138.CrossRefGoogle ScholarPubMed
Leitao, A, Cartaxeiro, C, Coelho, R, Cruz, B, Parkhouse, RME, Portugal, FC, Vigario, JD and Martins, CLV (2001) The non-haemadsorbing African swine fever virus isolate ASFV/NH/P68 provides a model for defining the protective anti-virus immune response. Journal of General Virology 82(Pt 3), 513523. https://doi.org/10.1099/0022-1317-82-3-513.CrossRefGoogle Scholar
Li, G, Fu, D, Zhang, G, Zhao, D, Li, M, Geng, X, Sun, D, Wang, Y, Chen, C, Jiao, P, Cao, L, Guo, Y and Rao, Z (2020) Crystal structure of the African swine fever virus structural protein p35 reveals its role for core shell assembly. Protein and Cell 11(8), 600605. https://doi.org/10.1007/s13238-020-00730-w.CrossRefGoogle ScholarPubMed
Ligue-Sabio, KDB, Lacaba, MFT, Mijares, JEC, Murao, LAE and Alviola, P (2025) Spatiotemporal patterns and risk factors for African swine fever-affected smallholder pig farms in Davao Region, Southern Philippines. Preventive Veterinary Medicine 239, 106495. https://doi.org/10.1016/j.prevetmed.2025.106495.CrossRefGoogle ScholarPubMed
Liu, J, Liu, B, Shan, B, Wei, S, An, T, Shen, G and Chen, Z (2020) Prevalence of African swine fever in China, 2018-2019. Journal of Medical Virology 92(8), 10231034. https://doi.org/10.1002/jmv.25638.CrossRefGoogle Scholar
Liu, Q, Ma, B, Qian, N, Zhang, F, Tan, X, Lei, J and Xiang, Y (2019a) Structure of the African swine fever virus major capsid protein p72. Cell Research 29(11), 953955. https://doi.org/10.1038/s41422-019-0232-x.CrossRefGoogle Scholar
Liu, S, Luo, Y, Wang, Y, Li, S, Zhao, Z, Bi, Y, Sun, J, Peng, R, Song, H, Zhu, D, Sun, Y, Li, S, Zhang, L, Wang, W, Sun, Y, Qi, J, Yan, J, Shi, Y, Zhang, X, Wang, P, Qiu, HJ and Gao, GF (2019b) Cryo-EM structure of the African swine fever virus. Cell Host & Microbe 26(6), 836–843.e833. https://doi.org/10.1016/j.chom.2019.11.004.CrossRefGoogle Scholar
Liu, W, He, X, Zhu, Y, Li, Y, Wang, Z, Li, P, Pan, J, Wang, J, Chu, B, Yang, G, Zhang, M, He, Q, Li, Y, Li, W and Zhang, C (2024) Identification of a conserved G-quadruplex within the E165R of African swine fever virus (ASFV) as a potential antiviral target. Journal of Biological Chemistry 300(7), 107453. https://doi.org/10.1016/j.jbc.2024.107453.CrossRefGoogle ScholarPubMed
Lokhandwala, S, Petrovan, V, Popescu, L, Sangewar, N, Elijah, C, Stoian, A, Olcha, M, Ennen, L, Bray, J, Bishop, RP, Waghela, SD, Sheahan, M, Rowland, RRR and Mwangi, W (2019) Adenovirus-vectored African swine fever virus antigen cocktails are immunogenic but not protective against intranasal challenge with Georgia 2007/1 isolate. Veterinary Microbiology 235, 1020. https://doi.org/10.1016/j.vetmic.2019.06.006.CrossRefGoogle Scholar
Lokhandwala, S, Waghela, SD, Bray, J, Sangewar, N, Charendoff, C, Martin, CL, Hassan, WS, Koynarski, T, Gabbert, L, Burrage, TG, Brake, D, Neilan, J and Mwangi, W (2017) Adenovirus-vectored novel African swine fever virus antigens elicit robust immune responses in swine. PLoS One 12(5), e0177007. https://doi.org/10.1371/journal.pone.0177007.CrossRefGoogle ScholarPubMed
Lopera-Madrid, J, Medina-Magues, LG, Gladue, DP, Borca, MV and Osorio, JE (2021) Optimization in the expression of ASFV proteins for the development of subunit vaccines using poxviruses as delivery vectors. Scientific Reports 11(1), 23476. https://doi.org/10.1038/s41598-021-02949-x.CrossRefGoogle ScholarPubMed
Lopera-Madrid, J, Osorio, JE, He, Y, Xiang, Z, Adams, LG, Laughlin, RC, Mwangi, W, Subramanya, S, Neilan, J, Brake, D, Burrage, TG, Brown, WC, Clavijo, A and Bounpheng, MA (2017) Safety and immunogenicity of mammalian cell derived and modified vaccinia ankara vectored African swine fever subunit antigens in swine. Veterinary Immunology and Immunopathology 185, 2033. https://doi.org/10.1016/j.vetimm.2017.01.004.CrossRefGoogle ScholarPubMed
Lopez, E, van Heerden, J, Bosch-Camos, L, Accensi, F, Navas, MJ, Lopez-Monteagudo, P, Argilaguet, J, Gallardo, C, Pina-Pedrero, S, Salas, ML, Salt, J and Rodriguez, F (2020) Live attenuated African swine fever viruses as ideal tools to dissect the mechanisms involved in cross-protection. Viruses 12(12), 1474. https://doi.org/10.3390/v12121474.CrossRefGoogle ScholarPubMed
Lyra, TM (2006) The eradication of African swine fever in Brazil, 1978-1984. Revue Scientifique Et Technique 25(1), 93103.Google ScholarPubMed
Maeda, K, West, K, Hayasaka, D, Ennis, FA and Terajima, M (2005) Recombinant adenovirus vector vaccine induces stronger cytotoxic T-cell responses than recombinant vaccinia virus vector, plasmid DNA, or a combination of these. Viral Immunology 18(4), 657667. https://doi.org/10.1089/vim.2005.18.657.CrossRefGoogle ScholarPubMed
Malmquist, WA (1962) Propagation, modification, and hemadsorption of African swine fever virus in cell cultures. American Journal of Veterinary Research 23, 241247.Google ScholarPubMed
Malmquist, WA and Hay, D (1960) Hemadsorption and cytopathic effect produced by African swine fever virus in swine bone marrow and buffy coat cultures. American Journal of Veterinary Research 21, 104108.Google ScholarPubMed
Malogolovkin, A, Burmakina, G, Titov, I, Sereda, A, Gogin, A, Baryshnikova, E and Kolbasov, D (2015) Comparative analysis of African swine fever virus genotypes and serogroups. Emerging Infectious Diseases 21(2), 312315. https://doi.org/10.3201/eid2102.140649.CrossRefGoogle ScholarPubMed
Manso-Ribeiro, JN-PJ, Lopez-Frazao, F and Sobral, M (1963) Vaccination against ASF. Bulletin de L’office International Des Epizooties 60, 921937.Google Scholar
Marin-Moraleda, D, Munoz-Basagoiti, J, Tort-Miro, A, Navas, MJ, Munoz, M, Vidal, E, Cobos, A, Martin-Mur, B, Meas, S, Motuzova, V, Chang, CY, Gut, M, Accensi, F, Pina-Pedrero, S, Nunez, JI, Esteve-Codina, A, Gavrilov, B, Rodriguez, F, Liu, L and Argilaguet, J (2024) Elucidating the onset of cross-protective immunity after intranasal vaccination with the attenuated African swine fever vaccine candidate BA71DeltaCD2. Vaccines (Basel) 12(5). https://doi.org/10.3390/vaccines12050517.Google ScholarPubMed
Matamoros, T, Alejo, A, Rodríguez, JM, Hernáez, B, Guerra, M, Fraile-Ramos, A and Andrés, G (2020) African swine fever virus protein pE199L mediates virus entry by enabling membrane fusion and core penetration. mBio 11(4). https://doi.org/10.1128/mBio.00789-20.CrossRefGoogle ScholarPubMed
Matson, BA (1960) An outbreak of African swine fever in Nyasaland. Bulletin of Epizootic Diseases of Africa 8, 305308.Google Scholar
Mazur-Panasiuk, N and Woźniakowski, G (2020) Natural inactivation of African swine fever virus in tissues: influence of temperature and environmental conditions on virus survival. Veterinary Microbiology 242, 108609. https://doi.org/10.1016/j.vetmic.2020.108609.CrossRefGoogle ScholarPubMed
Mazur-Panasiuk, N, Zmudzki, J and Wozniakowski, G (2019) African swine fever virus - persistence in different environmental conditions and the possibility of its indirect transmission. Journal of Veterinary Research 63(3), 303310. https://doi.org/10.2478/jvetres-2019-0058.CrossRefGoogle ScholarPubMed
Mendes, AM (1971) Algumas doenc¸ as dos animais em Angola e Moc¸ ambique e sua importância na higiene das carnes. Revista Portuguesa de Ciências Veterinárias 66(420), 271286.Google Scholar
Milian-Suazo, F, Gutierrez-Pabello, JA, Bojorquez-Narvaez, L, Anaya-Escalera, AM, Canto-Alarcon, GJ, Gonzalez-Enriquez, JL and Campos-Guillen, J (2011) IFN-g response to vaccination against tuberculosis in dairy heifers under commercial settings. Research in Veterinary Science 90(3), 419424. https://doi.org/10.1016/j.rvsc.2010.07.018.CrossRefGoogle ScholarPubMed
Misinzo, G, Magambo, J, Masambu, J, Yongolo, MG, Van Doorsselaere, J and Nauwynck, HJ (2011) Genetic characterization of African swine fever viruses from a 2008 outbreak in Tanzania. Transboundary and Emerging Diseases 58(1), 8692. https://doi.org/10.1111/j.1865-1682.2010.01177.x.CrossRefGoogle ScholarPubMed
Monteagudo, PL, Lacasta, A, Lopez, E, Bosch, L, Collado, J, Pina-Pedrero, S, Correa-Fiz, F, Accensi, F, Navas, MJ, Vidal, E, Bustos, MJ, Rodriguez, JM, Gallei, A, Nikolin, V, Salas, ML and Rodriguez, F (2017) BA71DeltaCD2: A new recombinant live attenuated African swine fever virus with cross-protective capabilities. Journal of Virology 91(21). https://doi.org/10.1128/JVI.01058-17.CrossRefGoogle ScholarPubMed
Montgomery, RE (1921) On a form of swine fever occurring in British east Africa (Kenya Colony). Journal of Comparative Pathology 34(3/4), 159–191, 243–262.10.1016/S0368-1742(21)80031-4CrossRefGoogle Scholar
Mulumba-Mfumu, LK, Goatley, LC, Saegerman, C, Takamatsu, HH and Dixon, LK (2016) Immunization of African indigenous pigs with attenuated genotype I African swine fever virus OURT88/3 induces protection against challenge with virulent strains of genotype I. Transboundary and Emerging Diseases 63(5), e323327. https://doi.org/10.1111/tbed.12303.CrossRefGoogle ScholarPubMed
Mulumba-Mfumu, LK, Saegerman, C, Dixon, LK, Madimba, KC, Kazadi, E, Mukalakata, NT, Oura, CAL, Chenais, E, Masembe, C, Stahl, K, Thiry, E and Penrith, ML (2019) African swine fever: update on Eastern, Central and Southern Africa. Transboundary and Emerging Diseases 66(4), 14621480. https://doi.org/10.1111/tbed.13187.Google ScholarPubMed
Mur, L, Atzeni, M, Martinez-Lopez, B, Feliziani, F, Rolesu, S and Sanchez-Vizcaino, JM (2016) Thirty-five-year presence of African swine fever in sardinia: history, evolution and risk factors for disease maintenance. Transboundary and Emerging Diseases 63(2), e165177. https://doi.org/10.1111/tbed.12264.CrossRefGoogle ScholarPubMed
Neilan, JG, Zsak, L, Lu, Z, Burrage, TG, Kutish, GF and Rock, DL (2004) Neutralizing antibodies to African swine fever virus proteins p30, p54, and p72 are not sufficient for antibody-mediated protection. Virology 319(2), 337342. https://doi.org/10.1016/j.virol.2003.11.011.CrossRefGoogle Scholar
Netherton, CL, Goatley, LC, Reis, AL, Portugal, R, Nash, RH, Morgan, SB, Gault, L, Nieto, R, Norlin, V, Gallardo, C, Ho, CS, Sánchez-Cordón, PJ, Taylor, G and Dixon, LK (2019) Identification and Immunogenicity of African Swine Fever Virus Antigens. Frontiers in Immunology 10, 1318. https://doi.org/10.3389/fimmu.2019.01318.CrossRefGoogle ScholarPubMed
Ni, Z, Chen, L, Yun, T, Xie, R, Ye, W, Hua, J, Zhu, Y and Zhang, C (2023) Inactivation performance of pseudorabies virus as African swine fever virus surrogate by four commercialized disinfectants. Vaccines (Basel) 11(3). https://doi.org/10.3390/vaccines11030579.Google ScholarPubMed
Niederwerder, MC, Stoian, AMM, Rowland, RRR, Dritz, SS, Petrovan, V, Constance, LA, Gebhardt, JT, Olcha, M, Jones, CK, Woodworth, JC, Fang, Y, Liang, J and Hefley, TJ (2019) Infectious dose of African swine fever virus when consumed naturally in liquid or feed. Emerging Infectious Diseases 25(5), 891897. https://doi.org/10.3201/eid2505.181495.CrossRefGoogle ScholarPubMed
Njau, EP, Machuka, EM, Cleaveland, S, Shirima, GM, Kusiluka, LJ, Okoth, EA and Pelle, R (2021) African swine fever virus (ASFV): biology, genomics and genotypes circulating in sub-Saharan Africa. Viruses 13(11), 2285. https://doi.org/10.3390/v13112285.CrossRefGoogle ScholarPubMed
O’Donnell, V, Holinka, LG, Krug, PW, Gladue, DP, Carlson, J, Sanford, B, Alfano, M, Kramer, E, Lu, Z, Arzt, J, Reese, B, Carrillo, C, Risatti, GR and Borca, MV (2015) African swine fever virus Georgia 2007 with a deletion of virulence-associated gene 9GL (B119L), when administered at low doses, leads to virus attenuation in swine and induces an effective protection against homologous challenge. Journal of Virology 89(16), 85568566. https://doi.org/10.1128/JVI.00969-15.CrossRefGoogle Scholar
O’Donnell, V, Holinka, LG, Sanford, B, Krug, PW, Carlson, J, Pacheco, JM, Reese, B, Risatti, GR, Gladue, DP and Borca, MV (2016) African swine fever virus Georgia isolate harboring deletions of 9GL and MGF360/505 genes is highly attenuated in swine but does not confer protection against parental virus challenge. Virus Research 221, 814. https://doi.org/10.1016/j.virusres.2016.05.014.CrossRefGoogle Scholar
Oh, SI, NA, B, Vn, B, Dao, DT, Cho, A, Lee, HG, Jung, YH, Do, YJ, Kim, E, Bok, EY, Hur, TY and Lee, HS (2023) Pathobiological analysis of African swine fever virus contact-exposed pigs and estimation of the basic reproduction number of the virus in Vietnam. Porcine Health Management 9(1), 30. https://doi.org/10.1186/s40813-023-00330-0.CrossRefGoogle ScholarPubMed
Oura, CAL, Denyer, MS, Takamatsu, H and Parkhouse, RME (2005) In vivo depletion of CD8+ T lymphocytes abrogates protective immunity to African swine fever virus. Journal of General Virology 86(Pt 9), 24452450. https://doi.org/10.1099/vir.0.81038-0.CrossRefGoogle ScholarPubMed
Oviedo, JM, Rodríguez, F, Gómez-Puertas, P, Brun, A, Gómez, N, Alonso, C and Escribano, JM (1997) High level expression of the major antigenic African swine fever virus proteins p54 and p30 in baculovirus and their potential use as diagnostic reagents. Journal of Virological Methods 64(1), 2735. https://doi.org/10.1016/s0166-0934(96)02140-4.CrossRefGoogle ScholarPubMed
Pavone, S, Iscaro, C, Dettori, A and Feliziani, F (2023) African swine fever: the state of the art in Italy. Animals (Basel) 13(19). https://doi.org/10.3390/ani13192998.Google ScholarPubMed
Penrith, ML, Vosloo, W, Jori, F and Bastos, AD (2013) African swine fever virus eradication in Africa. Virus Research 173(1), 228246. https://doi.org/10.1016/j.virusres.2012.10.011.CrossRefGoogle ScholarPubMed
Pereira de Oliveira, R, Hutet, E, Paboeuf, F, Duhayon, M, Boinas, F, Perez de Leon, A, Filatov, S, Vial, L and Le Potier, MF (2019) Comparative vector competence of the Afrotropical soft tick Ornithodoros moubata and Palearctic species, O. erraticus and O. verrucosus, for African swine fever virus strains circulating in Eurasia. PLoS One 14(11), e0225657. https://doi.org/10.1371/journal.pone.0225657.CrossRefGoogle Scholar
Perez-Nunez, D, Sunwoo, SY, Garcia-Belmonte, R, Kim, C, Vigara-Astillero, G, Riera, E, Kim, DM, Jeong, J, Tark, D, Ko, YS, You, YK and Revilla, Y (2022) Recombinant African swine fever virus arm/07/CBM/c2 lacking CD2v and A238L is attenuated and protects pigs against virulent Korean paju strain. Vaccines (Basel) 10(12). https://doi.org/10.3390/vaccines10121992.Google ScholarPubMed
Petrovan, V, Yuan, F, Li, Y, Shang, P, Murgia, MV, Misra, S, Rowland, RRR and Fang, Y (2019) Development and characterization of monoclonal antibodies against p30 protein of African swine fever virus. Virus Research 269, 197632. https://doi.org/10.1016/j.virusres.2019.05.010.CrossRefGoogle ScholarPubMed
Portugal, R, Leitão, A and Martins, C (2009) Apoptosis in porcine macrophages infected in vitro with African swine fever virus (ASFV) strains with different virulence. Archives of Virology 154(9), 14411450. https://doi.org/10.1007/s00705-009-0466-x.CrossRefGoogle ScholarPubMed
Quembo, CJ, Jori, F, Vosloo, W and Heath, L (2018) Genetic characterization of African swine fever virus isolates from soft ticks at the wildlife/domestic interface in Mozambique and identification of a novel genotype. Transboundary and Emerging Diseases 65(2), 420431. https://doi.org/10.1111/tbed.12700.CrossRefGoogle ScholarPubMed
Rajukumar, K, Senthilkumar, D, Venkatesh, G, Singh, F, Patil, VP, Kombiah, S, Tosh, C, Dubey, CK, Sen, A, Barman, NN, Chakravarty, A, Dutta, B, Pegu, SR, Bharali, A and Singh, VP (2021) Genetic characterization of African swine fever virus from domestic pigs in India. Transboundary and Emerging Diseases 68(5), 26872692. https://doi.org/10.1111/tbed.13986.CrossRefGoogle ScholarPubMed
Ramirez-Medina, E, Vuono, E, Silva, E, Rai, A, Valladares, A, Pruitt, S, Espinoza, N, Velazquez-Salinas, L, Borca, MV and Gladue, DP (2022) Evaluation of the deletion of MGF110-5L-6L on swine virulence from the pandemic strain of African swine fever virus and use as a DIVA marker in vaccine candidate ASFV-G-deltai177l. Journal of Virology 96, e0059722. https://doi.org/10.1128/jvi.00597-22.CrossRefGoogle ScholarPubMed
Rathakrishnan, A, Moffat, K, Reis, AL and Dixon, LK (2020) Production of recombinant African swine fever viruses: speeding up the process. Viruses 12(6), 615. https://doi.org/10.3390/v12060615.CrossRefGoogle ScholarPubMed
Redrejo-Rodríguez, M, Rodríguez, JM, Suárez, C, Salas, J and Salas, ML (2013) Involvement of the reparative DNA polymerase pol X of African swine fever virus in the maintenance of viral genome stability in vivo. Journal of Virology 87(17), 97809787. https://doi.org/10.1128/jvi.01173-13.CrossRefGoogle ScholarPubMed
Redrejo-Rodríguez, M and Salas, ML (2014) Repair of base damage and genome maintenance in the nucleo-cytoplasmic large DNA viruses. Virus Research 179, 1225. https://doi.org/10.1016/j.virusres.2013.10.017.CrossRefGoogle ScholarPubMed
Reis, AL, Abrams, CC, Goatley, LC, Netherton, C, Chapman, DG, Sanchez-Cordon, P and Dixon, LK (2016) Deletion of African swine fever virus interferon inhibitors from the genome of a virulent isolate reduces virulence in domestic pigs and induces a protective response. Vaccine 34(39), 46984705. https://doi.org/10.1016/j.vaccine.2016.08.011.CrossRefGoogle ScholarPubMed
Rock, KL and Shen, L (2005) Cross-presentation: underlying mechanisms and role in immune surveillance. Immunological Reviews 207, 166183. https://doi.org/10.1111/j.0105-2896.2005.00301.x.CrossRefGoogle ScholarPubMed
Rodríguez, JM, García-Escudero, R, Salas, ML and Andrés, G (2004) African swine fever virus structural protein p54 is essential for the recruitment of envelope precursors to assembly sites. Journal of Virology 78(8), 4299–1313. https://doi.org/10.1128/jvi.78.8.4299-4313.2004.CrossRefGoogle ScholarPubMed
Rodríguez, JM, Yáñez, RJ, Almazán, F, Viñuela, E and Rodriguez, JF (1993) African swine fever virus encodes a CD2 homolog responsible for the adhesion of erythrocytes to infected cells. Journal of Virology 67(9), 53125320. https://doi.org/10.1128/jvi.67.9.5312-5320.1993.CrossRefGoogle ScholarPubMed
Rowlands, RJ, Michaud, V, Heath, L, Hutchings, G, Oura, C, Vosloo, W, Dwarka, R, Onashvili, T, Albina, E and Dixon, LK (2008) African swine fever virus isolate, Georgia, 2007. Emerging Infectious Diseases 14(12), 18701874. https://doi.org/10.3201/eid1412.080591.CrossRefGoogle ScholarPubMed
Salas, ML and Andrés, G (2013) African swine fever virus morphogenesis. Virus Research 173(1), 2941. https://doi.org/10.1016/j.virusres.2012.09.016.CrossRefGoogle ScholarPubMed
Salguero, FJ, Gil, S, Revilla, Y, Gallardo, C, Arias, M and Martins, C (2008) Cytokine mRNA expression and pathological findings in pigs inoculated with African swine fever virus (E-70) deleted on A238L. Veterinary Immunology and Immunopathology 124(1–2), 107119. https://doi.org/10.1016/j.vetimm.2008.02.012.CrossRefGoogle ScholarPubMed
Salguero, FJ, Sánchez-Cordón, PJ, Núñez, A, Fernández de Marco, M and Gómez-Villamandos, JC (2005) Proinflammatory cytokines induce lymphocyte apoptosis in acute African swine fever infection. Journal of Comparative Pathology 132(4), 289302. https://doi.org/10.1016/j.jcpa.2004.11.004.CrossRefGoogle ScholarPubMed
Sanchez-Cordon, PJ, Chapman, D, Jabbar, T, Reis, AL, Goatley, L, Netherton, CL, Taylor, G, Montoya, M and Dixon, L (2017) Different routes and doses influence protection in pigs immunised with the naturally attenuated African swine fever virus isolate OURT88/3. Antiviral Research 138, 18. https://doi.org/10.1016/j.antiviral.2016.11.021.CrossRefGoogle ScholarPubMed
Sanchez-Cordon, PJ, Floyd, T, Hicks, D, Crooke, HR, McCleary, S, McCarthy, RR, Strong, R, Dixon, LK, Neimanis, A, Wikstrom-Lassa, E, Gavier-Widen, D and Nunez, A (2021) Evaluation of lesions and viral antigen distribution in domestic pigs inoculated intranasally with African Swine fever virus Ken05/Tk1 (Genotype X). Pathogens 10(6), 768. https://doi.org/10.3390/pathogens10060768.CrossRefGoogle ScholarPubMed
Sánchez-Cordón, PJ, Montoya, M, Reis, AL and Dixon, LK (2018) African swine fever: a re-emerging viral disease threatening the global pig industry. The Veterinary Journal 233, 4148. https://doi.org/10.1016/j.tvjl.2017.12.025.CrossRefGoogle ScholarPubMed
Sánchez, EG, Pérez-Núñez, D and Revilla, Y (2017) Mechanisms of entry and endosomal pathway of African swine fever virus. Vaccines (Basel) 5(4). https://doi.org/10.3390/vaccines5040042.Google ScholarPubMed
Sánchez, EG, Quintas, A, Nogal, M, Castelló, A and Revilla, Y (2013) African swine fever virus controls the host transcription and cellular machinery of protein synthesis. Virus Research 173(1), 5875. https://doi.org/10.1016/j.virusres.2012.10.025.CrossRefGoogle ScholarPubMed
Sang, H, Miller, G, Lokhandwala, S, Sangewar, N, Waghela, SD, Bishop, RP and Mwangi, W (2020) Progress toward development of effective and safe African swine fever virus vaccines. Frontiers in Veterinary Science 7, 84. https://doi.org/10.3389/fvets.2020.00084.CrossRefGoogle ScholarPubMed
Sereda, AD, Kazakova, AS, Namsrayn, SG, Vlasov, ME, Sindryakova, IP and Kolbasov, DV (2023) Subsequent immunization of pigs with African swine fever virus (ASFV) seroimmunotype IV vaccine strain FK-32/135 and by recombinant plasmid DNA containing the CD2v derived from MK-200 ASFV seroimmunotype III strain does not protect from challenge with ASFV seroimmunotype III. Vaccines (Basel) 11(5). https://doi.org/10.3390/vaccines11051007.Google Scholar
Sereda, AD, Vlasov, ME, Koltsova, GS, Morgunov, SY, Kudrjashov, DA, Sindryakova, IP, Kolbasova, OL, Lyska, VM, Koltsov, AY, Zhivoderov, SP, Pivova, EY, Baluishev, VM, Gogin, AE and Kolbasov, DV (2022) Immunobiological Characteristics of the attenuated African swine fever virus strain katanga-350. Viruses 14(8), 1630. https://doi.org/10.3390/v14081630.CrossRefGoogle ScholarPubMed
Simões, M, Freitas, FB, Leitão, A, Martins, C and Ferreira, F (2019) African swine fever virus replication events and cell nucleus: new insights and perspectives. Virus Research 270, 197667. https://doi.org/10.1016/j.virusres.2019.197667.CrossRefGoogle Scholar
Steyn (1932) East African virus disease in pigs. In: 18th Report of the Director of Veterinary Services and Animal Industry, Union of South Africa 1, pp. 99109.Google Scholar
Sun, E, Huang, L, Zhang, X, Zhang, J, Shen, D, Zhang, Z, Wang, Z, Huo, H, Wang, W, Huangfu, H, Wang, W, Li, F, Liu, R, Sun, J, Tian, Z, Xia, W, Guan, Y, He, X, Zhu, Y, Zhao, D and Bu, Z (2021) Genotype I African swine fever viruses emerged in domestic pigs in China and caused chronic infection. Emerging Microbes & Infections 10(1), 21832193. https://doi.org/10.1080/22221751.2021.1999779.CrossRefGoogle Scholar
Sun, E, Zhang, Z, Wang, Z, He, X, Zhang, X, Wang, L, Wang, W, Huang, L, Xi, F, Huangfu, H, Tsegay, G, Huo, H, Sun, J, Tian, Z, Xia, W, Yu, X, Li, F, Liu, R, Guan, Y, Zhao, D and Bu, Z (2021) Emergence and prevalence of naturally occurring lower virulent African swine fever viruses in domestic pigs in China in 2020. Science China Life Sciences 64(5), 752765. https://doi.org/10.1007/s11427-021-1904-4.CrossRefGoogle Scholar
Sunwoo, SY, Pérez-Núñez, D, Morozov, I, Sánchez, EG, Gaudreault, NN, Trujillo, JD, Mur, L, Nogal, M, Madden, D, Urbaniak, K, Kim, IJ, Ma, W, Revilla, Y and Richt, JA (2019) DNA-protein vaccination strategy does not protect from challenge with African swine fever virus Armenia 2007 strain. Vaccines (Basel) 7(1). https://doi.org/10.3390/vaccines7010012.Google Scholar
Tidona C, D (2011) The Springer Index of Viruses. New York: Springer.10.1007/978-0-387-95919-1CrossRefGoogle Scholar
Tran, XH, Phuong, LTT, Huy, NQ, Thuy, DT, Nguyen, VD, Quang, PH, Ngon, QV, Rai, A, Gay, CG, Gladue, DP and Borca, MV (2022) Evaluation of the safety profile of the ASFV vaccine candidate ASFV-G-DeltaI177L. Viruses 14(5), 896. https://doi.org/10.3390/v14050896.CrossRefGoogle ScholarPubMed
Tury, E, Ramos, JR and Urquiaga, R (1973) Pathological anatomic experiences in the African swine fever outbreak in Cuba in 1971. Acta Veterinaria Academiae Scientiarum Hungaricae 23(4), 389409.Google ScholarPubMed
Urbano, AC and Ferreira, F (2022) African swine fever control and prevention: an update on vaccine development. Emerging Microbes & Infections 11(1), 20212033. https://doi.org/10.1080/22221751.2022.2108342.CrossRefGoogle ScholarPubMed
van den Born, E, Olasz, F, Meszaros, I, Goltl, E, Olah, B, Joshi, J, van Kilsdonk, E, Segers, R and Zadori, Z (2025) African swine fever virus vaccine strain Asfv-G-∆I177l reverts to virulence and negatively affects reproductive performance. NPJ Vaccines 10(1), 46. https://doi.org/10.1038/s41541-025-01099-9.CrossRefGoogle ScholarPubMed
Wambura, PN, Masambu, J and Msami, H (2006) Molecular diagnosis and epidemiology of African swine fever outbreaks in Tanzania. Veterinary Research Communication 30(6), 667672. https://doi.org/10.1007/s11259-006-3280-x.CrossRefGoogle ScholarPubMed
Wang, A, Jiang, M, Liu, H, Liu, Y, Zhou, J, Chen, Y, Ding, P, Wang, Y, Pang, W, Qi, Y and Zhang, G (2021a) Development and characterization of monoclonal antibodies against the N-terminal domain of African swine fever virus structural protein, p54. International Journal of Biological Macromolecules 180, 203211. https://doi.org/10.1016/j.ijbiomac.2021.03.059.CrossRefGoogle Scholar
Wang, F, Zhang, H, Hou, L, Yang, C and Wen, Y (2021b) Advance of African swine fever virus in recent years. Research in Veterinary Science 136, 535539. https://doi.org/10.1016/j.rvsc.2021.04.004.CrossRefGoogle Scholar
Wang, N, Zhao, D, Wang, J, Zhang, Y, Wang, M, Gao, Y, Li, F, Wang, J, Bu, Z, Rao, Z and Wang, X (2019) Architecture of African swine fever virus and implications for viral assembly. Science 366(6465), 640644. https://doi.org/10.1126/science.aaz1439.CrossRefGoogle ScholarPubMed
Wang, Z, Zhang, J, Li, F, Zhang, Z, Chen, W, Zhang, X, Sun, E, Zhu, Y, Liu, R, He, X, Bu, Z and Zhao, D (2024) The attenuated African swine fever vaccine HLJ/18-7GD provides protection against emerging prevalent genotype II variants in China. Emerging Microbes & Infections 13(1), 2300464. https://doi.org/10.1080/22221751.2023.2300464.CrossRefGoogle ScholarPubMed
Wilkinson, PJ (1989) African swine fever virus. In Pensaert, M.C., Virus Infections of Porciness. Amsterdam: Elsevier, 1732.Google Scholar
Xia, N, Wang, H, Liu, X, Shao, Q, Ao, D, Xu, Y, Jiang, S, Luo, J, Zhang, J, Chen, N, Meurens, F, Zheng, W and Zhu, J (2020) African swine fever virus structural protein p17 inhibits cell proliferation through ER stress-ROS mediated cell cycle arrest. Viruses 13(1), 21. https://doi.org/10.3390/v13010021.CrossRefGoogle ScholarPubMed
Xie, Z, Liu, Y, Di, D, Liu, J, Gong, L, Chen, Z, Li, Y, Yu, W, Lv, L, Zhong, Q, Song, Y, Liao, X, Song, Q, Wang, H and Chen, H (2022) Protection evaluation of a five-gene-deleted African swine fever virus vaccine candidate against homologous challenge. Frontiers in Microbiology 13, 902932. https://doi.org/10.3389/fmicb.2022.902932.CrossRefGoogle ScholarPubMed
Yutin, N, Wolf, YI, Raoult, D and Koonin, EV (2009) Eukaryotic large nucleo-cytoplasmic DNA viruses: clusters of orthologous genes and reconstruction of viral genome evolution. Virology Journal 6, 223. https://doi.org/10.1186/1743-422X-6-223.CrossRefGoogle ScholarPubMed
Zhang, H, Zhao, S, Zhang, H, Qin, Z, Shan, H and Cai, X (2023) Vaccines for African swine fever: an update. Frontiers in Microbiology 14, 1139494. https://doi.org/10.3389/fmicb.2023.1139494.CrossRefGoogle ScholarPubMed
Zhao, D, Sun, E, Huang, L, Ding, L, Zhu, Y, Zhang, J, Shen, D, Zhang, X, Zhang, Z, Ren, T, Wang, W, Li, F, He, X and Bu, Z (2023) Highly lethal genotype I and II recombinant African swine fever viruses detected in pigs. Nature Communications 14(1), 3096. https://doi.org/10.1038/s41467-023-38868-w.CrossRefGoogle ScholarPubMed
Zhu, JJ, Ramanathan, P, Bishop, EA, O’Donnell, V, Gladue, DP and Borca, MV (2019) Mechanisms of African swine fever virus pathogenesis and immune evasion inferred from gene expression changes in infected swine macrophages. PLoS One 14(11), e0223955. https://doi.org/10.1371/journal.pone.0223955.CrossRefGoogle ScholarPubMed
Figure 0

Figure 1. Morphological structure of ASFV (Liu et al., 2019b). (Reproduced directly from Liu et al., 2019b Cell Host Microbe with permission from the Cell Press).

Figure 1

Figure 2. Schematic diagram of ASFV’s structure.

Figure 2

Figure 3. Global distribution of ASFV as of 29 June 2025.

Note: The number on the map represents the number of outbreaks of ASF that occurred in the area from 3 October 2018 to 29 June 2025.
Figure 3

Figure 4. Annotated genome structure of ASFV (Dixon et al., 2013; Hakizimana et al., 2021). (Reproduced directly from Dixon et al.2013 Virus Research with permission from the Elsevier and slightly modified according to Hakizimana et al., 2021.) A, The overall genomic structure of open reading frames (ORFs) in ASFV, represented by Georgia 2007/1. B, The arrangement of genes (open reading frames [ORFs]) in the genome of ASFV, represented by Georgia 2007/1. The direction and size of the arrows indicate the transcriptional direction and length of the ORFs, respectively. The different colours indicate the various functions of the ORFs. The black arrow indicates ORFs encoding enzymes and factors involved in genome replication, repair, or transcription. The grey arrow indicates ORFs that can be expressed as proteins. The pink arrow suggests ORFs associated with escape from the host immune system. The yellow arrow indicates ORFs with other predicted functions. The white arrow indicates ORFs with unknown functions. The turquoise, blue, green, brown, and mauve arrows indicate members of multigene families. Deletion of the genes shown in red can reduce virulence.