Next Article in Journal
The Efficiency of TB LAM Antigen Test to Xpert MTB/RIF Ultra Test for the Diagnosis of Tuberculous Pericarditis Using Pericardial Fluid Samples
Next Article in Special Issue
Ocular Manifestations of Flavivirus Infections
Previous Article in Journal
Characterization of Bovine Intraepithelial T Lymphocytes in the Gut
Previous Article in Special Issue
West Nile, Sindbis and Usutu Viruses: Evidence of Circulation in Mosquitoes and Horses in Tunisia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Rift Valley Fever Virus—Infection, Pathogenesis and Host Immune Responses

by
Niranjana Nair
,
Albert D. M. E. Osterhaus
,
Guus F. Rimmelzwaan
and
Chittappen Kandiyil Prajeeth
*,†
Research Center for Emerging Infections and Zoonoses, University of Veterinary Medicine Hannover, Foundation, 30559 Hannover, Germany
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Pathogens 2023, 12(9), 1174; https://doi.org/10.3390/pathogens12091174
Submission received: 21 August 2023 / Revised: 9 September 2023 / Accepted: 14 September 2023 / Published: 19 September 2023
(This article belongs to the Special Issue Current Advances in Flavivirus and Other Arboviruses)

Abstract

:
Rift Valley Fever Virus is a mosquito-borne phlebovirus causing febrile or haemorrhagic illness in ruminants and humans. The virus can prevent the induction of the antiviral interferon response through its NSs proteins. Mutations in the NSs gene may allow the induction of innate proinflammatory immune responses and lead to attenuation of the virus. Upon infection, virus-specific antibodies and T cells are induced that may afford protection against subsequent infections. Thus, all arms of the adaptive immune system contribute to prevention of disease progression. These findings will aid the design of vaccines using the currently available platforms. Vaccine candidates have shown promise in safety and efficacy trials in susceptible animal species and these may contribute to the control of RVFV infections and prevention of disease progression in humans and ruminants.

1. Introduction

Bunyaviruses are among the largest orders of RNA viruses comprising nine virus families and are transmitted by arthropods or rodents. Viruses belonging to the Phlebovirus genus of the Phenuiviridae infect multiple hosts and have a major impact on agriculture, animal husbandry and human health. [1,2,3,4,5]. Phleboviruses contain more than 60 different species and are predominantly transmitted by arthropods such as ticks, mosquitoes and sandflies [6]. Rift Valley Fever Virus (RVFV) is one of the phleboviruses and was first identified in Rift Valley of Africa in the early 1930s, initially as a causative agent of neonatal mortality and abortion storms in sheep and, subsequently, also as a cause of febrile illness in man [7,8]. Primarily domesticated ruminants such as sheep, goat, cattle and ungulates like camels were considered to be most susceptible, however, there is also serological evidence of infection in several other wild animals such as impalas, giraffes, pigs, warthogs, etc. [9,10,11,12,13].
RVFV is transmitted among ruminants by mosquitoes such as Aedes, Anopheles and Culex species, with large outbreaks occurring frequently during heavy rainfall [14,15,16]. Mosquito-borne transmission among humans is rare and the major mode of transmission is by direct contact with infected blood, carcass or by the consumption of raw milk from infected ruminants [17,18,19]. Signs of infection in ruminants generally start with fever and could manifest into respiratory illness, hepatic necrosis, abortions, stillbirths and death [20]. In humans, the infection is mostly asymptomatic or a self-limiting febrile illness. However, in severe cases, liver damage, haemorrhagic illnesses, encephalitis and neurological disorders have been reported [21]. Although miscarriages in pregnant women have been attributed to RVFV outbreaks in endemic areas, data supporting such clinical outcomes is currently lacking [22].
Following the first isolation of RVFV in 1930s, multiple RVF epizootic outbreaks have been reported. The outbreaks were initially limited to Kenya and later spread across the eastern parts of Africa. Over the past few decades, the disease was reported in most African countries, Saudi Arabia, Yemen and serological evidence was obtained from Turkey suggesting its spread to a wider geographical area [3,23,24]. Climate change has increased the average temperatures of several countries, providing a favourable condition for more extended vector breeding. Livestock trade and international travel have introduced the disease to previously unreported territories. The possibility of aerosol transmission poses RVFV as a high-risk pathogen [25]. The geographical expansion of affected areas, along with the lack of a highly efficient and safe vaccine, underscores the unmet need for novel intervention strategies. Obtaining a better understanding of the pathogenesis of RVFV infections and host–virus interaction may aid the development of novel prophylactic and therapeutic approaches. This review aims to provide an outlook on the host immune response against RVFV. We briefly review the characteristic features of RVFV, the replication cycle and host susceptibility. We also discuss the recent advancements in understanding the innate and adaptive immune responses to RVFV infection. Finally, we discuss vaccines that were used in the past and the ones that are being developed for future use against the virus.

2. Virus Biology

The RVF virion is approximately 100 nm in size and has an envelope decorated with surface glycoproteins. CryoEM reconstruction reveals a quasisymmetrical, icosahedral lattice with pentameric and hexameric glycoprotein capsomers [26]. Characteristic for all bunyaviruses, the negative-sense RNA genome consists of three gene segments, S, M and L, encoding several structural and non-structural proteins. The genomic S RNA codes for the nucleocapsid (N) protein and the anti-genomic RNA encodes non-structural NSs protein [27]. The M segment encodes the glycoprotein precursor Gc and Gn, the non-structural protein NSm and accessory proteins P78, P14 and P13. The L segment encodes the viral RNA dependent RNA polymerase (RdRp). Each genome segment is flanked by complementary 3′ and 5′ untranslated regions (UTRs), thereby forming a panhandle structure. The N protein is a hexameric unit with the N terminal end enabling oligomerization and a concave cleft in the centre for binding to viral RNA. The N protein-bound RNA is thus packaged into the virion, protected from RNA-degrading enzymes and innate immune sensors of the host [27,28,29]. The viral genome segments are encapsulated by the N protein along the length of genome and the RdRp at the UTRs, creating a ribonucleoprotein (RNP) complex, essential for replication and transcription. The structure and amino acid sequence of RdRp is highly conserved among the Bunyaviridae [30]. As a result, both experimental evidence and in silico modelling suggest that the viral RdRp is an ideal target for antiviral drugs because mutations in the conserved motifs that would confer drug resistance have proven to attenuate the virus [31,32]. The M segment contains five distinct sites for translation initiation, resulting in the synthesis of different polyproteins. The signal peptides influence the translocation into organelles, differential cleaving and processing of the polyproteins to generate functional proteins [33]. The accessory protein P78, formed by the translation from first AUG, is a glycoprotein localizing to the Golgi complex and playing a role in the dissemination of virus particles in mosquito models [33]. The glycoprotein precursor molecule could be generated from all translation initiation sites and is post translationally cleaved to form the Gn spike protein and Gc class II fusion protein. Gn glycoprotein has head and stem domains, with the head domain exposed on the capsomer surface [34]. The Gc protein contains three domains mainly comprised of β-sheets, with domain II functioning as the fusion loop [35]. The structural biology approach predicts that the Gn protein forms a ring structure in pentamers, thereby shielding the Gc fusion loop. Functionally, the location of Gn on Gc might be relevant for the stepwise entry of the virus into host cells, in which the interaction of Gn with cellular entry factors would trigger endocytosis, and a low pH within endosomes enables a Gc-mediated virus-membrane fusion [36]. The non-structural proteins NSm and NSs contribute to the virulence of RVFV. NSs proteins form amyloid filaments within the infected cells which aggregate into fibrils, leading to neurological symptoms in mice [37]. The protein also downregulates host factors PKR and Abl2 to enable viral translation, IFN-β suppression, as well as actin filament formation, cell migration and spreading [38]. Protein NSm is reported to be localized on the mitochondrial membrane and prevents apoptosis of the virus-infected cells [39]. NSm protein is responsible for virus replication and dissemination from the midgut of mosquitoes and its deletion impairs virus replication kinetics [40]. The structural and non-structural proteins thus facilitate host interaction, virus replication and mediate immune evasion.

3. Replication Cycle

The replication cycle starts with the attachment of the virus to the cell surface and subsequent entry into the host cell. The interaction of bunyaviruses with molecules such as DC-SIGN, L-SIGN, heparan sulphates and nonmuscle myosin heavy-chain IIA (NMMHC-IIA) expressed on the host cell surface have been reported and are likely to act as viral entry receptors [41,42,43,44,45]. More recently, low-density lipoprotein receptor-related protein 1 (Lrp1) was shown to be essential for RVFV infection [46]. Virus-receptor interaction triggers receptor clustering and subsequent endocytosis at a low pH, as observed with fluorescently tagged Uukuniemi virus (UUKV) [44]. Although caveolin-mediated endocytosis was proven to be utilized for virus entry, the role of macropinocytosis in RVFV entry cannot be ruled out [47,48]. Following endosomal uptake, the viral genome is released into the cytoplasm after pH-dependent membrane fusion, as was shown for Hantaan virus, Andes virus and RVFV [49,50]. For other bunyaviruses, like Bunyamwera virus (BUNV) and Hazara virus (HAZV), it was suggested that cellular cholesterol levels and an accumulation of potassium within endosomes play a role in membrane fusion, although it is not known whether RVFV is exploiting the same mechanism [51]. Post membrane fusion, the genome is released into the host cytoplasm and replication proceeds using the prime and realign mechanism by the RdRp and complementary RNA. Transcription is known to be initiated using cellular-capped RNA as primers, while translation is still not understood [52]. The proteins Gn-Gc heterodimerize in the endoplasmic reticulum (ER) lumen and are transported to the Golgi for post-translational modifications. Severe fever with thrombocytopenia syndrome virus (SFTSV), belonging to Phenuiviridae, was shown to upregulate the levels of autophagy-associated protein LC3-II and has been shown to exploit the autophagy machinery for the assembly and egress of nascent virion from host cells [53]. Treatment of RVFV-infected cells with sorafenib, a Raf kinase inhibitor, traps RVFV virion in the ER, hinders cellular secretory pathway thereby reducing viral egress [54]. Bunyaviruses exploit various mechanisms for their infection and replication, which is determined by host factors and viral determinants.

4. Host Susceptibility

RVFV infections can have high case fatality rates in certain species. Among ruminants, sheep and goat species are highly susceptible to RVFV infections and are fatal in up to 30% of young adult animals and in over 95% in neonatal and unborn animals [55]. Cattle and camels are less susceptible to infection. Wild animals have been found to be seropositive for RVFV antibodies, for instance 31.8% of wild ruminants surveyed in Kenya during the epizootic period in 2006–2007 were positive for anti-RVFV antibodies [11]. Another critical factor determining the susceptibility and clinical manifestation of the disease is the route of infection. For instance, intranasal administration of RVFV results in higher viremia and central nervous system (CNS) infections [56].
RVFV infection of sheep is characterized by necrosis in the liver, gall bladder, spleen, kidney, adrenal gland and gastrointestinal tract. Necrotic foci in the liver is unevenly distributed and often extends to the periportal zone, as seen in ruminants and humans [57,58]. Apoptotic features such as cell dissociation, hypereosinophilic cytoplasm, pyknosis, karyorrhexis, etc. are also observed in injured hepatocytes, with surrounding cells exhibiting vesicular degeneration. Lymphocytic infiltration is visible in liver and lungs, while lymphocytolysis is observed in lymphoid organs. Haemorrhage and oedema are seen in the spleen and lungs of infected sheep. Inclusion bodies were observed in the liver of neonatal sheep and goats. The gall bladder had noticeable necrosis, haemorrhage and oedema and neutrophil infiltration [57,58,59]. Humans present symptoms of ocular disease and encephalitis in addition to febrile disease and haemorrhage [60]. Pathological manifestations of the CNS including encephalitis is observed in humans and experimentally infected non-human primate and rodent models. A histopathological examination of infected human brains demonstrated encephalitis with an infiltration of lymphocytes and necrosis in neurons [61]. In pregnant sheep infected with RVFV during the gestation period, viral antigens were detected in trophoblasts, syncytial epithelium and endothelium of umbilical cord. This indicates that foetal infection takes place via the haemophagus zone during maternal blood phagocytosis or infecting foetal endothelium through the maternal epithelium and trophoblasts [19]. In contrast to the vertical transmission observed in ruminants, maternal–foetus transmission in humans has not been observed. Cells of the human trophoblast cell lines A3 and Jar could be infected with wild type and NSs-deficient virus strains. The RVFV strain 35/74 replicated efficiently in cytotrophoblasts of human placental explants, suggesting that transmission from human maternal trophoblasts to foetal chorionic villi via blood is possible [19,62]. Thus, RVFV can affect several organs including the gastrointestinal system, lungs, kidney, spleen, brain and/or the placenta of susceptible hosts.

5. RVFV Infection—Host Immunity and Evasion Strategies

5.1. Immune Evasion Mechanisms of RVFV

Virus invasion in the host is recognized by the intrinsic mechanism in cells prompting the immune system to launch a response against it. The host’s innate immune system has the capability of recognizing foreign molecules and pathogens based on common conserved features. Such molecular signatures known as pathogen-associated molecular patterns (PAMP), comprises of certain cell membrane components and genetic material of pathogens and are recognized by membrane-bound and intracellular pattern-recognition receptors (PRRs) of the host cells. Toll-like receptors (TLRs), RIG-I-like receptors (RLRs), C-type lectin receptors (CLRs), NOD-like receptors (NLRs) and AIM2-like receptors (ALRs) are some of the PRRs which recognize PAMPs and lead to the activation of the signalling cascade that acts to control invading pathogens. The N terminal domain of TLRs recognizes molecular patterns, while the C terminal domain is involved in signal transduction via the MyD88 pathway or the TRIF pathway [63]. TLRs 3, 7, 8, 9 recognize genetic material and are able to launch immune responses against these for pathogen clearance. The retinoic acid-inducible gene-I (RIG-I) identifies uncapped 5′-triphosphate ssRNA and short dsRNA formed as a result of viral replication, while the melanoma differentiation-associated gene 5 (MDA5) recognizes long dsRNA, thereby playing a major role in distinguishing virus and host cell genetic material. The binding of viral RNA with RLRs results in mitochondrial antiviral signalling protein (MAVS) recruitment, TBK-1 phosphorylation or NF-κB activation, which acts as a transcription factor for the interferons (IFN) and interferon-stimulated genes (ISG) [64]. IFNs interact with the IFN receptor in autocrine or paracrine manner, thereby phosphorylating STAT protein, leading to the transcriptional activation of interferon-stimulated genes under the control of an interferon-stimulated response element or gamma interferon activated sequence [65].
The RVFV genome is recognized by the RIG-I molecule and activates the downstream signalling process, including the production of interferon-inducible transmembrane proteins (IFITM) -2 and -3, which restrict the virus life cycle in early stages post internalization and prior to replication in vitro [66,67]. As an immune evasion mechanism, RVFV modifies 5′ termini and escapes RIG-I recognition [68]. Protein atypical RIO Kinase 3 (RIOK3) is an essential regulator of the IFN-β pathway and NF-κB pathway. The constitutively expressed form is crucial in IFN-β production, while the alternative splice variant is involved in NF-κB generation [69,70]. RVFV triggers inflammasomes as indicated by the interaction of NLRP3 and MAVS in infected cells [71]. RVFV infection in liver cells led to the NSs protein-dependent generation of reactive oxygen species, p65 production and triggering the NF-κB pathway [72]. The NSs proteins in the RVFV MP12 strain activate the ATM DNA damage pathway, leading to cell cycle arrest at the S phase, while NSs of ZH548 arrest the cell cycle at the G0/G1 phase. The cell cycle arrest halts host cellular processes, thereby favouring virus replication [73,74]. The exploitation of the DNA damage pathway is also utilised by other viruses for facilitating replication, including human parvovirus B19, Epstein–Barr virus, Porcine epidemic diarrhoea virus, SFTSV, etc. [75,76,77,78]. Similarly, the abrogation of interferon signalling which is critical for antiviral immunity is observed in related phleboviruses. SFTSV infection is highly reduced upon IFN-γ treatment prior or post infection. The IFN-γ signalling is disrupted by NSs proteins through sequestration and downregulation of STAT1 expression [79]. NSs proteins of sandfly fever viruses prevent the phosphorylation of Jak-1 and Stat-1 and inhibit the JAK-STAT pathway for the production of ISG [80]. In RVFV, the NSs protein interacts with the host Sin3A-associated protein 30 (SAP30) component of the histone deacetylase complex, thereby maintaining the IFN-β signalling in a transcriptionally silent state and promoting virus replication. NSs mutants with amino acid deletions which are essential for SAP30 binding abrogated its interaction with the IFN-β promoter sequence and enabled survival of the infected cells [81]. The host protein PKR has a promoter region activated by IFN that is responsible for activating immune responses in paracrine signalling. The NSs degrades double-stranded RNA (dsRNA)-dependent protein kinase (PKR), thereby ensuring that innate immunity is blocked in uninfected cells. PKR is also responsible for phosphorylating eukaryotic initiation factor 2α (eIF2α) to enable efficient host mRNA translation, which prevents viral translation. PKR degradation by NSs also functions to prevent the phosphorylation of eIF2α and ensures viral translation and shutdown of the host’s immune defence [82]. Incoming RVFV virions trigger canonical autophagy pathways in several cell lines. RVFV infection induces autophagy in mouse embryonic fibroblasts, primary mouse hepatocytes, primary rat mixed neuroglial cultures and human osteosarcoma cell lines (U2OS). Moreover, treatment of these cell lines with autophagy-activating drugs prior to infection showed antiviral activity against RVFV [83]. In contrast, RVFV infection in differentiated macrophage cells (THP-1PMA) resulted in the interaction of viral nucleoprotein and host sequestome 1 (SQSTM1), triggering autophagosome formation and aiding viral replication. However, in Huh7 cells or human umbilical vein endothelial cells (HUVEC), RVFV infection triggered autophagy, which was dispensable for virus replication [84]. Hence, RVFV-triggered autophagy has a pro-viral or anti-viral role, depending on the cell type infected. Exosomes are a category of extracellular vesicles formed as a result of endocytosis and subsequent fusion with the cell membrane [85]. Exosomes formed as a result of RVFV infection containing viral RNA could activate the RIG-I pathway, leading to Interferon-β (IFN-β) production and upregulating autophagic flux [86]. Thus, various intrinsic cellular defence pathways can influence the innate immune response to infection (Figure 1).

5.2. Innate Immune Response to RVFV Infection

RVFV infections induce strong cytokine and chemokine responses in target organs, which are not only critical for the recruitment of innate immune cells to the sites of infection but are also crucial for priming an effective adaptive immune response. Cytokine responses to RVFV infection have been sparsely studied in goats, sheep and cattle. Goats infected with the ZH501 strain had negligible levels of type I IFN production, due to the suppression of interferon production by the NSs gene. Nevertheless, IL-12 and IFN-γ levels were higher in infected goats compared to uninfected goats. Pro-inflammatory cytokines TNF-α, IL-6, IL-1β expression increased two days post infection (dpi) and contributed to the virus clearance from infected goats [87].
Human monocyte-derived macrophages (MDM) could be productively infected in vitro with wild type ZH501 and NSs-deficient (ΔNSs) strains of RVFV. As opposed to the wild type virus, the ΔNSs strain induced pro-inflammatory cytokines IFN-α2, IFN-β and TNF-α which restricted virus replication and cytopathic effect (CPE) in human MDM. [88]. The infection of murine bone marrow-derived macrophage (BMDM) with RVFV strains ZH501 or MP12 resulted in a lower expression of pro-inflammatory cytokines IFN-β, IFN-γ, TNF-α than after infection with recombinant MP12 with an in-frame deletion of 69% of the NSs gene (rMP12-C13). In general, the production of inflammatory cytokines was downregulated in ZH501-infected cells, while for other cytokines such as IL-3, IL-4, IL-10 and IL-17, the expression was comparable to rMP12-C13 and MP-12 infected cells. The diminished cytokine response to wild type virus can be mostly attributed to the immune suppressive capabilities of NSs [89]. Similar observations were also made in the human choriocarcinoma cell line Jar and the trophoblast cell line A3, in which inflammatory cytokines such as TNF-α, IL-6 and IL-8 were highly expressed [62]. Serum samples collected from infected humans during the outbreak in 2010–2011 in South Africa were tested for cytokine levels. Higher level of the anti-inflammatory IL-10 was observed in fatal cases than in the survivors and negative control subjects [90]. The level of pro-inflammatory cytokines IL-6 and IL-8 was elevated in infected subjects and correlated with liver damage, while levels of CCL5/RANTES responsible for attracting immune cells were reduced in fatal cases and correlated with haemorrhage. CXCL-9, CXCL-10 and MCP-1 levels were higher in all infected samples, compared to the control group. Thus, the balance between anti- and pro-inflammatory cytokines and chemokines largely defines the disease outcome of the infection [90].
In experimental mouse models, the genetic background of the mouse strains used determines the response to and the outcome of RVFV infection. MBT/Pas mice displayed a higher susceptibility to RVFV infection than BALB/cByJ mice, which is surprising because MBT/Pas mice had relatively high serum IFN-α levels. However, their IFNAR1 receptor expression on leukocytes was lower, which abolished the protective effect of type I IFN [91,92]. Interestingly, the site of RVFV infection is important in determining the disease manifestations. Intranasally infected mice were found to be at higher risk of high viral loads in the brain and subsequent development of encephalitis than mice infected by injection in the footpad. Also, the extent of spreading to peripheral organs was lower. The infiltration of leukocytes, the production of pro-inflammatory cytokines and a lack of virus clearance were the leading factors associated with encephalitis in mice [93].
The pathogenesis and response to RVFV infection by aerosols was also studied in non-human primates (NHP), namely African green monkeys (AGM), rhesus macaques, cynomolgus macaques and common marmosets [94]. African green monkeys and marmosets showed signs typical for RVFV, including anorexia, weight loss and neurological distress at 9–12 dpi [94]. Lethally infected monkeys had high virus loads in the CNS, while peripheral organs had a lower viral load. The delayed increase in monocytes, dendritic cells and the pro-inflammatory cytokines IFN-γ, IL-8, sCD40L in the blood of lethally infected monkeys results in an inefficient immune response and subsequent morbidity. A sublethal infection led to the upregulation of inflammatory cytokines within 3 dpi, providing protection against severe disease [95]. Thus, the timing of innate response induction is crucial in determining the survival of infected organisms [95].

5.3. Adaptive Immune Response

Sero-epidemiological studies conducted during epidemic or inter-epizootic phases in RVFV endemic regions showed that humans and ruminants have RVFV-specific antibodies [96,97,98]. In humans naturally exposed to RVFV, high titers of virus neutralizing (VN) antibodies directed against the glycoproteins were detected, in addition to low frequencies of N-, Gc- and Gn-specific T cells [99]. Virus-specific cellular and antibody responses were also observed in goats twenty-one days after experimental infection with the RVFV-ZH501 strain [87]. Also in sheep, virus-neutralizing antibody responses were observed upon RVFV infection [100]. Serum samples obtained from cattle, goats and sheep collected two months after a RVFV outbreak contained virus-specific antibodies, as demonstrated by ELISA. In this study, 65% of RT-PCR positive sera were still seronegative, suggesting that these animals were acutely infected during the outbreak [101]. Finally, non-human primates experimentally infected with RVFV developed an antibody response [94]. Thus, in various susceptible species such as ruminants, humans, wild ungulates and camels, RVFV infection induces virus-specific antibody responses [11,102,103,104]. A sublethal dose of RVFV administered in African green monkeys also led to the expansion of both cytotoxic and helper T lymphocytes in peripheral blood [95]. The induction of long-lasting, robust antibody- and cell-mediated immunity affords protection against subsequent infection in previously infected hosts.
Studying adaptive immune responses to wild type strains in the experimental infection of rodent models is often challenging due to rapid disease progression and high mortality within a few days after infection. Moreover, the adaptive immunity to wild type RVFV infection is not highly explored, compared to the response induced by viral proteins or attenuated strains. Hence, humoral and cellular responses to the respective RVFV viral proteins have been characterized using expression systems or recombinant protein. One of the first studies on RVFV-induced humoral response utilized baculovirus vector expressing glycoproteins for immunizing mice. The immunized mice developed an antibody response against the glycoproteins, as evident by the presence of antibodies and the survival of passively immunized mice [105]. Rabbits immunized with Gn glycoprotein mounted a specific antibody response. Their B cells were used to generate hybridoma cells that produced VN antibodies which afforded protection against the challenge infection in naïve recipient mice [106]. The immunization of sheep with an expression plasmid encoding the ectodomain of Gn elicited an antibody response which afforded protection, as judged by an accelerated viral clearance and reduced disease severity [107]. In naturally infected humans, Gn but not Gc glycoprotein was the major viral protein inducing a sustained VN-antibody response, while both Gn and Gc induced T cell responses [99]. The immunization of sheep with a recombinant Equine herpesvirus-1 expressing the RVFV Gn and Gc proteins elicited a serum VN-antibody response [108]. The immunization of mice with recombinant nucleocapsid protein along with Alhydrogel induced an N-specific Th2 response [109].
The importance of adaptive immunity in controlling brain pathology was studied in C57BL/6J mice depleted of B cells, CD4+ T cells or CD8+ T cells using antibodies by infecting subcutaneously with ∆NSs strain of virus. B cell-deficient mice succumbed to the challenge at 21 dpi, indicating the importance of B lymphocytes in reducing virus infection [110]. Surprisingly, CD8+ T cells were not essential for protective immunity as the CD8+ T cell-depleted mice did not succumb to disease after challenge infection. CD4+ T cell depletion resulted in high virus RNA copy numbers in the liver and brain and reduced antibody titer compared to the mock-depleted mice. CD4+ T cell-depleted mice had a higher viral load and pro-inflammatory cytokine levels in the brain, including CCL2, CCL4, CCL5, CCL7, CCL12, CXCL1, CXCL10, CXCL-13, TNF-α and IL-1β [110]. A recent study also confirmed that the NSs-deleted RVFV strain is capable of causing encephalitis in the absence of CD4+ T cells working synergistically with monocytes and CD8+ T cells [111]. Moreover, NSG mice deficient in T-, B-, natural killer (NK)-cells and macrophage-mediated immunity had viral replication in organs including the brain, with one animal developing significant clinical signs [112]. These observations lead to the conclusion that CD4+ T cells control progression to late onset encephalitis in mice.

6. RVFV Vaccine Development

Since the discovery of RVFV as a causative agent of disease in animals and humans, several attempts have been made to develop safe and effective vaccines. One of the first vaccine candidates tested was based on the Smithburn strain, which was accidentally isolated while studying Yellow Fever in Uganda by infecting mice intracerebrally with mosquito extracts. Later, it was found that some of these mosquitoes were vectors for RVFV, and the disease pathology of a few infected mice were similar to RVF manifestations [113]. Follow-up studies on the Smithburn strain suggested that the lower number of passages leads to a pantropic virus, affecting the liver predominantly, and higher passages form a neurotropic virus against which the pantropic virus induced higher antibody titers and protection [114]. Alpacas in South Africa were vaccinated with the Smithburn vaccine as a precautionary measure against RVFV, although it was previously unreported in the species. A few of the vaccinated individuals died and histopathological examination revealed meningoencephalitis and RVFV antigens in the brain. A sequence analysis confirmed the identity of the Smithburn strain isolated from the cerebrum of the diseased alpacas and excluded natural RVFV infection as the cause of death [115].

6.1. Formalin-Inactivated Vaccines

Vaccine candidates NDBR-103 and TSI-GSD-200 are formalin-inactivated virus preparations which were administered previously in high-risk groups. Even though a virus-neutralizing antibody titer was generated post vaccination, few individuals that received NDBR-103 developed Guillain-Barré Syndrome and the safety of the vaccine candidate was questioned subsequently [116,117]. Individuals working in high-risk conditions with RVFV vaccinated with TSI-GSD-200 showed a robust humoral response with sustained antibody titers. Studies on the longevity of TSI-GSD-200-induced antibody responses showed that antibody titers waned 6–12 months post vaccination, which could be overcome by booster vaccinations. The major disadvantage of the requirement of booster vaccinations is the inability to induce immediate protective antibody responses during outbreaks of RVFV infections [118,119]. Nevertheless, RVFV-specific T cell responses were detectable in immunized individuals, even 24 years post vaccination [120].

6.2. MP12

The RVFV strain MP12 was developed from the non-fatal ZH548 strain, which was passaged in suckling mouse brains, propagated in foetal rhesus lungs and human diploid fibroblasts and later mutated through alternate passaging in the absence and presence of 5-fluorouracil [121]. The MP12 vaccine strain differs from the parent virus strain ZH548 by 25 mutations, rendering it highly attenuated and temperature-sensitive [122,123]. Numerous studies have been conducted on the MP12 strain to analyse its efficacy, safety, threat of transmission or reversion to virulence. The vaccine was found to be safe in vaccinated sheep with no adverse effects and no potential transmission to mosquito vectors of four different species. MP12-induced antibodies persisted even 24 months post vaccination [124]. A single-dose MP12 vaccination was safe in alpacas and elicited an antibody response [125]. An intramuscular injection of rhesus macaques with the MP12 strain did not cause clinical signs and induced virus-specific serum IgM and IgG antibody responses, which persisted even up to 126 days post vaccination [126]. The safety of the vaccine in pregnant ewes and lambs was tested by inoculating ewes at their second trimester and neonatal lambs. The administration of MP12 did not cause lesions in peripheral organs and induced VN-antibody responses. Neonatal lambs inoculated with MP12 had virus present in the liver, spleen and kidney up to day 4 post inoculation, albeit at low titers. Newborn lambs of vaccinated ewes lacked MP-12 antibodies in peripheral blood at birth, prior to colostrum feeding. Colostrum feeding introduced maternal antibodies in the lambs, indicating acquired immunity through nursing [123]. The immunization of sheep with MP12 during the early stages of pregnancy has caused adverse effects including foetal malformation and teratogenicity [127]. Also, in four-month-old calves, the injection of MP12 caused adverse effects, including bronchopneumonia in five of ten vaccinated animals, and lesions in the liver, lung, gall bladder and trachea up to 10 days post inoculation. Viral antigen could also be detected from the liver of a vaccinated calf [128]. The efficacy of the vaccine was demonstrated in macaques by intravenous or aerosol challenge infection with the strain ZH-501 [126,129]. MP12 was also tested as a vaccine candidate in phase 1 and 2 clinical trials. Also, in human study subjects, MP12 proved to be immunogenic and induced VN antibodies, which persisted up to 5 years. [130,131].

6.3. Clone 13

Clone 13 is a live attenuated RVFV strain derived from the 74HB59 strain by plaque purification. The clone 13 has a 69% deletion in-frame within the NSs gene, with intact N- and C-terminals, and contains nucleotide substitutions in the N and NSs gene [132]. Clone 13 was evaluated for its safety and efficacy against virus challenge in susceptible species such as sheep, cattle and goats. In sheep, the administration of clone 13 was relatively safe as no abortions were reported in pregnant ewes in the early or later stages of pregnancy. The vaccinated sheep developed VN-antibody responses and were protected from challenge infection [133]. Also, in seven-month-old calves, vaccination with clone 13 induced protective immunity [134]. Field studies conducted in Senegal and Tanzania in herds of ruminants confirmed the safety and immunogenicity of clone 13 [135,136]. However, adverse effects including abortions, stillbirths, and malformations have been reported in ewes vaccinated at 50 days of gestation, indicating the vertical transmission of clone 13 during early stages of pregnancy [137]. Thus, although clone 13 proved to be efficacious, its safety especially in pregnant ruminants is still a matter of debate.

6.4. RVFV-4s

More recently, live attenuated RVFV vaccines were developed with a genome consisting of four gene segments (RVFV-4s) by splitting the Gn and Gc segments and deleting the NSs gene [138]. Two versions of the vaccine candidate were designed: a vaccine for veterinary use with RVFV strain-35/74 as backbone, and a vaccine for humans with Clone 13 as the backbone [139]. Vaccination with RVFV-4s induced VN antibodies primarily against the Gn proteins and afforded young lambs protection against challenge infection with the wild type virus [140]. The safety of the vaccine was tested by administering RVFV-4s in pregnant ewes at 50 days of gestation. The vaccine induced VN antibodies and did not result in infection in foetuses or ewes [141]. Additionally, the veterinary RVFV-4s vaccine candidate did not disseminate, shed or spread into the environment in a study cohort and did not revert to a virulent strain when lymph node suspensions of immunized lambs were administered to naïve lambs. A single vaccine dose was sufficient to protect lambs, goat kids and calves from virus infection [142]. Both the human and veterinary RVFV-4s vaccine candidates were tested in common marmosets and induced VN-antibody responses [139]. The RVFV-4s vaccine is highly efficacious and safe in all tested species and is a promising candidate for a veterinary and human RVF vaccine.

6.5. DNA Vaccines

Various other vaccine candidates have been developed by modifying virus strains, the deletion of the NSs and NSm genes and producing expression plasmids encoding RVFV viral proteins [143,144]. A DNA vaccine encoding the Gn and Gc glycoproteins (pWRG7077-RVFV-NSm) was highly immunogenic in BALB/c mice. Moreover, the challenge of three-time immunized mice with the ZH501 strain enabled survival, compared to the control mice [145]. A plasmid vector consisting of the Gn gene coupled to the murine complement protein C3d (Gn-C3d) elicited a strong immune response in mice and was effective in protecting mice against ZH501 challenge. The passive immunization of naïve mice with sera from vaccinated mice one hour prior to infection protected these against ZH501 challenge [146].

6.6. Viral Vectored Vaccines

Viral vector systems have also been explored as RVFV vaccine candidates. A recombinant Venezuelan Equine Encephalitis Virus (VEEV) expressing Gn proteins fused to the E2 protein on the cell surface was immunogenic in mice after a single vaccination and afforded protection against challenge infection with the RVFV ZH501 strain [147]. Modified Vaccinia Virus Ankara (MVA) have previously been utilized as vaccine candidates against human immunodeficiency virus, Japanese encephalitis virus, P. falciparum, M. tuberculosis and other pathogens [148,149,150]. rMVA-expressing RVFV glycoproteins or nucleoprotein provided sterile protective immunity in mice against challenge infection with RVFV. Protective immunity was a cumulative effect of RVFV-specific antibody and T cell response [151]. Vaccination with Lumpy skin disease virus (LSDV) recombinantly expressing RVFV glycoprotein induced protective immunity in Merino sheep against RVFV challenge infection [152]. A chimpanzee adenovirus vectored vaccine consisting of Gn and Gc glycoproteins with adjuvants Matrix-M™ and AddaVax™ elicited strong humoral immunity, even at eight weeks post vaccination, and a CD8+ T cell response two weeks post vaccination in mice [153]. The safety and efficacy of the vaccine candidate was established in pregnant ewes, while low level viremia were detected in placenta and plasma of pregnant does post challenge [154]. Recently, a novel method was developed with bacterial outer membrane vesicles, decorated with RVFV Gn head domain. Immunization in mice induced IgG antibodies and antigen-specific T cell responses [155]. Thus, numerous strategies and platforms to develop RVFV vaccines are under investigation to test their safety and efficacy, as summarized in Table 1.

7. Conclusions

Rift Valley Fever is a continuing threat to livestock and human health. Global warming could provide a fertile breeding ground for mosquitoes and interaction between humans and animals might increase the risk of transmission of the virus [158]. Sero-epidemiological evidence indicates that endemic regions are expanding, which underscores the need for close surveillance and vigilance. Considering the losses in agriculture and the public health threat, a one health approach to combat this virus seems warranted and the availability of safe and efficacious vaccines for veterinary and medical use is urgently needed and, therefore, RVFV is on the WHO R&D blueprint list [159].
Despite the fact that RVFV was already isolated in the early 1930s, our understanding of its epidemiology, pathogenesis and immunity is limited. The inter epizootic maintenance of virus and reservoir hosts are not entirely known. A recent study implicated black rats as reservoirs of RVFV because they could be infected persistently for one month without disease manifestations [160]. A RT-PCR analysis of the bat species Pipistrellus deserti in Egypt identified partial sequences of the virus, suggesting its role as a potential reservoir [161]. Further studies focusing on interspecies interaction and virus prevalence could improve our understanding of the RVFV transmission cycle and inter-epidemic virus maintenance. A phylogenetic characterization of virus isolates indicated that distinct lineages are highly similar in amino acid sequences. However, virus replication rates and mortality rates of different strains are highly variable [162]. Insights into the nature of genome variation and the impact on virus pathogenesis would be of interest. The high sequence similarity is also advantageous for the development of targeted therapeutics, as antivirals or vaccines developed against one strain would be effective against other RVFV strains.
RVFV replication and evasion of innate host immune responses is facilitated by the NSs protein, which inhibits interferon response and is an important virulence factor. Induction of the interferon response is crucial for the innate and adaptive immune response, even though the interferon response is unable to control virus replication in susceptible hosts directly. Moreover, the severity of the disease also depends on the genetics of the organism, for instance, different strains of mice respond distinctly to infection, where some are more susceptible to lethal infections [163]. The clinical outcome of infection also depends on the age of host, with foetuses and younger organisms being more susceptible to infection and prone to fatal outcomes, and time of induction of immune response. In the absence of a functional NSs protein, the IFN type I response of the host is efficient in controlling (mutant) virus replication to a certain extent and preventing the development of severe disease. Furthermore, a balance between pro- and anti-inflammatory cytokine levels determines disease severity. During natural infections with wild type RVFV, an efficient pro-inflammatory cytokine response may not be induced, resulting in a more severe disease. Sustained virus-specific antibody- and cell-mediated immune responses afford protection against subsequent infections and should be the target for vaccination strategies.
Various vaccine candidates have been developed to protect susceptible hosts from RVFV infections. Formalin-inactivated vaccines caused side effects such as teratogenicity and neurological complications, while some live attenuated vaccines were under-attenuated or pose a risk for reversion to virulence. Other vaccine candidates like subunit vaccines and virus-like particles are still in the early stages of development.
Newer and more promising vaccine candidates have been designed using segmentation of the viral genome and complete deletion of non-structural NSs and/or NSm genes. These vaccine candidates proved to be highly immunogenic, have an excellent safety profile and are currently under investigation in clinical trials. To obtain a better understanding of the mode of action and correlates of vaccine-induced protection, further research is warranted. The use of highly effective vaccines in livestock and humans in endemic regions may mitigate the impact RVFV has on agriculture and public health as an effective “One health” approach.

Author Contributions

Conceptualization, writing—review and editing, A.D.M.E.O., G.F.R. and C.K.P.; writing—original draft preparation, N.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)—398066876/GRK 2485/2. This publication was supported by the DFG - 491094227 “Open Access Publication Funding” and the University of Veterinary Medicine Hannover, Foundation.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Koch, J.; Xin, Q.; Tischler, N.D.; Lozach, P.-Y. Entry of Phenuiviruses into Mammalian Host Cells. Viruses 2021, 13, 299. [Google Scholar] [CrossRef] [PubMed]
  2. Sun, M.; Ji, Y.; Li, G.; Shao, J.; Chen, R.; Gong, H.; Chen, S.; Chen, J. Highly adaptive Phenuiviridae with biomedical importance in multiple fields. J. Med. Virol. 2022, 94, 2388–2401. [Google Scholar] [CrossRef] [PubMed]
  3. Kwaśnik, M.; Rożek, W.; Rola, J. Rift Valley fever—A growing threat to humans and animals. J. Vet. Res. 2021, 65, 7–14. [Google Scholar] [CrossRef]
  4. Ciota, A.T. The role of co-infection and swarm dynamics in arbovirus transmission. Virus Res. 2019, 265, 88–93. [Google Scholar] [CrossRef]
  5. Kormelink, R.; Verchot, J.; Tao, X.; Desbiez, C. The Bunyavirales: The Plant-Infecting Counterparts. Viruses 2021, 13, 842. [Google Scholar] [CrossRef]
  6. Matsuno, K.; Kajihara, M.; Nakao, R.; Nao, N.; Mori-Kajihara, A.; Muramatsu, M.; Qiu, Y.; Torii, S.; Igarashi, M.; Kasajima, N.; et al. The Unique Phylogenetic Position of a Novel Tick-Borne Phlebovirus Ensures an Ixodid Origin of the Genus Phlebovirus. mSphere 2018, 3, e00239-18. [Google Scholar] [CrossRef] [PubMed]
  7. Daubney, R.; Hudson, J.R. Enzootic Hepatitis or Rift Valley Fever. An Undescribed Virus Disease of Sheep Cattle and Man from East Africa. J. Pathol. Bacteriol. 1931, 34, 545–579. [Google Scholar] [CrossRef]
  8. Schwentker, F.F.; Rivers, T.M. Rift Valley Fever in Man: Report of a Fatal Laboratory Infection Complicated by Thrombophlebitis. J. Exp. Med. 1934, 59, 305–313. [Google Scholar] [CrossRef]
  9. Lubisi, B.A.; Ndouvhada, P.N.; Neiffer, D.; Penrith, M.L.; Sibanda, D.; Bastos, A. Seroprevalence of Rift valley fever in South African domestic and wild suids (1999–2016). Transbound. Emerg. Dis. 2020, 67, 811–821. [Google Scholar] [CrossRef]
  10. Jori, F.; Alexander, K.A.; Mokopasetso, M.; Munstermann, S.; Moagabo, K.; Paweska, J.T. Serological Evidence of Rift Valley Fever Virus Circulation in Domestic Cattle and African Buffalo in Northern Botswana (2010–2011). Front. Vet. Sci. 2015, 2, 63. [Google Scholar] [CrossRef]
  11. Britch, S.C.; Binepal, Y.S.; Ruder, M.G.; Kariithi, H.M.; Linthicum, K.J.; Anyamba, A.; Small, J.L.; Tucker, C.J.; Ateya, L.O.; Oriko, A.A.; et al. Rift Valley Fever Risk Map Model and Seroprevalence in Selected Wild Ungulates and Camels from Kenya. PLoS ONE 2013, 8, e66626. [Google Scholar] [CrossRef] [PubMed]
  12. Atuman, Y.J.; Kudi, C.A.; Abdu, P.A.; Okubanjo, O.O.; Wungak, Y.; Ularamu, H.G.; Abubakar, A. Serological Evidence of Antibodies to Rift Valley Fever Virus in Wild and Domestic Animals in Bauchi State, Nigeria. Vet. Med. Int. 2022, 2022, 6559193. [Google Scholar] [CrossRef] [PubMed]
  13. Gakuya, F.; Akoko, J.; Wambua, L.; Nyamota, R.; Ronoh, B.; Lekolool, I.; Mwatondo, A.; Muturi, M.; Ouma, C.; Nthiwa, D.; et al. Evidence of co-exposure with Brucella spp, Coxiella burnetii, and Rift Valley fever virus among various species of wildlife in Kenya. PLoS Negl. Trop. Dis. 2022, 16, e0010596. [Google Scholar] [CrossRef]
  14. El Mamy, A.B.O.; Baba, M.O.; Barry, Y.; Isselmou, K.; Dia, M.L.; Hampate, B.; Diallo, M.Y.; El Kory, M.O.B.; Diop, M.; Lo, M.M.; et al. Unexpected Rift Valley Fever Outbreak, Northern Mauritania. Emerg. Infect. Dis. 2011, 17, 1894–1896. [Google Scholar] [CrossRef] [PubMed]
  15. Jäckel, S.; Eiden, M.; EL Mamy, B.O.; Isselmou, K.; Vina-Rodriguez, A.; Doumbia, B.; Groschup, M.H. Molecular and Serological Studies on the Rift Valley Fever Outbreak in Mauritania in 2010. Transbound. Emerg. Dis. 2013, 60 (Suppl. S2), 31–39. [Google Scholar] [CrossRef]
  16. Fang, Y.; Khater, E.I.M.; Xue, J.-B.; Ghallab, E.H.S.; Li, Y.-Y.; Jiang, T.-G.; Li, S.-Z. Epidemiology of Mosquito-Borne Viruses in Egypt: A Systematic Review. Viruses 2022, 14, 1577. [Google Scholar] [CrossRef]
  17. Grossi-Soyster, E.N.; Lee, J.; King, C.H.; LaBeaud, A.D. The influence of raw milk exposures on Rift Valley fever virus transmission. PLoS Negl. Trop. Dis. 2019, 13, e0007258. [Google Scholar] [CrossRef]
  18. Msimang, V.; Thompson, P.N.; van Vuren, P.J.; Tempia, S.; Cordel, C.; Kgaladi, J.; Khosa, J.; Burt, F.J.; Liang, J.; Rostal, M.K.; et al. Rift Valley Fever Virus Exposure amongst Farmers, Farm Workers, and Veterinary Professionals in Central South Africa. Viruses 2019, 11, 140. [Google Scholar] [CrossRef]
  19. Oymans, J.; Wichgers Schreur, P.J.; Van Keulen, L.; Kant, J.; Kortekaas, J. Rift Valley fever virus targets the maternal-foetal interface in ovine and human placentas. PLoS Negl. Trop. Dis. 2020, 14, e0007898. [Google Scholar] [CrossRef]
  20. Wright, D.; Kortekaas, J.; Bowden, T.A.; Warimwe, G.M. Rift Valley fever: Biology and epidemiology. J. Gen. Virol. 2019, 100, 1187–1199. [Google Scholar] [CrossRef]
  21. Gerken, K.N.; LaBeaud, A.D.; Mandi, H.; Jackson, M.L.; Breugelmans, J.G.; King, C.H. Paving the way for human vaccination against Rift Valley fever virus: A systematic literature review of RVFV epidemiology from 1999 to 2021. PLoS Negl. Trop. Dis. 2022, 16, e0009852. [Google Scholar] [CrossRef] [PubMed]
  22. Baudin, M.; Jumaa, A.M.; Jomma, H.J.E.; Karsany, M.S.; Bucht, G.; Näslund, J.; Ahlm, C.; Evander, M.; Mohamed, N. Association of Rift Valley fever virus infection with miscarriage in Sudanese women: A cross-sectional study. Lancet Glob. Health 2016, 4, e864–e871. [Google Scholar] [CrossRef] [PubMed]
  23. Mansfield, K.L.; Banyard, A.C.; McElhinney, L.; Johnson, N.; Horton, D.L.; Hernández-Triana, L.M.; Fooks, A.R. Rift Valley fever virus: A review of diagnosis and vaccination, and implications for emergence in Europe. Vaccine 2015, 33, 5520–5531. [Google Scholar] [CrossRef]
  24. Tezcan-Ulger, S.; Kurnaz, N.; Ulger, M.; Aslan, G.; Emekdas, G. Serological evidence of Rift Valley fever virus among humans in Mersin province of Turkey. J. Vector Borne Dis. 2019, 56, 373–379. [Google Scholar] [CrossRef]
  25. Grossi-Soyster, E.N.; LaBeaud, A.D. Rift Valley Fever: Important Considerations for Risk Mitigation and Future Outbreaks. Trop. Med. Infect. Dis. 2020, 5, 89. [Google Scholar] [CrossRef] [PubMed]
  26. Freiberg, A.N.; Sherman, M.B.; Morais, M.C.; Holbrook, M.R.; Watowich, S.J. Three-Dimensional Organization of Rift Valley Fever Virus Revealed by Cryoelectron Tomography. J. Virol. 2008, 82, 10341–10348. [Google Scholar] [CrossRef] [PubMed]
  27. Raymond, D.D.; Piper, M.E.; Gerrard, S.R.; Smith, J.L. Structure of the Rift Valley fever virus nucleocapsid protein reveals another architecture for RNA encapsidation. Proc. Natl. Acad. Sci. USA 2010, 107, 11769–11774. [Google Scholar] [CrossRef]
  28. Ferron, F.; Li, Z.; Danek, E.I.; Luo, D.; Wong, Y.; Coutard, B.; Lantez, V.; Charrel, R.; Canard, B.; Walz, T.; et al. The Hexamer Structure of the Rift Valley Fever Virus Nucleoprotein Suggests a Mechanism for its Assembly into Ribonucleoprotein Complexes. PLoS Pathog. 2011, 7, e1002030. [Google Scholar] [CrossRef]
  29. Mottram, T.J.; Li, P.; Dietrich, I.; Shi, X.; Brennan, B.; Varjak, M.; Kohl, A. Mutational analysis of Rift Valley fever phlebovirus nucleocapsid protein indicates novel conserved, functional amino acids. PLoS Negl. Trop. Dis. 2017, 11, e0006155. [Google Scholar] [CrossRef]
  30. Wang, X.; Hu, C.; Ye, W.; Wang, J.; Dong, X.; Xu, J.; Li, X.; Zhang, M.; Lu, H.; Zhang, F.; et al. Structure of Rift Valley Fever Virus RNA-Dependent RNA Polymerase. J. Virol. 2022, 96, e0171321. [Google Scholar] [CrossRef]
  31. Alamri, M.A.; Mirza, M.U.; Adeel, M.M.; Ashfaq, U.A.; Qamar, M.T.U.; Shahid, F.; Ahmad, S.; Alatawi, E.A.; Albalawi, G.M.; Allemailem, K.S.; et al. Structural Elucidation of Rift Valley Fever Virus L Protein towards the Discovery of Its Potential Inhibitors. Pharmaceuticals 2022, 15, 659. [Google Scholar] [CrossRef] [PubMed]
  32. Borrego, B.; Moreno, S.; López-Valiñas, Á.; de la Losa, N.; Weber, F.; Núñez, J.I.; Brun, A. Identification of Single Amino Acid Changes in the Rift Valley Fever Virus Polymerase Core Domain Contributing to Virus Attenuation In Vivo. Front. Cell. Infect. Microbiol. 2022, 12, 875539. [Google Scholar] [CrossRef] [PubMed]
  33. Kreher, F.; Tamietti, C.; Gommet, C.; Guillemot, L.; Ermonval, M.; Failloux, A.-B.; Panthier, J.-J.; Bouloy, M.; Flamand, M. The Rift Valley fever accessory proteins NSm and P78/NSm-GN are distinct determinants of virus propagation in vertebrate and invertebrate hosts. Emerg. Microbes Infect. 2014, 3, e71. [Google Scholar] [CrossRef]
  34. Wu, Y.; Zhu, Y.; Gao, F.; Jiao, Y.; Oladejo, B.O.; Chai, Y.; Bi, Y.; Lu, S.; Dong, M.; Zhang, C.; et al. Structures of phlebovirus glycoprotein Gn and identification of a neutralizing antibody epitope. Proc. Natl. Acad. Sci. USA 2017, 114, E7564–E7573. [Google Scholar] [CrossRef]
  35. Dessau, M.; Modis, Y. Crystal structure of glycoprotein C from Rift Valley fever virus. Proc. Natl. Acad. Sci. USA 2013, 110, 1696–1701. [Google Scholar] [CrossRef] [PubMed]
  36. Halldorsson, S.; Li, S.; Li, M.; Harlos, K.; Bowden, T.A.; Huiskonen, J.T. Shielding and activation of a viral membrane fusion protein. Nat. Commun. 2018, 9, 349. [Google Scholar] [CrossRef]
  37. Léger, P.; Nachman, E.; Richter, K.; Tamietti, C.; Koch, J.; Burk, R.; Kummer, S.; Xin, Q.; Stanifer, M.; Bouloy, M.; et al. NSs amyloid formation is associated with the virulence of Rift Valley fever virus in mice. Nat. Commun. 2020, 11, 3281. [Google Scholar] [CrossRef]
  38. Bamia, A.; Marcato, V.; Boissière, M.; Mansuroglu, Z.; Tamietti, C.; Romani, M.; Simon, D.; Tian, G.; Niedergang, F.; Panthier, J.-J.; et al. The NSs Protein Encoded by the Virulent Strain of Rift Valley Fever Virus Targets the Expression of Abl2 and the Actin Cytoskeleton of the Host, Affecting Cell Mobility, Cell Shape, and Cell-Cell Adhesion. J. Virol. 2020, 95, 10–1128. [Google Scholar] [CrossRef]
  39. Won, S.; Ikegami, T.; Peters, C.J.; Makino, S. NSm Protein of Rift Valley Fever Virus Suppresses Virus-Induced Apoptosis. J. Virol. 2007, 81, 13335–13345. [Google Scholar] [CrossRef]
  40. Kading, R.C.; Crabtree, M.B.; Bird, B.H.; Nichol, S.T.; Erickson, B.R.; Horiuchi, K.; Biggerstaff, B.J.; Miller, B.R. Deletion of the NSm Virulence Gene of Rift Valley Fever Virus Inhibits Virus Replication in and Dissemination from the Midgut of Aedes aegypti Mosquitoes. PLoS Negl. Trop. Dis. 2014, 8, e2670. [Google Scholar] [CrossRef]
  41. de Boer, S.M.; Kortekaas, J.; de Haan, C.A.M.; Rottier, P.J.M.; Moormann, R.J.M.; Bosch, B.J. Heparan Sulfate Facilitates Rift Valley Fever Virus Entry into the Cell. J. Virol. 2012, 86, 13767–13771. [Google Scholar] [CrossRef] [PubMed]
  42. Hofmann, H.; Li, X.; Zhang, X.; Liu, W.; Kühl, A.; Kaup, F.; Soldan, S.S.; González-Scarano, F.; Weber, F.; He, Y.; et al. Severe Fever with Thrombocytopenia Virus Glycoproteins Are Targeted by Neutralizing Antibodies and Can Use DC-SIGN as a Receptor for pH-Dependent Entry into Human and Animal Cell Lines. J. Virol. 2013, 87, 4384–4394. [Google Scholar] [CrossRef] [PubMed]
  43. Léger, P.; Tetard, M.; Youness, B.; Cordes, N.; Rouxel, R.N.; Flamand, M.; Lozach, P.-Y. Differential Use of the C-Type Lectins L-SIGN and DC-SIGN for Phlebovirus Endocytosis. Traffic 2016, 17, 639–656. [Google Scholar] [CrossRef] [PubMed]
  44. Lozach, P.-Y.; Kühbacher, A.; Meier, R.; Mancini, R.; Bitto, D.; Bouloy, M.; Helenius, A. DC-SIGN as a Receptor for Phleboviruses. Cell Host Microbe 2011, 10, 75–88. [Google Scholar] [CrossRef] [PubMed]
  45. Sun, Y.; Qi, Y.; Liu, C.; Gao, W.; Chen, P.; Fu, L.; Peng, B.; Wang, H.; Jing, Z.; Zhong, G.; et al. Nonmuscle Myosin Heavy Chain IIA Is a Critical Factor Contributing to the Efficiency of Early Infection of Severe Fever with Thrombocytopenia Syndrome Virus. J. Virol. 2014, 88, 237–248. [Google Scholar] [CrossRef] [PubMed]
  46. Ganaie, S.S.; Schwarz, M.M.; McMillen, C.M.; Price, D.A.; Feng, A.X.; Albe, J.R.; Wang, W.; Miersch, S.; Orvedahl, A.; Cole, A.R.; et al. Lrp1 is a host entry factor for Rift Valley fever virus. Cell 2021, 184, 5163–5178.e24. [Google Scholar] [CrossRef]
  47. Harmon, B.; Schudel, B.R.; Maar, D.; Kozina, C.; Ikegami, T.; Tseng, C.-T.K.; Negrete, O.A. Rift Valley Fever Virus Strain MP-12 Enters Mammalian Host Cells via Caveola-Mediated Endocytosis. J. Virol. 2012, 86, 12954–12970. [Google Scholar] [CrossRef]
  48. Filone, C.M.; Hanna, S.L.; Caino, M.C.; Bambina, S.; Doms, R.W.; Cherry, S. Rift Valley Fever Virus Infection of Human Cells and Insect Hosts Is Promoted by Protein Kinase C Epsilon. PLoS ONE 2010, 5, e15483. [Google Scholar] [CrossRef]
  49. de Boer, S.M.; Kortekaas, J.; Spel, L.; Rottier, P.J.M.; Moormann, R.J.M.; Bosch, B.J. Acid-Activated Structural Reorganization of the Rift Valley Fever Virus Gc Fusion Protein. J. Virol. 2012, 86, 13642–13652. [Google Scholar] [CrossRef]
  50. Willensky, S.; Bar-Rogovsky, H.; Bignon, E.A.; Tischler, N.D.; Modis, Y.; Dessau, M. Crystal Structure of Glycoprotein C from a Hantavirus in the Post-fusion Conformation. PLoS Pathog. 2016, 12, e1005948. [Google Scholar] [CrossRef]
  51. Charlton, F.W.; Hover, S.; Fuller, J.; Hewson, R.; Fontana, J.; Barr, J.N.; Mankouri, J. Cellular cholesterol abundance regulates potassium accumulation within endosomes and is an important determinant in bunyavirus entry. J. Biol. Chem. 2019, 294, 7335–7347. [Google Scholar] [CrossRef]
  52. Olschewski, S.; Cusack, S.; Rosenthal, M. The Cap-Snatching Mechanism of Bunyaviruses. Trends Microbiol. 2020, 28, 293–303. [Google Scholar] [CrossRef] [PubMed]
  53. Yan, J.-M.; Zhang, W.-K.; Yan, L.-N.; Jiao, Y.-J.; Zhou, C.-M.; Yu, X.-J. Bunyavirus SFTSV exploits autophagic flux for viral assembly and egress. Autophagy 2022, 18, 1599–1612. [Google Scholar] [CrossRef] [PubMed]
  54. Brahms, A.; Mudhasani, R.; Pinkham, C.; Kota, K.; Nasar, F.; Zamani, R.; Bavari, S.; Kehn-Hall, K. Sorafenib Impedes Rift Valley Fever Virus Egress by Inhibiting Valosin-Containing Protein Function in the Cellular Secretory Pathway. J. Virol. 2017, 91, 10–1128. [Google Scholar] [CrossRef] [PubMed]
  55. Ikegami, T.; Makino, S. The Pathogenesis of Rift Valley Fever. Viruses 2011, 3, 493–519. [Google Scholar] [CrossRef]
  56. Kroeker, A.L.; Smid, V.; Embury-Hyatt, C.; Collignon, B.; Pinette, M.; Babiuk, S.; Pickering, B. Increased Susceptibility of Cattle to Intranasal RVFV Infection. Front. Vet. Sci. 2020, 7, 137. [Google Scholar] [CrossRef] [PubMed]
  57. Odendaal, L.; Davis, A.S.; Venter, E.H. Insights into the Pathogenesis of Viral Haemorrhagic Fever Based on Virus Tropism and Tissue Lesions of Natural Rift Valley Fever. Viruses 2021, 13, 709. [Google Scholar] [CrossRef]
  58. Odendaal, L.; Clift, S.J.; Fosgate, G.T.; Davis, A.S. Lesions and Cellular Tropism of Natural Rift Valley Fever Virus Infection in Adult Sheep. Vet. Pathol. 2019, 56, 61–77. [Google Scholar] [CrossRef]
  59. Odendaal, L.; Fosgate, G.T.; Romito, M.; Coetzer, J.A.W.; Clift, S.J. Sensitivity and specificity of real-time reverse transcription polymerase chain reaction, histopathology, and immunohistochemical labeling for the detection of Rift Valley fever virus in naturally infected cattle and sheep. J. Vet. Diagn. Investig. 2014, 26, 49–60. [Google Scholar] [CrossRef]
  60. Schwarz, M.M.; Connors, K.A.; Davoli, K.A.; McMillen, C.M.; Albe, J.R.; Hoehl, R.M.; Demers, M.J.; Ganaie, S.S.; Price, D.A.; Leung, D.W.; et al. Rift Valley Fever Virus Infects the Posterior Segment of the Eye and Induces Inflammation in a Rat Model of Ocular Disease. J. Virol. 2022, 96, e0111222. [Google Scholar] [CrossRef]
  61. Alrajhi, A.A.; Al-Semari, A.; Al-Watban, J. Rift Valley Fever Encephalitis. Emerg. Infect. Dis. 2004, 10, 554–555. [Google Scholar] [CrossRef] [PubMed]
  62. Gwon, Y.D.; Mahani, S.A.N.; Nagaev, I.; Mincheva-Nilsson, L.; Evander, M. Rift Valley Fever Virus Propagates in Human Villous Trophoblast Cell Lines and Induces Cytokine mRNA Responses Known to Provoke Miscarriage. Viruses 2021, 13, 2265. [Google Scholar] [CrossRef]
  63. Lester, S.N.; Li, K. Toll-Like Receptors in Antiviral Innate Immunity. J. Mol. Biol. 2014, 426, 1246–1264. [Google Scholar] [CrossRef]
  64. Thompson, M.R.; Kaminski, J.J.; Kurt-Jones, E.A.; Fitzgerald, K.A. Pattern Recognition Receptors and the Innate Immune Response to Viral Infection. Viruses 2011, 3, 920–940. [Google Scholar] [CrossRef]
  65. Tolomeo, M.; Cavalli, A.; Cascio, A. STAT1 and Its Crucial Role in the Control of Viral Infections. Int. J. Mol. Sci. 2022, 23, 4095. [Google Scholar] [CrossRef]
  66. Mudhasani, R.; Tran, J.P.; Retterer, C.; Radoshitzky, S.R.; Kota, K.P.; Altamura, L.A.; Smith, J.M.; Packard, B.Z.; Kuhn, J.H.; Costantino, J.; et al. IFITM-2 and IFITM-3 but Not IFITM-1 Restrict Rift Valley Fever Virus. J. Virol. 2013, 87, 8451–8464. [Google Scholar] [CrossRef] [PubMed]
  67. Confort, M.P.; Duboeuf, M.; Thiesson, A.; Pons, L.; Marziali, F.; Desloire, S.; Ratinier, M.; Cimarelli, A.; Arnaud, F. IFITMs from Naturally Infected Animal Species Exhibit Distinct Restriction Capacities against Toscana and Rift Valley Fever Viruses. Viruses 2023, 15, 306. [Google Scholar] [CrossRef] [PubMed]
  68. Habjan, M.; Andersson, I.; Klingström, J.; Schümann, M.; Martin, A.; Zimmermann, P.; Wagner, V.; Pichlmair, A.; Schneider, U.; Mühlberger, E.; et al. Processing of Genome 5′ Termini as a Strategy of Negative-Strand RNA Viruses to Avoid RIG-I-Dependent Interferon Induction. PLoS ONE 2008, 3, e2032. [Google Scholar] [CrossRef]
  69. Bisom, T.C.; White, L.A.; Lanchy, J.-M.; Lodmell, J.S. RIOK3 and Its Alternatively Spliced Isoform Have Disparate Roles in the Innate Immune Response to Rift Valley Fever Virus (MP12) Infection. Viruses 2022, 14, 2064. [Google Scholar] [CrossRef]
  70. White, L.A.; Bisom, T.C.; Grimes, H.L.; Hayashi, M.; Lanchy, J.-M.; Lodmell, J.S. Tra2beta-Dependent Regulation of RIO Kinase 3 Splicing During Rift Valley Fever Virus Infection Underscores the Links Between Alternative Splicing and Innate Antiviral Immunity. Front. Cell. Infect. Microbiol. 2022, 11, 799024. [Google Scholar] [CrossRef]
  71. Ermler, M.E.; Traylor, Z.; Patel, K.; Schattgen, S.A.; Vanaja, S.K.; Fitzgerald, K.A.; Hise, A.G. Rift Valley fever virus infection induces activation of the NLRP3 inflammasome. Virology 2014, 449, 174–180. [Google Scholar] [CrossRef] [PubMed]
  72. Narayanan, A.; Amaya, M.; Voss, K.; Chung, M.; Benedict, A.; Sampey, G.; Kehn-Hall, K.; Luchini, A.; Liotta, L.; Bailey, C.; et al. Reactive oxygen species activate NFκB (p65) and p53 and induce apoptosis in RVFV infected liver cells. Virology 2014, 449, 270–286. [Google Scholar] [CrossRef] [PubMed]
  73. Baer, A.; Austin, D.; Narayanan, A.; Popova, T.; Kainulainen, M.; Bailey, C.; Kashanchi, F.; Weber, F.; Kehn-Hall, K. Induction of DNA Damage Signaling upon Rift Valley Fever Virus Infection Results in Cell Cycle Arrest and Increased Viral Replication. J. Biol. Chem. 2012, 287, 7399–7410. [Google Scholar] [CrossRef] [PubMed]
  74. Austin, D.; Baer, A.; Lundberg, L.; Shafagati, N.; Schoonmaker, A.; Narayanan, A.; Popova, T.; Panthier, J.J.; Kashanchi, F.; Bailey, C.; et al. p53 Activation following Rift Valley Fever Virus Infection Contributes to Cell Death and Viral Production. PLoS ONE 2012, 7, e36327. [Google Scholar] [CrossRef] [PubMed]
  75. Luo, Y.; Kleiboeker, S.; Deng, X.; Qiu, J. Human Parvovirus B19 Infection Causes Cell Cycle Arrest of Human Erythroid Progenitors at Late S Phase That Favors Viral DNA Replication. J. Virol. 2013, 87, 12766–12775. [Google Scholar] [CrossRef]
  76. Kudoh, A.; Daikoku, T.; Sugaya, Y.; Isomura, H.; Fujita, M.; Kiyono, T.; Nishiyama, Y.; Tsurumi, T. Inhibition of S-Phase Cyclin-Dependent Kinase Activity Blocks Expression of Epstein-Barr Virus Immediate-Early and Early Genes, Preventing Viral Lytic Replication. J. Virol. 2004, 78, 104–115. [Google Scholar] [CrossRef]
  77. Su, M.; Shi, D.; Xing, X.; Qi, S.; Yang, D.; Zhang, J.; Han, Y.; Zhu, Q.; Sun, H.; Wang, X.; et al. Coronavirus Porcine Epidemic Diarrhea Virus Nucleocapsid Protein Interacts with p53 To Induce Cell Cycle Arrest in S-Phase and Promotes Viral Replication. J. Virol. 2021, 95, e0018721. [Google Scholar] [CrossRef]
  78. Liu, S.; Liu, H.; Kang, J.; Xu, L.; Zhang, K.; Li, X.; Hou, W.; Wang, Z.; Wang, T. The Severe Fever with Thrombocytopenia Syndrome Virus NSs Protein Interacts with CDK1 To Induce G2 Cell Cycle Arrest and Positively Regulate Viral Replication. J. Virol. 2020, 94, 10–1128. [Google Scholar] [CrossRef]
  79. Ning, Y.-J.; Mo, Q.; Feng, K.; Min, Y.-Q.; Li, M.; Hou, D.; Peng, C.; Zheng, X.; Deng, F.; Hu, Z.; et al. Interferon-γ-Directed Inhibition of a Novel High-Pathogenic Phlebovirus and Viral Antagonism of the Antiviral Signaling by Targeting STAT1. Front. Immunol. 2019, 10, 1182. [Google Scholar] [CrossRef]
  80. Moalem, Y.; Malis, Y.; Voloshin, K.; Dukhovny, A.; Hirschberg, K.; Sklan, E.H. Sandfly Fever Viruses Attenuate the Type I Interferon Response by Targeting the Phosphorylation of JAK-STAT Components. Front. Immunol. 2022, 13, 865797. [Google Scholar] [CrossRef]
  81. Le May, N.; Mansuroglu, Z.; Léger, P.; Josse, T.; Blot, G.; Billecocq, A.; Flick, R.; Jacob, Y.; Bonnefoy, E.; Bouloy, M. A SAP30 Complex Inhibits IFN-β Expression in Rift Valley Fever Virus Infected Cells. PLoS Pathog. 2008, 4, e13. [Google Scholar] [CrossRef] [PubMed]
  82. Ikegami, T.; Narayanan, K.; Won, S.; Kamitani, W.; Peters, C.J.; Makino, S. Rift Valley Fever Virus NSs Protein Promotes Post-Transcriptional Downregulation of Protein Kinase PKR and Inhibits eIF2α Phosphorylation. PLoS Pathog. 2009, 5, e1000287. [Google Scholar] [CrossRef] [PubMed]
  83. Moy, R.H.; Gold, B.; Molleston, J.M.; Schad, V.; Yanger, K.; Salzano, M.-V.; Yagi, Y.; Fitzgerald, K.A.; Stanger, B.Z.; Soldan, S.S.; et al. Antiviral Autophagy Restricts Rift Valley Fever Virus Infection and Is Conserved from Flies to Mammals. Immunity 2014, 40, 51–65. [Google Scholar] [CrossRef] [PubMed]
  84. Zhu, X.; Guan, Z.; Fang, Y.; Zhang, Y.; Guan, Z.; Li, S.; Peng, K. Rift Valley Fever Virus Nucleoprotein Triggers Autophagy to Dampen Antiviral Innate Immune Responses. J. Virol. 2023, 97, e0181422. [Google Scholar] [CrossRef]
  85. Kalluri, R.; LeBleu, V.S. The biology, function, and biomedical applications of exosomes. Science 2020, 367, eaau6977. [Google Scholar] [CrossRef] [PubMed]
  86. Alem, F.; Olanrewaju, A.A.; Omole, S.; Hobbs, H.E.; Ahsan, N.; Matulis, G.; Brantner, C.A.; Zhou, W.; Petricoin, E.F.; Liotta, L.A.; et al. Exosomes originating from infection with the cytoplasmic single-stranded RNA virus Rift Valley fever virus (RVFV) protect recipient cells by inducing RIG-I mediated IFN-B response that leads to activation of autophagy. Cell Biosci. 2021, 11, 220. [Google Scholar] [CrossRef]
  87. Nfon, C.K.; Marszal, P.; Zhang, S.; Weingartl, H.M. Innate Immune Response to Rift Valley Fever Virus in Goats. PLoS Negl. Trop. Dis. 2012, 6, e1623. [Google Scholar] [CrossRef]
  88. McElroy, A.K.; Nichol, S.T. Rift Valley fever virus inhibits a pro-inflammatory response in experimentally infected human monocyte derived macrophages and a pro-inflammatory cytokine response may be associated with patient survival during natural infection. Virology 2011, 422, 6–12. [Google Scholar] [CrossRef]
  89. Roberts, K.K.; Holbrook, M.R.; Hill, T.E.; Freiberg, A.N.; Davis, M.N. Cytokine response in mouse bone marrow derived macrophages after infection with pathogenic and non-pathogenic Rift Valley fever virus. J. Gen. Virol. 2015, 96, 1651–1663. [Google Scholar] [CrossRef]
  90. van Vuren, P.J.; Shalekoff, S.; Grobbelaar, A.A.; Archer, B.N.; Thomas, J.; Tiemessen, C.T.; Paweska, J.T. Serum levels of inflammatory cytokines in Rift Valley fever patients are indicative of severe disease. Virol. J. 2015, 12, 159. [Google Scholar] [CrossRef]
  91. Lathan, R.; Simon-Chazottes, D.; Jouvion, G.; Godon, O.; Malissen, M.; Flamand, M.; Bruhns, P.; Panthier, J.-J. Innate Immune Basis for Rift Valley Fever Susceptibility in Mouse Models. Sci. Rep. 2017, 7, 7096. [Google Scholar] [CrossRef] [PubMed]
  92. Valle, T.Z.D.; Billecocq, A.; Guillemot, L.; Alberts, R.; Gommet, C.; Geffers, R.; Calabrese, K.; Schughart, K.; Bouloy, M.; Montagutelli, X.; et al. A New Mouse Model Reveals a Critical Role for Host Innate Immunity in Resistance to Rift Valley Fever. J. Immunol. 2010, 185, 6146–6156. [Google Scholar] [CrossRef] [PubMed]
  93. Cartwright, H.N.; Barbeau, D.J.; Doyle, J.D.; Klein, E.; Heise, M.T.; Ferris, M.T.; McElroy, A.K. Genetic diversity of collaborative cross mice enables identification of novel rift valley fever virus encephalitis model. PLoS Pathog. 2022, 18, e1010649. [Google Scholar] [CrossRef] [PubMed]
  94. Hartman, A.L.; Powell, D.S.; Bethel, L.M.; Caroline, A.L.; Schmid, R.J.; Oury, T.; Reed, D.S. Aerosolized Rift Valley Fever Virus Causes Fatal Encephalitis in African Green Monkeys and Common Marmosets. J. Virol. 2014, 88, 2235–2245. [Google Scholar] [CrossRef] [PubMed]
  95. Wonderlich, E.R.; Caroline, A.L.; McMillen, C.M.; Walters, A.W.; Reed, D.S.; Barratt-Boyes, S.M.; Hartman, A.L. Peripheral Blood Biomarkers of Disease Outcome in a Monkey Model of Rift Valley Fever Encephalitis. J. Virol. 2018, 92, 10–1128. [Google Scholar] [CrossRef] [PubMed]
  96. Tshilenge, G.M.; Dundon, W.G.; De Nardi, M.; Mfumu, L.K.M.; Rweyemamu, M.; Kayembe-Ntumba, J.-M.; Masumu, J. Seroprevalence of Rift Valley fever virus in cattle in the Democratic Republic of the Congo. Trop. Anim. Health Prod. 2019, 51, 537–543. [Google Scholar] [CrossRef]
  97. de Glanville, W.A.; Nyarobi, J.M.; Kibona, T.; Halliday, J.E.B.; Thomas, K.M.; Allan, K.J.; Johnson, P.C.D.; Davis, A.; Lankester, F.; Claxton, J.R.; et al. Inter-epidemic Rift Valley fever virus infection incidence and risks for zoonotic spillover in northern Tanzania. PLoS Negl. Trop. Dis. 2022, 16, e0010871. [Google Scholar] [CrossRef]
  98. Salekwa, L.P.; Wambura, P.N.; Matiko, M.K.; Watts, D.M. Circulation of Rift Valley Fever Virus Antibody in Cattle during Inter-Epizootic/Epidemic Periods in Selected Regions of Tanzania. Am. J. Trop. Med. Hyg. 2019, 101, 459–466. [Google Scholar] [CrossRef]
  99. Wright, D.; Allen, E.R.; Clark, M.H.; Gitonga, J.N.; Karanja, H.K.; Hulswit, R.J.; Taylor, I.; Biswas, S.; Marshall, J.; Mwololo, D.; et al. Naturally Acquired Rift Valley Fever Virus Neutralizing Antibodies Predominantly Target the Gn Glycoprotein. iScience 2020, 23, 101669. [Google Scholar] [CrossRef]
  100. Busquets, N.; Xavier, F.; Martín-Folgar, R.; Lorenzo, G.; Galindo-Cardiel, I.; del Val, B.P.; Rivas, R.; Iglesias, J.; Rodríguez, F.; Solanes, D.; et al. Experimental Infection of Young Adult European Breed Sheep with Rift Valley Fever Virus Field Isolates. Vector-Borne Zoonotic Dis. 2010, 10, 689–696. [Google Scholar] [CrossRef]
  101. Chengula, A.A.; Kasanga, C.J.; Mdegela, R.H.; Sallu, R.; Yongolo, M. Molecular detection of Rift Valley fever virus in serum samples from selected areas of Tanzania. Trop. Anim. Health Prod. 2014, 46, 629–634. [Google Scholar] [CrossRef] [PubMed]
  102. Memish, Z.A.; Masri, M.A.; Anderson, B.D.; Heil, G.L.; Merrill, H.R.; Khan, S.U.; Alsahly, A.; Gray, G.C. Elevated Antibodies Against Rift Valley Fever Virus Among Humans with Exposure to Ruminants in Saudi Arabia. Am. J. Trop. Med. Hyg. 2015, 92, 739–743. [Google Scholar] [CrossRef]
  103. Swai, E.S.; Sindato, C. Seroprevalence of Rift Valley fever virus infection in camels (dromedaries) in northern Tanzania. Trop. Anim. Health Prod. 2015, 47, 347–352. [Google Scholar] [CrossRef] [PubMed]
  104. Maganga, G.D.; Ndong, A.L.A.; Okouyi, C.S.M.; Mandanda, S.M.; N’Dilimabaka, N.; Pinto, A.; Agossou, E.; Cossic, B.; Akue, J.-P.; Leroy, E.M. Serological Evidence for the Circulation of Rift Valley Fever Virus in Domestic Small Ruminants in Southern Gabon. Vector-Borne Zoonotic Dis. 2017, 17, 443–446. [Google Scholar] [CrossRef]
  105. Schmaljohn, C.S.; Parker, M.D.; Ennis, W.H.; Dalrymple, J.M.; Collett, M.S.; Slizich, J.A.; Schmaljohn, A.L. Baculovirus expression of the M genome segment of Rift Valley fever virus and examination of antigenic and immunogenic properties of the expressed proteins. Virology 1989, 170, 184–192. [Google Scholar] [CrossRef]
  106. Allen, E.R.; Krumm, S.A.; Raghwani, J.; Halldorsson, S.; Elliott, A.; Graham, V.A.; Koudriakova, E.; Harlos, K.; Wright, D.; Warimwe, G.M.; et al. A Protective Monoclonal Antibody Targets a Site of Vulnerability on the Surface of Rift Valley Fever Virus. Cell Rep. 2018, 25, 3750–3758.e4. [Google Scholar] [CrossRef]
  107. Chrun, T.; Lacôte, S.; Urien, C.; Jouneau, L.; Barc, C.; Bouguyon, E.; Contreras, V.; Ferrier-Rembert, A.; Peyrefitte, C.N.; Busquets, N.; et al. A Rift Valley fever virus Gn ectodomain-based DNA vaccine induces a partial protection not improved by APC targeting. npj Vaccines 2018, 3, 14. [Google Scholar] [CrossRef]
  108. Said, A.; Elmanzalawy, M.; Ma, G.; Damiani, A.M.; Osterrieder, N. An equine herpesvirus type 1 (EHV-1) vector expressing Rift Valley fever virus (RVFV) Gn and Gc induces neutralizing antibodies in sheep. Virol. J. 2017, 14, 154. [Google Scholar] [CrossRef]
  109. van Vuren, P.J.; Tiemessen, C.T.; Paweska, J.T. Anti-Nucleocapsid Protein Immune Responses Counteract Pathogenic Effects of Rift Valley Fever Virus Infection in Mice. PLoS ONE 2011, 6, e25027. [Google Scholar] [CrossRef]
  110. Dodd, K.A.; McElroy, A.K.; Jones, M.E.B.; Nichol, S.T.; Spiropoulou, C.F. Rift Valley Fever Virus Clearance and Protection from Neurologic Disease Are Dependent on CD4+ T Cell and Virus-Specific Antibody Responses. J. Virol. 2013, 87, 6161–6171. [Google Scholar] [CrossRef]
  111. Harmon, J.R.; Spengler, J.R.; Coleman-McCray, J.D.; Nichol, S.T.; Spiropoulou, C.F.; McElroy, A.K. CD4 T Cells, CD8 T Cells, and Monocytes Coordinate to Prevent Rift Valley Fever Virus Encephalitis. J. Virol. 2018, 92, 10–1128. [Google Scholar] [CrossRef] [PubMed]
  112. Michaely, L.M.; Rissmann, M.; Keller, M.; König, R.; von Arnim, F.; Eiden, M.; Rohn, K.; Baumgärtner, W.; Groschup, M.; Ulrich, R. NSG-Mice Reveal the Importance of a Functional Innate and Adaptive Immune Response to Overcome RVFV Infection. Viruses 2022, 14, 350. [Google Scholar] [CrossRef] [PubMed]
  113. Smithburn, K.C.; Haddow, A.J.; Gillett, J.D. Rift Valley fever; isolation of the virus from wild mosquitoes. Br. J. Exp. Pathol. 1948, 29, 107–121. [Google Scholar] [PubMed]
  114. Smithburn, K.C. Rift Valley fever; the neurotropic adaptation of the virus and the experimental use of this modified virus as a vaccine. Br. J. Exp. Pathol. 1949, 30, 1–16. [Google Scholar]
  115. Anthony, T.; van Schalkwyk, A.; Romito, M.; Odendaal, L.; Clift, S.J.; Davis, A.S. Vaccination with Rift Valley fever virus live attenuated vaccine strain Smithburn caused meningoencephalitis in alpacas. J. Vet. Diagn. Investig. 2021, 33, 777–781. [Google Scholar] [CrossRef]
  116. Kark, J.D.; Aynor, Y.; Peters, C.J. A rift Valley fever vaccine trial: I. Side effects and serologic response over a six-month follow-up. Am. J. Epidemiol. 1982, 116, 808–820. [Google Scholar] [CrossRef]
  117. Niklasson, B.; Peters, C.; Bengtsson, E.; Norrby, E. Rift Valley fever virus vaccine trial: Study of neutralizing antibody response in humans. Vaccine 1985, 3, 123–127. [Google Scholar] [CrossRef]
  118. Pittman, P.R.; Liu, C.; Cannon, T.L.; Makuch, R.S.; Mangiafico, J.A.; Gibbs, P.H.; Peters, C.J. Immunogenicity of an inactivated Rift Valley fever vaccine in humans: A 12-year experience. Vaccine 1999, 18, 181–189. [Google Scholar] [CrossRef]
  119. Rusnak, J.M.; Gibbs, P.; Boudreau, E.; Clizbe, D.P.; Pittman, P. Immunogenicity and safety of an inactivated Rift Valley fever vaccine in a 19-year study. Vaccine 2011, 29, 3222–3229. [Google Scholar] [CrossRef]
  120. Harmon, J.R.; Barbeau, D.J.; Nichol, S.T.; Spiropoulou, C.F.; McElroy, A.K. Rift Valley fever virus vaccination induces long-lived, antigen-specific human T cell responses. npj Vaccines 2020, 5, 17. [Google Scholar] [CrossRef]
  121. Caplen, H.; Peters, C.J.; Bishop, D.H. Mutagen-directed Attenuation of Rift Valley Fever Virus as a Method for Vaccine Development. J. Gen. Virol. 1985, 66 Pt 10, 2271–2277. [Google Scholar] [CrossRef]
  122. Vialat, P.; Muller, R.; Vu, T.H.; Prehaud, C.; Bouloy, M. Mapping of the mutations present in the genome of the Rift Valley fever virus attenuated MP12 strain and their putative role in attenuation. Virus Res. 1997, 52, 43–50. [Google Scholar] [CrossRef]
  123. Morrill, J.; Carpenter, L.; Taylor, D.; Ramsburg, H.; Quance, J.; Peters, C. Further evaluation of a mutagen-attenuated Rift Valley fever vaccine in sheep. Vaccine 1991, 9, 35–41. [Google Scholar] [CrossRef]
  124. Miller, M.M.; Bennett, K.E.; Drolet, B.S.; Lindsay, R.; Mecham, J.O.; Reeves, W.K.; Weingartl, H.M.; Wilson, W.C. Evaluation of the Efficacy, Potential for Vector Transmission, and Duration of Immunity of MP-12, an Attenuated Rift Valley Fever Virus Vaccine Candidate, in Sheep. Clin. Vaccine Immunol. 2015, 22, 930–937. [Google Scholar] [CrossRef]
  125. Rissmann, M.; Ulrich, R.; Schröder, C.; Hammerschmidt, B.; Hanke, D.; Mroz, C.; Groschup, M.; Eiden, M. Vaccination of alpacas against Rift Valley fever virus: Safety, immunogenicity and pathogenicity of MP-12 vaccine. Vaccine 2017, 35, 655–662. [Google Scholar] [CrossRef]
  126. Morrill, J.C.; Peters, C.J. Protection of MP-12-Vaccinated Rhesus Macaques Against Parenteral and Aerosol Challenge with Virulent Rift Valley Fever Virus. J. Infect. Dis. 2011, 204, 229–236. [Google Scholar] [CrossRef]
  127. Ikegami, T. Rift Valley fever vaccines: An overview of the safety and efficacy of the live-attenuated MP-12 vaccine candidate. Expert Rev. Vaccines 2017, 16, 601–611. [Google Scholar] [CrossRef]
  128. Wilson, W.C.; Bawa, B.; Drolet, B.S.; Lehiy, C.; Faburay, B.; Jasperson, D.C.; Reister, L.; Gaudreault, N.N.; Carlson, J.; Ma, W.; et al. Evaluation of lamb and calf responses to Rift Valley fever MP-12 vaccination. Vet. Microbiol. 2014, 172, 44–50. [Google Scholar] [CrossRef]
  129. Morrill, J.C.; Peters, C.J. Mucosal Immunization of Rhesus Macaques with Rift Valley Fever MP-12 Vaccine. J. Infect. Dis. 2011, 204, 617–625. [Google Scholar] [CrossRef]
  130. Pittman, P.R.; McClain, D.; Quinn, X.; Coonan, K.M.; Mangiafico, J.; Makuch, R.S.; Morrill, J.; Peters, C.J. Safety and immunogenicity of a mutagenized, live attenuated Rift Valley fever vaccine, MP-12, in a Phase 1 dose escalation and route comparison study in humans. Vaccine 2016, 34, 424–429. [Google Scholar] [CrossRef]
  131. Pittman, P.R.; Norris, S.L.; Brown, E.S.; Ranadive, M.V.; Schibly, B.A.; Bettinger, G.E.; Lokugamage, N.; Korman, L.; Morrill, J.C.; Peters, C.J. Rift Valley fever MP-12 vaccine Phase 2 clinical trial: Safety, immunogenicity, and genetic characterization of virus isolates. Vaccine 2016, 34, 523–530. [Google Scholar] [CrossRef] [PubMed]
  132. Muller, R.; Saluzzo, J.-F.; Lopez, N.; Dreier, T.; Turell, M.; Smith, J.; Bouloy, M. Characterization of clone 13, a naturally attenuated avirulent isolate of Rift Valley fever virus, which is altered in the small segment. Am. Soc. Trop. Med. Hyg. 1995, 53, 405–411. [Google Scholar] [CrossRef] [PubMed]
  133. Dungu, B.; Louw, I.; Lubisi, A.; Hunter, P.; von Teichman, B.F.; Bouloy, M. Evaluation of the efficacy and safety of the Rift Valley Fever Clone 13 vaccine in sheep. Vaccine 2010, 28, 4581–4587. [Google Scholar] [CrossRef]
  134. von Teichman, B.; Engelbrecht, A.; Zulu, G.; Dungu, B.; Pardini, A.; Bouloy, M. Safety and efficacy of Rift Valley fever Smithburn and Clone 13 vaccines in calves. Vaccine 2011, 29, 5771–5777. [Google Scholar] [CrossRef] [PubMed]
  135. Sindato, C.; Karimuribo, E.D.; Swai, E.S.; Mboera, L.E.G.; Rweyemamu, M.M.; Paweska, J.T.; Salt, J. Safety, Immunogenicity and Antibody Persistence of Rift Valley Fever Virus Clone 13 Vaccine in Sheep, Goats and Cattle in Tanzania. Front. Vet. Sci. 2021, 8, 779858. [Google Scholar] [CrossRef]
  136. Lo, M.M.; Mbao, V.; Sierra, P.; Thiongane, Y.; Diop, M.; Donadeu, M.; Dungu, B. Safety and immunogenicity of Onderstepoort Biological Products’ Rift Valley fever Clone 13 vaccine in sheep and goats under field conditions in Senegal. Onderstepoort J. Vet. Res. 2015, 82, 857. [Google Scholar] [CrossRef]
  137. Makoschey, B.; van Kilsdonk, E.; Hubers, W.R.; Vrijenhoek, M.P.; Smit, M.; Schreur, P.J.W.; Kortekaas, J.; Moulin, V. Rift Valley Fever Vaccine Virus Clone 13 Is Able to Cross the Ovine Placental Barrier Associated with Foetal Infections, Malformations, and Stillbirths. PLoS Negl. Trop. Dis. 2016, 10, e0004550. [Google Scholar] [CrossRef]
  138. Schreur, P.J.W.; Oreshkova, N.; Moormann, R.J.M.; Kortekaas, J. Creation of Rift Valley Fever Viruses with Four-Segmented Genomes Reveals Flexibility in Bunyavirus Genome Packaging. J. Virol. 2014, 88, 10883–10893. [Google Scholar] [CrossRef]
  139. Schreur, P.J.W.; Mooij, P.; Koopman, G.; Verstrepen, B.E.; Fagrouch, Z.; Mortier, D.; van Driel, N.; Kant, J.; van de Water, S.; Bogers, W.M.; et al. Safety and immunogenicity of four-segmented Rift Valley fever virus in the common marmoset. npj Vaccines 2022, 7, 54. [Google Scholar] [CrossRef]
  140. Schreur, P.J.W.; Kant, J.; van Keulen, L.; Moormann, R.J.; Kortekaas, J. Four-segmented Rift Valley fever virus induces sterile immunity in sheep after a single vaccination. Vaccine 2015, 33, 1459–1464. [Google Scholar] [CrossRef]
  141. Schreur, P.J.W.; van Keulen, L.; Kant, J.; Kortekaas, J. Four-segmented Rift Valley fever virus-based vaccines can be applied safely in ewes during pregnancy. Vaccine 2017, 35, 3123–3128. [Google Scholar] [CrossRef] [PubMed]
  142. Schreur, P.J.W.; Oreshkova, N.; van Keulen, L.; Kant, J.; van de Water, S.; Soós, P.; Dehon, Y.; Kollár, A.; Pénzes, Z.; Kortekaas, J. Safety and efficacy of four-segmented Rift Valley fever virus in young sheep, goats and cattle. NPJ Vaccines 2020, 5, 65. [Google Scholar] [CrossRef] [PubMed]
  143. Weingartl, H.M.; Nfon, C.K.; Zhang, S.; Marszal, P.; Wilson, W.C.; Morrill, J.; Bettinger, G.E.; Peters, C.J. Efficacy of a recombinant Rift Valley fever virus MP-12 with NSm deletion as a vaccine candidate in sheep. Vaccine 2014, 32, 2345–2349. [Google Scholar] [CrossRef] [PubMed]
  144. Bird, B.H.; Albariño, C.G.; Hartman, A.L.; Erickson, B.R.; Ksiazek, T.G.; Nichol, S.T. Rift Valley Fever Virus Lacking the NSs and NSm Genes Is Highly Attenuated, Confers Protective Immunity from Virulent Virus Challenge, and Allows for Differential Identification of Infected and Vaccinated Animals. J. Virol. 2008, 82, 2681–2691. [Google Scholar] [CrossRef] [PubMed]
  145. Spik, K.; Shurtleff, A.; McElroy, A.K.; Guttieri, M.C.; Hooper, J.W.; Schmaljohn, C. Immunogenicity of combination DNA vaccines for Rift Valley fever virus, tick-borne encephalitis virus, Hantaan virus, and Crimean Congo hemorrhagic fever virus. Vaccine 2006, 24, 4657–4666. [Google Scholar] [CrossRef]
  146. Bhardwaj, N.; Heise, M.T.; Ross, T.M. Vaccination with DNA Plasmids Expressing Gn Coupled to C3d or Alphavirus Replicons Expressing Gn Protects Mice against Rift Valley Fever Virus. PLoS Negl. Trop. Dis. 2010, 4, e725. [Google Scholar] [CrossRef]
  147. Gorchakov, R.; Volkova, E.; Yun, N.; Petrakova, O.; Linde, N.S.; Paessler, S.; Frolova, E.; Frolov, I. Comparative analysis of the alphavirus-based vectors expressing Rift Valley fever virus glycoproteins. Virology 2007, 366, 212–225. [Google Scholar] [CrossRef]
  148. Goonetilleke, N.P.; McShane, H.; Hannan, C.M.; Anderson, R.J.; Brookes, R.H.; Hill, A.V.S. Enhanced Immunogenicity and Protective Efficacy Against Mycobacterium tuberculosis of Bacille Calmette-Guérin Vaccine Using Mucosal Administration and Boosting with a Recombinant Modified Vaccinia Virus Ankara. J. Immunol. 2003, 171, 1602–1609. [Google Scholar] [CrossRef]
  149. Nam, J.-H.; Chae, S.-L.; Cho, H.-W. Immunogenicity of a Recombinant MVA and a DNA Vaccine for Japanese Encephalitis Virus in Swine. Microbiol. Immunol. 2002, 46, 23–28. [Google Scholar] [CrossRef]
  150. Amara, R.R.; Villinger, F.; Altman, J.D.; Lydy, S.L.; O’neil, S.P.; Staprans, S.I.; Montefiori, D.C.; Xu, Y.; Herndon, J.G.; Wyatt, L.S.; et al. Control of a mucosal challenge and prevention of AIDS by a multiprotein DNA/MVA vaccine. Vaccine 2002, 20, 1949–1955. [Google Scholar] [CrossRef]
  151. López-Gil, E.; Lorenzo, G.; Hevia, E.; Borrego, B.; Eiden, M.; Groschup, M.; Gilbert, S.C.; Brun, A. A Single Immunization with MVA Expressing GnGc Glycoproteins Promotes Epitope-specific CD8+-T Cell Activation and Protects Immune-competent Mice against a Lethal RVFV Infection. PLoS Negl. Trop. Dis. 2013, 7, e2309. [Google Scholar] [CrossRef] [PubMed]
  152. Wallace, D.; Ellis, C.; Espach, A.; Smith, S.; Greyling, R.; Viljoen, G. Protective immune responses induced by different recombinant vaccine regimes to Rift Valley fever. Vaccine 2006, 24, 7181–7189. [Google Scholar] [CrossRef] [PubMed]
  153. Warimwe, G.M.; Lorenzo, G.; Lopez-Gil, E.; Reyes-Sandoval, A.; Cottingham, M.G.; Spencer, A.J.; Collins, K.A.; Dicks, M.D.; Milicic, A.; Lall, A.; et al. Immunogenicity and efficacy of a chimpanzee adenovirus-vectored Rift Valley Fever vaccine in mice. Virol. J. 2013, 10, 349. [Google Scholar] [CrossRef]
  154. Stedman, A.; Wright, D.; Schreur, P.J.W.; Clark, M.H.A.; Hill, A.V.S.; Gilbert, S.C.; Francis, M.J.; van Keulen, L.; Kortekaas, J.; Charleston, B.; et al. Safety and efficacy of ChAdOx1 RVF vaccine against Rift Valley fever in pregnant sheep and goats. npj Vaccines 2019, 4, 44. [Google Scholar] [CrossRef]
  155. Zhang, S.; Yan, F.; Liu, D.; Li, E.; Feng, N.; Xu, S.; Wang, H.; Gao, Y.; Yang, S.; Zhao, Y.; et al. Bacterium-Like Particles Displaying the Rift Valley Fever Virus Gn Head Protein Induces Efficacious Immune Responses in Immunized Mice. Front. Microbiol. 2022, 13, 799942. [Google Scholar] [CrossRef] [PubMed]
  156. Botros, B.; Omar, A.; Elian, K.; Mohamed, G.; Soliman, A.; Salib, A.; Salman, D.; Saad, M.; Earhart, K. Adverse response of non-indigenous cattle of European breeds to live attenuated Smithburn Rift Valley fever vaccine. J. Med. Virol. 2006, 78, 787–791. [Google Scholar] [CrossRef]
  157. Kamal, S. Pathological studies on postvaccinal reactions of Rift Valley fever in goats. Virol. J. 2009, 6, 94. [Google Scholar] [CrossRef]
  158. Couper, L.I.; Farner, J.E.; Caldwell, J.M.; Childs, M.L.; Harris, M.J.; Kirk, D.G.; Nova, N.; Shocket, M.; Skinner, E.B.; Uricchio, L.H.; et al. How will mosquitoes adapt to climate warming? eLife 2021, 10, e69630. [Google Scholar] [CrossRef]
  159. World Health Organization. Prioritizing Diseases for Research and Development in Emergency Contexts. Available online: https://www.who.int/activities/prioritizing-diseases-for-research-and-development-in-emergency-contexts (accessed on 15 August 2023).
  160. Stoek, F.; Rissmann, M.; Ulrich, R.; Eiden, M.; Groschup, M.H. Black rats (Rattus rattus) as potential reservoir hosts for Rift Valley fever phlebovirus: Experimental infection results in viral replication and shedding without clinical manifestation. Transbound. Emerg. Dis. 2021, 69, 1307–1318. [Google Scholar] [CrossRef]
  161. Saeed, O.S.; El-Deeb, A.H.; Gadalla, M.R.; El-Soally, S.A.G.; Ahmed, H.A.H. Genetic Characterization of Rift Valley Fever Virus in Insectivorous Bats, Egypt. Vector-Borne Zoonotic Dis. 2021, 21, 1003–1006. [Google Scholar] [CrossRef]
  162. Bird, B.H.; Khristova, M.L.; Rollin, P.E.; Ksiazek, T.G.; Nichol, S.T. Complete Genome Analysis of 33 Ecologically and Biologically Diverse Rift Valley Fever Virus Strains Reveals Widespread Virus Movement and Low Genetic Diversity due to Recent Common Ancestry. J. Virol. 2007, 81, 2805–2816. [Google Scholar] [CrossRef] [PubMed]
  163. Cartwright, H.N.; Barbeau, D.J.; McElroy, A.K. Rift Valley Fever Virus Is Lethal in Different Inbred Mouse Strains Independent of Sex. Front. Microbiol. 2020, 11, 1962. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Immunity to RVFV infection in host cells and evasion strategies adopted by the virus. RVFV uses Lrp1 as one of the attachment factors and enters host cell by endocytosis. Low pH in endosome triggers membrane fusion, release of virus genome and proteins. Viral RNA is sensed by RIG-I/MDA5 which interacts with MAVS and activates TBK1 signalling. NF-κB is phosphorylated and translocates to the nucleus, resulting in IFN production. MAVS also triggers NLRP3 inflammasome formation. IFN interacts with IFNAR receptor in autocrine or paracrine manner, phosphorylating STAT protein, which acts as transcription factor for ISGs and PKR. PKR phosphorylates eIF2α, thereby initiating eukaryotic translation. Viral proteins released from endosome including nucleoprotein and NSs affect host immune response. N protein interacts with SQSTM1 to generate autophagosome. Exosomes originating from RVFV-infected cells can also trigger autophagy. NSs protein suppresses innate response by several methods, including cell cycle arrest and inhibiting STAT phosphorylation, preventing IFNAR signalling. NSs forms fibrils which localize in nucleus, interact with SAP30 and prevent IFN transcription. NSs also degrades PKR, halting eukaryotic translation and promoting virus RNA translation. Created with BioRender.com (accessed on 20 August 2023).
Figure 1. Immunity to RVFV infection in host cells and evasion strategies adopted by the virus. RVFV uses Lrp1 as one of the attachment factors and enters host cell by endocytosis. Low pH in endosome triggers membrane fusion, release of virus genome and proteins. Viral RNA is sensed by RIG-I/MDA5 which interacts with MAVS and activates TBK1 signalling. NF-κB is phosphorylated and translocates to the nucleus, resulting in IFN production. MAVS also triggers NLRP3 inflammasome formation. IFN interacts with IFNAR receptor in autocrine or paracrine manner, phosphorylating STAT protein, which acts as transcription factor for ISGs and PKR. PKR phosphorylates eIF2α, thereby initiating eukaryotic translation. Viral proteins released from endosome including nucleoprotein and NSs affect host immune response. N protein interacts with SQSTM1 to generate autophagosome. Exosomes originating from RVFV-infected cells can also trigger autophagy. NSs protein suppresses innate response by several methods, including cell cycle arrest and inhibiting STAT phosphorylation, preventing IFNAR signalling. NSs forms fibrils which localize in nucleus, interact with SAP30 and prevent IFN transcription. NSs also degrades PKR, halting eukaryotic translation and promoting virus RNA translation. Created with BioRender.com (accessed on 20 August 2023).
Pathogens 12 01174 g001
Table 1. Representative RVFV vaccine candidates developed over the years.
Table 1. Representative RVFV vaccine candidates developed over the years.
Vaccine CandidateType of VaccineImmune ResponseSpecies TestedCommentsReferences
SmithburnLive-attenuatedNeutralizing antibodyCow, goat, alpacaAdverse effects: abortion, pathologic manifestation[134,156,157]
NDBR-103Formalin inactivatedNeutralizing antibodyHumansSide effect: reported case of Guillain-Barré Syndrome[116,117]
TSI-GSD-200Formalin inactivatedNeutralizing antibody, T cellHumansBooster vaccinations required[118,119,120]
MP-12Live-attenuatedNeutralizing antibodySheep, goat, alpaca, rhesus macaque, humansAdverse effect for foetus when administered in early pregnancy, phase 1 and 2 clinical trials [123,124,125,126,128,129,130,131]
Clone-13Live-attenuatedNeutralizing antibodySheep, goat, cattleVertical transmission[133,134,135,136,137]
RVFV-4sLive-attenuatedNeutralizing antibodySheep, goat, cow, common marmosetNo evidence of vertical transmission[139,140,141,142]
arMP-12-ΔNSm21/384 1Live-attenuatedNeutralizing antibody, IFN-γ responseSheep [143]
rZH501-ΔNSs:GFP, rZH501-ΔNSs:GFP-ΔNSm 2Live-attenuatedNeutralizing antibodyRats [144]
pWRG7077-RVFV-NSmDNA-vectoredNeutralizing antibodyMice [145]
Gn-C3dDNA-vectoredNeutralizing antibodyMiceProtects passively immunized mice against RVFV-ZH501 challenge[146]
RVFV-BLP 3 (Gn head)Bacterium-like particleAntibody and T cell responseMice [155]
rLSDV-RVFV Virus-vectored Neutralizing antibodySheep [152]
rVEEV-RVFVVirus-vectoredNeutralizing antibodyMouse [147]
rMVA-RVFVVirus-vectoredNeutralizing antibody, T cellMouse [151]
ChAdOx1-RVFV 4Virus vectoredNeutralizing antibodyMouse, sheep, goatViremia in placenta of pregnant doe challenged with RVFV-35/74[153,154]
1 arMP-12-ΔNSm21/384- recombinant MP12 with M gene deletion from nucleotide 21 to 384. 2 rZH501-ΔNSs:GFP, rZH501-ΔNSs:GFP-ΔNSm- recombinant ZH501 with deletion of NSs gene or NSs and NSm genes and coupled with GFP. 3 BLP- bacterium like particle. 4 ChAdOx1-chimpanzee adenovirus vector.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nair, N.; Osterhaus, A.D.M.E.; Rimmelzwaan, G.F.; Prajeeth, C.K. Rift Valley Fever Virus—Infection, Pathogenesis and Host Immune Responses. Pathogens 2023, 12, 1174. https://doi.org/10.3390/pathogens12091174

AMA Style

Nair N, Osterhaus ADME, Rimmelzwaan GF, Prajeeth CK. Rift Valley Fever Virus—Infection, Pathogenesis and Host Immune Responses. Pathogens. 2023; 12(9):1174. https://doi.org/10.3390/pathogens12091174

Chicago/Turabian Style

Nair, Niranjana, Albert D. M. E. Osterhaus, Guus F. Rimmelzwaan, and Chittappen Kandiyil Prajeeth. 2023. "Rift Valley Fever Virus—Infection, Pathogenesis and Host Immune Responses" Pathogens 12, no. 9: 1174. https://doi.org/10.3390/pathogens12091174

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

Article Metrics

Back to TopTop