Next Article in Journal
Nipah Virus Impairs Autocrine IFN Signaling by Sequestering STAT1 and STAT2 into Inclusion Bodies
Next Article in Special Issue
Novel Epitopes Mapping of African Swine Fever Virus CP312R Protein Using Monoclonal Antibodies
Previous Article in Journal
Hyperinflammatory Response in COVID-19: A Systematic Review
Previous Article in Special Issue
HoBi-like Pestivirus Is Highly Prevalent in Cattle Herds in the Amazon Region (Northern Brazil)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Bovine Herpesvirus 1 Latency-Reactivation Cycle, a Chronic Problem in the Cattle Industry

Department of Veterinary Pathobiology, College of Veterinary Medicine, Oklahoma State University, Stillwater, OK 74078, USA
*
Author to whom correspondence should be addressed.
Viruses 2023, 15(2), 552; https://doi.org/10.3390/v15020552
Submission received: 26 January 2023 / Revised: 10 February 2023 / Accepted: 11 February 2023 / Published: 16 February 2023

Abstract

:
Bovine alphaherpesvirus 1 (BoHV-1) is a persistent and recurring disease that affects cattle worldwide. It is a major contributor to bovine respiratory disease and reproductive failure in the US. A major complication of BoHV-1 arises from the lifelong latent infection established in the sensory ganglia of the peripheral nervous system following acute infection. Lifelong latency is marked by periodic reactivation from latency that leads to virus transmission and transient immunosuppression. Physiological and environmental stress, along with hormone fluctuations, can drive virus reactivation from latency, allowing the virus to spread rapidly. This review discusses the mechanisms of the latency/reactivation cycle, with particular emphasis on how different hormones directly regulate BoHV-1 gene expression and productive infection. Glucocorticoids, including the synthetic corticosteroid dexamethasone, are major effectors of the stress response. Stress directly regulates BoHV-1 gene expression through multiple pathways, including β-catenin dependent Wnt signaling, and the glucocorticoid receptor. Related type 1 nuclear hormone receptors, the androgen and progesterone receptors, also drive BoHV-1 gene expression and productive infection. These receptors form feed-forward transcription loops with the stress-induced Krüppel-like transcription factors KLF4 and KLF15. Understanding these molecular pathways is critical for developing novel therapeutics designed to block reactivation and reduce virus spread and disease.

1. BoHV-1 Is a World-Wide Pathogen That Causes Several Diseases

Bovine alphaherpesvirus 1 (BoHV-1) is a member of the Alphaherpesvirinae subfamily and genus Varicellovirus [1,2]. The genome is 135.3 kilobase pairs (kbp) with a high guanine and cytosine (GC) content and is arranged as a class D herpesvirus genome [3]. The viral genome contains a unique long (UL, 104 kbp) and unique short (US, 10 kbp) region with the latter flanked on both sides by inverted internal and external repeats [4] (Figure 1A). Approximately 70 open reading frames (ORFs) have been identified, most of which are functional homologues to genes present in other alphaherpesviruses [4,5,6].
All BoHV-1 strains belong to one single viral species but are divided into three subtypes based on antigenic and genomic analysis: BoHV-1.1, -1.2a, and -1.2b [7]. Subtype 1 virus isolates are the causative agents for infectious bovine rhinotracheitis (IBR) and are typically found in the respiratory tract and aborted fetuses. Subtype 1 strains are common in Europe and North and South America. Subtype 2a infections are associated with a broad range of clinical symptoms in the respiratory and genital tract, for example, infectious pustular vulvovaginitis (IPV), infectious balanoposthitis (IPB), and abortions [8]. Subtype 2a was prevalent in Brazil and Europe prior to 1970 [8]. Subtype 2b has been isolated in Australia and Europe and is associated with respiratory disease, IPV, and IPB but not abortions [8,9].

2. BoHV-1 Is an Important Pathogen in Cattle

Natural BoHV-1 infection occurs by contact with the virus in mucosal membranes of the upper respiratory or genital tracts of cattle [10]. Virus entry into the respiratory tract can occur via aerosol or by direct contact with virus present in nasal secretions [11]. Genital transmission generally requires direct contact at mating; however, virus has been isolated from semen samples used for artificial insemination [8,12]. Two major syndromes are caused by BoHV-1 infections: (i) IBR which occurs in the respiratory tract and (ii) IPV (cows) or IPB (bulls) which occurs in the genital tract. IBR is associated with clinical symptoms including pyrexia, apathy, increased respiratory rate with persistent harsh cough, and anorexia. Adult dairy cows also exhibit a severe drop in milk production. Primary infection occurs in the nasal and tracheal turbinate, while mucopurulent discharge from the nostrils and eyes is associated with pustular lesions in the nasal mucosa and conjunctivitis [4,13,14]. Multiple factors influence the clinical symptoms and severity of the disease caused by BoHV-1: strain-specific virulence, infected tissue type, secondary bacterial infection, host age and resistance factors.
Notably, BoHV-1 is the most frequently diagnosed cause of viral abortion in North American cattle [15]. Although current BoHV-1 modified live vaccines are generally effective at preventing disease in cattle, infection or vaccination with a modified live vaccine can also induce abortion storms in pregnant cows [15,16,17,18,19]. BoHV-1, including commercially available modified live vaccines, can cause abortion storms where 50% of cows in herds are affected. Published studies concluded that unbred heifers vaccinated with a “killed” BoHV-1 vaccine have fewer abnormal estrus cycles and higher pregnancy rates when compared to heifers vaccinated with a modified live BoHV-1 vaccine [15,16,17,18,19]. Furthermore, current modified live vaccines reactivate from latency with nearly the same frequency as virulent strains present in nature, which complicates the use of these vaccines in cow and calf herds.
BoHV-1 infection is a major risk factor for bovine respiratory disease (BRD), a poly-microbial disease that can culminate in mortality due to bacterial pneumonia. BRD is the most economically important disease affecting beef and dairy cattle, accounting for approximately 75% of morbidities and >50% of mortalities in feedlot cattle [20,21,22,23,24]. Mannheimia haemolytica (MH) is a commensal bacterium [25] that belongs to the normal flora in the upper respiratory tract of healthy ruminants [26]. This commensal relationship is disrupted following stress or co-infections [27]; consequently, MH becomes the predominant organism that causes bronchopneumonia in many BRD cases [28,29,30,31]. BoHV-1 infection frequently causes upper respiratory tract disease [14,32], and co-infection of calves with MH consistently causes life-threatening pneumonia [33]. BoHV-1 enhances interactions between the MH leukotoxin and bovine peripheral blood mononuclear cells, including neutrophils [34,35]. Acute BoHV-1 infection impairs cell-mediated immunity [36,37,38,39], CD8+ T cell recognition of infected cells [40,41,42,43], and CD4+ T cell functions. BoHV-1 infection also erodes mucosal surfaces of the upper respiratory tract, which promotes MH colonization in the lower respiratory tract [28,29,30,31]. Notably, a BoHV-1 entry protein is a BRD susceptibility gene for Holstein calves [44], supporting the premise this virus is an important BRD cofactor.

3. BoHV-1 Acute Infection

Acute infection induces apoptosis, inflammation, and high levels of virus shedding [13,14]. Lesions may be present, but subclinical viral shedding may also occur. Viral gene expression occurs in three well-defined phases: immediate early (IE), early (E), or late (L). IE gene expression begins immediately following virus entry and requires only host factors and virus tegument proteins present in the infecting virion. IE transcription unit 1 (IEtu1) encodes two transcriptional regulatory proteins (bICP0 and bICP4) which stimulate the productive infection [45,46,47] (Figure 1A). IEtu2 encodes the bICP22 protein [45]. These proteins are required for efficient viral gene expression. A tegument protein, VP16, stimulates IE transcription by interacting with sequences present in all IE promoters [48,49]. IE proteins drive E gene expression, which are generally translated into non-structural proteins that promote viral DNA replication. L proteins comprise the infectious virus capsid and tegument, which are required for virion assembly and egress. Viral particles produced during acute infection enter the peripheral nervous system via cell-cell spread into sensory neurons. A productive infection will periodically reoccur (reactivation from latency) as virus is shed from neurons periodically during the lifetime of the host.
Figure 1. Schematic of BoHV-1 genes encompassing the LR gene. (A) The BoHV-1 genome contains a unique long (L), a unique short (S), and two repeats denoted by white rectangles. Positions of IE transcripts and LR transcript (LR-RNA) [45,46,50] are denoted. IE/4.2 is the IE transcript that encodes bICP4. IE/2.9 is the IE transcript that encodes bICP0. The IEtu1 promoter activates the expression of IE/4.2 and IE/2.9 and is denoted by a black rectangle (IEtu1 pro). E/2.6 is the early transcript that encodes bICP0, and a separate early promoter regulates the expression of the E/2.6 transcript (E pro). Exon 2 (e2) of bICP0 contains all the protein-coding sequences. The origin of replication (ORI) separates IEtu1 from IEtu2. The IEtu2 promoter (IEtu2 pro) drives the expression of the bICP22 protein. Solid lines in the transcript denote exons (e1, e2, e3) and dashed lines introns. (B) Partial restriction map of LR gene. The LR gene contains two open reading frames (ORF-1 and ORF-2) [51,52,53]. Reading frame B (RF-B) and reading frame C (RF-C) do not contain an initiating methionine. Asterisks denote the positions of in-frame stop codons. Start sties for lytic and latent LR-RNA are denoted. The position of the bICP0 termination codon (single red *) and LR-micro-RNAs are also denoted. The *** denote the position of the three in frame stop codons for ORF2.
Figure 1. Schematic of BoHV-1 genes encompassing the LR gene. (A) The BoHV-1 genome contains a unique long (L), a unique short (S), and two repeats denoted by white rectangles. Positions of IE transcripts and LR transcript (LR-RNA) [45,46,50] are denoted. IE/4.2 is the IE transcript that encodes bICP4. IE/2.9 is the IE transcript that encodes bICP0. The IEtu1 promoter activates the expression of IE/4.2 and IE/2.9 and is denoted by a black rectangle (IEtu1 pro). E/2.6 is the early transcript that encodes bICP0, and a separate early promoter regulates the expression of the E/2.6 transcript (E pro). Exon 2 (e2) of bICP0 contains all the protein-coding sequences. The origin of replication (ORI) separates IEtu1 from IEtu2. The IEtu2 promoter (IEtu2 pro) drives the expression of the bICP22 protein. Solid lines in the transcript denote exons (e1, e2, e3) and dashed lines introns. (B) Partial restriction map of LR gene. The LR gene contains two open reading frames (ORF-1 and ORF-2) [51,52,53]. Reading frame B (RF-B) and reading frame C (RF-C) do not contain an initiating methionine. Asterisks denote the positions of in-frame stop codons. Start sties for lytic and latent LR-RNA are denoted. The position of the bICP0 termination codon (single red *) and LR-micro-RNAs are also denoted. The *** denote the position of the three in frame stop codons for ORF2.
Viruses 15 00552 g001

4. BoHV-1 Establishes Life-Long Latency in Neurons and Lymphoid Tissue

If acute infection is initiated within the oral, nasal, or ocular cavity, sensory neurons in trigeminal ganglia (TG) are a primary site for latency [51,52,53,54]. In acutely infected sensory neurons, lytic cycle viral gene expression and virus production occur [55]. Lytic cycle viral gene expression is then extinguished, a significant number of infected neurons survive, and surviving infected neurons harbor viral genomes; these steps are operationally defined as the establishment of latency. The maintenance of latency is defined as long-term repression of lytic cycle viral gene expression and virus production. The only viral gene abundantly expressed during latency is the latency-related (LR) gene [52,53,56,57]; LR gene functions are discussed below.
Pharyngeal tonsils (PT) are continually exposed to viral and bacterial pathogens. For example, viruses that infect the nasal cavity drain into PT and can potentially infect PT cells. Furthermore, the lacrimal duct system transmits passage of tear film from ocular surfaces to the nasal cavity; consequently, the BoHV-1 can drain into the PT. BoHV-1 DNA is consistently detected in PT of latently infected cattle [58,59,60,61]. Viral DNA from other 𝛼-herpesvirinae subfamily members is also detected in the PT of their respective natural hosts. These include pseudorabies virus [62,63], equine herpesvirus 4 [64], canine herpesvirus 1 [65], and human alphaherpesvirus 1 (HSV-1) but not HSV-2 [66]. For these studies, their respective hosts were not shedding detectable levels of infectious virus nor were they exhibiting disease symptoms suggesting they were latently infected. We previously demonstrated that BoHV-1 lytic cycle viral gene expression is detected in PT using in situ hybridization when latently infected calves were given an IV injection of the synthetic corticosteroid dexamethasone (DEX) [58]. Consequently, we suggest that virus shedding during reactivation from latency via PT is crucial for virus transmission.

5. LR Gene Products Modulate the Latency-Reactivation Cycle

5.1. Identification and Characterization of Viral Gene Products Expressed during Latency

The only BoHV-1 gene abundantly expressed during latency is LR-RNA [67,68,69]. In situ hybridization revealed that most LR-RNA is localized in the nucleus of latently infected neurons; however, LR-RNA is also detected in the cytoplasm [67,68,70,71]. A subset of LR-RNA is polyadenylated and alternatively spliced [72,73]. The start site for LR-RNA during productive infection is 24 bases downstream of the TATA box (Figure 1B) [74]. A different start site is utilized during latency that is 200–300 nucleotides upstream of the lytic site [73]. The LR gene promoter is contained within a 980 bp PstI fragment. This promoter is more active in neuronal cells when compared to non-neuronal cells and promoter activity is downregulated by DEX [75,76].
Two open reading frames (ORF) (ORF1 and ORF2) and two reading frames without an initiating methionine (RF-B and RF-C) are present in the LR gene [67,68,69,70,71] (Figure 1B). LR-RNA splicing in TG of calves generates a poly A+ mRNA where ORF2 is intact in TG at day one after the infection of the calves [72,77,78]. Alternative splicing of LR-RNA in TG generates ORF2 coding sequences at the amino terminus whereas the C-terminus contains portions of ORF1 or RF-B coding sequences at 7 or 15 dpi, respectively [72,77,78]. Full-length ORF1 would also be expressed from an unspliced LR transcript. In sharp contrast to TG, LR-RNA splicing in productively infected bovine cells is different indicating neuronal-specific splicing occurs in TG of infected calves.
ORF2 is a multi-functional protein and a master regulator of the latency-reactivation cycle. For example, ORF2 impairs apoptosis in transiently transfected mouse neuroblastoma cells (Neuro-2A) suggesting this function promotes survival of infected cells, including neurons [79]. Yeast two-hybrid studies revealed ORF2 interacts with cellular regulatory proteins, for example, Notch family members, β-catenin, and CCAAT enhancer binding protein alpha (c/EBP-α) [80,81,82,83,84]. These interactions were confirmed in transiently transfected Neuro-2A cells. Interestingly, Neuro-2A cells consistently support ORF2 protein expression and expression of ORF2 isoforms generated by splicing from an expression vector; conversely, other cell types examined do not support protein expression. C/EBP-α and Notch-1 activate bICP0 expression, an important viral trans-activator [85,86], suggesting ORF2 impairs promoter activity via interactions with C/EBP-α and Notch1. The ability of Notch family members to impair neurite formation in Neuro-2A cells was also blocked when ORF2 was present [87], suggesting ORF2 enhances neuronal differentiation. Two LR-encoded micro-RNAs interfered with bICP0 protein expression [88] (see Figure 1B for locations of LR-encoded micro-RNAs). These viral micro-RNAs may also reduce expression of additional viral or cellular proteins. While ORF2 plays a prominent role in the BoHV-1 latency-reactivation cycle, other LR gene products are also predicted to be important.
A mutant BoHV-1 virus that contains three stop codons at the beginning of ORF2 (LR mutant virus) does not express ORF2 or RF-C but expresses low levels of ORF1 [57,89,90]. Calf studies revealed that the LR mutant virus exhibits reduced virus shedding in ocular tissue, TG, and tonsils during acute infection. During establishment of latency, higher levels of apoptosis are detected in TG neurons of calves infected with the LR mutant when compared to calves infected with wt or the rescued LR mutant. This was unexpected because the LR mutant virus does not grow as well as wt BoHV-1 or the rescued LR mutant virus in calves. Strikingly, the LR mutant virus does not shed virus following DEX treatment, which triggers reactivation from latency in all calves infected with wt BoHV-1 or the rescued LR mutant [57,60]. Thus, ORF2 expression correlates with efficient establishment and maintenance of latency [91].
A recombinant human alphaherpesvirus 1 (HSV-1) that contains the BoHV-1 LR promoter and coding sequences inserted into an HSV-1 mutant lacking sequences encoding the first 1.5 kb of the latency-associated transcript (LAT) was constructed; this virus is referred to as CJLAT. CJLAT exhibits enhanced reactivation from latency in mice and rabbits when compared to the parental wild-type HSV-1 McKrae strain [92]. Notably, CJLAT also increases the incidence of fatal encephalitis in mice and rabbits relative to wild-type HSV-1. The introduction of three stop codons at the N-terminus of ORF2 and following insertion into the same site of the HSV-1 LAT mutant generated a virus that behaved like the parental LAT null mutant, suggesting ORF2 expression enhances reactivation from latency and pathogenesis [93]. Additional studies revealed CJLAT sustains corneal scarring and neovascularization after acute infection of rabbits, suggesting recurrent disease occurred [94]. These studies indicate that the multi-functional ORF2 enhances the HSV-1 latency-reactivation cycle by co-opting certain important cellular regulatory proteins and signaling pathways. HSV-1 LAT is predicted to express multiple micro-RNAs, two small non-coding RNAs, and a long-coding RNA but no functional promoter [51,52]. Consequently, ORF2 expression may have induced immune responses during acute infection and latency, which exacerbates ocular scarring. Finally, ORF2 expression may also enhance inflammation in the nervous system, culminating in encephalitis.

5.2. DEX Promotes BoHV-1 Productive Infection and Consistently Initiates Reactivation from Latency

DEX, a synthetic corticosteroid hormone that mimics the effects of stress, consistently induces reactivation from latency in calves or rabbits latently infected with BoHV-1 [57,68,85,95,96,97,98,99,100,101]. Stressful stimuli activate the hypothalamic-pituitary-adrenal axis leading to the release of glucocorticoids (GCs) from the adrenal gland, reviewed in [102]. GCs have a pleiotropic effect on cell growth and survival, development, and metabolism. GCs enter a cell and bind to a cytoplasmic localized glucocorticoid receptor (GR) or a mineralocorticoid receptor (MR), reviewed in [103]. MR or GR homodimers bound to DEX are translocated to the nucleus where they bind consensus GR response elements (GRE), remodel chromatin, and induce transcription in the absence of de novo protein synthesis. Consequently, stress is an immediate early response that does not require de novo protein synthesis and stimulates transcription within minutes. DEX promotes productive infection in bovine kidney or Neuro-2A cells transfected with BoHV-1 genomic DNA [104,105]. This is independent of virion VP16 because viral DNA was transfected into cells; thus, the virion tegument containing VP16 was not initially present in the transfected cell. VP16 is expressed early during reactivation but is not initially present.
Corticosteroids also modulate the serum and glucocorticoid-regulated protein kinases (SGK), members of the AGC family of serine/threonine protein kinases, reviewed in [106]. SGK1 mRNA and protein levels are rapidly increased following stressful stimuli. Inhibition of SGK activity demonstrated a marked reduction in BoHV-1 productive infection and viral protein synthesis [107]. Importantly, SGK inhibition did not block GR and DEX-dependent activation of the BoHV-1 IEtu1. This demonstrates multiple independent roles for stress-induced viral replication.
During the early stages of DEX-induced reactivation in TG, bICP0 and VP16 are detected before bICP4 and bICP22 [108,109,110]. Conversely, bICP4 mRNA, but not bICP0, bICP22, and VP16, is abundantly expressed within 30 min after DEX treatment in PT of latently infected calves [111]. Since DEX reproducibly initiates reactivation from latency in the natural host, we have searched for viral and cellular factors that mediate reactivation from latency, as discussed below.

6. Regulation of Wnt/β-Catenin Signaling Pathway during Latency and DEX-Induced Reactivation

RNA sequencing studies revealed that more than 100 genes linked to the canonical Wnt/β-catenin signaling pathway are differentially expressed in TG during latency when compared to 30 min and 3 h after an intravenous DEX injection was given to latently infected calves to initiate reactivation from latency [81]. The canonical Wnt/β-catenin signaling pathway is more active during latency when compared to TG from uninfected calves or DEX-induced reactivation from latency (Figure 2A). Notably, four Wnt agonists are expressed at significantly higher levels during latency. The two genes with the highest levels of differential expression are the guanine nucleotide-binding protein alpha Q (GNAQ) [112,113] and Wnt ligand 16 (Wnt16). The third Wnt agonist is bone morphogenetic protein receptor 2 (BMPR2), a serine/threonine protein kinase, required for dorsoventral patterning of TG, peripheral innervation, and survival of sensory neurons [114,115,116]. Finally, a serine/threonine cellular protein kinase (Akt3) is differentially expressed [117] (see Figure 2A). Akt3 is more important than Akt1 and Akt2 with respect to blocking stroke-induced neuronal injury [118], blocking neuronal apoptosis, and triggering axonal development [117]. These differentially expressed cellular genes promote β-catenin stabilization and nuclear localization of β-catenin, in part because Wnt 16 or other Wnt proteins enhance expression of Disheveled, which impairs the association of β-catenin with the destruction pathway. The β-catenin destruction pathway is comprised of GSK-3b (glycogen 3 kinase 3b), Axin, APC (Adenomatous Polyposis Coli), and CKIa (Casein kinase I isoform alpha). Nuclear β-catenin forms an active transcription complex with T-cell factor/lymphoid enhancer factor (TCF/LEF) [119,120]. Generally, the canonical Wnt/β-catenin signaling pathway promotes cellular growth and constitutively activated β-catenin is linked to tumor cell development [121]. In sharp contrast, the canonical Wnt/β-catenin signaling pathway enhances neurogenesis and neuronal survival in the nervous system [120,121,122,123].
Recent studies revealed Akt1 and Akt 2 impair stress-induced transcription, whereas Akt3 enhances neurite formation in Neuro-2A [124]. Akt1 and Akt2 are expressed in TG during the latency-reactivation cycle, but their RNA levels do not change dramatically after DEX treatment. Furthermore, ORF2 stimulates β-catenin-dependent transcription in transfected cells [125] and promotes neuronal survival during latency. Finally, over-expression of β-catenin and a small non-coding RNA spanning part of ORF-2 coding sequences impair stress-induced transcription [126]. Collectively, these studies revealed that maintaining latency is an active process mediated, in part, by activation of the canonical Wnt/β-catenin signaling pathway.
Expression of five known Wnt antagonists, dickkopf-1 (DKK1), dickkopf-1 like protein (DKKL1), Wnt inhibitory factor 1 (Wif-1), secreted frizzled-related protein 4 (SFRP4), SFRP5, and two intracellular Wnt inhibitors (SOX3 and SOX18) are induced within 3 h after DEX treatment to trigger reactivation from latency (Figure 2B). The canonical Wnt/β-catenin signaling pathway is negatively regulated by the DKK family of secreted proteins. In general, Wnt antagonists specifically bind to Wnt co-receptors, the low-density lipoprotein receptor-related proteins (LRP), and these interactions block Wnt family members from binding to the receptor/coreceptor complex [121,127,128]. SFRP family members bind to Wnt family members and block the Wnt pathway. Interestingly, SFRP4 also induces apoptosis [129,130], in part by blocking Akt signaling pathways [131]. A destruction complex interacts with β-catenin in the cytoplasm, culminating in reduced β-catenin steady-state protein levels (Figure 2B). Initially, β-catenin is phosphorylated by CKIa and then GSK-3β phosphorylates β-catenin. β-catenin phosphorylation targets the protein for ubiquitination and degradation by the proteasome. SOX3 and SOX18 impair β-catenin-dependent transcription independent of the β-catenin destruction pathway [132]. In summary, soluble Wnt antagonists induced by corticosteroids, including DEX, stabilize the β-catenin degradation complex and promote β-catenin phosphorylation and degradation. Induction of apoptosis, via impairment of the Wnt/Akt signaling pathways, may increase α-herpesvirus replication [133] and reactivation from latency [134].

7. GR and Stress-Induced Transcription Factors Stimulate Key Viral Promoters

GR interactions with target DNA are mediated in part by the glucocorticoid response element (GRE, 5′-GGTACANNNTGTTCT-3′). Over 100 potential GREs were identified in the BoHV-1 genome, including two in the immediate early transcription unit 1 (IEtu1) promoter, a regulatory element that drives ICP4 and ICP0 expression following virus entry into a cell and at the onset reactivation (Figure 3A). These critical regulatory proteins drive viral gene expression and are required for efficient infection and reactivation from latency. Indeed, GR and DEX strongly activate the IEtu1 promoter, and transactivation is significantly reduced when either GRE is mutated [104]. Interestingly, one of the two GREs is an exact match to the consensus sequence, and mutating this site shows a more dramatic reduction in GR-induced activation relative to the other GRE, which contains a single mismatch from the consensus. Mutating both sites completely abolishes GR and DEX-induced promoter activation.
During early stages of reactivation from latency, novel stress-induced cellular transcription factors were identified in TG [96], suggesting these cellular transcription factors play an important role during the reactivation from latency. These stress-induced transcription factors include promyelocytic leukemia zinc finger (PLZF), Slug, SPDEF (Sam-pointed domain containing Ets transcription factor), Krüppel-like transcription factor 4 (KLF4), KLF6, KLF15, and GATA6. These stress-induced transcription factors stimulate productive infection and key viral promoters [96,105,135,136,137,138,139]. The finding that 4 KLF family members (KLF4, KLF6, KLF15, and PLZF) are stimulated during DEX-induced reactivation from latency is important because KLF family members belong to the Sp1 transcription factor family and certain KLF family members interact with GC-rich motifs, including Sp1 binding sites, reviewed in [140,141]. The BoHV-1 genome is GC-rich, and many viral promoters contain Sp1 consensus binding sites and other GC-rich motifs, suggesting KLF binding sites are frequently found in the viral genome [140]. Expression of key viral regulatory proteins (bICP0, bICP4, bICP22, and VP16), but not glycoprotein C or E, are detected in TG neurons within 90 min after DEX treatment [108,109,110].
GR interacts with several of these stress-induced transcription factors forming feed-forward regulatory networks to enhance BoHV-1 gene expression. KLF4 and KLF15 are of particular importance. GR and KLF15 cooperatively transactivate the IEtu1 promoter in the presence of DEX by 40-fold, which is significantly higher than GR + DEX (~8-fold) or KLF15 alone (~3-fold) [139]. Transactivation is dependent upon the two GREs located in the DEX-responsive region of the IEtu1 promoter. As with GR + DEX, the consensus GRE is largely responsible for GR and KLF15-mediated transactivation, as a mutation of this sequence significantly reduces promoter activity. Additionally, a KLF-like binding site is present in this region; however, mutating this site only results in an approximate 30% reduction in promoter activity. It is notable that GR and KLF15 directly interact, as demonstrated by coimmunoprecipitation, suggesting cooperative transactivation is the result of GR-KLF15 protein complexes.
The BoHV-1 genome was analyzed for similar motifs with GREs and a KLF-like site, and 13 sites were identified. Not all these sites were in promoters, but GR-dependent transactivation has been documented in genes up to 5kb from the GRE [142]. These 13 sites were analyzed for GR and KLF15-dependent transactivation. Interestingly, none of the sites were strongly transactivated by either GR + DEX or KLF15. However, four of these sites were significantly transactivated by GR and KLF15 cooperatively in the presence of DEX. These sites included a UL52 intergenic sequence, a bICP4 sequence, the IEtu2, and part of the unique short region. Of these, the UL52, bICP4 site, and IEtu1 are also transactivated by GR in combination with the related KLF4 protein. The UL52 intergenic sequence was strongly transactivated by GR and KLF15. UL52 is the primase subunit of the herpesvirus helicase-primase complex; how this GR-KLF15 responsive region affects virus gene expression or productive infection is unknown.
IEtu1 drives bICP0 and bICP4 expression during the immediate early stages of BoHV-1 infection from a single primary transcript that is alternatively spliced (Figure 1). Additionally, a second promoter (bICP0 E pro) drives bICP0 expression during the early stage of infection [143]. This promoter is strongly transactivated by GR along with KLF4 and KLF15; however, DEX is either inhibitory, in the case of KLF4, or does not further stimulate activation in the case of KLF15 [105]. Ligand-independent GR activation is well documented, particularly with certain transcriptional coactivators [144,145]. The bICP0 E promoter is quite complex (Figure 3B). For example, it is highly GC-rich (~80%), contains numerous potential KLF and Sp1 binding sites, and has two half-GREs [105,138,139]. KLF4 and KLF15 belong to the Sp1/KLF superfamily of zinc-finger transcription factors. There is a similarity between Sp1 binding sites (GGGCGG) and certain KLF4 binding sites (GGGAGGG), and both KLF4 and KLF15 can bind Sp1 sites. KLF4 additionally binds to certain CACCC elements, reviewed in [140,141]. In the bICP0 E promoter, three separate and independent enhancer domains were identified, ranging from nucleotides 943 to 638 (EP-943), nucleotides 638 to 328 (EP-638), and nucleotides 328 to 172 (EP-328). Each domain is transactivated by GR and KLF4, while only EP-638 is transactivated by GR with KLF15 [105,146]. Both half-GREs are in the EP-943 fragment and are dispensable for transactivation. This is not surprising given the unliganded nature of GR-mediated transactivation of the bICP0 E promoter. However, an Sp1 and KLF4 site in this fragment are both required for transactivation by GR and KLF4. The EP-328 fragment also contains an Sp1 and a KLF4 site, both of which are required for transactivation. The EP-638 fragment contains two separate Sp1 sites, a KLF4 site, and a large triple Sp1 site (GGGCGGGGGCGGGGGCGGG). The triple Sp1 site is required for GR and KLF4 transactivation. These sites are directly responsible for GR and KLF4 mediated transactivation, as mutations reduced GR and KLF4 occupancy of enhancer fragments. It is notable that while GR expression plasmids were required for the transactivation of these enhancer fragments, chromatin immunoprecipitation (ChIP) experiments identified enhancer occupancy by endogenous GR, in the absence of over-expressing GR. Previous work demonstrated two potential isoforms of GR by western blot analysis, which migrate at approximately 90 and 120 kDa [139]. Only the 90kDa isoform is present in untransfected cells, while both are observed with GR overexpression. Since the expression vector only expresses GR-a, which migrates at 120 kd, it is clear GR-a, not the 90 kd GR protein, transactivates the bICP0 E promoter. Notably, both isoforms are detected following coimmunoprecipitation with KLF15 antibodies. The role that different GR isoforms play in the transactivation of BoHV-1 promoters is the subject of ongoing studies.

8. Androgen and Progesterone Receptors Stimulate IE and E Promoters Cooperatively with KLF4 and KLF15

In addition to GR, the related androgen (AR) and progesterone receptors (PR) activate the BoHV-1 IEtu1 and bICP0 E promoters cooperatively with KLF4 and KLF15 [136,147,148]. AR and PR are type I nuclear hormone receptors activated by dihydrotestosterone (DHT) and progesterone (P4), respectively, allowing them to translocate to the nucleus and regulate transcription, similar to GR [149,150,151]. Importantly, both AR and PR recognize similar DNA sequences as GR and directly associate with KLF4 [147,148,152]. The AR strongly transactivates the IEtu1 promoter when stimulated with DHT, and this activation is increased by KLF15. In contrast, AR and KLF4 cooperatively transactivate the IEtu1 principally in the absence of DHT [147]. The PR also strongly activates the IEtu1 promoter when stimulated with P4, in cooperation with KLF15 [136]. IEtu1 transactivation by AR and PR is dependent upon two GREs, and their proximity to the transcription initiation site is important. When the 796 nucleotides between the TATA box and GREs are deleted (Figure 3A), cooperative transactivation between KLF15 and AR or PR is dramatically increased. The AR and PR also transactivate the bICP0 E promoter cooperatively with KLF4. As with GR, treatment with ligand (DHT or P4) impairs transactivation. The Sp1 and KLF4 binding sites in this promoter (Figure 3B) are required for KLF4 cooperative transactivation with AR and PR, suggesting a similar mechanism as GR. Furthermore, mutating these sites significantly reduced the occupancy of the promoter by all three transcription factors.
The responsiveness of IEtu1 and bICP0 E promoters to stress and sex hormone signaling suggests these cellular signaling pathways mediate key aspects of the BoHV-1 life cycle in vivo. For example, stress or mating correlates with heightened virus transmission, suggesting an evolutionary response by BoHV-1. While it is not clear whether AR agonists (DHT for example) trigger reactivation from latency, P4 sporadically induces reactivation from latency in female calves [153] or rabbits [154] latently infected with BoHV-1. Collectively, these studies revealed sex hormones have the potential to stimulate viral gene expression and induce reactivation from latency.

Author Contributions

Conceptualization, C.J., Writing—Original Draft Preparation, C.J. and J.B.O.; Writing—Review & Editing, C.J. and J.B.O.; Visualization, C.J. and J.B.O.; Supervision, C.J.; Project Administration, C.J.; Funding Acquisition, C.J. and J.B.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by USDA NIFA grant numbers [2018-06668, 2021-67015-34463, and 2022-08080], National Institute of Neurological Disorders and Stroke grant number (R01NS111167). NIH grant number (P20GM103648). The Sitlington Endowment also contributed to this research.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data is included in publications that were cited.

Acknowledgments

This research was supported by grants to C.J. from the USDA National Institute of Food and Agriculture (NIFA) (2018-06668 and 2021-67015-34463) and National Institute of Neurological Disorders and Stroke (R01NS111167), funds from the Oklahoma Center for Respiratory and Infectious Diseases (National Institutes of Health Centers for Biomedical Research Excellence Grant # P20GM103648), and funds from the Sitlington Endowment. J.B. Ostler is, in part, supported by a grant from USDA NIFA (2022-08080).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Roizman, B.; Knipe, D.M. Herpes Simplex Viruses and Their Replication. In Fields Virology, 4th ed.; Knipe, D.M., Howley, P.M., Eds.; Lippincott, Williams & Wilkins: New York, NY, USA, 2001; Volume 2, pp. 2399–2459. [Google Scholar]
  2. Brown, F. The classification and nomenclature of viruses: Summary of results of meetings of the International Committee on Taxonomy of Viruses in Edmonton, Canada 1987. Intervirology 1989, 30, 181–186. [Google Scholar] [CrossRef]
  3. Plummer, G.; Goodheart, C.R.; Henson, D.; Bowling, C.P. A comparative study of the DNA density and behavior in tissue cultures of fourteen different herpesviruses. Virology 1969, 39, 134–137. [Google Scholar] [CrossRef]
  4. Muylkens, B.; Thiry, J.; Kirten, P.; Schynts, F.; Thiry, E. Bovine herpesvirus 1 infection and infectious bovine rhinotracheitis. Vet. Res. 2007, 38, 181–209. [Google Scholar] [CrossRef] [Green Version]
  5. Biswas, S.; Bandyopadhyay, S.; Dimri, U.; Patra, P.H. Bovine herpesvirus-1 (BHV-1)—A re-emerging concern in livestock: A revisit to its biology, epidemiology, diagnosis, and prophylaxis. Vet. Q. 2013, 33, 68–81. [Google Scholar] [CrossRef] [Green Version]
  6. Robinson, K.E.; Meers, J.; Gravel, J.L.; McCarthy, F.M.; Mahony, T.J. The essential and non-essential genes of Bovine herpesvirus 1. J. Gen. Virol. 2008, 89, 2851–2863. [Google Scholar] [CrossRef]
  7. Metzler, A.E.; Matile, H.; Gassmann, U.; Engels, M.; Wyler, R. European isolates of bovine herpesvirus 1: A comparison of restriction endonuclease sites, polypeptides, and reactivity with monoclonal antibodies. Arch. Virol. 1985, 85, 57–69. [Google Scholar] [CrossRef]
  8. van Oirschot, J.T. Bovine herpesvirus 1 in semen of bulls and the risk of transmission: A brief review. Vet. Q. 1995, 17, 29–33. [Google Scholar] [CrossRef] [Green Version]
  9. D’Arce, R.C.; Almeida, R.S.; Silva, T.C.; Franco, A.C.; Spilki, F.; Roehe, P.M.; Arns, C.W. Restriction endonuclease and monoclonal antibody analysis of Brazilian isolates of bovine herpesviruses types 1 and 5. Vet. Microbiol. 2002, 88, 315–324. [Google Scholar] [CrossRef] [PubMed]
  10. Taylor, R.E.; Seal, B.S.; St Jeor, S. Isolation of infectious bovine rhinotracheitis virus from the soft-shelled tick, Ornithodoros coriaceus. Science 1982, 216, 300–301. [Google Scholar] [CrossRef] [PubMed]
  11. Mars, M.H.; de Jong, M.C.; van Maanen, C.; Hage, J.J.; van Oirschot, J.T. Airborne transmission of bovine herpesvirus 1 infections in calves under field conditions. Vet. Microbiol. 2000, 76, 1–13. [Google Scholar] [CrossRef] [PubMed]
  12. Kupferschmied, H.U.; Kihm, U.; Bachmann, P.; Muller, K.H.; Ackermann, M. Transmission of IBR/IPV virus in bovine semen: A case report. Theriogenology 1986, 25, 439–443. [Google Scholar] [CrossRef]
  13. Jones, C.; Chowdhury, S. A review of the biology of bovine herpesvirus type 1 (BHV-1), its role as a cofactor in the bovine respiratory disease complex, and development of improved vaccines. Adv. Anim. Health 2007, 8, 187–205. [Google Scholar] [CrossRef] [PubMed]
  14. Jones, C.; Chowdhury, S. Bovine herpesvirus type 1 (BHV-1) is an important cofactor in the bovine respiratory disease complex. In Veterinary Clinics of North America, Food Animal Practice, Bovine Respiratory Disease; Broderson, B., Cooper, V., Eds.; Elsevier: New York, NY, USA, 2010; Volume 26, pp. 303–321. [Google Scholar]
  15. Chase, C.; Fulton, R.W.; O’Toole, D.; Gillette, B.; Daly, R.F.; Perry, G.; Clement, T. Bovine herpesvirus 1 modified live vaccines for cattle reproduction: Balancing protection with undesired effects. Vet. Microbiol. 2017, 206, 69–77. [Google Scholar] [CrossRef] [PubMed]
  16. O’Toole, D.; Miller, M.M.; Cavender, J.; Cornish, T. Pathology in Practice. Vet. Med. Today 2012, 241, 189–191. [Google Scholar] [CrossRef] [PubMed]
  17. O’Toole, D.; Van Campen, H. Abortifacient vaccines and bovine herpesvirus-1. J. Am. Vet. Med. Assoc. 2010, 237, 259–260. [Google Scholar]
  18. Miller, J.; Van der Maaten, M.J. Early embryonic death in heifers after inoculation with bovine herpesvirus-1 and reactivation of latency virus in reproductive tissues. Am. J. Vet. Res. 1987, 48, 1555–1558. [Google Scholar]
  19. Perry, G.; Zimmerman, A.D.; Daly, R.F.; Butterbaugh, R.E.; Rhoades, J.; Schultz, D.; Harmon, A.; Chase, C. The effects of vaccination on serum hormone concentrations and conception rates in synchronized naive beef heifers. Theriongenology 2013, 79, 200–205. [Google Scholar] [CrossRef]
  20. Johnson, K.; Pendell, D.L. Market impacts of reducing the prevalence of bovine respiratory disease in United States beef cattle feedlots. Front. Vet. Sci. 2017, 4, 189. [Google Scholar] [CrossRef] [Green Version]
  21. Edwards, A.J. Respiratory diseases of feedlot cattle in teh central USA. Bov. Pract. 1996, 30, 5–7. [Google Scholar]
  22. Griffin, D. Economic impact associated with respiratory disease in beef cattle. Vet. Clin. N. Am. Food Anim. Pract. 1997, 13, 367–377. [Google Scholar] [CrossRef]
  23. Kapil, S.; Basaraba, R.J. Infectious bovine rhinotracheitis, parainfluenza-3, and respiratory coronavirus. Bovine respiratory disease update. Vet. Clin. N. Am. Food Anim. Pract. 1997, 13, 455–461. [Google Scholar] [CrossRef] [PubMed]
  24. NASS. Agricultural Statistics Board; U.S. Department of Agriculture: Washington, DC, USA, 1996.
  25. Songer, J.G.; Post, K.W. (Eds.) The Genera Mannheimia and Pasteurella; Elsevier Saunders: St. Louis, MO, USA, 2005. [Google Scholar]
  26. Frank, G. Bacteria as etiologic agents in bovine respiratory disease. In Proceedings of Bovine Respiratory Disease: A Symposium; Texas A&M University Press: College Station, TX, USA, 2010; Volume 2, pp. 381–394. [Google Scholar]
  27. Rice, J.A.; Carrasco-Medina, L.; Hodgins, D.; Shewen, P. Mannheimia haemolytica and bovine respiratory disease. Anim. Health Res. Rev. 2008, 8, 117–128. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Highlander, S.K.; Fedorova, N.D.; Dusek, D.M.; Panciera, R.; Alvarez, L.E.; Renehart, C. Inactivation of Pasteurella (Mannheimia) haemolytica leukotoxin causes partial attenuation of virulence in a calf challenge model. Infect. Immun. 2000, 68, 3916–3922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Highlander, S.K. Molecular genetic analysis of virulence in Mannheimia (Pasteurella) haemolytica. Front. Biosci. 2001, D1128–D1150. [Google Scholar] [CrossRef] [Green Version]
  30. Zecchinon, L.; Fett, T.; Desmecht, D. How Mannheimia haemolytica defeats host defense through a kiss of death mechanism. Vet. Res. 2005, 36, 133–156. [Google Scholar] [CrossRef] [Green Version]
  31. Hodgins, D.C.; Shewen, P.E. (Eds.) Pneumonic Pasteurellosis of Cattle; Oxford University Press: Cape Town, South Africa, 2004; Volume 3, pp. 1677–1684. [Google Scholar]
  32. Hodgson, P.D.; Aich, P.; Manuja, A.; Hokamp, K.; Roche, F.M.; Brinkman, F.S.L.; Potter, A.; Babiuk, L.A.; Griebel, P.J. Effect of stress on viral-bacterial synergy in bovine respiratoryt disease: Novel mechanisms to regulate inflammation. Comp. Funct. Genom. 2005, 6, 244–250. [Google Scholar] [CrossRef] [Green Version]
  33. Yates, W.D.; Babiuk, L.A.; Jericho, K.W. Viral-bacterial pneumonia in calves: Duration of the interaction between bovine herpesvirus 1 and Pasteurella haemolytica. Can. J. Comp. Med. 1983, 47, 257–264. [Google Scholar]
  34. Rivera-Rivas, J.J.; Kisiela, D.; Czuprynski, C.J. Bovine herpesvirus type 1 infection of bovine bronchial epithelial cells increases neutrophil adhesion and activation. Vet. Immunol. Immunopathol. 2009, 131, 167–176. [Google Scholar] [CrossRef]
  35. Leite, F.; Kuckleburg, C.; Atapattu, D.; Schulz, R.; Czuprynski, C.J. BHV-1 infection and inflammatory cytokines amplify the interaction between Mannheimia haemolytica lukotoxin with bovine peripheral blood mononuclear cells in vitro. Vet. Immunol. Immunopathol. 2004, 99, 193–202. [Google Scholar] [CrossRef]
  36. Carter, J.J.; Weinberg, A.D.; Pollard, A.; Reeves, R.; Magnuson, J.A.; Magnuson, N.S. Inhibition of T-lymphocyte mitogenic responses and effects on cell functions by bovine herpesvirus 1. J. Virol. 1989, 63, 1525–1530. [Google Scholar] [CrossRef] [Green Version]
  37. Griebel, P.; Ohmann, H.B.; Lawman, M.J.; Babiuk, L.A. The interaction between bovine herpesvirus type 1 and activated bovine T lymphocytes. J. Gen. Virol. 1990, 71, 369–377. [Google Scholar] [CrossRef] [PubMed]
  38. Griebel, P.; Qualtiere, L.; Davis, W.C.; Gee, A.; Ohmann, H.B.; Lawman, M.J.; Babiuk, L.A. T lymphocyte population dynamics and function following a primary bovine herpesvirus type-1 infection. Viral Immunol. 1987, 1, 287–304. [Google Scholar] [CrossRef] [PubMed]
  39. Griebel, P.J.; Qualtiere, L.; Davis, W.C.; Lawman, M.J.; Babiuk, L.A. Bovine peripheral blood leukocyte subpopulation dynamics following a primary bovine herpesvirus-1 infection. Viral Immunol. 1987, 1, 267–286. [Google Scholar] [CrossRef] [PubMed]
  40. Koppers-Lalic, E.A.; Reits, E.A.J.; Ressing, M.E.; Lipinska, A.D.; Abele, R.; Koch, J.; Rezende, M.M.; Admiraal, P.; van Leeuwen, D.; Bienkowsaka-Szewczyc, K.; et al. Varicelloviruses avoid T cell recognition by UL49.5-mediated inactivation of the transporter associated with antigen processing. Proc. Natl. Acad. Sci. USA 2005, 102, 5144–5149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Nataraj, C.; Eidmann, S.; Hariharan, M.J.; Sur, J.H.; Perry, G.A.; Srikumaran, S. Bovine herpesvirus 1 downregulates the expression of bovine MHC class I molecules. Viral Immunol. 1997, 10, 21–34. [Google Scholar] [CrossRef]
  42. Hariharan, M.J.; Nataraj, C.; Srikumaran, S. Down regulation of murine MHC class I expression by bovine herpesvirus 1. Viral Immunol. 1993, 6, 273–284. [Google Scholar] [CrossRef]
  43. Hinkley, S.; Hill, A.B.; Srikumaran, S. Bovine herpesvirus-1 infection affects the peptide transport activity in bovine cells. Virus Res. 1998, 53, 91–96. [Google Scholar] [CrossRef]
  44. Neibergs, H.L.; Seabury, C.M.; Wojtowicz, A.J.; Wang, Z.; Scraggs, E.; Kiser, J.N.; Neupane, M.; Womack, J.E.; Van Eenennaam, A.; Hagevortm, G.R.; et al. Susceptibility loci revealed for bovine respiratory disease complex in pre-weaned holstein calves. BMC Genom. 2014, 15, 1164. [Google Scholar] [CrossRef] [Green Version]
  45. Wirth, U.; Vogt, B.; Schwyzer, M. The three major immediate-early transcripts of bovine herpesvirus 1 arise from two divergent and spliced transcription units. J. Virol. 1991, 65, 195–205. [Google Scholar] [CrossRef] [Green Version]
  46. Wirth, U.V.; Fraefel, C.; Vogt, B.; Vlcek, C.; Paces, V.; Schwyzer, M. Immediate-early RNA 2.9 and early RNA 2.6 of bovine herpesvirus 1 are 3’ coterminal and encode a putative zinc finger transactivator protein. J. Virol. 1992, 66, 2763–2772. [Google Scholar] [CrossRef] [Green Version]
  47. Fraefel, C.; Zeng, J.; Choffat, Y.; Engels, M.; Schwyzer, M.; Ackermann, M. Identification and zinc dependence of the bovine herpesvirus 1 transactivator protein BICP0. J. Virol. 1994, 68, 3154–3162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Misra, V.; Bratanich, A.C.; Carpenter, D.; O’Hare, P. Protein and DNA elements involved in transactivation of the promoter of the bovine herpesvirus (BHV) 1 IE-1 transcription unit by the BHV alpha gene trans-inducing factor. J. Virol. 1994, 68, 4898–4909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Misra, V.; Walker, S.; Hayes, S.; O’Hare, P. The bovine herpesvirus alpha gene trans-inducing factor activates transcription by mechanisms different from those of its herpes simplex virus type 1 counterpart VP16. J. Virol. 1995, 69, 5209–5216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Wirth, U.; Gunkel, K.; Engels, M.; Schwyzer, M. Spatial and temporal distribution of bovine herpesvirus 1 transcripts. J. Virol. 1989, 63, 4882–4889. [Google Scholar] [CrossRef] [Green Version]
  51. Jones, C. Alphaherpesvirus Latency: Its Role in Disease and Survival of the Virus in Nature. In Advances in Virus Research; Maramorosch, K., Murphy, F.A., Shatkin, A.J., Eds.; Academic Press: Cambridge, MA, USA, 1998; Volume 51, pp. 81–133. [Google Scholar]
  52. Jones, C. Herpes Simplex Virus Type 1 and Bovine Herpesvirus 1 Latency. Clin. Microbiol. Rev. 2003, 16, 79. [Google Scholar] [CrossRef] [Green Version]
  53. Jones, C.; Geiser, V.; Henderson, G.; Jiang, Y.; Meyer, F.; Perez, S.; Zhang, Y. Functional analysis of bovine herpesvirus 1 (BHV-1) genes expressed during latency. Vet. Micro 2006, 113, 199–210. [Google Scholar] [CrossRef]
  54. Jones, C. Regulation of innate immune responses by bovine herpesvirus 1 and infected cell protein 0. Viruses 2009, 1, 255–275. [Google Scholar] [CrossRef] [Green Version]
  55. Schang, L.; Jones, C. Analysis of bovine herpesvirus 1 transcripts during a primary infection of trigeminal ganglia of cattle. J. Virol. 1997, 71, 6786–6795. [Google Scholar] [CrossRef] [Green Version]
  56. Jones, C. Latency of Bovine Herpesvirus 1 (BoHV-1) in Sensory Neurons. In Herpesviridae; InTech: Budapest, Hungary, 2016. [Google Scholar] [CrossRef] [Green Version]
  57. Inman, M.; Lovato, L.; Doster, A.; Jones, C. A Mutation in the Latency-Related Gene of Bovine Herpesvirus 1 Disrupts the Latency Reactivation Cycle in Calves. J. Virol. 2002, 76, 6771. [Google Scholar] [CrossRef] [Green Version]
  58. Winkler, M.T.C.; Doster, A.; Jones, C. Persistence and reactivation of bovine herpesvirus-1 (BHV-1) in tonsils of latently infected cattle. Persistence and reactivation of bovine herpesvirus 1 in the tonsil of latently infected calves. J. Virol. 2000, 74, 5337–5346. [Google Scholar] [CrossRef]
  59. Mweene, A.S.; Okazaki, K.; Kida, H. Detection of viral genome in non-neural tissues of cattle experimentally infected with bovine herpesvirus 1. Jpn. J. Res. 1996, 44, 165–174. [Google Scholar]
  60. Perez, S.; Inman, M.; Doster, A.; Jones, C. Latency-related gene encoded by bovine herpesvirus 1 promotes virus growth and reactivation from latency in tonsils of infected calves. J. Clin. Micro 2005, 43, 393–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Winkler, M.T.; Doster, A.; Jones, C. Bovine herpesvirus 1 can infect CD4(+) T lymphocytes and induce programmed cell death during acute infection of cattle. J. Virol. 1999, 73, 8657–8668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Cheung, A.K. Investigation of pseudorabies virus DNA and RNA in trigeminal ganglia and tonsil tissues of latently infected swine. Am. J. Vet. Res. 1995, 56, 45–50. [Google Scholar] [PubMed]
  63. Sabo, A.; Rajanci, A. Latent pseudorabies virus infection in pigs. Acta Virol. 1976, 20, 208–214. [Google Scholar]
  64. Borchers, K.; Wolfinger, U.; Ludwig, H. Latency-associated transcript of equine herpesvirus 4 in trigeminal ganglia of naturally infected horses. J. Gen. Virol. 1999, 80, 2165–2171. [Google Scholar] [CrossRef] [Green Version]
  65. Miyoshi, M.; Ishii, Y.; Takiguchi, M.; Takada, A.; Yasuda, J.; Hashimoto, A.; Okasaki, K.; Kida, H. Detection of canine herpesvirus DNA in ganglionic neurons and the lymph node lymphocytes of latently infected dogs. J. Vet. Med. Sci. 1999, 61, 375–379. [Google Scholar] [CrossRef] [Green Version]
  66. Sahin, F.; Gerceker, D.; Karasartova, D.; Ozsan, T.M. Detection of herpes simplex virus type 1 in addition to Epstein–Bar virus in tonsils using a new multiplex polymerase chain reaction assay. Diagn. Microbiol. Infect. Dis. 2007, 57, 47–51. [Google Scholar] [CrossRef]
  67. Kutish, G.; Mainprize, T.; Rock, D. Characterization of the latency-related transcriptionally active region of the bovine herpesvirus 1 genome. J. Virol. 1990, 64, 5730–5737. [Google Scholar] [CrossRef] [Green Version]
  68. Rock, D.; Lokensgard, J.; Lewis, T.; Kutish, G. Characterization of dexamethasone-induced reactivation of latent bovine herpesvirus 1. J. Virol. 1992, 66, 2484–2490. [Google Scholar] [CrossRef] [Green Version]
  69. Rock, D.L.; Beam, S.L.; Mayfield, J.E. Mapping bovine herpesvirus type 1 latency-related RNA in trigeminal ganglia of latently infected rabbits. J. Virol. 1987, 61, 3827–3831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Rock, D.L.; Nesburn, A.B.; Ghiasi, H.; Ong, J.; Lewis, T.L.; Lokensgard, J.R.; Wechsler, S.L. Detection of latency-related viral RNAs in trigeminal ganglia of rabbits latently infected with herpes simplex virus type 1. J. Virol. 1987, 61, 3820–3826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Jones, C.; da Silva, L.F.; Sinani, D. Regulation of the latency-reactivation cycle by products encoded by the bovine herpesvirus 1 (BHV-1) latency-related gene. J. Neurovirol. 2011, 17, 535–545. [Google Scholar] [CrossRef]
  72. Devireddy, L.R.; Jones, C. Alternative splicing of the latency-related transcript of bovine herpesvirus 1 yields RNAs containing unique open reading frames. J. Virol. 1998, 72, 7294–7301. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Hossain, A.; Schang, L.M.; Jones, C. Identification of gene products encoded by the latency-related gene of bovine herpesvirus 1. J. Virol. 1995, 69, 5345–5352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Bratanich, A.C.; Hanson, N.D.; Jones, A.C. The latency-related gene of bovine herpesvirus 1 inhibits the activity of immediate-early transcription unit 1. Virology 1992, 191, 988–991. [Google Scholar] [CrossRef] [PubMed]
  75. Bratanich, A.C.; Jones, C.J. Localization of cis-acting sequences in the latency-related promoter of bovine herpesvirus 1 which are regulated by neuronal cell type factors and immediate-early genes. J. Virol. 1992, 66, 6099–6106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Jones, C.; Delhon, G.; Bratanich, A.; Kutish, G.; Rock, D. Analysis of the transcriptional promoter which regulates the latency- related transcript of bovine herpesvirus 1. J. Virol. 1990, 64, 1164–1170. [Google Scholar] [CrossRef] [Green Version]
  77. Liu, Y.; Jones, C. Regulation of Notch-mediated transcription by a bovine herpesvirus 1 encoded protein (ORF2) that is expressed in latently infected sensory neurons. J. Neurovirol. 2016, 22, 518–528. [Google Scholar] [CrossRef]
  78. Devireddy, L.; Zhang, Y.; Jones, C. Cloning and initial characterization of an alternatively spliced transcript encoded by the bovine herpes virus 1 latency related (LR) gene. J. Neurovirol. 2003, 9, 612–622. [Google Scholar] [CrossRef]
  79. Shen, W.; Jones, C. Open reading frame 2, encoded by the latency-related gene of bovine herpesvirus 1, has antiapoptotic activity in transiently transfected neuroblastoma cells. J. Virol. 2008, 82, 10940–10945. [Google Scholar] [CrossRef] [Green Version]
  80. Workman, A.; Sinani, D.; Pittayakhajonwut, D.; Jones, C. A Protein (ORF2) Encoded by the Latency Related Gene of Bovine Herpesvirus 1 Interacts with Notch1 and Notch3. J. Virol. 2011, 85, 2536–2546. [Google Scholar] [CrossRef] [Green Version]
  81. Workman, A.L.; Zhu, L.; Keel, B.N.; Smith, T.P.L.; Jones, C. The Wnt signaling pathway is differentially expressed during the bovine herpesvirus 1 latency-reactivation cycle: Evidence that two proteinkinases associated with neuronal survival, Akt3 and BMPR2, are expressed at higher levels during latency. J. Virol. 2018, 92, e01937-17. [Google Scholar] [CrossRef] [Green Version]
  82. Zhu, L.; Workman, A.; Jones, C. A potential role for a beta-catenin coactivator (high mobility group AT-hook 1 protein) during the latency-reactivation cycle of bovine herpesvirus 1. J. Virol. 2017, 91, e02132-16. [Google Scholar] [CrossRef] [Green Version]
  83. Meyer, F.; Jones, C. The cellular transcription factor, CCAAT enhancer-binding protein alpa (C/EBP-a) has the potential to activate the bovine herpesvirus 1 immediate early transcription unit 1 promoter. J. Neurovirol. 2009, 15, 1–8. [Google Scholar] [CrossRef]
  84. Meyer, F.; Perez, S.; Geiser, V.; Sintek, M.; Inman, M.; Jones, C. A protein encoded by the bovine herpes virus 1 (BHV-1) latency related gene interacts with specific cellular regulatory proteins, including the CCAAT enhancer binding protein alpha (C/EBP-a). J. Virol. 2007, 81, 59–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Workman, A.; Perez, S.; Doster, A.; Jones, C. Dexamethasone treatment of calves latently infected with bovine herpesvirus 1 (BHV-1) leads to activation of the bICP0 early promoter, in part by the cellular transcription factor C/EBP-alpha. J. Virol. 2009, 83, 8800–8809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Sinani, D.; Jones, C. Localization of sequences in a protein (ORF2) encoded by the latency-related gene of bovine herpesvirus 1 that inhibits apoptosis and interferes with Notch1-mediated trans-activation of the bICP0 promoter. J. Virol. 2011, 85, 12124–12133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Sinani, D.; Frizzo da Silva, L.; Jones, C. A bovine herpesvirus 1 protein expressed in latently infected neurons (ORF2) promotes neurite sprouting in the presence of activated Notch1 or Notch3. J. Virol. 2013, 87, 1183–1192. [Google Scholar] [CrossRef] [Green Version]
  88. Jaber, T.; Workman, A.; Jones, C. Small noncoding RNAs encoded within the bovine herpesvirus 1 latency-related gene can reduce steady-state levels of infected cell protein 0 (bICP0). J. Virol. 2010, 84, 6297–6307. [Google Scholar] [CrossRef] [Green Version]
  89. Jiang, Y.; Inman, M.; Zhang, Y.; Posadas, N.A.; Jones, C. A mutation in the latency related gene of bovine herpesvirus 1 (BHV-1) inhibits protein expression of a protein from open reading frame 2 (ORF-2) and an adjacent reading frame during productive infection. J. Virol. 2004, 78, 3184–3189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Meyer, F.; Perez, S.; Jiang, Y.; Zhou, Y.; Henderson, G.; Jones, C. Identification of a novel protein encoded by the latency-related gene of bovine herpesvirus 1. J. Neurovirol. 2007, 13, 569–578. [Google Scholar] [CrossRef] [PubMed]
  91. Lovato, L.; Inman, M.; Henderson, G.; Doster, A.; Jones, C. Infection of cattle with a bovine herpesvirus 1 (BHV-1) strain that contains a mutation in the latency related gene leads to increased apoptosis in trigeminal ganglia during the transition from acute infection to latency. J. Virol. 2003, 77, 4848–4857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Perng, G.-C.; Maguen, B.; Jin, L.; Mott, K.R.; Osorio, N.; Slanina, S.M.; Yukht, A.; Ghiasi, H.; Nesburn, A.B.; Inman, M.; et al. Wechsler. A gene capable of blocking apoptosis can substitute for the herpes simplex virus type 1 latency-associated transcript gene and restore wild-type reactivation levels. J. Virol. 2002, 76, 1224–1235. [Google Scholar] [CrossRef] [Green Version]
  93. Mott, K.R.; Osorio, N.; Jin, L.; Brick, D.J.; Naito, J.; Cooper, J.; Henderson, G.; Inman, M.; Jones, C.; Wechsler, S.L.; et al. The bovine herpesvirus-1 LR ORF2 is critical for this gene’s ability to restore the high wild-type reactivation phenotype to a herpes simplex virus-1 LAT null mutant. J. Gen. Virol. 2003, 84, 2975–2985. [Google Scholar] [CrossRef]
  94. Barsam, C.A.; Brick, D.J.; Jones, C.; Wechsler, S.L.; Perng, G.C. A viral model for corneal scarring and neovascularization following ocular infection of rabbits with a herpes simplex virus type 1 (HSV-1) mutant. Cornea 2005, 24, 460–466. [Google Scholar] [CrossRef]
  95. Jones, C.; Newby, T.J.; Holt, T.; Doster, A.; Stone, M.; Ciacci-Zanella, J.; Webster, C.J.; Jackwood, M.W. Analysis of latency in cattle after inoculation with a temperature sensitive mutant of bovine herpesvirus 1 (RLB106). Vaccine 2000, 18, 3185–3195. [Google Scholar] [CrossRef]
  96. Workman, A.; Eudy, J.; Smith, L.; da Silva, L.F.; Sinani, D.; Bricker, H.; Cook, E.; Doster, A.; Jones, C. Cellular transcription factors induced in trigeminal ganglia during dexamethasone-induced reactivation from latency stimulate bovine herpesvirus 1 productive infection and certain viral promoters. J. Virol. 2012, 86, 2459–2473. [Google Scholar] [CrossRef] [Green Version]
  97. Brown, G.; Field, H. Experimental reactivation of bovine herpesvirus 1 (BHV-1) by means of corticosteroids in an intranasal rabbit model. Arch. Virol. 1990, 112, 81–101. [Google Scholar] [CrossRef]
  98. Ackermann, M.; Peterhans, E.; Wyler, R. DNA of bovine herpesvirus type 1 in the trigeminal ganglia of latently infected calves. Am. J. Vet. Res. 1982, 43, 36–40. [Google Scholar]
  99. Hage, J.; Glas, R.; Westra, H.; Maris-Veldhuis, M.; Van Oirschot, J.; Rijsewijk, F. Reactivation of latent bovine herpesvirus 1 in cattle seronegative to glycoproteins gB and gE. Vet. Microbiol. 1998, 60, 87–98. [Google Scholar] [CrossRef] [PubMed]
  100. Homan, E.; Easterday, B. Experimental latent and recrudescent bovine herpesvirus-1 infections in calves. Am. J. Vet. Res. 1983, 44, 309–313. [Google Scholar] [PubMed]
  101. Sheffy, B.E.; Davies, D.H. Reactivation of a Bovine Herpesvirus After Corticosteroid Treatment. Proc. Soc. Exp. Biol. Med. 1972, 140, 974–976. [Google Scholar] [CrossRef] [PubMed]
  102. Smoak, K.L.J.; Cidlowski, A. Mechanisms of glucocorticoid receptor signaling during inflammation. Mech. Aging Dev. 2004, 125, 697–706. [Google Scholar] [CrossRef] [PubMed]
  103. Oakley, R.H.; Cidlowski, J.A. The biology of the glucocorticoid receptor: New signaling mechanisms in health and disease. J. Allergy Clin. Immunol. 2013, 132, 1033–1044. [Google Scholar] [CrossRef] [Green Version]
  104. Kook, I.; Henley, C.; Meyer, F.; Hoffmann, F.G.; Jones, C. Bovine herpesvirus 1 productive infection and immediate early transcription unit 1 promoter are stimulated by the synthetic corticosteroid dexamethasone. Virology 2015, 484, 377–385. [Google Scholar] [CrossRef] [Green Version]
  105. El-mayet, F.; Sawant, L.; Thunuguntla, P.; Zhao, J.; Jones, C. Two pioneer transcription factors, Krüppel-like transcription factor 4 and glucocorticoid receptor, cooperatively transactivate the bovine herpesvirus 1 ICP0 early promoter and stimulate productive infection. J. Virol. 2020, 94, e01670-19. [Google Scholar] [CrossRef]
  106. Pearce, L.R.; Komander, D.; Alessi, D.R. The nuts and bolts of AGC protein kinases. Nat. Rev. Mol. Cell Biol. 2010, 11, 9–22. [Google Scholar] [CrossRef]
  107. Kook, I.; Jones, C. The serum and glucocorticoid-regulated protein kinases (SGK) stimulate bovine herpesvirus 1 and herpes simplex virus 1 productive infection. Virus Res. 2016, 222, 106–112. [Google Scholar] [CrossRef] [Green Version]
  108. Frizzo da Silva, L.I.K.; Doster, A.; Jones, C. Bovine herpesvirus 1 regulatory proteins, bICP0 and VP16, are readily detected in trigeminal ganglionic neurons expressing the glucocorticoid receptor during the early stages of reactivation from latency. J. Virol. 2013, 87, 11214–11222. [Google Scholar] [CrossRef] [Green Version]
  109. Kook, I.; Doster, A.; Jones, C. Bovine herpesvirus 1 regulatory proteins are detected in trigeminal ganglionic neurons during the early stages of stress-induced escape from latency. J. Neurovirol. 2015, 21, 585–591. [Google Scholar] [CrossRef] [PubMed]
  110. Guo, J.; Li, Q.; Jones, C. The bovine herpesvirus 1 regulatory proteins, bICP4 and bICP22, are expressed during the escape from latency. J. Neuovirol. 2019, 25, 42–49. [Google Scholar] [CrossRef] [PubMed]
  111. Toomer, G.; Workman, A.; Harrison, K.S.; Stayton, E.; Hoyt, P.R.; Jones, C. Stress Triggers Expression of Bovine Herpesvirus 1 Infected Cell Protein 4 (bICP4) RNA during Early Stages of Reactivation from Latency in Pharyngeal Tonsil. J. Virol. 2022, in press. [CrossRef]
  112. Liu, T.; Liu, X.; Wang, H.; Moon, R.T.; Malbon, C.C. Activation of a rat frizzled-1 promotes Wnt signaling and differentiation of mouse F9 teratocarcinoma cells via pathways that require Galpha(q) and Galpha(o) function. J. Biol. Chem. 1999, 274, 33539–33544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Liu, X.; Robbins, J.S.; Kimmel, A.R. Rapid, Wnt-induced changes in GSK3beta association that regulate beta-catenin stabilization are mediated by Galpha proteins. Curr. Biol. 2005, 15, 1989–1997. [Google Scholar] [CrossRef] [Green Version]
  114. Guha, U.; Gomes, W.; Samantra, J.; Guptam, M.; Rice, F.; Kessler, J. Target-derived BMP signaling limits sensory neuron number and the extent of peripheral innervation in vivo. Development 2004, 131, 1175–1186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Hodge, L.; Klassen, M.; Han, B.; Yiu, G.; Hurrel, J.; Howell, A.; Rosseau, G.; Lemaigre, F.; Tressier-Lavigne, M.; Wang, F. Retrograde BMP signaling regulates trigeminal sensory neuron identitities and the formation of precise face maps. Neuron 2007, 55, 572–586. [Google Scholar] [CrossRef] [Green Version]
  116. Farkas, L.; Jasai, J.; Unsicker, K.; Krieglstein, K. Characterization of bone morphogenetic protein family members as neurotropic factors for cultured sensory neurons. Neuroscience 1999, 92, 227–235. [Google Scholar] [CrossRef]
  117. Diez, H.; Garrido, J.J.; Wandosell, F. Specific Roles of Akt iso forms in apoptosis and axon growth regulation in nuerons. PLoS ONE 2012, 74, e32715. [Google Scholar]
  118. Xie, R.; Cheng, M.; Li, M.; Xiong, X.; Daadi, M.; Sapolsky, R.; Zhao, H. Akt isoforms differentially protect against stroke-induced neuronal injury by regulating mTOR activities. J. Cereb. Blood Flow Metab. 2013, 33, 1875–1885. [Google Scholar] [CrossRef] [Green Version]
  119. Clevers, H.; Nusse, R. Wnt/Β−catenin signaling and disease. Cell 2012, 149, 1192–1205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Salinas, P.C. Wnt signaling in the vertebrate central nervous system: From axon guidance to synaptic function. Cold Spring Harb. Perpect. Biol. 2012, 4, a008003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Nusse, R.; Clever, H. Wnt/beta-catenin signaling, diseases, and emerging therapeutic modalities. Cell 2017, 169, 985–999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Lambert, C.; Cisternas, P.; Inestrosa, N.C. Role of Wnt signaling in central nervous system injury. Mol. Neurobiol. 2016, 53, 2297–2311. [Google Scholar] [CrossRef] [PubMed]
  123. Rosso, S.B.; Inestrosa, N.C. Wnt signaling in neuronal maturation and synaptogenesis. Cell. Neurosci. 2013, 7, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Zhao, J.; Zhu, L.; Wijesekera, N.; Jones, C. Specific Akt family members impair stress mediated transactivation of viral promoters and enhance neuronal differentiation: Important functions for maintaining latency. J. Virol. 2020, 94, e00901–e00920. [Google Scholar] [CrossRef]
  125. Liu, Y.; Hancock, M.; Workman, A.; Doster, A.; Jones, C. Beta-catenin, a transcription factor activated by canonical Wnt signaling, is expressed in sensory neurons of calves latently infected with bovine herpesvirus 1. J. Virol. 2016, 90, 3148–3159. [Google Scholar] [CrossRef] [Green Version]
  126. Zhao, J.; Wijesekera, N.; Jones, C. Inhibition of Stress-Induced Viral Promoters by a Bovine Herpesvirus 1 Non-Coding RNA and the Cellular Transcription Factor, beta-Catenin. Int. J. Molec. Sci. 2021, 22, 519. [Google Scholar] [CrossRef]
  127. Zorn, A.M. Wnt signaling: Antagonistic Dicckopf. Curr. Biol. 2001, 11, R592–R595. [Google Scholar] [CrossRef] [Green Version]
  128. Niehrs, C. Function and biological roles of the dickkopf family of Wnt modulators. Oncogene 2006, 25, 7469–7481. [Google Scholar] [CrossRef] [Green Version]
  129. Magana, R.; Giles, N.; Adcroft, K.; Keeney, D.; Wood, F.; Fear, M.; Dharmarajan, A. Secreted frizzled related protein-4 (sFRP4) promotes epidermal differentiation and apoptosis. Biochem. Biophys. Res. Comm. 2008, 377, 606–611. [Google Scholar] [CrossRef] [PubMed]
  130. Bhuvanalakshmi, G.; Arfuso, F.; Millward, M.; Dharmarajan, A.; Warrier, S. Secreted frizzled-related protein 4 inhibits glioma stem-like cells by reversing epithelial to mesenchymal transition, inducing apoptosis and decreasing cancer stem cell properties. PLoS ONE 2015, 10, e0127517. [Google Scholar]
  131. Lacher, M.; Siegenthaler, A.; Jager, R.; Yan, X.; Hett, S.; Xuan, L.; Saurer, S.; Lareu, R.; Dharmarajan, M.; Friis, R. Role of DDC-4/sFRP-4, a secreted frizzled-related protein, as the onser of apoptosis in mammary involution. Cell Death Differ. 2003, 10, 528–538. [Google Scholar] [CrossRef] [PubMed]
  132. Kormish, J.; Sinner, D.; Zorn, A.M. Interactions between SOX factors and Wnt/beta-catenin signaling in development and disease. Dev. Dyn. 2010, 239, 56–68. [Google Scholar] [PubMed] [Green Version]
  133. Prasad, A.; Remick, J.; Zeichner, S.L. Activation of human herpesvirus replication by apoptosis. J. Virol. 2013, 87, 10641–10650. [Google Scholar] [CrossRef] [Green Version]
  134. Du, T.; Zhou, G.; Roizman, B. Induction of apoptosis accelerates reactivation from latent HSV-1 in ganglionic organ cultures and replication in cell cultures. Proc. Natl. Acad. Sci. USA 2012, 109, 14616–14621. [Google Scholar] [CrossRef] [Green Version]
  135. El-mayet, F.; El-Habbaa, A.S.; El-Bagoury, G.F.; Sharawi, S.S.A.; El-Nahas, E.M.; Jones, C. Transcription Factors Have the Potential to Synergistically Stimulate Bovine Herpesvirus 1 Transcription and Reactivation from Latency. In Transcriptional Regulation; Kais, G., Ed.; INTECH: Rejeka, Croatia, 2018; Volume 1, pp. 36–54. [Google Scholar]
  136. El-mayet, F.S.; El-Habbaa, A.; D’Offay, J.; Jones, C. Synergistic activation of bovine herpesvirus 1 productive infection and viral regulatory promoters by the progesterone receptor and Krüppel-like transcription factor 15. J. Virol. 2019, 93, e01519-18. [Google Scholar] [CrossRef] [Green Version]
  137. Ostler, J.B.; Sawant, L.; Harrison, K.S.; Jones, C. Regulation of neurotropic herpesvirus productive infection and latency-reactivation cycle by glucocorticoid receptor and stress-induced transcription factors. In Hormones, Regulators, and Viruses; Litwack, G., Ed.; Elsiever: Amsterdam, The Netherlands, 2021; Volume 117. [Google Scholar]
  138. Sawant, L.; Ostler, J.B.; Jones, C. A Pioneer Transcription Factor and Type I Nuclear Hormone Receptors Synergistically Activate the Bovine Herpesvirus 1 Infected Cell Protein 0 (ICP0) Early Promoter. J. Virol. 2021, 95, e00768-21. [Google Scholar] [CrossRef]
  139. El-Mayet, F.S.; Sawant, L.; Thungunutla, P.; Jones, C. Combinatorial effects of the glucocorticoid receptor and Krüppel-like transcription factor 15 on bovine herpesvirus 1 transcription and productive infection. J. Virol. 2017, 91, e00904-17. [Google Scholar] [CrossRef] [Green Version]
  140. Kaczynski, J.; Cook, T.; Urrutia, R. Sp1- and Kruppel-like transcription factors. Genome Biol. 2003, 4, 206.201–206.208. [Google Scholar] [CrossRef] [Green Version]
  141. Bieker, J.J. Krüppel-like factors: Three fingers in many pies. J. Biol. Chem. 2001, 276, 34355–34358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Polman, J.A.E.; Welten, J.E.; Bosch, D.S.; de Jonge, R.T.; Balog, J.; van der Maarel, S.M.; de Kloet, E.R.; Datson, N.A. A genome-wide signature of glucocorticoid receptor binding in neuronal PC12 cells. BMC Neurosci. 2012, 13, 118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Workman, A.; Jones, C. Bovine herpesvirus 1 (BoHV1) productive infection and bICP0 early promoter activity are stimulated by E2F1. J. Virol. 2010, 84, 6308–6317. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Hapgood, J.P.; Avenant, C.; Moliki, J.M. Glucocorticoid-independent modulation of GR activity: Implications for immunotherapy. Pharm. Ther. 2016, 165, 93–113. [Google Scholar] [CrossRef] [Green Version]
  145. Scheschowitsch, K.; Leite, J.A.; Assreuy, J. New Insights in Glucocorticoid Receptor Signaling-More Than Just a Ligand-Binding Receptor. Front. Endocrinol. 2017, 8, 16. [Google Scholar] [CrossRef] [Green Version]
  146. Biddie, S.C.; Hager, G.L. Glucocorticoid receptor dynamics and gene regulation. Stress 2009, 12, 193–205. [Google Scholar] [CrossRef]
  147. Sawant, L.; Thunuguntla, P.; Jones, C. Cooperative activation of bovine herpesvirus 1 productive infection and viral regulatory promoters by androgen receptor and Krüppel-like transcription factors 4 and 15. Virology 2021, 552, 63–72. [Google Scholar] [CrossRef]
  148. Sawant, L.; Wijesekera, N.; Jones, C. Pioneer transcription factors, progesterone receptor and Krüppel like transcription factor 4, cooperatively stimulate the bovine herpesvirus 1 ICP0 early promoter and productive late protein expression. Virus Res. 2020, 288, 198115. [Google Scholar] [CrossRef]
  149. Azeez, J.M.; Susmi, T.R.; Remadevi, V.; Ravindran, V.; Sasikumar Sujatha, A.; Ayswarya, R.N.S.; Sreeja, S. New insights into the functions of progesterone receptor (PR) isoforms and progesterone signaling. Am. J. Cancer Res. 2021, 11, 5214–5232. [Google Scholar]
  150. Claessens, F.; Joniau, S.; Helsen, C. Comparing the rules of engagement of androgen and glucocorticoid receptors. Cell. Mol. Life Sci. 2017, 74, 2217–2228. [Google Scholar] [CrossRef] [Green Version]
  151. Rundlett, S.E.; Miesfeld, R.L. Quantitative differences in androgen and glucocorticoid receptor DNA binding properties contribute to receptor-selective transcriptional regulation. Mol. Cell. Endocrinol. 1995, 109, 1–10. [Google Scholar] [CrossRef] [PubMed]
  152. Pecci, A.; Ogara, M.F.; Sanz, R.T.; Vicent, G.P. Choosing the right partner in hormone-dependent gene regulation: Glucocorticoid and progesterone receptors crosstalk in breast cancer cells. Front. Endocrinol. 2022, 13, 1037177. [Google Scholar] [CrossRef] [PubMed]
  153. El-Mayet, F.S.; Toomer, G.; Ostler, J.B.; Harrison, K.S.; Santos, V.C.; Wijesekera, N.; Stayton, E.; Ritchey, J.; Jones, C. Progesterone Sporadically Induces Reactivation from Latency in Female Calves but Proficiently Stimulates Bovine Herpesvirus 1 Productive Infection. J. Virol. 2022, 96, e0213021. [Google Scholar] [CrossRef] [PubMed]
  154. El-Mayet, F.S.; Sawant, L.; Wijesekera, N.; Jones, C. Progesterone increases the incidence of bovine herpesvirus 1 reactivation from latency and stimulates productive infection. Virus Res. 2020, 276, 197803. [Google Scholar] [CrossRef]
Figure 2. Summary of latency and DEX-induced reactivation on regulation of the canonical Wnt/β-catenin signaling pathway. (A) Key differentially expressed cellular genes identified in TG neurons during latency are shown in blue. (B) Summary of how DEX inhibits the Wnt/β-catenin signaling pathway in TG neurons, which correlates with reactivation from latency. Key cellular genes differentially expressed are shown in red. This schematic was generated using BioRender.
Figure 2. Summary of latency and DEX-induced reactivation on regulation of the canonical Wnt/β-catenin signaling pathway. (A) Key differentially expressed cellular genes identified in TG neurons during latency are shown in blue. (B) Summary of how DEX inhibits the Wnt/β-catenin signaling pathway in TG neurons, which correlates with reactivation from latency. Key cellular genes differentially expressed are shown in red. This schematic was generated using BioRender.
Viruses 15 00552 g002
Figure 3. Schematic of BoHV-1 IEtu1 and bICP0 E promoters. (A) Key features of the IEtu1 promoter are depicted. The site of transcription initiation is indicated by an arrow immediately downstream of the TATA box and TATGARAT motif. The KLF-like CACCC and two GREs are also indicated. The TATA box is separated from GRE2 by 796 nucleotides. (B) Key features of the bICP0 E promoter are depicted. The site of transcription initiation is indicated by an arrow immediately downstream of the TATA box. The three enhancer domains are indicated: EP-328 (nucleotides 172 to 328), EP-638 (nucleotides 328 to 638), and EP-943 (nucleotides 638 to 943). Relative positions of Sp1, KLF4, and KLF-like sites are indicated. (C) Legend of transcription factor binding sites and consensus nucleotide sequences for binding DNA.
Figure 3. Schematic of BoHV-1 IEtu1 and bICP0 E promoters. (A) Key features of the IEtu1 promoter are depicted. The site of transcription initiation is indicated by an arrow immediately downstream of the TATA box and TATGARAT motif. The KLF-like CACCC and two GREs are also indicated. The TATA box is separated from GRE2 by 796 nucleotides. (B) Key features of the bICP0 E promoter are depicted. The site of transcription initiation is indicated by an arrow immediately downstream of the TATA box. The three enhancer domains are indicated: EP-328 (nucleotides 172 to 328), EP-638 (nucleotides 328 to 638), and EP-943 (nucleotides 638 to 943). Relative positions of Sp1, KLF4, and KLF-like sites are indicated. (C) Legend of transcription factor binding sites and consensus nucleotide sequences for binding DNA.
Viruses 15 00552 g003
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

Ostler, J.B.; Jones, C. The Bovine Herpesvirus 1 Latency-Reactivation Cycle, a Chronic Problem in the Cattle Industry. Viruses 2023, 15, 552. https://doi.org/10.3390/v15020552

AMA Style

Ostler JB, Jones C. The Bovine Herpesvirus 1 Latency-Reactivation Cycle, a Chronic Problem in the Cattle Industry. Viruses. 2023; 15(2):552. https://doi.org/10.3390/v15020552

Chicago/Turabian Style

Ostler, Jeffery B., and Clinton Jones. 2023. "The Bovine Herpesvirus 1 Latency-Reactivation Cycle, a Chronic Problem in the Cattle Industry" Viruses 15, no. 2: 552. https://doi.org/10.3390/v15020552

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