Next Article in Journal
Genetic Therapy of Fuchs Endothelial Corneal Dystrophy: Where Are We? A Review
Previous Article in Journal
Whole-Exome Sequencing of Discordant Monozygotic Twins for Congenital Scoliosis: A Family Case Study
Previous Article in Special Issue
Resistance Mutations in CLL: Genetic Mechanisms Shaping the Future of Targeted Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Transcriptional Activation Mechanisms and Target Genes of the Oncogene Product Tax of Human T-Cell Leukemia Virus Type 1

1
Department of Biomedical Sciences, School of Biological and Environmental Sciences, Kwansei Gakuin University, 1 Gakuen Uegahara, Sanda 669-1330, Hyogo, Japan
2
Department of Obstetrics and Gynecology, University of Colorado School of Medicine, Anschutz Medical Campus, 12700 East 19th Avenue, Aurora, CO 80045, USA
3
Division of Molecular Genetics, Cancer Research Institute, Kanazawa University, Kakuma-machi, Kanazawa 920-1192, Ishikawa, Japan
4
Department of Morphological Biology, Ohu University School of Dentistry, 31-1 Misumido Tomitamachi, Koriyama 963-8611, Fukushima, Japan
*
Author to whom correspondence should be addressed.
Genes 2025, 16(10), 1221; https://doi.org/10.3390/genes16101221
Submission received: 2 September 2025 / Revised: 2 October 2025 / Accepted: 7 October 2025 / Published: 15 October 2025

Abstract

Human T-cell leukemia virus type 1 (HTLV-1) is the causative agent of adult T-cell leukemia/lymphoma (ATL). The trans-activator protein Tax of HTLV-1 is thought to play a crucial role in the early-stage transformation of the virus-infected cells. Tax is a multi-functional protein and modulates cellular signaling pathways that promote proliferation and survival of HTLV-1-infected cells, primarily through the trans-activation of cellular target genes. Tax interacts with a variety of host cell factors including signal transducers and transcription factors, leading to the activation of transcription factors such as CREB, NF-κB, and SRF and activates both its own promoter and those of a variety of host cellular genes. Tax activates its own promoter mainly through CREB and host cellular genes through NF-κB, SRF, and CREB. Accumulating evidence indicates that the Tax-mediated trans-activation of target genes through NF-κB plays an essential role in the transformation of HTLV-1 infected cells. However, the repertoire of Tax target genes, especially those crucial for leukemogenesis, are not known in detail. In this review, we summarize transcriptional activation mechanisms and target genes of Tax, especially focusing on transformation, to facilitate understanding of the underlying mechanisms of leukemogenesis induced by HTLV-1 infection.

1. Introduction

Human T-cell leukemia virus type 1 (HTLV-1) is the first retrovirus discovered in humans [1,2] and is the causative agent of the unique type of leukemia/lymphoma, adult T-cell leukemia/lymphoma (ATL) [3,4,5,6]. ATL was first reported in Japan [7] and is characterized by adult onset leukemic cells with a mature helper-T phenotype (CD4+, CD25+) and indented nuclei, frequent skin infiltration of leukemic cells, hypercalcemia, and a striking geographic clustering of the patients in the south-western part of Japan [8]. Other ATL endemic areas include sub-Saharan Africa, the Caribbean coast, South America, and parts of the Middle East and Australia [9]. Geographic clustering suggested infectious agents as the cause of ATL. Accordingly, a type C retrovirus particle was identified by electron microscopy in an ATL cell line, and all ATL patients, without exception, possessed antibodies for the virus, establishing that the infection of HTLV-1 is the cause of ATL [3,5]. HTLV-1 has also been isolated from fresh and cultured lymphocytes of a patient with cutaneous T-cell lymphoma, but causality in that disease has not been addressed [1]. The isolated HTLV-1 virus by itself cannot infect target cells efficiently and is mostly transmitted from cell to cell [10,11,12,13,14]. The virus is mainly transmitted from mother to child through the virus-infected cells in breast milk. About 5–10 million people worldwide are estimated to be infected with HTLV-1 and approximately 2–5% of virus-infected people develop ATL after a long latency (40–50 years) [9]. HTLV-1 infection has also been shown to cause HTLV-1-associated myelopathy (HAM)/tropical spastic paraparesis (TSP) and other inflammatory diseases [6,15,16,17,18].
Animal retroviruses transform host cells by two mechanisms, as previously reported (Figure 1) [19,20,21]. One is the transduction of viral oncogenes (v-onc) derived from cellular proto-oncogenes. The other is the insertion of the viral genome, which contains a promoter to drive its own gene expression, within or in close proximity to proto-oncogenes of host cells, leading to an increased expression of those genes.
The HTLV-1 genome exhibits features and an organization in common with other retroviruses, such as long terminal repeats (LTRs), gag, pro, pol, and env genes, but, unlike oncogenic animal retroviruses, it does not contain viral oncogenes derived from cellular proto-oncogenes [22]. In addition, the virus integration sites are not accumulated in the vicinity of certain host proto-oncogenes [23]. Thus, the two mechanisms of oncogenic animal retroviruses do not apply to HTLV-1-mediated oncogenesis. The HTLV genome has a unique pX region (protein coding region X) between env and 3′ LTR, which codes for several regulatory proteins such as the trans-activator of pX region (Tax), Rex, p30, p13, and p12 (Figure 2) [5,24,25,26].
Tax functions as a transcriptional regulator and plays a crucial role in the expression of viral genes, and Rex is important in determining the expression of structural proteins by suppressing the splicing of transcribed viral mRNA. An additional regulatory protein, HTLV-1 basic leucine zipper factor (HBZ), is encoded in the anti-sense strand, which plays a crucial role in the persistence of virus-infected cells. Among these viral regulatory proteins, Tax is thought to play crucial roles in the early-stage transformation of HTLV-1-infected cells by trans-activating cellular genes regulating proliferation and survival (Figure 3) [5].
Tax is a multifunctional protein that interacts with a variety of cellular factors and modulates transcription pathways to activate viral and host cell gene expression, promotes cell cycle progression, facilitates cell survival, and induces genome instability [5,6,27,28,29,30,31,32]. Tax facilitates virus replication by functioning as a transcriptional activator for its own promoter in the 5′ LTR through activation of the CREB/ATF family of host cell transcription factors [33,34]. Tax also trans-activates a variety of cellular genes involved in cell proliferation and survival through the activation of cellular transcription factors, such as NF-κB and SRF, and is thought to play a crucial role in the early stage of the transformation of virus-infected cells [5,6,27,28,29,30,31,32]. Among cellular transcription factors activated by Tax, NF-κB is thought to be critical for the transformation of the virus-infected cells [28,35,36,37]. Indeed, the ability of Tax to immortalize peripheral blood lymphocytes (PBLs) is dependent on the NF-κB pathway and the presence of interleukin (IL-2) [38,39,40,41,42,43]. However, Tax protein is highly immunogenic, and its expression is maintained at extremely low levels in ATL cells [44], likely to evade host immune surveillance. In addition, the sustained activation of NF-κB by Tax induces cellular senescence [37,45,46]. To maintain the expression of Tax at low levels and to avoid cellular senescence induced by the sustained activation of NF-κB, the HTLV-1 bZIP factor (HBZ), encoded by the anti-sense strand, counteracts many functions of Tax to permit the persistence of virus-infected cells [27,37,45,47,48,49,50,51,52]. HBZ suppresses Tax-mediated activation of the LTR by interacting with CREB/ATF family transcription factors, reducing the expression of Tax to escape immune surveillance. HBZ also suppresses Tax-mediated activation of the canonical NF-κB pathway by suppressing the DNA-binding capacity of p65 and by facilitating the degradation of p65 [49]. p30 and p13 also suppress HTLV-1 gene expression by sequestering Tax coding mRNA in the nucleus and interfering with the Tax-CBP/p300 interaction [53,54,55,56,57,58].
Although Tax plays a central role in the promotion of and proliferation of virus-infected cells, not all virus carriers progress to ATL. About 2–5% of the virus-infected population develop ATL, and additional genetic changes are thought to be required for malignant transformation. Tax induces genome instability by the trans-activation of target genes and inhibition of the cell cycle and DNA damage checkpoints through interacting with the mediators of those signal transduction pathways. This leads to the accumulation of mutations in the virus-infected cells [59,60,61], which may contribute to malignant transformation [62,63].
Since Tax immortalization of peripheral blood T cells depends on its ability to activate NF-κB [40,41,42,43], the Tax-mediated activation of cellular genes through NF-κB is thought to be crucial for the transformation of host T cells by HTLV-1 infection [28,35,36,37]. However, the repertoire of Tax target genes, especially those crucial for leukemogenesis, are not known in detail [31]. In this review, we will summarize transcriptional activation mechanisms and target genes of Tax to facilitate the understanding of the mechanisms of leukemogenesis induced by HTLV-1 infection. Other integrated reviews have focused on the protein–protein interactions of Tax and their consequent downstream effects [27,28,30,31,32,36,64].

2. Roles of Tax in HTLV-1-Mediated Transformation

A variety of experimental data support the notion that Tax plays a central role in the early-stage transformation of virus-infected cells [6,27,28,29,30,31,32,65]. Virus vector-mediated introduction of Tax in normal peripheral blood T cells immortalized the cells dependent on the T-cell growth factor IL-2 [38,39]. Immortalization of peripheral blood T lymphocytes by HTLV-1 Tax mutants, introduced by the retrovirus vector or a functional molecular clone, shows that the ability of Tax to activate the NF-κB pathway is necessary and sufficient for the immortalization of normal T cells, whilst the CREB pathway is dispensable [40,41,42,43]. The introduction of Tax into immortalized mouse and rat fibroblast cell lines causes phenotypic transformation and renders the rat cell line tumorigenic in nude mice [66,67,68]. Tax has the ability to promote cell cycle progression depending on its ability to activate the NF-κB pathway [69,70,71]. Tax-transgenic mice developed a variety of tumors including mesenchymal tumors [72,73,74] and leukemia/lymphoma [75,76,77,78], depending on the promoter used [79,80]. In a murine ATL model, triggering Tax proteolysis cures Tax-driven ATL in mice by reducing the tumor-initiating cells [81]. Although Tax expression is very low in vivo, Tax is expressed in a small fraction of leukemic cells and spontaneously switches between on and off states [82]. The transient expression of Tax confers anti-apoptotic activity by upregulating anti-apoptotic gene expression, and this anti-apoptotic effect continues even if Tax expression diminished. This activation creates an antiapoptotic environment that extends to nearby Tax-negative cells, allowing the whole population of cells to survive [82]. The knockdown of Tax rapidly induced apoptosis in most cells, indicating that the transient expression of Tax is critical for maintaining the whole population of the virus-infected cells [82]. Similarly, primary ATL cells from patients with acute or chronic ATL express very low levels of Tax mRNA and protein, but survival of these primary ATL cells is dependent on continued Tax expression [83]. Accordingly, the downregulation of Tax reduced NF-κB activity and resulted in p53/PML activation and apoptosis [83]. Simian T-lymphotropic Virus (STLV) is the etiological agent of ATL in non-human primates, serving as a crucial animal model for studying HTLV-1 pathogenesis [84]. Tax proteins of HTLV-1 and STLV-1 are highly homologous and fulfill similar functions [85,86], supporting the crucial roles of Tax in pathogenicity, although subtle differences in function exist, such as the HTLV-1 Tax showing higher transforming activity than STLV-1 Tax, partly due to its more efficient activation of the non-canonical NF-κB pathway [84]. In addition, recent pilot clinical studies show the effectiveness of the Tax peptide-pulsed dendritic cell (DC) vaccine for treatment of ATL [87,88,89], suggesting that Tax is indeed expressed in ATL cells in vivo and serves as a target of immunotherapy. These observations suggest that Tax may also play important roles in later stages of leukemogenesis induced by HTLV-1 infection.

3. Tax Activates Host Cell Transcription Factors

Tax transcriptionally trans-activates not only its own promoter to augment the virus gene expression [33,34,90,91] but also a variety of host cell genes to promote cell proliferation and survival [5,6,27,28,29,30,31,32]. Tax interacts with a variety of cellular factors to activate at least three transcription factors, cAMP responsive element (CRE)-binding factor (CREB), nuclear factor (NF)-κB, and serum-responsive factor (SRF), to regulate target genes [5]. Intriguingly, while CREB is important for the activation of HTLV-1 gene expression, it is the Tax-induced activation of NF-κB that plays a critical role in the immortalization of host T cells [28,31,37].

3.1. cAMP Responsive Element-Binding Factor (CREB)/Activating Transcription Factor (ATF)

Tax increases HTLV-1 gene expression by activating the viral promoter in the long terminal repeat (LTR) [33,34,90,91]. Tax trans-activates the HTLV-1 LTR primarily via the 21 bp-motifs present in three tandem repeats in the U3 region of LTR [92,93,94]. Tax has a marginal affinity for DNA and does not directly bind to the target sequences [95]. The 21 bp-motifs contain a cyclic AMP response element (CRE), which is activated by cAMP through the CRE binding protein (CREB) [96]. CREB is a member of a family of basic leucine zipper (bZIP) transcription factors in which the leucine zipper is utilized for the formation of homo- or heterodimers, and the basic region mediates binding to the target DNA, cAMP response elements (CRE). CREB forms homo- or heterodimers with other bZIP transcription factors, such as activating transcription factors (ATFs). The activity of CREB is normally regulated by phosphorylation by protein kinase A or Ca2+/calmodulin-dependent protein kinases on Ser133 (pCREB), resulting in the recruitment of the transcriptional coactivator CREB-binding protein (CBP)/p300 for transcriptional activation.
Tax binds indirectly to the 21 bp-motifs through CREB and stabilizes CREB binding to DNA (Figure 4) [95,97]. Tax also binds to ATF2 and CREB-2 (ATF-4) and stabilizes their binding to the 21 bp-motifs to activate transcription [98,99,100]. Tax binds to a basic segment in the bZIP of CREB and ATF, thereby enhancing their affinity for the target sequence [101]. Although Tax enhances CREB binding to the 21 bp-motifs, Tax does not enhance the binding of CREB to other cellular CREs, such as the somatostatin CRE [102,103]. Tax-mediated enhancement of CREB binding to the 21 bp-motifs is dependent on the GC-rich flanking sequences of the viral CRE, with which Tax interacts directly to stabilize CREB binding (Figure 4) [104,105]. Accordingly, Tax does not activate all CRE-containing cellular genes. SR-related protein TAXREB803/SRL300 binds to this GC-rich flanking sequence and contributes to the activation by Tax [106].
Tax bound to CREB recruits CBP/p300, independent of CREB phosphorylation, to activate transcription [95,107,108,109,110]. The interaction between Tax and CBP/p300 is mediated through the kinase-inducible domain (KID)-like domain of Tax [111,112] and multiple regions of CBP/p300, including the KIX, C/H1, and SRC-1-interacting domains (Figure 4) [108,113,114,115]. Tax also interacts with the transducers of regulated CREB 1 (TORC1), TORC2, and TORC3 and enhances p300 recruitment [116,117].
CREB is highly phosphorylated in HTLV-1-infected T-cell lines, and phospho-CREB plays an important role in activation of the LTR by Tax [118]. Tax also binds PCAF and activates the transcription independent of its HAT activity (Figure 4) [119]. Recruitment of CBP/p300 facilitates the acetylation of histones around the viral promoter and loosens chromatin structure in co-operation with the histone chaperone, nucleosome assembly protein 1 (NAP1), to promote transcription (Figure 4) [120,121]. Tax also recruits the chromatin remodeling factor BRG1 to facilitate transcription [122]. In addition, Tax binds to histone deacetylase HDAC1 and inhibits binding to and/or promotes its dissociation from the HTLV-1 promoter [123], thereby attenuating suppression of transcription by deacetylation. Myocyte enhancer factor (MEF)-2 stabilizes Tax/CREB binding to target sequences by binding to the HTLV-1 LTR [124].

3.2. Nuclear Factor-Kappa B (NF-κB)

In contrast to the HTLV-1 LTR, which is mainly activated by Tax through CREB, many cellular genes important for the transformation of virus-infected cells are activated by Tax via the transcription factor NF-κB.
NF-κB plays pivotal roles in important cellular processes such as cell survival, proliferation, and immune responses. The NF-κB family is composed of RelA/p65, RelB, c-Rel, p50, and p52, which are structurally related. The NF-κB proteins activate target gene expression by binding to NF-κB response elements in the promoter regions of target genes as heterodimers (RelA/p50, c-Rel/p50, and RelB/p52) [125]. Whereas RelA/p50 preferentially binds to the sequence 5′-GGGRNNYYCC-3′ (R = purine, N = any nucleotide, and Y = pyrimidine), RelB/p52 recognizes the sequence 5′-RGGAGAYTTR-3′ that is not bound by the RelA:p50 dimer [126].
In the absence of activating stimuli, the NF-κB proteins are retained in the cytoplasm by binding to inhibitory molecules, the inhibitor of NF-κB (IκBα, IκBβ, and IκBε) for RelA/p50 and c-Rel/p50, and p100 (precursor of p52) for RelB/p52, and are kept transcriptionally inactive. The activation of NF-κB proteins are mediated through the canonical and non-canonical pathways [125,127].
Classical NF-κB proteins (RelA/p50 and c-Rel/p50 heterodimers) are activated through the canonical pathway. Upon activating stimuli, the IκB kinase (IKK) complex is activated and phosphorylates IκB, which leads to ubiquitination and proteasome-mediated degradation [128]. This releases NF-κB and unmasks a nuclear localization signal, resulting in the translocation of NF-κB into the nucleus and activation of target genes [125,129,130,131]. The IKK complex is composed of two catalytic subunits (IKKα and IKKβ) and a regulatory subunit (NF-κB essential modulator (NEMO) or IKKγ) [128].
In contrast, RelB/p52 is activated through the non-canonical NF-κB pathway. RelB/p100 (p52 precursor) is retained in the cytoplasm through the action of the ankyrin motif of p100. In response to stimuli, NF-κB-inducing kinase (NIK) activates IKKα, which phosphorylates p100. This phosphorylation marks p100 for ubiquitination and limited proteolysis to generate mature p52, leading to the translocation of RelB/p52 into the nucleus and activation of target genes [127,132,133].
Tax activates the NF-κB pathway through a variety of mechanisms [28,37]. One way Tax activates the canonical NF-κB pathway is through activation of the IKK complex, leading to the phosphorylation and degradation of IκB through the ubiquitin–proteasome pathway (Figure 5) [134,135,136]. This process involves a variety of mechanisms and is explained in detail later.
Tax binds to the ankyrin motifs of IκBα, which is required for interaction with NF-κB, and inhibits binding to or the release of NF-κB and facilitates the proteasome-mediated degradation of IκBα (Figure 5A) [137,138]. Tax also binds to IκBβ and facilitates its degradation [139]. In contrast to IκBα, which is induced by the activation of NF-κB as a negative feedback mechanism, IκBβ expression is not induced by Tax. Hence, in Tax-expressing cells, IκBβ expression is scarcely detected, resulting the activation of c-Rel [139]. Tax also binds to ankyrin motifs of IκBγ (an alternative transcript of the p105 gene) and releases NF-κB p65 to activate transcription [140]. Tax binds to p105, precursor of p50, and also two subunits of the 20S proteasome, HsN3 and HC9, and facilitates the interaction between p105 and HC9, likely promoting the processing of p105 to p50 and accumulation of p65/p50 in the nucleus [141,142,143]. In Tax-expressing cells, expression of p100, the precursor of p52, and c-Rel are enhanced; consequently, RelB/p52 and c-Rel/p50 are the principal components of NF-κB activity [144,145].
Tax also binds to the Rel homology domain of p65 and c-Rel and enhances transcriptional activity (Figure 5A) [146,147]. Consistent with this, Tax co-localizes with p65/p50, CBP/p300, and RNA pol II [148,149].
Tax activates IKK through a variety of mechanisms. Tax activates IKKα and IKKβ through direct interaction with the NF-κB essential modulator (NEMO) (IKKγ), which was isolated by complementation cloning using a flat cell variant of HTLV-1 Tax-transformed rat fibroblasts unresponsive to all NF-κB-activating stimuli tested (Figure 5A) [150,151,152,153,154,155,156]. Consistent with this, somatic mutagenesis assays demonstrated that a Jurkat T-cell line lacking the expression of NEMO was compromised in the activation of NF-κB, and exogenous introduction of NEMO restores the activation of NF-κB by Tax in the mutant cells [155]. In addition, the Tax point mutant M22 (G137A and L138S), which lacks the ability to activate NF-κB, is also compromised in binding to NEMO [152]. Efficient activation of NF-κB also involves the phosphorylation of RelA/p65 at Ser529 and Ser536 by IKKα [157]. Tax also activates kinases upstream of IKK to facilitate its activation. Tax binds to TGF-β-activating kinase 1 (TAK1) and TAK1-binding protein 2 (TAB2) and stimulates TAK1 kinase activity (Figure 5A) [158,159]. Since Tax also binds to the IKK complex through NEMO, Tax recruits TAK1 to the IKK complex to stimulate its kinase activity [158]. In addition, Tax binds to and activates MAP/ERK kinase kinase 1 (MEKK1), which also facilitates the activation of IKKβ [160].
Upon T-cell activation, IKK complexes are transiently recruited to the plasma membrane-associated lipid raft microdomains for the activation of NF-κB in promoting T-cell proliferation [161]. Tax modulates the localization of the IKK complex from cytoplasm to Golgi-associated lipid raft microdomains dependent on the presence of NEMO [162,163], where Optineurin and Tax-binding protein 1 (TAXBP1) cooperate to activate NF-κB [164]. TANK-binding kinase 1 (TBK1) also co-localizes with the Tax and IKK complex in the lipid raft microdomains and plays a crucial role in the activation of NF-κB and Stat3 [165].
In addition to activating the IKK complex by direct interaction, Tax also facilitates IKK complex activation through the formation of polyubiquitin chains, which function as a platform to recruit the IKK complex and regulatory factors [166]. Tax undergoes a variety of post-translational modifications including Lysine 63 (K63)-linked polyubiquitination, which provides an important regulatory mechanism that promotes the Tax-mediated interaction with the IKK complex and activation of NF-κB [167]. The E3/E4 ubiquitin conjugation factor UBE4B interacts with Tax and induces the K48- and K63-linked polyubiquitination of Tax [168].
Tax stimulates the ubiquitin E3 ligase ring finger protein 8 (RNF8) to assemble lysine 63 (K63)-linked polyubiquitin chains to which NEMO and TAB2/3 bind, thereby recruiting IKK and TAK1 (Figure 5A). This recruitment to K63-linked polyubiquitin chains facilitates the activation of TAK1 by cross-autophosphorylation, leading to TAK1-mediated IKK phosphorylation and activation [169]. Tax also recruits the (Met1-linked) ubiquitin E3 ligase, linear ubiquitin chain assembly complex (LUBAC), to the IKK complex, generating Lys63- and Met1-linked hybrid polyubiquitin chains (Figure 5A). This hybrid-chain-dependent oligomerization of the IKK complex leads to trans-autophosphorylation-mediated IKK activation [170]. The possibility that Tax itself functions as a putative ubiquitin E3 ligase for the activation of IKK via synthesis of mixed-linkage polyubiquitin chains is also reported [171].
Tax maintains persistent NF-κB activation by a number of mechanisms such as counteracting negative regulators of the NF-κB pathway, cross-talk between canonical and non-canonical pathways, and the establishment of feed-forward signaling loops [28,30].
Tax also activates the non-canonical NF-κB pathway by multiple mechanisms (Figure 5B). Processing of the RelB heterodimeric partner p52 precursor p100 to p52 is enhanced in HTLV-1-transformed cells. Tax associates with p100 [172]. Tax not only activates IKKα and IKKβ but also activates NF-κB-inducing kinase (NIK) [136], which activates IKKα to phosphorylate p100 to induce processing to p52 (Figure 5B). Tax, via IKKγ, and NIK recruit IKKα to p100 to facilitate the phosphorylation of p100 on serine residues, leading to ubiquitination by the β-transducin repeat-containing protein (β-TrCP) ubiquitin ligase and subsequent processing of p100 to p52 mediated by the 26S proteasome (Figure 5B) [173,174]. Although β-TrCP is essential for NIK-induced p100 ubiquitination and processing [175], the Tax-induced processing of p100 is only partially dependent on β-TrCP [176].

3.3. SRF

Tax activates immediate early response genes such as c-fos through activation of the serum response factor (SRF), which binds the serum response elements (SRE, CArG box). Tax binds not only to NF-κB but also to SRF to activate transcription [146]. Tax also interacts with the ternary complex factor (TCF), which binds to Ets box sequences (GGAA/T), and TCF binding is necessary for the Tax-mediated activation of the SRE [177]. Tax binding to SRF also enhances its binding affinity for divergent SRE sequences [178].

4. Cellular Target Genes of Tax

Target genes of Tax identified to date, in addition to the viral LTR, include those coding for cytokines, their receptors, growth signal transducers, proto-oncogenes, transcription factors, cell cycle regulators, apoptosis-related genes, and cell attachment molecules. Early transformation of virus-infected cells seems to be mediated through the co-operation of a repertoire of target genes, which promote cell proliferation and cell survival.
JPX-9 is a subline of the human T-cell line Jurkat, in which expression of the Tax gene is driven by the metallothionein promoter, enabling the induction of Tax expression by the addition of heavy metal ions, such as Cd2+ and Zn2+, with little, if any, expression in their absence [179]. This cell line has been successfully utilized in many laboratories to identify the target genes of Tax and to examine the effects of Tax [180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207].

4.1. Cytokines, Their Receptors, and Cell Surface Molecules

Cytokines are low-molecular-weight proteins (5~20 kDa) secreted from a variety of cell types that include chemokines, interleukins (IL), interferons (IFN), lymphokines, and tumor necrosis factors (TNFs). The T-cell growth factor IL-2 [208,209,210,211,212,213] and its receptor IL-2Rα [208,209,214,215,216,217,218,219,220] are among the first cellular target genes of Tax to be identified. The Tax-mediated induction of both genes is primarily mediated through the NF-κB pathway. Induction of both IL-2 and its receptor suggested autocrine growth stimulation as a possible mechanism underlying leukemogenesis induced by HTLV-1 Tax. However, IL-2Rα alone cannot transmit a growth signal into cells. IL-2R is composed of three components, IL-2Rα, IL-2Rβ, and IL-2Rγ. IL-2Rα contributes to the formation of a high affinity receptor, and IL-2Rβ and IL-2Rγ are required to transmit growth signals by coupling to the tyrosine kinases Janus kinase 1 (JAK1) and JAK3, respectively [221]. Tax also induces expression of the IL-2Rγ gene [222], suggesting the possibility that Tax increases the expression of the functional high affinity receptor for IL-2 [223]. ATL cell lines established from ATL patients are all dependent on IL-2 for their isolation and culture. However, IL-2-dependent ATL cell lines become IL-2-independent during further culturing [224]. These observations suggest the possible involvement of an IL-2/IL-2R autocrine mechanism in the early phase of transformation by HTLV-1 infection. In support of this notion, dual infection of the threadworm Strongyloides stercoralis (S. stercoralis) with HTLV-1 induces polyclonal expansion of the HTLV-1-infected cells by stimulating the IL-2/IL-2R system [225].
Tax also activates the c-sis/platelet-derived growth factor-B [226], IL-9 [227,228], IL-15/IL-15R [229,230,231], and IL-21/IL-21R genes [232,233], which are involved in the activation, cell growth, and survival of T cells, suggesting that the Tax-mediated activation of these genes may also contribute to the proliferation of HTLV-1-infected T cells. The Tax-mediated activation of these genes is also mainly mediated through the NF-κB pathway [229,230,231,232].
HTLV-1-infected T-cell lines and fresh leukemic cells isolated from patients with ATL express high levels of TGF-β1 mRNA and secrete TGF-β1. Tax activates the TGF-β1 promoter through activator protein 1 (AP-1) binding sites [234]. In addition, tumors from Tax transgenic mice, under the control of HTLV-1 LTR [72], express high levels of TGF-β1 mRNA and protein. Moreover, TGF-β1 stimulated the growth of the tumor cells, suggesting that the Tax-mediated induction of TGF-β1 may play a role in tumorigenesis [235].
Tax activates the gp34 gene [181], which codes for tumor necrosis factor ligand superfamily member 4, the ligands for OX40, a member of the TNF receptor family [236]. OX40 is expressed on activated T cells and is a costimulatory T-cell molecule known to be upregulated in HTLV-1-infected cells [189,237,238]. Tax activation of both genes is mediated through the NF-κB pathway [237,239]. The Tax-mediated induction of OX40 and its ligand gp34 may transmit growth signals and facilitate the proliferation of HTLV-1-infected cells [4,223]. Intriguingly, the Tax of HTLV-1 (Tax1) but not of HTLV-2 (Tax2), which is not associated with pathogenesis of ATL, induces OX40L expression through the activation of NF-κB2/p100/p52 [240]. Expression levels of OX40 are also associated with disease progression in patients with HAM/TSP [238].
Tax induces the aberrant expression of CD40, a member of the tumor necrosis factor receptor (TNFR) family that plays an important role in lymphocyte activation and differentiation [241]. Tax activates the CD40 gene through the NF-κB signaling pathway. Since Tax also induces expression of the CD40 ligand (CD40L) and ligation of CD40 on T cells with recombinant CD40L elicited NF-κB activation, the CD40/CD40L pathway may form a positive regulatory loop in NF-κB activation in HTLV-1-infected T cells [242].
HTLV-1-transformed T cells and fresh ATL leukemic cells produce vascular endothelial growth factor (VEGF) and basic fibroblast growth factor (bFGF), which induce angiogenesis. Tax also activates the VEGF promoter [243].
ATL is often accompanied by hypercalcemia, and the expression of parathyroid hormone-related peptide (PTHrP) is increased in all the ATL patients examined [244]. PTHrP increases serum calcium concentration by the same mechanism as the parathyroid hormone (PTH). PTHrP expression is increased in asymptomatic carriers without hypercalcemia but is not elevated in HTLV-1-irrelevant leukemic cells [240]. Tax activates the PTHrP gene [245] and PTHrP P2 promoter through Ets1, AP1/c-jun, Sp1, and the NF-κB pathway [246,247,248,249]. Expression of the PTHrP receptor (PTH1R) is also increased in HTLV-1-infected lymphocytes, indicating a potential autocrine role for PTHrP [250]. However, whether PTHrP signaling facilitates the growth of HTLV-1-infected cells has yet to be elucidated.
Chemokines and their receptors play important roles in the migration and tissue localization of various lymphocyte subsets. C-C chemokine receptor type 4 (CCR4) is the receptor for CC chemokines CCL2 (MCP-1), CCL4 (MIP-1), CCL5 (RANTES), CCL17 (TARC), and CCL22 (macrophage-derived chemokine: MDC). CCR4 expression is enhanced in primary ATL cells and HTLV-1-immortalized T cells. Primary ATL cells show efficient migration in response to the CCR4 ligands CCL17/TARC and CCL22/MDC in chemotaxis assays. Since CCR4 is known to be involved in T-cell migration into skin, Tax-mediated CCR4 expression may explain the frequent dermal infiltration of ATL cells [251,252]. Tax induces CCR4 expression through the induction of Fos-related antigen 2 (Fra-2), which forms Fra-2/JunB and Fra-2/JunD heterodimers to activate an AP-1 site in the CCR4 promoter. Fra-2 also induces the expression of c-Myb, BCL-6, and MDM2, thereby promoting cell growth [253]. CXCR7 is an atypical chemokine receptor, frequently expressed by tumor cells, known to promote cell growth and survival. Tax induces the expression of CXCR7 through NF-κB and promotes the growth and survival of HTLV-1-infected T cells [186].
HTLV-1 infection causes not only ATL but also several inflammatory diseases such as HTLV-associated myelopathy/tropical spastic paraparesis (HAM/TSP). Tax activates a variety of genes involved in inflammation and immune response in addition to cell growth, such as IL-1α [254,255,256,257,258], IL-1β [256,259], IL-4 [260], IL-5 [203,227,261], IL-6/IL-6R [188,254,256,262,263,264,265,266], IL-8 [184,205,267], IL-9 [266], IL-10 [83,268,269], IL-12 [270], IL-13 [227,271,272,273], IL-17/IL-17R [274,275], IL-21/IL-21R [232,233], IFN-γ [270,271,276], TNF-α [256,270,277,278,279], TNF-β or lymphotoxin [256,280,281], granulocyte–macrophage colony-stimulating factor (GM-CSF) or colony-stimulating factor 2 (CSF2) [209,218,282,283], granulocyte colony-stimulating factor (G-CSF) [283], monocyte chemoattractant protein-1 (MCP-1) [270,284,285], MCP-3 [270], C-C motif chemokine ligand 1 (CCL1) [286], CCL3/macrophage inflammatory protein 1α (MIP-1α) [184,284,285,287], CCR4 ligand CCL4/MIP-1β [288,289], CCR4 ligand CCL5/regulated on activation, normal T cell expressed and secreted (RANTES) [184,284,290], CCL11/Eotaxin [270], CCL20/MIP-3α and its receptor CCR6 [291], CCR4 ligand CCL22 [292], chemokine, C motif, ligand 1 (XCL1, SCM1) [184], chemokine, CXC motif, ligand 10 (CXCL10) (IP-10) [184,284,293,294], CXCL12 (SDF-1, PBSF) [295], C-C chemokine receptor type 9 (CCR9) [198], C-X-C chemokine receptor type 4 (CXCR4) [296], and CXCR7 [186]. CXCR7 is an atypical chemokine receptor frequently expressed by tumor cells and known to promote cell growth and survival. Tax induces the expression of CXCR7 through NF-κB and promotes the growth and survival of HTLV-1-infected T cells [186]. Most of these genes are activated by Tax through the NF-κB pathway. The Tax-mediated activation of these genes is likely to be involved in the onset of HTLV-1-associated pathogenesis such as inflammatory diseases and/or ATL. Consistent with this, the addition of IL-10 to Tax-depleted primary ATL cells rescues the survival of the Tax-depleted cells, suggesting that Tax-induced expression of IL-10 plays a critical role in the survival of HTLV-1-infected cells [83]. Expression of repulsive guidance molecule A (RGMa), which is known to contribute to neuronal damage, is upregulated in HTLV-1-infected cells from patients with HAM, and Tax induces the expression of RGMa [297]. Tax also enhances the expression of T-box transcription factor (T-bet) and induces the Th1-like state, consequently promoting the production of IFN-γ, which enhances inflammation [276]. Tax-induced RGMa and T-bet may also contribute to the pathogenesis of HAM.
HTLV-1-infected cells show an increased expression of vascular cell adhesion molecule 1 (VCAM-1, CD106), and Tax activates the VCAM-1 promoter through an NF-κB binding site [204]. Tax also induces the expression of leucocyte function-associated antigen-3 (LFA-3) [191]. Tax-induced VCAM-1 acts in synergy with leucocyte function-associated antigen-3 (LFA-3) to promote the interaction between T cells and increase T-cell proliferation [298].
Cell adhesion molecule 1 (CADM-1), also known as the tumor suppressor in lung cancer 1 (TSLC1) or immunoglobulin superfamily member 4 (IGSF4), is a cell adhesion molecule that is expressed in most cell types, except for peripheral blood lymphocytes, and is thought to function as a tumor suppressor in a variety of cancers [299,300]. Expression of CADM1 is enhanced in HTLV-1-infected T-cell lines and primary ATL cells, and the percentage of CADM1(+) cells in HTLV-1 carriers and ATL patients correlates well with viral loads, indicating that CADM1 serves as a diagnostic cell surface marker for ATL [301]. The roles of CADM1 in ATL pathogenesis have yet to be elucidated.
Osteopontin (OPN) is a ligand for CD44, which promotes cancer cell growth, invasion, and metastasis. CD44 associates with a variety of growth factor receptors and enhances ligand binding, thereby promoting cell proliferation. Tax induces the expression of both OPN and CD44 through AP-1 and non-canonical NF-κB pathways, respectively [302,303,304], suggesting the involvement of OPN/CD44 in the growth of HTLV-1-infected cells.
Tax induces the expression of CD137. CD137 is transferred by trogocytosis to CD137L-expressing APC, where the CD137-CD137L complex is internalized and degraded, resulting in a reduced CD137-mediated T-cell co-stimulation and a weakened cellular immune response, which may facilitate the escape of the virus from immune surveillance [305].
Other cell surface molecules reported to be upregulated by Tax include E-cadherin [306], CD44 [303], CD83 [193], tumor necrosis factor receptor superfamily member 9 (TNFRSF9/4–1BB/CD137/ILA) [307], tumor necrosis factor ligand superfamily member 7 (CD70) [308], early activation marker of lymphocytes CD69 [309], and major histocompatibility complex, class I [310].
Tax transgenic mice under the control of HLV-1 LTR develop mesenchymal tumors including neurofibroma [73]. Tumors and tumor cell lines derived from these mice produce nerve growth factors (NGFs), and Tax trans-activates the NGF promoter through a Tax-responsive element exhibiting a 92% homology to the 21 bp repeats of the HTLV-1 LTR. The NGF receptor is also expressed in transgenic tumor cells, suggesting that Tax may activate an autocrine mechanism through the upregulation of NGF [311].
Enkephalins are opioid peptides that function as neurotransmitters and neuroimmunomodulators. Tax activates the proenkephalin gene in glial cells and activates the proenkephalin promoter through the increased binding of c-Fos/c-Jun proteins to the AP-1 site [312,313,314].
Activation of cytokines, their receptors, and other cell surface molecules, which transmit growth signals, may facilitate the proliferation of HTLV-1-infected cells at early stages of expansion of the virus-infected cells.
The genes among those mentioned above, which have been reported to be induced by Tax, are summarized in Table 1.

4.2. Proto-Oncogenes, Transcription Factors, and Growth Signal Transducers

Tax activates a variety of proto-oncogenes and growth signal transducers, including early response genes such as c-Fos [180,315,316], Fra-1 [206], v-jun avian sarcoma virus 17 oncogene homolog (c-Jun) [206], JunD [206], early growth response 1 (Egr-1) [206,317,318], Egr-2 [206,318], ETR101 [202], butyrate response factor 1 (BRF1) [319], and c-Myc [320]. Dimerization between the Jun (c-Jun, JunB, and JunD), Fos (c-Fos, FosB, Fra1, and Fra2), and activating transcription factor (ATF) protein families generate AP-1 transcription activity, which plays an important role in tumorigenesis [321]. Accordingly, primary ATL cells and HTLV-1-infected cells exhibit constitutively active AP-1 [322,323,324]. Tax activates the Egr-1 and Egr-2 genes through SRF [318]. Tax also activates the Egr-1 gene through the NF-κB pathway and stabilizes the Egr-1 protein by direct interaction. Elevated Egr-1 in turn facilitates the translocation of p65 into the nucleus, leading to increased NF-κB activation [325]. These results suggest a positive feedback loop in Egr-1 activation and NF-κB activation in Tax-expressing cells.
Tax increases c-Rel and p105 mRNA expression [326], although it may not be directly mediated by Tax, suggesting that the induction of expression of these genes may also contribute to the activation of NF-κB by Tax. Tax-mediated induction of the c-Rel gene is mediated through the activation of an intrinsic κB enhancer element in the c-Rel promoter by RelA/p65 [134].
The Pim family of serine/threonine kinases (Pim-1, -2, and -3) promote tumorigenesis by enhancing cell survival. Tax induces the expression of Pim-1 and Pim-3 through NF-κB, which may contribute to the cell growth and survival of HTLV-1-infected cells [327,328]. Conversely, Pim-1 and Pim-3 bind to Tax and reduce its expression, thereby creating a negative feedback loop. Since Tax protein is highly immunogenic, Tax-mediated Pim-1 and Pim-3 induction may contribute to the escape of HTLV-1-infected cells from immune surveillance by reducing Tax protein expression [328]. Pim kinases may serve as a target for the treatment of ATL [329].
The signal transducer and activator of transcription (STAT) is a family of transcription factors which transmit cytokine signals from membrane receptors coupled with JAK kinases into the nucleus. The activation of the JAK-STAT pathway and the expression of STAT1, STAT3, and STAT5 mRNAs are induced by mitogen stimulation of normal human peripheral blood mononuclear cells. In contrast, HTLV-1-infected T cells show constitutive activation of the JAK-STAT pathway [330,331]. In addition, HTLV-1-infected T-cell lines also show an enhanced expression of STAT1, STAT3, and STAT5 genes, and Tax induces the expression of STAT1 and STAT5 genes in a human T-cell line [332]. IL-2 stimulation enhanced the DNA-binding activity of STAT3 and STAT5 in a HTLV-1-transformed cell line concomitant with its proliferation. The induction of STAT1 and STAT5 by Tax may enhance cytokine-induced proliferation of the virus-infected T cells, thereby playing a role in the IL-2-dependent transformation of T cells in response to HTLV-1 infection.
Interferon regulatory factor (IRF-4) was shown to be constitutively produced in HTLV-1-infected cells but not in continuously growing non-Tax-expressing T-cell lines. In contrast to other IRFs, IRF-4 is not induced by INFs but is transiently expressed upon growth stimulation of T lymphocytes, such as anti-CD3 and phorbol 12-myristate 13-acetate (PMA)/ionomycin stimulation. Tax activates the IRF-4 promoter through the NF-κB and ATF pathway, suggesting that IRF4 induction in HTLV-1-infected cells is mediated through Tax activation of the IRF-4 gene transcription [333,334]. Intriguingly, an activating mutation of IRF4 is frequently observed in ATL [335]. IRF5 is also constitutively expressed in HTLV-1-infected T cells and Tax activates the IRF5 promoter [336]. Tax can indirectly activate the interferon-responsive enhancer element of the 2′,5′-oligoadenylate synthetase gene in an NF-κB pathway-dependent manner [337].
The Src family of protein-tyrosine kinases (Lck, Fyn, and Lyn) is associated with growth signal transduction, including cytokine signaling. Expression of the Lyn gene is detected in HTLV-1-infected and Tax-expressing T-cell lines but not in normal T cells [338]. Tax induces expression of the Lyn gene and activates the Lyn promoter in T cells, suggesting that the induction of Lyn gene transcription in HTLV-1-infected cells is mediated through Tax [339]. Lyn is associated with IL-2Rβ and JAK3, and dominant negative Lyn reduced the phosphorylation of JAK3 and STAT5, suggesting the involvement of Lyn in IL-2R signaling. These results suggest that the induction of Lyn expression by Tax may facilitate IL-2R signaling, although Lyn is not essential for IL-2R signal transduction [340].
Expression of key genes associated with T-cell activation was compared between 20 asymptomatic carriers (ACs) of HTLV-1 and 20 healthy control subjects using RT-qPCR [341]. Expression (mRNA) of glycogen synthase kinase-3β (GSK3β), mitogen-activated protein kinase kinase kinase 14 (MAP3K14 or NIK), phospholipase C γ 1 (PLCG1), and mitogen-activated protein kinase kinase kinase-7 (MAP3K7 or TAK1) were significantly higher in asymptomatic carriers (ACs) compared to healthy subjects, suggesting the involvement of these genes in the early stage of ATL development.
Activation of these early response and signal transducer genes may transmit growth signals in the absence of growth simulation, thereby contributing to the growth of HTLV-1-infected T cells.
Nur77 is a macrophage-expressed nuclear receptor that regulates the inflammatory response [342]. HTLV-1-infected T cells showed a higher expression of Nur77, and Tax induces Nur77 expression in JPX-9 cells [196]. Promoter analysis shows that the induction is mediated by CREB/ATF-related transcription factors [196,201]. Induction of Nur77 expression may be relevant in HTLV-1-related inflammatory diseases.
Expression of the Ellis Van Creveld (EVC), a positive mediator for three hedgehog (Hh) signaling molecules Desert (DHH), Indian (IHH), and Sonic (SHH), is upregulated in ATL and HTLV-1-infected cells. The downregulation of EVC leads to apoptotic cell death in ATL cell lines, indicating that HH activation is involved in the regulation of leukemic cell survival [343].
An inducibly expressed atypical IκB protein NF-κB-ζ (IκB-ζ) associates with the NF-κB subunit in the nucleus and positively regulates its transcriptional activity [344]. IκB-ζ is constitutively expressed in HTLV-1-infected T-cell lines and ATL cells, and Tax trans-activates the IκB-ζ gene, mainly through NF-κB [345]. The over-expression of IκB-ζ induces the expression of NF-κB and interferon-regulatory genes, such as B cell CLL/lymphoma 3 (BCL3), guanylate-binding protein 1, and signal transducer and activator of transcription 1 (STAT1) [345]. Interestingly, IκB-ζ interacts with Tax and inhibits the Tax-mediated activation of NF-κB, AP-1, and HTLV-1 transcription [345], suggesting that IκB-ζ may be involved in the pathogenesis of HTLV-1-associated diseases.
The comparison of gene expression profiles between 10 ATL patients, 10 ACs, and 10 normal controls revealed that Evi-1 and Foxo-1 levels were significantly higher in ATL patients [346]. Evi-1 and Foxo-1 are important factors in maintaining hematopoietic and embryonic stem cells, suggesting a role for both in HTLV-1-mediated transformation.
B-cell/CLL lymphoma 6 (BCL6) expression is enhanced in HTLV-1-infected T-cell lines, and Tax induces BCL6 expression. Knockdown or pharmacological inhibition of BCL6 reduced the cell growth and survival of HTLV-1-infected T-cell lines. BCL6 might be considered a novel target for ATL treatment [347].
Expression of Exportin 1 (XPO1), also known as chromosomal region maintenance 1 (CRM1), is upregulated in HTLV-1-infected T cells, and the knockdown of XPO1 reduced cell proliferation. The XPO1 inhibitor KPT-330 also reduced proliferation, increased DNA damage, and induced G1 cell cycle arrest and caspase-dependent apoptosis, suggesting the use of KPT-330 in clinical trials targeting ATL [348].
Expression of OSBP-related protein 4 (ORP4L), also known as oxysterol-binding protein 2 (OSBP2), is upregulated in ATL cells due to the loss of miR-31 expression caused by HTLV-1 infection. ORP4L knockout blocked T-cell leukemogenesis in Tax transgenic mouse models, whereas ORP4L expression in T cells resulted in T-cell leukemia in mice. ORP4L interacts with PI3Kδ to promote PI(3,4,5)P3 generation, thereby contributing to the activation of AKT, NF-κB, and T-cell leukemogenesis. Consistent with this, ORP4L knockout or pharmacological inhibition reduces the growth of ATL cells in patient-derived xenograft ATL models. These results point to ORP4L as a therapeutic target of ATL [349].
Notch signaling plays an important role during neuronal development. Notch is a cell surface protein and, upon binding of the ligand, such as Jagged (JAG1), the Notch intra-cellular domain (NICD) is cleaved and translocates into the nucleus to regulate gene expression. Notch signaling is constitutively activated in ATL cells, as determined by gene array analysis [350]. The Notch ligand, JAG1, is overexpressed in primary ATL cells compared to normal peripheral blood mononuclear cells (PBMCs). Tax induces JAG1 through the NF-κB pathway. Inhibition of JAG1 by knockdown or antibodies reduces Notch1 downstream signaling and limits the cell migration of transformed ATL cells [351], suggesting that JAG1 can be a new target to dampen Notch1 signaling in HTLV-1-transformed cells. F-box and WD repeat domain containing 7 (Fbw7), a component of E3 ubiquitin ligase, recognizes NICD and facilitates its degradation. Intriguingly, more than 30% of ATL patients possess activating mutations in Notch including those leading to reduced Fbw7-mediated degradation and the stabilization of NICD [352]. In addition, 25% of ATL patients possess mutations in Fbw7, which eliminates its interaction with NICD, resulting in increased protein stability and constitutive Notch1 signaling [353]. Notch signaling is activated in ATL cells irrespective of the presence of activating mutations in Notch itself or its repressor Fbw7, and lncRNA H19/miR-675 is highly expressed in ATL cells, contributing to Notch signaling activation [351]. In addition, Tax directly binds to Fbw7 in the nucleus and competes with its substrate to stabilize NICD [354]. Tax also increases NICD expression by stabilizing the half-life of NICD mRNA. Tax interacts with NICD and recombination signal Jκ binding protein (RBP-Jκ) and stabilizes binding of the ternary complex to the RBP-Jκ target sequence, thereby enhancing Notch signaling [355]. Inhibition of Notch signaling by knockdown or γ-secretase inhibitors reduces tumor cell proliferation and tumor formation in an ATL-xenograft mouse model [350,352,355], suggesting that the Tax-mediated enhancement of Notch signaling plays an important role in the proliferation of HTLV-1-infected cells.
HTLV-1 Tax induces and associates with Crk-associated substrate lymphocyte-type (Cas-L). Cas-L is a docking protein that is heavily tyrosine phosphorylated by the engagement of Integrin β1 in T cells. Fresh ATL cells and HTLV-1-infected T-cell lines show an elevated level of Cas-L. Tax induces expression and tyrosine phosphorylation of Cas-L in the Jurkat human T-cell line, resulting in increased motile behavior and response to integrins [185]. Tax also associates with Cas-L and exogenous Cas-L-inhibited Tax-mediated activation of NF-κB, suggesting that Cas-L might regulate Tax-mediated NF-κB activation [185]. Tax transgenic mice spontaneously develop rheumatoid arthritis (RA). Splenocytes from Tax transgenic mice show higher migratory activity with enhanced expression and phosphorylation of Cas-L than non-transgenic control [356]. In addition, in a mouse model of collagen-induced arthritis, Cas-L (−/−) mice showed a reduced severity of arthritis as compared to Cas-L (+/+) mice with a reduced immune response to collagen [357]. These observations suggest the involvement of Cas-L in the pathogenesis of RA.
Tax induces the expression of membrane-associated guanylate kinase, WW and PDZ domains-containing 3 (MAGI3) in Rat-1 cells [358]. MAGI3 has multiple PDZ domains and interacts with Tax in 293T cells, dependent on a PDZ domain-binding motif (PBM). PBM is not present in the Tax2 of HTLV-2 and is thought to be important for transforming the activity of Tax1 [359]. The interaction of Tax1 with MAGI-3 alters their respective subcellular localization and correlates well with the transforming activity of Tax, suggesting that the interaction of Tax with MAGI3 may play a role in the pathogenesis of HTLV-1-associated diseases [358]. Tax1 PBM also binds other PDZ domain-containing proteins such as Dlg1, Scribble, and MAGI-1 [360,361,362] and plays an important role in the IL-2 independent growth induction of a T-cell line [363], sustained T-cell proliferation in HTLV-1-infected humanized mice [364], and Akt pathway activation [365].
Expression of the calpain 2 gene is increased in HTLV-1-infected cells, and the introduction of the pX gene into an NK-like cell line induced calpain 2 gene expression, suggesting that Tax activates the calpain 2 gene [366]. Calpain 2 localizes in lipid rafts in T cells, suggesting the involvement of calpain 2 in TCR signaling [367].
The genes among those mentioned above, which have been reported to be induced by Tax, are summarized in Table 2.

4.3. Cell Cycle Regulators

Cell cycle progression is governed by cyclin-dependent kinases (CDKs), whose activity is regulated by the expression of partner cyclins [368]. Growth stimulation induces the expression of D-type cyclins, which bind to and activate CDK4 and CDK6 [369,370,371]. CDK4 and CDK6 in turn phosphorylate and inactivate p130, a member of the pRB protein family, which is bound to repressor-type E2F (E2F4 and E2F5) to suppress growth-related E2F target genes [370,371,372]. This relieves repression and induces the expression of E2F targets including cyclin E and activator-type E2F (E2F1, E2F2, and E2F3a). Cyclin E binds to and activates CDK2, which further phosphorylates and inactivates pRB family members to enhance E2F activity [370,371,372]. Induced activator E2Fs replace repressor-type E2F to activate target genes. This enables passage through the restriction point in the late G1 phase and entry into the S phase [370,371,372]. Tax activates a variety of cell cycle regulatory genes to promote cell cycle progression.
Kit 225 is a human T-cell line whose growth is dependent on IL-2 and can be growth arrested by starvation of the IL-2 without significant cell death. This enables the examination of effects of Tax on cell cycle progression. In contrast, the growth of Jurkat human T cells is independent of IL-2 and does not enter a resting state in response to IL-2 deprivation. Tax can induce cell cycle progression in IL-2-starved Kit 225 cells with concomitant induction of cyclin D2 and the activation of E2F [71]. Tax can also stimulate cell cycle progression in phytohemagglutinin (PHA)-stimulated peripheral blood lymphocytes (PBLs), indicating that Tax can induce cell cycle progression in human normal T cells. Intriguingly, the Tax-mediated cell cycle progression of PHA-PBLs is observed in Tax-expressing cells but not in Tax-negative cells [71], suggesting that the cell cycle progression induced by Tax is a cell intrinsic event and not mediated through humoral factors under these experimental conditions. The expression of Tax in IL-2-starved Kit 225 cells increased the expression of cyclin D2, cyclin E, E2F1, CDK2, CDK4, and CDK6 and reduced that of the CDK inhibitors p19INK4d and p27Kip1 [373]. The increased expression and activity of cyclin D2, CDK4, and CDK6 leads to the enhanced phosphorylation of RB family proteins, activation of E2F, and cell cycle progression [373]. The Tax-mediated induction of cyclin D1, cyclin D2, and E2F1 have also been reported in other experimental settings. HTLV-1-infected T-cell lines and ATL cells show the enhanced expression of cyclin D2, and Tax induces cyclin D2 gene expression [374]. Introduction of Tax into the IL-2-dependent murine T-cell line CTLL-2 enables IL-2-independent growth with the concomitant induction of cyclin D1 and cyclin D2 [375], and the introduction of Tax into CEM human T cells, which lack the expression of p16INK4a, increased E2F1 expression [376]. Tax-mediated induction of cyclin D2 and CDK6 expression is mediated through the activation of transcription through NF-κB-like sequences in their promoters and is dependent on the NF-κB pathway [377,378]. Tax-mediated activation of the cyclin D1 gene is mediated through CRE in co-operation with CREB, the transducer of regulated CREB 2 (TORC2) and p300 [117]. Intriguingly, Tax cannot activate E2F or induce cell cycle progression in fibroblasts, in which the over-expression of E2F1 could induce cell cycle progression [71], suggesting that the Tax-mediated activation of E2F and cell cycle progression are cell-type-specific effects of Tax. Knockdown of cyclin D2 and CDK6 suppresses cell cycle progression induced by Tax in Kit 225 cells. Tax cannot induce cyclin D2 or CDK6 expression or activate the cyclin D2 or CDK6 promoters in fibroblasts, and over-expression of cyclin D2 and CDK6 induces cell cycle progression in fibroblasts [378]. These results suggest that cell-type-specific activation of the cyclin D2 and CDK6 genes through the NF-κB pathway is important for Tax-mediated cell cycle progression.
In addition to the induction of partner cyclins, the activation of CDKs requires phosphorylation at the activation segment T-loop by CDK-activating kinase (CAK) [379]. CAK is composed of CDK7, Cyclin H, and Menage a trois homolog 1 (Mat1) [380,381], among which CDK7 is the catalytic component. Accordingly, CDK7 plays an essential role in cell cycle progression by activating other CDKs (CDK1, 2, 4, and 6) involved in cell cycle progression. CAK, with its catalytic component CDK7, is also a component of transcription factor II H (TFIIH), which is required for the initiation of transcription by RNA polymerase II. RNA polymerase II recruited to target genes is paused at the transcriptional start site by anchoring of the C-terminal domain (CTD) of RNA polymerase II. To initiate transcription, RNA polymerase II needs to be released from anchoring by phosphorylation of the CTD catalyzed by CDK7 in TFIIH [382,383]. Accordingly, CDK7 plays an essential role in the initiation of transcription by RNA polymerase II. Tax induces the expression of CDK7 but not its partner Cyclin H [384]. In spite of this, Tax enhances phosphorylation of CDK2 and the CTD of RNA polymerase II [384], which are reported to be mediated by CDK7. Moreover, the knockdown of CDK7 suppressed the induction of target gene expression and cell cycle progression induced by Tax [384]. These observations suggest a unique relationship between CDK7 and Cyclin H, opposite to that of other CDKs and cyclins in the control of its activity. Since a basic mechanism of Tax-mediated promotion of cell cycle progression is the activation of target genes and CDKs, the Cdk7 gene could be a crucial target of Tax in the proliferation of HTLV-1-infected cells.
RNA polymerase II released from transcription initiation sites still pauses at a promoter proximal site [385]. This pause is released by the function of the positive transcription elongation factor (P-TEFb), which contains Cdk9/cyclin T as a catalytic component. Tax induces the expression of the transcription elongation factor for RNA polymerase II 2 (ELL2) in HTLV-1-infected T cells [386]. Tax also interacts with ELL2 and synergistically activates the HTLV-1 LTR [386,387]. These results suggest that Tax also facilitates the elongation of transcription by modulating RNA polymerase II progression.
Primary ATL cells and Tax-expressing cells exhibit higher amounts of cyclin-dependent kinase inhibitor 1A (CDKN1A, p21Cip1) mRNA and protein compared to Tax-negative cells [388,389,390]. This is likely due to activation of the CDKN1A gene by Tax, since Tax activates the p21Cip1 promoter through an E2A transcription factor binding site, the CREB/ATF pathway, and two TGF-β-inducible Sp1 binding sites, independently of p53 [389,391,392]. Although p21Cip1 binds to and inhibits the CDK2/cyclin complex, it facilitates the assembly of cyclin D and CDK4 or CDK6 and enhances kinase activity [393,394]. Indeed, in HTLV-1-infected cells, p21Cip1 is in a complex with cyclin D2/CDK4, which retains kinase activity to phosphorylate pRB [395]. Although Tax induces the expression of p21Cip1, p21Cip1 is phosphorylated through the Tax-mediated activation of PI3/AKT and retained in the cytoplasm in an inactive state. In addition, in ATL cells, the expression of p21 is not increased, likely due in part to methylation of the gene [396].
Tax also promotes cell cycle progression by directly interacting with cell cycle regulators. Tax associates with cyclin D3/CDK6 [397]. Tax binds to and activates CDK4, thereby facilitating the phosphorylation of pRB [398]. Tax also associates with pRB and facilitates its proteasomal degradation [399]. Tax binds to and inactivates CDK inhibitors p16INK4a and p15INK4b, leading to the activation of CDK4 [400,401,402].
Although Tax facilitates progression through the G1 and S phase transition [403], Tax delays progression through S-G2-M by directly binding to and prematurely activating the Cdc20-associated anaphase-promoting complex (APC/CCdc20), an E3 ubiquitin ligase that controls the metaphase to anaphase transition [404,405]. This leads to premature degradation of cyclin B in S-G2 phases and may explain the mitotic abnormalities observed in HTLV-1-infected cells.
The genes among those mentioned above, which have been reported to be induced by Tax, are summarized in Table 3.

4.4. Genes Related to Apoptosis and Cell Survival

Apoptosis is induced by a cascade of activation of caspases which degrade cellular components, including DNA by activating DNase. The activation of caspase 9, an initiator caspase, is initiated by the release of cytochrome c from the mitochondria by destabilization of the mitochondrial outer membrane. Cytochrome c in turn forms the apoptosome, together with apoptotic protease-activating factor 1 (Apaf-1), in the cytosol to activate caspase 9. Caspase 9 then activates caspase 3 to initiate a cascade of activation of caspases 3, 6, and 7. Accordingly, stability of the outer membrane of the mitochondria is a key determinant of cell survival versus the induction of apoptosis. Anti-apoptotic Bcl-2 family proteins stabilize the outer membrane of the mitochondria and promote cell survival, while pro-apoptotic Bcl-2 family proteins destabilize it to promote apoptosis. Death receptors such as Fas activate caspase 8, an initiator caspase, upon the binding of a death ligand, such as the Fas ligand, to activate caspase 3 to initiate the caspase activation cascade.
Tax not only facilitates cell cycle progression but also cell survival by activating a variety of anti-apoptotic genes, indicated by changes in the gene expression profiles of HTLV-1-immortalized T cells, determined by DNA micro array, compared to normal T cells [406]. Many of the differentially expressed genes identified in HTLV-1-immortalized T cells are involved in the control of apoptosis, suggesting that deregulation of the apoptotic signaling network may play a role in the HTLV-1-mediated oncogenesis. Activation of anti-apoptotic genes by Tax is also mainly mediated through the NF-κB pathway.
One way Tax suppresses apoptosis is through stabilization of the mitochondrial outer membrane through the induction of anti-apoptotic Bcl-2 family members. Expression of Tax, in the IL-2-dependent murine T-cell line CTLL-2, alleviates apoptosis induced by deprivation of IL-2 concomitant with the induction of Bcl-xL expression via the NF-κB pathway [407,408]. Consistent with this, the Bcl-2/Bcl-xL inhibitor navitoclax in combination with the JAK inhibitor ruxolitinib reduces the IL-2-dependent ex vivo proliferation of peripheral blood mononuclear cells (PBMCs) from ATL patients in smoldering/chronic stages [409]. Induction of Tax in JPX-9 cells also suppresses receptor-mediated apoptosis induced by the anti-Fas antibody and tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) and inhibits chemical-induced apoptosis (pyrrolidine dithiocarbamate, etoposide, and staurosporine) [182]. Tax-mediated induction of Bcl2-related protein A1 (BFL1) through the NF-κB pathway also contributes to cell survival [410].
Tax also suppresses apoptosis by inhibiting the activation of caspases. Inhibitors of apoptosis (IAPs) act to block the activation of caspases. Expression of baculoviral IAP repeat-containing 3 (BIRC3, cIAP2, and HIAP-1) is upregulated in HTLV-1-infected T-cell lines and ATL cells, and the induction of cIAP2 by Tax is mediated through NF-κB. Knockdown of cIAP2 increased apoptosis of HTLV-1-infected cells [411]. Expression of cIAP2 is also induced in HTLV-1 infected CD8(+) T cells through the NF-κB pathway, rendering the cells resistant to Fas-mediated cell death [412]. The Tax-mediated induction of X-chromosome-linked inhibitor of apoptosis (XIAP) through NF-κB also contributes to cell survival [183]. Survivin belongs to the IAP family and suppresses the activation of caspases. Expression of survivin is increased in HTLV-1-infected T-cell lines and primary ATL cells, with levels correlating with ATL prognosis [413,414]. Introduction of Tax into the IL-2-dependent T-cell line CTLL-2 induces the expression of survivin and protects cells from apoptosis, conferring IL-2-independent growth [415]. Tax increases the expression of cellular FLICE-inhibitory protein (c-FLIP), which binds to and inhibits Fas-associated proteins with the death domain (FADD) through NF-κB, thereby suppressing Fas-mediated apoptosis [416,417,418].
Ubiquitin-editing enzyme A20 not only suppresses the activation of NF-κB but also suppresses apoptosis by inhibiting caspase activation. In endothelial cells, A20 suppresses apoptosis induced by TNF, Fas, and natural killer cells by inhibiting the activation of caspase 8 [419]. Tax induces the expression of A20 through NF-κB in Jurkat T cells [420], and A20 is abundantly expressed in primary ATL cells [421]. A20 stably interacts with caspase 8 and FADD in HTLV-1-infected cells and suppresses caspase activation and apoptosis independent of its ubiquitin-editing catalytic activity [421].
Tax also facilitates cell survival through protein interaction. Tax induces lysine 63 (K63)-linked polyubiquitination of the myeloid cell leukemia sequence 1 (Mcl-1), an anti-apoptotic Bcl-2 family member, depending on TNF receptor-associated factor 6 (TRAF6). This stabilizes the Mcl-1 protein, thereby playing an important role in the survival of HTLV-1-transformed cells [422].
Multidrug resistance protein 1 (MDR1), also known as ATP-binding cassette sub-family B member 1 (ABCB1) or permeability glycoprotein (P-glycoprotein, P-gp), is a membrane glycoprotein that pumps many foreign substances out of cells, including anticancer drugs, thereby playing an important role in chemoresistance. HTLV-1-infected cell lines and primary ATL cells show enhanced MDR1 expression, and Tax activates the MDR1 promoter through an NF-IL6-binding site [423,424,425], suggesting the possibility that Tax-mediated induction of MDR1 may play a role in the drug resistance of ATL cells. Expression of Tax in S1T cells, a leukemic non-Tax-producing T-cell clone established from an ATL patient, conferred resistance to doxorubicin, etoposide, and vindesine but not to cisplatin. Expression of the lung resistance-related protein (LRP) gene was increased, but those of MDR1, multidrug resistance-associated protein 1 (MRP1), and multi-specific organic anion transporter and canalicular (cMOAT) genes were not significantly changed [426], suggesting that Tax-mediated drug resistance of ATL cells is due to LRP and not MDR1 in this context.
Activation of these anti-apoptotic genes by Tax, mainly through the NF-κB pathway, and the enhancement of their activity are thought to play crucial roles in the survival of HTLV-1-infected cells. ATL is known to be extremely resistant to chemotherapy, which may be due, at least in part, to the Tax-mediated promotion of cell survival.
Intriguingly, Tax can also activate some pro-apoptotic genes and facilitate apoptosis under some experimental conditions. Expression of Tax in Jurkat T cells induces significant apoptosis concomitant with increased Fas ligand (FasL) and TNF-related apoptosis-inducing ligand (TRAIL) gene expression [190,427,428]. Tax activates the FasL promoter through the CREB and NF-κB pathways [190] and the TRAIL gene via NF-κB [428].
The genes among those mentioned above, which have been reported to be induced by Tax, are summarized in Table 4.

4.5. Human Telomerase Reverse Transcriptase (hTERT)

Most somatic cells have a finite replicative life span (40~50 cell divisions) determined by the length of telomere DNA, which is shortened every cell division due to the end replication problem. Hence, the expression of human telomerase reverse transcriptase (hTERT) is required for the immortalization of cancer cells to maintain telomere length.
Although initial evidence suggested Tax represses hTERT promoter activity [429], later studies indicate that Tax activates the hTERT gene. Infection of primary T cells with HTLV-1 induces hTERT expression, and Tax activates the hTERT promoter through the NF-κB pathway [430]. Intriguingly, T cells are unique among somatic cells in that growth stimulation, such as treatment with IL-2, induces the expression of hTERT [431]. Tax also induces expression of the hTERT gene in human peripheral blood leukocytes. Interestingly, hTERT gene expression is tightly linked with progression from the G1 to S phase of the cell cycle, suggesting that it is regulated by transcription factor(s) associated with the G1 to S phase transition [432]. Consistent with this, in HTLV-1-positive CD4(+) cells cloned from patients, hTERT expression parallels Tax expression and cell cycling [433]. Transcription-independent activation of hTERT enzymatic activity by Tax is also suggested to be mediated through the activation of the NF-κB pathway [434]. Induction of hTERT by Tax is thought to be important for the immortalization of HTLV-1-infected T cells.

4.6. Genes Involved in Suppression of DNA Repair and Induction of Genome Instability

Although Tax plays a crucial role in the early-stage transformation of HTLV-1-infected T cells, malignant transformation of the virus-infected cells is thought to require additional genetic changes. Indeed, ATL cells possess a variety of gene mutations involved in growth signal transduction, the maintenance of cell survival, and other pathways [59,60,61]. Tax increases the frequency of cellular DNA mutations and causes genomic instability by the induction of target gene expression and protein interactions to generate mutations required for malignant transformation of the virus-infected cells [435]. Tax-expressing cells show defective DNA repair, especially base-excision repair of oxidative damage [436], and are more likely to undergo gene amplification than control cells [437]. Tax also increases genetic instability by inducing DNA double strand breaks (DDSBs) during DNA replication and switches DNA repair from homologous recombination (HR) to the non-homologous end-joining (NHEJ) pathway, which is error prone, through activation of the NF-κB pathway [438]. These effects of Tax increase genetic and genome instability, which may be responsible for the accumulation of mutations, leading to malignant transformation [63].
Tax activates the proliferating cell nuclear antigen (PCNA) gene and increases PCNA protein expression [439,440,441], likely by relieving repression of the PCNA promoter [442]. Tax-expressing cells show a reduced capacity for nucleotide excision repair (NER), which correlates with the ability of Tax to activate the PCNA gene. Moreover, over-expression of PCNA reduces NER [440,441]. These observations suggest that Tax-mediated over-expression of PCNA reduces NER and predisposes cells to accumulate mutations, leading to the transformation of HTLV-1-infected cells.
Nitric oxide synthase (NOS) catalyzes the production of nitric oxide (NO). The inducible nitric oxide synthase (iNOS) gene is over-expressed in HTLV-1-infected and Tax-expressing cells, and Tax activates the iNOS promoter [443,444]. Tax induces iNOS expression in the human monoblast cell line, U937 [445]. Tax expression causes DNA double strand breaks (DSBs), concomitant with increased levels of NO and the induction of iNOS. Conversely, the inhibition of NO production dramatically reduces DSBs [446,447]. Taken together, these observations suggest that Tax promotes DSBs through, at least in part, activation of the iNOS gene. Tax also induces DNA damage by increasing reactive oxygen species (ROS) [448,449] and promotes cell cycle progression in the presence of DNA damage by suppressing checkpoints, thereby increasing genetic instability [450].
R-loops are structures consisting of a duplex of transcribed RNA and template DNA and an unpaired non-templated DNA strand. R-loops can be formed by invasion of nascent RNA into the DNA duplex to displace the non-template strand. Hence, highly transcribed regions tend to form R-loops. R-loop excision by nucleotide excision repair (NER) promotes DNA double-strand breaks (DSBs). The Tax-mediated constitutive activation of NF-κB increases R-loop accumulation and DNA double-strand breaks. Accordingly, defects in nucleotide excision repair are selected in adult T-cell leukemia [451,452].
Tax induces and activates WIP1 phosphatase, which dephosphorylates γH2AX and marks the location of damaged DNA, thereby mitigating the cellular response to DNA damage [453]. Consistent with this, Tax+Wip1−/− mice show a statistically significant reduced prevalence of tumorigenesis compared to their Tax+Wip1+/+ counterparts [454].
Tax also facilitates genome instability by protein interaction. Tax interacts with checkpoint kinase 2 (Chk2) and co-localizes with Chk2 and p53-binding protein 1 (p53BP1) [455]. Tax also interacts with ataxia-telangiectasia mutated (ATM), Chk1, and Chk2 and attenuates the DNA damage response [456,457,458]. Tax associates with minichromosome maintenance proteins (MCMs) and accelerates the replication timing, thereby generating replicative stress [459]. Tax binds to and inhibits topoisomerase I, suggesting the possibility that Tax affects cellular processes such as the transcription and maintenance of genomic stability in which DNA topoisomerase I is involved [460,461]. Tax forms microscopically visible nuclear speckles, termed Tax-speckle structures (TSS), depending on the activity of RNF8, which sequesters DNA damage response factors such as breast cancer 1 (BRCA1), DNA-dependent protein kinase (DNA-PK), and mediator of DNA damage checkpoint protein 1 (MDC1), thereby compromising DSB repair [462].
The genes among those mentioned above, which have been reported to be induced by Tax, are summarized in Table 5.

4.7. MicroRNAs (miRNAs)

MicroRNAs (miRNAs) are short non-coding RNAs that bind complementary mRNAs and silence gene expression by RNA interference. Transcribed precursor miRNAs are processed by ribonucleases DROSHA and DICER to generate mature miRNAs and are assembled into RNA-induced silencing complexes (RISCs) with Argonaute proteins. RISCs bind to mRNAs through complementary miRNA and degrade or inhibit the translation of target mRNAs [463,464,465,466]. Since the length of miRNAs is relatively short (around 22 base), a miRNA can affect multiple targets with similar sequences and effect their growth-promoting and tumor-suppressive functions. miRNAs targeting growth-promoting genes can function as tumor suppressors, while those targeting tumor suppressive genes can function as onco-miRNA. Several miRNAs are reported to be deregulated in HTLV-1-infected and ATL cells, and the involvement of miRNA deregulation in HTLV-1-mediated transformation has been suggested [29,467].
Several miRNAs including miR-34a, miR-146a, miR-155, miR-93, and miR-130b are upregulated in HTLV-1-infected T cells, mostly through the NF-κB pathway. miR-34a is a direct target gene of the tumor suppressor p53 and ectopic miR-34a expression induces apoptosis, cell cycle arrest, or senescence. Hence miR-34a expression is suppressed in many types of cancers [468,469,470]. Intriguingly, however, miR-34a expression is increased in HTLV-1-infected cell lines and ATL cells compared to normal CD4+ T cells [471]. The regulatory region of miR-34a contained NF-κB binding motifs and a pharmacological inhibitor of NF-κB, Bay 11-7082, reduced miR-34a expression, suggesting that increased expression of miR-34a in HTLV-1-infected cells may be due to Tax-mediated activation of the NF-κB pathway. Introduction of an miR-34a mimic reduces levels of the pro-apoptotic Bcl-2 family member Bax, whereas sequestration of miR-34a with a sponge construct increases cell death. These findings suggest that the increased expression of miR-34a may support the survival of HTLV-1-infected cells in part by suppressing Bax expression. MiR-146a and miR-155 are also reported to be upregulated in HTLV-1-infected cells through the Tax-mediated activation of the NF-κB pathway [472,473], and knockdown of miR-146a or miR-155 reduced growth of HTLV-1-infected cells. In liver cancer stem cells, increased expression of miR-155 targets tumor protein p53-induced nuclear protein 1 (p53INP1), thereby contributing to the growth of the cells [474]. The expression of miR-93 and miR-130b is also increased in HTLV-1-transformed T-cell lines and ATL cells [475]. Both of these miRNAs target tumor protein 53-induced nuclear protein 1 (TP53INP1), and its expression is reduced in HTLV-1-transformed cells. Accordingly, the knockdown of miR-93 and miR-130b increases TP53INP1 expression and apoptosis, implicating the involvement of miR-93 and miR-130b in proliferation and survival of HTLV-1-infected cells.
Tax also affects miRNA biogenesis by interacting with Drosha in the nucleus, facilitating its proteasome-mediated degradation [476]. Reduced expression of Drosha somehow facilitates the replication of HTLV-1.
The miRNAs among those mentioned above, which have been reported to be induced by Tax, are summarized in Table 6.

4.8. Epigenetic Genes

Tax also has epigenetic effects and ATL cells show characteristic epigenetic changes [477,478]. ATL genomes are characterized by the prominent methylation of CpG islands in promoter regions, and the levels of hyper methylation correlated with ATL development and progression [479]. HTLV-1-infected cells show a differentially methylated position such as thymocyte-expressed molecule (THEMIS), leukocyte-associated immunoglobulin-like receptor 1 (LAIR1), and ring-type E3 ubiquitin transferase 130 (RNF130), which negatively regulate T-cell receptor (TCR) signaling [479]. Cyclin-dependent kinase inhibitor 2A (CDKN2A), Kruppel-like factor 4 (KLF4), early growth response 3 (EGR3), and bone morphogenetic protein 6 (BMP6) are also hyper-methylated in ATL cells [480,481,482]. In addition to CBP/p300, P/CAF, and HDACs, Tax interacts with multiple histone-modifying enzymes such as histone deacetylase NAD-dependent deacetylase sirtuin-1 (SIRT1), histone methyltransferases SUV39H1, and SET and MYND domain-containing protein (SMYD3) [483,484,485]. SIRT1 represses HTLV-1 viral gene expression but does not affect NF-κB-mediated transcriptional activation [483]. Binding of SUV39H1 to Tax represses the transcriptional activity of Tax, suggesting negative feedback [484]. The binding of SMYD3 to Tax enhances the Tax-mediated activation of NF-κB [485]. HTLV-1 gene expression is also under the control of epigenetic regulation. The HTLV-1 provirus genome has a CCCTC-binding factor (CTCF) binding site (CTCF-BS) at a sharp border in the pX region in T cells naturally infected with HTLV-1 [486]. CTCF is a zinc finger protein that binds to an insulator region in genomic DNA and plays a fundamental role in controlling higher order chromatin structure and gene expression. The binding of CTCF to the CTCF-BS of HTLV-1 provirus genome induces the hyper-methylation of DNA with repressive epigenetic marks upstream of CTCF-BS and the hypo-methylation of DNA with activating epigenetic marks downstream of CTCF-BS [486,487]. This leads to the suppression of expression of the positive strand, including Tax, while allowing the expression of the negative strand coding for HBZ. Introduction of the viral CTCF-BS into the host cellular genome also modulates cellular gene expression by forming loops between the viral CTCF-BS and cellular CTCF-BS [488,489], which may modulate host cell gene expression and facilitate transformation induced by HTLV-1 infection. Tax is able to activate transcription from the LTR, even when it is heavily methylated. The methyl-CpG-binding domain 2 (MBD2) protein interacts with Tax, and MBD2 enhances the transcriptional activity of Tax against the methylated LTR [490].
Tax-mediated induction of epigenetic regulators provides an additional means by which Tax can modulate gene transcription in HTLV-1-infected cells.
Protein arginine methyltransferase 5 (PRMT5) catalyzes the formation of monomethylarginine or symmetric dimethylarginine in a variety of proteins including histones. Symmetric methylation of histone H4 on Arg3 by PRMT5 generates repressive marks on chromatin, which recruits DNA methyltransferase DNMT3A to the promoter, followed by DNA methylation and gene silencing [491]. PRMT5 expression is upregulated during HTLV-1-mediated T-cell transformation. Knockdown of PRMT5 expression or pharmacological inhibition of PRMT5 activity in HTLV-1-infected T cells resulted in increased viral gene expression and decreased cellular proliferation [492,493], suggesting involvement of PRMT5 in HTLV-1-mediated T-cell transformation.
The enhancer of zeste homolog 2 (EZH2) is the catalytic component of Polycomb repressive complex 2 (PRC2), which methylates Lys27 of HisH3 and generates a repressive mark on chromatin. Expression of EZH2 is enhanced in HTLV-1-infected cells, likely due to the Tax-mediated activation of the EZH2 promoter through the NF-κB pathway [494,495]. PRC2-mediated trimethylation at histone H3Lys27 (H3K27me3) is enhanced in ATL cells, concomitant with ATL-specific gene expression changes that included several tumor suppressors, transcription factors, epigenetic modifiers, miRNAs, and developmental genes [495]. EZH2 enhances TGF-β signaling by facilitating the association between Smad3 and Smad4. In particular, EZH2 enhances the TGF-β-induced expression of Foxp3, the master regulator of regulatory T cells (Tregs), suggesting the involvement of EZH2 in the differentiation of regulatory T cells [495]. EZH2 is also involved in the repression of N-Myc downstream-regulated gene (NDRG) in HTLV-1-infected T cells [496].
Coactivator-associated arginine methyltransferase 1 (CARM1/PRMT4) methylates Arg17 of histone H3, which serves as an epigenetic mark for transcriptional activation. Tax associates with CARM1, and CARM1 facilitates the Tax-mediated activation of HTLV-1 LTR [497], suggesting that CARM1 enhances Tax trans-activation of the HTLV-1 LTR through a direct interaction between CARM1 and Tax. Intriguingly, Tax also activated the CARM1 gene through the NF-κB pathway and enhances methylation of Arg17 of the histone H3 on the target gene IL-2Rα. Moreover, knockdown of CARM1 reduced the Tax-mediated expression of IL-2Rα and cyclin D2 genes and cell cycle progression [498], suggesting that the Tax-mediated induction of CARM1 gene expression contributes to Tax-mediated target gene expression and cell cycle promotion.
The genes among those mentioned above, which have been reported to be induced by Tax, are summarized in Table 7.

4.9. Genes Involved in Cellular Metabolism

Cancer cells require increased metabolism to support rapid cell proliferation.
Disialoganglioside (GD2) is expressed in tumors of neuroectodermal origin. ATL cells express GD2 after in vitro culture in the presence of IL-2, concomitant with the expression of Tax. PBLs infected with Tax-expressing retrovirus also express GD2, indicating that Tax is responsible for GD2 induction. Tax increases β-1,4-N-acetylgalactosaminyltransferase (GM2/GD2 synthase) mRNA, suggesting that the induction of GD2 by Tax is due to the new synthesis of GD2 as a result of increased transcription of the D2 synthase gene [494,499].
Thioredoxin (TRX) plays an important role in cellular protection against oxidative stress through thiol-reducing activity [500]. TRX was originally identified as a growth factor derived from ATL cells [501,502], which facilitates autocrine cell growth [503]. Tax-mediated apoptosis in activated T cells requires an enhanced intracellular prooxidant state [504]. Tax induces the expression of TRX by activating the TRX promoter [505]. The Tax-induced expression of TRX may contribute to reduced oxidative stress induced by Tax.
Dihydroorotate dehydrogenase (DHODH) catalyzes the fourth reaction in the de novo pyrimidine biosynthesis pathway. HTLV-1-infected T-cell lines show a higher expression of DHODH, which is enhanced by Tax. Knockdown or pharmacological inhibition of DHODH decreases HTLV-1-infected T-cell growth and apoptosis, suggesting DHODH as a potential therapeutic target for treating ATL [506].
The genes among those mentioned above, which have been reported to be induced by Tax, are summarized in Table 8.

4.10. Genes Involved in Viral Transmission

Cell-free transmission of HTLV-1 is inefficient and HTLV-1 is usually transmitted from cell to cell. Tax-expressing HTLV-1-infected cells show an increased expression of intercellular adhesion molecule 1 (ICAM-1, CD54) [192,507], and Tax-induction in JPX-9 cells increased the expression of ICAM-1 [191,192]. Tax activates the ICAM-1 promoter through NF-κB s and CRE-like sites [191,508,509]. ICAM-1 and its receptor LFA-1 are involved in the syncytium formation induced in the co-culture of HTLV-1-positive and HTLV-1-negative human T-cell lines [507], suggesting that induction of ICAM-1 by Tax may facilitate infection of HTLV-1 from the virus-producing cells to uninfected cells. This cell-to-cell contact, termed the virological synapse, is accompanied by the re-orientation of the microtubule-organizing center (MTOC) in the infected T cell toward the cell contact region with the noninfected target cell. Tax and Tax-induced ICAM-1 co-operate to induce re-orientation of the MTOC through the cyclic AMP-binding protein-dependent pathway and Ras-MEK-ERK signaling pathway, respectively [510].
TNF-α-induced protein 2 (TNFAIP2), also known as M-Sec, plays a critical role in the cell-to-cell transmission of the virus. When purified and briefly cultured, CD4+ T cells of HTLV-1 carriers express M-Sec, which is mediated by Tax. Knockdown or pharmacological inhibition of M-Sec reduces viral infection in multiple co-culture conditions and an in vivo mouse model of HTLV-1 infection. Knockdown or inhibition of M-Sec reduces not only plasma membrane protrusions and the migratory activity of cells, but also decreases large clusters of Gag, a viral structural protein required for the formation of viral particles. Taken together, these results suggest that M-Sec, induced by Tax, mediates efficient cell-to-cell viral infection [511].
HTLV-1-infected CD4 T cells secrete a potent chemoattractant, leukotriene B4 (LTB4), to facilitate the encounter between virus-infected and target cells. LTB4 secretion is dependent on the Tax-mediated induction of phospholipase A2γ. Inhibition of LTB4 secretion or knockdown of the LTB4 receptor on target cells reduces T-cell recruitment, cellular contact formation, and virus propagation in vitro and in vivo, indicating critical roles of LTB4 in HTLV-1 transmission [512].
HTLV-1 is also transmitted vertically via breast milk. Lactoferrin facilitates the replication of HTLV-1 in lymphocytes derived from asymptomatic HTLV-1 carriers and the transmission to cord blood lymphocytes in vitro. Lactoferrin can transactivate the HTLV-1 long terminal repeat promoter [513]. In addition, HTLV-1 infection can induce lactoferrin gene expression, and Tax trans-activates the lactoferrin gene promoter. Mutual interaction between HTLV-1 and lactoferrin would facilitate and enhance the milk-borne transmission of this virus [514].
Galectins are a class of proteins that bind specifically to β-galactoside sugars, such as N-acetyllactosamine. Galectin-1 is a galactoside-binding protein which is secreted from activated T lymphocytes and increases cell-to-cell and cell-to-matrix contact. Expression of galectin-1 is enhanced in HTLV-1-infected T cells, and Tax activates the galectin-1 gene through the NF-κB pathway. The addition of galectin-1 increases cell fusion in a co-culture assay with HTLV-1-infected cells and potentiates the infectivity of HTLV-1 envelope pseudo-typed HIV-1 virions [515]. These observations suggest that Tax-mediated galectin-1 expression may facilitate HTLV-1 infection. Galectin-3 is a β-galactoside-binding lectin, and its expression is associated with various physiological and pathological processes, including cell growth, tumor transformation, and metastasis. Galectin-3 expression is enhanced in HTLV-1-infected human T-cell lines, and Tax activates the galectin-3 promoter through the CREB and NF-κB pathways. Galectin-3 activates IL-2 production in Jurkat T cells [516].
The viral gene product p8, a proteolytically cleaved product of p12, also plays an important role in viral transmission by increasing cellular conduits and virus transmission [12,517].
The genes among those mentioned above, which have been reported to be induced by Tax, are summarized in Table 9.

4.11. Genes Involved in Invasion and Infiltration

As mentioned above, since CCR4 is involved in T-cell migration into the skin, enhanced CCR4 expression in HTLV-1-infected cells may explain the frequent infiltration of ATL cells into skin [251,252].
Tax induces the expression of activated cell adhesion molecules (ALCAMs) through the canonical NF-κB pathway. ALCAM blockade or knockdown reduces the migration of HTLV-1-infected lymphocytes through a layer of endothelial cells of the blood–brain barrier (BBB), suggesting involvement of ALCAM in HAM/TSP pathogenesis [518].
Tax activates the vimentin gene through the NF-κB pathway [519,520]. Vimentin is an intermediate filament protein whose expression is induced upon epithelial–mesenchymal transition (EMT). The role of Tax-induced expression of vimentin in T cells is yet to be elucidated.
HTLV-1-infected T cells express the fucosyltransferase (Fuc-T) VII gene involved in the biosynthesis of the leukocyte sialyl Lewis X, which may be related to tissue infiltration in patients with ATL. Tax activates the Fuc-T VII promoter through the CREB pathway [521,522].
One characteristic of ATL is the infiltration of various tissues by circulating leukemic cells. Matrix metalloproteinases (MMPs), which mediate the degradation of the basement membrane and extracellular matrix, play an important role in metastasis and tumor cell dissemination. HTLV-1-infected T cells express high levels of MMP-7 and MMP-9 compared with uninfected T cells. Tax activates the MMP-7 and MMP-9 genes through JunD/AP-1 and NF-κB/SP-1, respectively [523,524]. Plasma levels of MMP-9 correlated with organ involvement in ATL patients, suggesting that the enhanced expression of MMP-9 may be responsible for the invasiveness of ATL cells [524]. The tissue inhibitor of matrix metalloproteinases-1 (TIMP-1) has a growth factor-like activity for erythroid cells and potentially HTLV-1-infected T cells. HTLV-1-infected cell lines express high levels of TIMP-1 mRNA, and Tax activates the TIMP-1 promoter in a human T-cell line (Jurkat) through the AP-1 binding site [525]. Tax-mediated induction of TIMP-1 may facilitate the growth of HTLV-1-infected cells.
HTLV-1-transformed cells and ATL cells show an enhanced expression of the actin-bundling protein fascin, and Tax activates the fascin gene through the NF-κB pathway [526,527]. Fascin also localizes to invadopodia, membrane protrusions formed at the adherent cell surface that facilitate extracellular matrix (ECM) invasion, providing a potential molecular mechanism for how fascin increases the invasiveness of cancer cells.
Recruitment of HTLV-1-infected T lymphocytes into the central nervous system (CNS) is an essential step in the development of virus-associated neuroinflammatory diseases. Collapsin response mediator protein 2 (CRMP2), originally identified as a semaphorin-signaling transducer in the nervous system, is also present in T cells and regulates T-cell polarization and migration [528]. Infection of T-cells with HTLV-1 elevates CRMP2 expression through Tax-mediated activation and increases T lymphocyte migration [529]. These observations suggest the involvement of CRMP2 in the infiltration of HTLV-1-infected lymphocytes into various tissues.
Fibronectin (FN) is a glycoprotein, which is a major component of the extracellular matrix. HTLV-1- infected T cells show enhanced expression of FN, and Tax trans-activates the FN promoter in Jurkat cells through the NF-κB pathway [530].
The genes among those mentioned above, which have been reported to be induced by Tax, are summarized in Table 10.

5. Tax-Mediated Repression of Gene Expression

Tax not only activates transcription but also indirectly represses gene expression, which may have negative effects on the HTLV-1-mediated transformation.

5.1. Tax Inactivates the Tumor Suppressor p53

The tumor suppressors p53 and pRB play major roles in tumorigenesis, and in almost all cancers, the functions of both are disabled [531]. Tax not only inactivates pRB to facilitate cell growth but also inactivates p53 to maintain cell viability. HTLV-1-transformed cells contain higher amounts of wildtype p53, but its transcriptional activity is suppressed. Tax stabilizes and inactivates the p53 function by interfering with the activity of the N-terminal activation domain [532]. Tax-induced p53 inactivation is dependent upon the ability of Tax to induce NF-κB and can be inhibited by a constitutively active IκB mutant in the human Jurkat T-cell line. This Tax-mediated p53 inactivation is not due to p300 squelching [533,534], since the over-expression of p300 does not restore p53 activity in the presence of Tax. Tax cannot inhibit p53 function in p65/RelA knockout MEFs, indicating critical roles of p65/RelA in the Tax-mediated inactivation of p53 [535]. Intriguingly, the expression of exogenous p300 in H1299 cells allows for the full recovery of p53 trans-activation in the presence of Tax, suggesting that Tax utilizes distinct cell-type-specific mechanisms to inhibit p53 function [536]. HTLV-1-transformed cells show an interaction between p65/RelA and p53 that is increased by Tax and correlates with Tax-mediated p53 inhibition. In HTLV-1-transformed cells, p53 and p65/RelA form a complex on the p53 target murine double minute 2 (MDM2) promoter. These observations suggest that binding of the p65/RelA subunit of NF-κB to p53 compromises its transcriptional activity [537]. Tax-induced p65/RelA Ser-536 phosphorylation through IKKβ plays a crucial role in this process [538]. Tax also activates AKT and facilitates the IKKβ phosphorylation of p65/RelA, which is critical for NF-κB-mediated p53 inhibition [539]. Repression of p73β, a family member of p53, by Tax through competition for the cellular coactivator CBP/p300 is also reported [113].

5.2. Pro-Apoptotic Genes

Bcl-2-associated X protein (Bax) is a member of the pro-apoptotic Bcl-2 family, which facilitates apoptosis by destabilizing the mitochondrial outer membrane. Tax suppresses Bax gene expression through repression of the Bax promoter [540]. The Bax gene is a target of p53, raising the possibility that repression of the Bax promoter by Tax may be mediated, at least in part, by the suppression of p53 activity. Bax mRNA is a target of miR-34a, which is induced by Tax through the NF-κB pathway [471], which may also explain the reduced level of Bax gene expression in HTLV-1-infected cells.
HTLV-1-infected leukemic cells show reduced expressions of proapoptotic BH3-only proteins, such as the Bcl-2-interacting mediator of cell death (Bim) and BH3-interacting domain death agonist (Bid). Tax downregulates Bim and Bid expression at the transcriptional level through increased hypoxia-inducible factor-1α (HIF-1α) [541], suggesting that Tax may also promote cell survival by suppressing expression of the proapoptotic BH3-only proteins Bim and Bid.
FOXO3a is a forkhead family transcription factor which plays an important role in the induction of apoptosis by upregulating BIM and PUMA and downregulating FLIP. Expression of Tax in CD4+ T cells by the infection of HTLV-1 or lentiviral-mediated introduction induces AKT activation, leading to the phosphorylation and inactivation of FOXO3a. Inhibition of FOXO3a enables the long-term survival of a population of highly activated, terminally differentiated CD4+ T cells. Pharmacological inhibition of AKT activity compromises the persistence of terminally differentiated CD4+ T cells, suggesting the roles of Tax-mediated suppression of FOXO3a in the long-term maintenance of Tax+CD4+ T lymphocytes during the early-stage transformation of HTLV-1-infected cells [542].
N-myc downstream-regulated gene 2 (NDRG2) is a candidate tumor suppressor that is frequently downregulated in HTLV-1-infected cells. Tax induces the expression of NDRG2 through the NF-κB pathway. However, NDRG2 expression is suppressed in HTLV-1-infected cell lines and various types of ATL cells, likely due to methylation of the NDRG2 promoter. EZH2, a member of the polycomb family, is increased in ATL cells, likely through the activation of NF-κB by Tax [494,495]. EZH2 binds to the NDRG2 promoter and induces DNA methylation of the NDRG2 promoter. EZH2-mediated downregulation of NDRG2 may contribute to the enhanced activation of signaling pathways during ATL development [496].

5.3. Signal Transducers

HTLV-1-transformed cell lines and ATL cells show elevated levels of cAMP, a biological feature of regulatory T-cell populations [543,544]. Continued expression of Tax, but not transient expression, elevates cAMP concentrations with concomitant reduction in the cAMP-degrading phosphodiesterase 3B (PDE3B) [545]. Decreased expression of PDE3B is associated with inhibitory histone modifications at the PDE3B locus [545].
In contrast to Src family tyrosine kinases Fyn and Lyn, which are induced by Tax [339,546], Lck expression is reduced in HTLV-1-infected cells [39]. Tax decreases Lck mRNA levels in Jurkat T cells and inhibits Lck promoter activity through an E-box [187]. Hence, Tax transcriptionally represses Lck gene expression. In addition to Lck, the expression of Zap-70, which relays T-cell activation signals, is also suppressed in HTLV-1-infected T cells, and Tax suppresses Zap-70 expression in Jurkat T cells [546].
Src homology-2-containing protein-tyrosine phosphatase 1 (SHP-1), also known as protein-tyrosine phosphatase non-receptor type 6 (PTPN6), is a candidate tumor suppressor that counteracts the action of tyrosine kinases, which play central roles in growth factor signal transduction. Expression of SHP-1 is repressed in HTLV-1-transformed T-cell lines and fresh ATL cells. Consistent with this, Tax represses SHP-1 P2 promoter activity. NF-κB plays an important role in both basal P2 promoter activity and Tax-induced promoter silencing (TIPS). NF-κB dissociates from the SHP-1 P2 promoter following the binding of Tax and HDAC1, suggesting that Tax recruits HDAC1 to the SHP-1 P2 promoter and forms an inhibitory complex that results in the deacetylation and dissociation of NF-κB from the promoter and attenuation of SHP-1 expression [547]. Moreover, in HTLV-1-transformed cells, the SHP-1 P2 promoter is heavily CpG methylated and silenced [548]. These observations suggest that TIPS is mediated by dissociation of transcription factors from promoters, followed by subsequent DNA methylation.
Tax transgenic mice under the control of the HTLV-1 LTR develop neurofibroma [73]. The neurofibromatosis type I (NF1) gene is a tumor suppressor gene involved in familial cancer syndrome neurofibromatosis type I (NF1). Tax represses NF1 gene expression through suppression of the NF1 promoter, thereby allowing the development of neurofibromas without the need for direct mutations in the NF1 gene [549].

5.4. Cell Cycle Regulators

Tax directly binds to inactivates CDK inhibitors p16Ink4a and p15Ink4b [400,402]. In contrast, Tax inhibits p18Ink4c activity by suppressing its expression. HTLV-1-infected T-cell lines express very low levels of p18Ink4c compared with HTLV-1-negative cell lines [388], and Tax represses p18Ink4c expression through an E-box [402]. This may facilitate the activation of CDK4 and CDK6 to promote cell cycle progression.
Tax represses the cyclin A promoter through a CREB/ATF site in ts13 adherent cells and continuously growing Jurkat T cells [550]. However, in IL-2-starved Kit 225 cells, in which Tax can promote cell cycle progression through the NF-κB pathway, Tax induces the cyclin A gene expression [373], consistent with the roles of cyclin A in the S-G2/M progression of the cell cycle.

5.5. Transcription Factors

Helicase-like transcription factor (HLTF) is a member of the SWI/SNF family, which have helicase and ATPase activities and regulate the transcription of genes by altering the chromatin structure. Tax represses HLTF transcription via EZH2 and also facilitates HLTF proteasomal degradation by direct interaction [551]. Restoring normal levels of HLTF expression enhances autophagy and increases the generation of defective virions, suggesting a role for reduced HLTF expression mediated by Tax in the normal replication of HTLV-1 [551].
Type I interferons (IFNα and IFNβ) are key effectors of the innate immune response against viral infection. Interferon regulatory factors (IRFs) play central roles in the activation of IFN genes. Virus infection is detected by pattern recognition receptors (PRRs), including various types of Toll-like receptors (TLRs) and activates IRFs by phosphorylation through IRF kinases, such as TANK-binding kinase 1 (TBK1), transforming growth factor beta-activated kinase 1 (TAK1) and IKKε. Apparently, contradictory observations have been reported for the effects on Tax on IFN gene expression. Tax activates TAK1, and knockdown of TAK1 reduces IFN-regulated gene expression [552]. Tax associates with IKKε and TBK1, which phosphorylate and activate IRF3 and enhance IFNβ promoter activity [553]. In contrast, Tax binds to and inhibits TBK1, thereby suppressing IRF3 phosphorylation and induction of the IFN genes [554], supporting the roles of Tax in suppression of the IFN response in HTLV-1-infected cells. ZNF268 is a zinc finger protein of the Kruppel-related family, which interacts with TBK1 to facilitate IRF3 activation during viral infection. ZNF268 mRNA was repressed in HTLV-1-infected cells, and Tax represses ZNF268 expression through the CREB/ATF pathway [555]. Expression of Tax in Jurkat T cells induces ICSAT/Pip/LSIRF, which negatively regulates the activity of interferon-regulated genes [556], supporting the concept that Tax negatively regulates the IFN response.
B-cell CLL/Lymphoma 11B (BCL11B) is a zinc finger transcription factor which is expressed in all T-cell compartments and is indispensable for T-lineage development. HTLV-1-infected T-cell lines show a reduced expression of BCL11B, and Tax suppresses the expression of BCL11B [557], suggesting an inhibitory role for BCL11B in the oncogenesis of HTLV-1-infected T cells.
GATA-binding protein 3 (GATA3) is a transcription factor whose expression is high in naïve and Th2 T cells but low in Th1 T cells. GATA3 mRNA expression is decreased in HTLV-1 carriers and in patients with HTLV-1-related diseases [558]. Tax suppresses the GATA3 promoter through interactions with the repressor ZEB and Sp1 [559]. Intriguingly, genetic mutations in GATA3 are reported [59,60,61], suggesting that the loss of function of GATA3 may favor the pathogenesis of ATL.
Tax represses the transcriptional activity of MyoD through competitive binding to the KIX domain of p300 with MyoD [560], providing a potential mechanism for the Tax-mediated repression of the basic helix-loop-helix transcription factors.
The c-Myb proto-oncogene is preferentially expressed in hematopoietic lineages and highly expressed in several leukemia types. The B-myb gene was identified on the basis of its homology with c-Myb. B-Myb plays an important role in the transition from the G1 to S phase of the cell cycle. Both Tax and c-Myb interact with the KIX domain of CBP, and Tax and c-Myb antagonize each other in transcriptional activation, suggesting that Tax suppresses the transcriptional activity of c-Myb by competing for CBP [561]. However, Tax mutant M22, which can bind to CBP, cannot repress c-Myb, and Tax mutant V89A, which cannot bind to CBP, can still suppress c-Myb, indicating that competition for CBP/p300 is not the underlying mechanism of Tax-mediated repression of c-Myb [562]. Tax suppression of c-Myb leads to the trans-repression of the c-Myb promoter through the Myb responsive elements in Jurkat T cells [562]. Not only c-Myb but also B-Myb gene expression is suppressed in HTLV-1-transformed cells, and the induction of Tax in JPX-9 results in the suppression of both the c-Myb and B-Myb gene expression [562,563]. The ability of Tax to repress c-Myb parallels that to activate NF-κB, as shown by Tax mutants. In addition, IκBα and NEMO knockout compromise the Tax-mediated repression of c-Myb. NF-κB-mediated repression of c-Myb seems to be due to the sequestration of CBP/p300 by RelA, since mutant p300, lacking the C/H1 and KIX domains, and unable to bind RelA, retains the ability to stimulate c-Myb transcription and prevents NF-κB-mediated repression [564].

5.6. Others

HTLV-1-infected cells show defects in DNA repair, especially base-excision repair of oxidative damage [436]. DNA polymerase β (Pol β) is involved in the base-excision repair (BER) [565], and Tax downregulated its expression [566]. A consensus bHLH binding site in the DNA polymerase β promoter mediates repression by Tax [567]. This may contribute to the increased mutation rate in Tax-expressing cells.
The pre-T-cell receptor α (pTCRα) gene encodes a polypeptide which associates with the TCRα chain and CD3 molecules to form the pre-TCR complex. The surface expression of the pre-TCR is pTCRα dependent, and signaling through this complex triggers an early αβ T-cell developmental checkpoint inside the thymus, known as β-selection. Tax suppresses the E47-mediated activation of the pTCRα promoter and the pTCRα gene expression, which is dependent on its ability to interact with p300. The expression of Tax in human immature thymocytes decreases the pTCRα gene transcription. These observations indicate that Tax, by silencing E47, downregulates pTCRα gene transcription during early thymocyte development [568].
The expressions of miR-149, miR-873, and miR-223 have been reported to be downregulated in HTLV-1-infected cells [472,569]. miR-149 and miR-873 target p300 and P/CAF, which play crucial roles in the Tax-mediated activation of target genes.
The genes among those mentioned above, which have been reported to be downregulated by Tax, are summarized in Table 11.

6. Conclusions

Tax activates cellular transcription factors through the interaction with a plethora of cellular factors and induces the expression of a variety of genes involved in cell proliferation, cell survival, immortalization, immune response, invasion, genome instability, and others, thereby contributing to the early-stage transformation of HTLV-1-infected cells (Figure 6). Induction of many of the genes crucial for transformation is mediated through the NF-κB pathway. Tax also promotes transformation through protein–protein interactions. Genome instability induced by Tax leads to the accumulation of mutations, which contribute to malignant transformation.

7. Future Perspectives

Tax is a pleiotropic protein with multiple functions such as transcriptional activation, modulation of signal transduction pathway, promotion of cell proliferation, suppression of apoptosis and promotion of cell survival, production of ROS, induction of genomic instability, which leads to the accumulation of genetic mutations, and the induction of epigenetic changes and others. These effects co-operatively or additively contribute to the early-stage transformation of virus-infected cells by facilitating cell proliferation, survival, and immune evasion, thereby enabling persistency and the gradual increase in the virus-infected cells. These effects of Tax are exerted through the direct interaction of Tax with cellular factors and indirectly through the induction of target gene expression. Further understanding of each of these functions of Tax and identification of whole repertoire of transcriptional target genes of Tax will deepen our understanding of the mechanisms of leukemogenesis induced by HTLV-1 infection and will pave the way to develop new therapeutic approaches.
The receptors for HTLV-1 identified to date include the glucose transporter GLUT-1 [570], neuropilin 1 (NRP1) [571,572], and heparan sulfate proteoglycans [572,573,574], which are widely expressed among cell types. Consistent with this, HTLV-1 can infect cell types other than T cells, such as B cells, monocytes, endothelial cells, and fibroblasts in vitro. However, HTLV-1 infection specifically causes T-cell malignancy in vivo. The mechanism underlying T-cell-specific transformation by HTLV-1 has yet to be elucidated and would be an important issue to address in future studies. Intriguingly, there seems to be T-cell-specific effects of Tax in transcriptional regulation. Tax can promote cell cycle progression in resting T cells concomitant with the induction of cyclin D2 and CDK6 expression through the NF-κB pathway [71,373,378]. In contrast, Tax cannot induce cyclin D2 or CDK6 nor induce cell proliferation in fibroblasts, whereas the over-expression of cyclin D2 and CDK6 can induce cell cycle progression [378]. In addition, Tax inactivates p53 through the activation of NF-κB in T-cells, whereas in adherent cells, Tax acts via sequestration of CBP/p300 [535,536,537]. These observations suggest the presence of cell-type-specific effects of Tax in regulation of the NF-κB pathway. Elucidation of the entire picture of the cell-type-specific effects of Tax and the underlying mechanisms would be critical to understand cell-type-specific transformation induced by HTLV-1 infection.

Author Contributions

Conceptualization, M.S., R.N. and K.O.; resources, Y.Z. and M.F.; writing—original draft preparation, M.S.; writing—review and editing, R.I., A.P.B., K.K., K.A. and K.O.; visualization, R.N.; supervision, K.O.; project administration, K.O. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

R.N. is supported by scholarships from TOBE MAKI Scholarship Foundation, Yoshida Scholar-ship Foundation (Doctor 21), and research grants from ANRI fellowship and The Japan Science Society (The Sasagawa Scientific Research Grant). Y.Z. is supported by scholarships from Tohka Education Culture Exchange Foundation. M.F. is supported by a scholarship from the Ministry of Education, Culture, Sports, Science and Technology in Japan.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ABCB1ATP-binding cassette sub-family B member 1
ACasymptomatic carrier
ALCAMactivated cell adhesion molecule
AP-1activator protein 1
Apaf-1apoptotic protease activating factor 1
APCCdc20Cdc20-associated anaphase promoting complex
ATFactivating transcription factor
ATLadult T-cell leukemia/lymphoma
ATMataxia-telangiectasia mutated
BaxBcl-2-associated X protein
BBBblood–brain barrier
BCLB cell CLL/lymphoma
Bcl-xLBcl2-related protein, long isoform
BERbase excision repair
β-TrCPβ-transducin repeat-containing protein
bFGFbasic fibroblast growth factor
BFL1BCL2-related protein A1
BH3Bcl-2 homology 3
bHLHbasic helix-loop-helix
BidBH3-interacting domain death agonist
BimBcl-2-interacting mediator of cell death
BIRC3baculoviral IAP repeat-containing 3
BMPbone morphogenetic protein
BRF1butyrate response factor 1
BRG1BRM/SWI2-related gene 1
bZIPbasic leucine zipper
CADM-1cell adhesion molecule 1
CAKCDK activating kinase
cAMPcyclic adenosine monophosphate
CARM1coactivator-associated arginine methyltransferase 1
Cas-LCrk-associated substrate lymphocyte type
CBPCREB binding protein
CCLCC chemokine
CCRC-C chemokine receptor
CDcluster of differentiation
CDKcyclin-dependent kinase
CDKN2Acyclin-dependent kinase inhibitor 2A
c-FLIPcellular FLICE-inhibitory protein
cMOATmultispecific organic anion transporter, canalicular
CREcAMP responsive element
CREBCRE binding factor
CRM1chromosomal region maintenance 1
CRMP2collapsin response mediator protein 2
CTCFCCCTC-binding factor
CTCF-BSCTCF binding site
CTDC-terminal domain
CXCLchemokine, CXC motif, ligand
CXCRC-X-C chemokine receptor
DNA-PKDNA-dependent protein kinase
DCdendritic cell
DDSBDNA double strand break
DG2disialoganglioside
DHHdesert hedgehog
DHODHdihydroorotate dehydrogenase
Egrearly growth response
EVCEllis Van Creveld
EZH2enhancer of zeste homolog 2
FADDfas-associated protein with death domain
FasLfas ligand
Fbw7F-box and WD repeat domain containing 7
FNfibronectin
Fuc-Tfucosyltransferase
GATA3GATA-binding protein 3
G-CSFgranulocyte colony-stimulating factor
GLUT1glucose transporter 1
GM2/GD2 synthaseβ-1,4-N-acetylgalactosaminyltransferase
GSK3βglycogen synthase kinase-3β
HAMHTLV-1-associated myelopathy
HBZHTLV-1 basic leucine zipper factor
HDAC1histone deacetylase 1
Hhhedgehog
HIF-1αhypoxia-inducible factor-1α
HLTFhelicase-like transcription factor
HRhomologous recombination
hTERThuman telomerase reverse transcriptase
HTLV-1human T-cell leukemia virus type 1
IAPinhibitors of apoptosis
ICAM-1intercellular adhesion molecule 1
IFNinterferon
IGSFimmunoglobulin superfamily
IHHIndian hedgehog
ILinterleukin
IκBinhibitor of NF-κB
IKKIκB kinase
iNOSinducible nitric oxide synthase
IRFinterferon regulatory factor
JAG1jagged
JAKJanus kinase
KIDkinase-inducible domain
KIXKID interacting domain
KLF4Kruppel-like factor 4
LAIR1leukocyte-associated immunoglobulin-like receptor 1
LFAlymphocyte function-associated antigen
LRPlung resistance-related protein
LTRlong terminal repeat
MAGImembrane-associated guanylate kinase, WW and PDZ domains-containing
MAP3Kmitogen-activated protein kinase kinase kinase
Mat1menage a trois homolog 1
MBD2methyl-CpG-binding domain 2
Mcl-1myeloid cell leukemia sequence 1
MCPmonocyte chemoattractant protein
MDC1mediator of DNA damage checkpoint protein 1
MDM2murine double minute 2
MDR1multidrug resistance protein 1
MEKK1MAP/ERK kinase kinase 1
MHCmajor histocompatibility complex, class I
MIPmacrophage inflammatory protein
miRNAmicroRNA
MMPmatrix metalloproteinase
MRP1multidrug resistance-associated protein 1
MTOCmicrotubule-organizing center
NAP1nucleosome assembly protein 1
NDRGN-Myc downstream-regulated gene
NEMONF-κB essential modulator
NERnucleotide excision repair
NF1neurofibromatosis type I
NF-κBnuclear factor κB
NGFnerve growth factor
NHEJnon-homologous end joining
NICDnotch intra-cellular domain
NIKNF-κB-inducing kinase
NOnitric oxide
NODnonobese diabetic
NOSnitric oxide synthase
NRP1neuropilin 1
OPNosteopontin
ORP4LOSBP-related protein 4
OSBP2oxysterol-binding protein 2
p53BP1p53-binding protein 1
PBLperipheral blood lymphocyte
PBMPDZ domain-binding motif
PBMCperipheral blood mononuclear cell
P/CAFp300/CBP-associated factor
PCNAproliferating cell nuclear antigen
PDE3Bphosphodiesterase 3B
P-gppermeability glycoprotein
PHAphytohemagglutinin
PI3Kphosphatidylinositol 3-kinases
PI(3,4,5)P3phosphatidylinositol (3,4,5)-trisphosphate
PLA2G4Cphospholipase A2, group IVC
PLCG1phospholipase C γ 1
PMAphorbol 12-myristate 13-acetate
PRC2polycomb repressive complex 2
PRMT5protein arginine methyltransferase 5
pTCRαpre-T-cell receptor α
PTHrPparathyroid hormone-related peptide
P-TEFbpositive transcription elongation factor
PTPN6protein-tyrosine phosphatase, nonreceptor-type, 6
PUMAp53-upregulated modulator of apoptosis
pXprotein coding region X
RArheumatoid arthritis
RANTESregulated on activation, normal T cell expressed and secreted
RBP-Jκrecombination signal Jκ binding protein
RGMarepulsive guidance molecule A
RISCRNA-induced silencing complex
RNF8ring finger protein 8
RNF130ring-type E3 ubiquitin transferase 130
ROSreactive oxygen species
SCIDsevere combined immunodeficiency
SHHSonic hedgehog
SHP-1Src homology-2-containing protein-tyrosine phosphatase 1
SIRT1sirtuin-1
SMYD3SET and MYND domain containing protein
SREserum response elements
SRFserum-responsive factor
STATsignal transducer and activator of transcription
STLVsimian T-lymphotropic virus
TAB2TAK1-binding protein 2
TAK1TGF-β-activating kinase 1
Taxtrans-activator of pX region
TAXBP1Tax binding protein 1
TBK1TANK-binding kinase 1
TCFternary complex factor
TCRT-cell receptor
TFtranscription factor
TFIIHtranscription factor II H
TGF-βtransforming growth factor β
THEMISthymocyte-expressed molecule
TIMP-1tissue inhibitor of matrix metalloproteinases-1
TIPStax-induced promoter silencing
TNFtumor necrosis factor
TNFAIP2TNF-α-induced protein 2
TNFRSFtumor necrosis factor receptor superfamily
TORCtransducers of regulated CREB
TPp53INP1tumor protein p53-induced nuclear protein 1
TRAF6TNF receptor-associated factor 6
TRAILtumor necrosis factor-related apoptosis-inducing ligand
Tregregulatory T-cell
TRXthioredoxin
TSLC1tumor suppressor in lung cancer 1
TSPtropical spastic paraparesis
TSStax-speckle structures
VCAM-1vascular cell adhesion molecule 1
VEGFvascular endothelial growth factor
v-oncviral oncogene
WIP1wildtype p53-induced phosphatase 1
XIAPX-chromosome-linked inhibitor of apoptosis
Zap-70ζ-chain-associated protein kinase
ZNF268zinc finger protein 268
XPO1exportin 1

References

  1. Poiesz, B.J.; Ruscetti, F.W.; Gazdar, A.F.; Bunn, P.A.; Minna, J.D.; Gallo, R.C. Detection and isolation of type C retrovirus particles from fresh and cultured lymphocytes of a patient with cutaneous T-cell lymphoma. Proc. Natl. Acad. Sci. USA 1980, 77, 7415–7419. [Google Scholar] [CrossRef]
  2. Yoshida, M.; Miyoshi, I.; Hinuma, Y. Isolation and characterization of retrovirus from cell lines of human adult T-cell leukemia and its implication in the disease. Proc. Natl. Acad. Sci. USA 1982, 79, 2031–2035. [Google Scholar] [CrossRef]
  3. Hinuma, Y.; Nagata, K.; Hanaoka, M.; Nakai, M.; Matsumoto, T.; Kinoshita, K.I.; Shirakawa, S.; Miyoshi, I. Adult T-cell leukemia: Antigen in an ATL cell line and detection of antibodies to the antigen in human sera. Proc. Natl. Acad. Sci. USA 1981, 78, 6476–6480. [Google Scholar] [CrossRef]
  4. Uchiyama, T. Human T cell leukemia virus type I (HTLV-I) and human diseases. Annu. Rev. Immunol. 1997, 15, 15–37. [Google Scholar] [CrossRef]
  5. Yoshida, M. Discovery of HTLV-1, the first human retrovirus, its unique regulatory mechanisms, and insights into pathogenesis. Oncogene 2005, 24, 5931–5937. [Google Scholar] [CrossRef]
  6. Forlani, G.; Shallak, M.; Accolla, R.S.; Romanelli, M.G. HTLV-1 infection and pathogenesis: New insights from cellular and animal models. Int. J. Mol. Sci. 2021, 22, 8001. [Google Scholar] [CrossRef] [PubMed]
  7. Uchiyama, T.; Yodoi, J.; Sagawa, K.; Takatsuki, K.; Uchino, H. Adult T-cell leukemia: Clinical and hematologic features of 16 cases. Blood 1977, 50, 481–492. [Google Scholar] [CrossRef] [PubMed]
  8. Takatsuki, K. Discovery of adult T-cell leukemia. Retrovirology 2005, 2, 16. [Google Scholar] [CrossRef] [PubMed]
  9. Gessain, A.; Cassar, O. Epidemiological aspects and world distribution of HTLV-1 infection. Front. Microbiol. 2012, 3, 388. [Google Scholar] [CrossRef]
  10. Igakura, T.; Stinchcombe, J.C.; Goon, P.K.; Taylor, G.P.; Weber, J.N.; Griffiths, G.M.; Tanaka, Y.; Osame, M.; Bangham, C.R. Spread of HTLV-I between lymphocytes by virus-induced polarization of the cytoskeleton. Science 2003, 299, 1713–1716. [Google Scholar] [CrossRef]
  11. Pais-Correia, A.M.; Sachse, M.; Guadagnini, S.; Robbiati, V.; Lasserre, R.; Gessain, A.; Gout, O.; Alcover, A.; Thoulouze, M.I. Biofilm-like extracellular viral assemblies mediate HTLV-1 cell-to-cell transmission at virological synapses. Nat. Med. 2010, 16, 83–89. [Google Scholar] [CrossRef]
  12. Van Prooyen, N.; Gold, H.; Andresen, V.; Schwartz, O.; Jones, K.; Ruscetti, F.; Lockett, S.; Gudla, P.; Venzon, D.; Franchini, G. Human T-cell leukemia virus type 1 p8 protein increases cellular conduits and virus transmission. Proc. Natl. Acad. Sci. USA 2010, 107, 20738–20743. [Google Scholar] [CrossRef]
  13. Pique, C.; Jones, K.S. Pathways of cell-cell transmission of HTLV-1. Front. Microbiol. 2012, 3, 378. [Google Scholar] [CrossRef]
  14. Gross, C.; Thoma-Kress, A.K. Molecular mechanisms of HTLV-1 cell-to-cell transmission. Viruses 2016, 8, 74. [Google Scholar] [CrossRef]
  15. Osame, M.; Usuku, K.; Izumo, S.; Ijichi, N.; Amitani, H.; Igata, A.; Matsumoto, M.; Tara, M. HTLV-I associated myelopathy, a new clinical entity. Lancet 1986, 1, 1031–1032. [Google Scholar] [CrossRef]
  16. Gessain, A.; Barin, F.; Vernant, J.C.; Gout, O.; Maurs, L.; Calender, A.; de The, G. Antibodies to human T-lymphotropic virus type-I in patients with tropical spastic paraparesis. Lancet 1985, 2, 407–410. [Google Scholar] [CrossRef]
  17. Bangham, C.R.; Araujo, A.; Yamano, Y.; Taylor, G.P. HTLV-1-associated myelopathy/tropical spastic paraparesis. Nat. Rev. Dis. Primers 2015, 1, 15012. [Google Scholar] [CrossRef] [PubMed]
  18. Martin, F.; Taylor, G.P.; Jacobson, S. Inflammatory manifestations of HTLV-1 and their therapeutic options. Expert. Rev. Clin. Immunol. 2014, 10, 1531–1546. [Google Scholar] [CrossRef] [PubMed]
  19. Varmus, H. Retroviruses. Science 1988, 240, 1427–1435. [Google Scholar] [CrossRef] [PubMed]
  20. Leib-Mosch, C.; Brack-Werner, R.; Salmons, B.; Schmidt, J.; Strauss, P.G.; Hehlmann, R.; Erfle, V. The significance of retroviruses in oncology. Onkologie 1990, 13, 405–414. [Google Scholar] [CrossRef]
  21. Burmeister, T. Oncogenic retroviruses in animals and humans. Rev. Med. Virol. 2001, 11, 369–380. [Google Scholar] [CrossRef]
  22. Seiki, M.; Hattori, S.; Hirayama, Y.; Yoshida, M. Human adult T-cell leukemia virus: Complete nucleotide sequence of the provirus genome integrated in leukemia cell DNA. Proc. Natl. Acad. Sci. USA 1983, 80, 3618–3622. [Google Scholar] [CrossRef]
  23. Seiki, M.; Eddy, R.; Shows, T.B.; Yoshida, M. Nonspecific integration of the HTLV provirus genome into adult T-cell leukaemia cells. Nature 1984, 309, 640–642. [Google Scholar] [CrossRef]
  24. Seiki, M.; Hikikoshi, A.; Taniguchi, T.; Yoshida, M. Expression of the pX gene of HTLV-I: General splicing mechanism in the HTLV family. Science 1985, 228, 1532–1534. [Google Scholar] [CrossRef]
  25. Ciminale, V.; Pavlakis, G.N.; Derse, D.; Cunningham, C.P.; Felber, B.K. Complex splicing in the human T-cell leukemia virus (HTLV) family of retroviruses: Novel mRNAs and proteins produced by HTLV type I. J. Virol. 1992, 66, 1737–1745. [Google Scholar] [CrossRef]
  26. Bai, X.T.; Nicot, C. Overview on HTLV-1 p12, p8, p30, p13: Accomplices in persistent infection and viral pathogenesis. Front. Microbiol. 2012, 3, 400. [Google Scholar] [CrossRef] [PubMed]
  27. Giam, C.Z.; Semmes, O.J. HTLV-1 infection and adult T-cell leukemia/lymphoma-A tale of two proteins: Tax and HBZ. Viruses 2016, 8, 161. [Google Scholar] [CrossRef] [PubMed]
  28. Harhaj, E.W.; Giam, C.Z. NF-κB signaling mechanisms in HTLV-1-induced adult T-cell leukemia/lymphoma. FEBS J. 2018, 285, 3324–3336. [Google Scholar] [CrossRef] [PubMed]
  29. Fochi, S.; Ciminale, V.; Trabetti, E.; Bertazzoni, U.; D’Agostino, D.M.; Zipeto, D.; Romanelli, M.G. NF-κB and microRNA deregulation mediated by HTLV-1 Tax and HBZ. Pathogens 2019, 8, 290. [Google Scholar] [CrossRef]
  30. Mohanty, S.; Harhaj, E.W. Mechanisms of oncogenesis by HTLV-1 Tax. Pathogens 2020, 9, 543. [Google Scholar] [CrossRef]
  31. Ernzen, K.J.; Panfil, A.R. Regulation of HTLV-1 transformation. Biosci. Rep. 2022, 42, BSR20211921. [Google Scholar] [CrossRef]
  32. Bellon, M.; Nicot, C. HTLV-1 Tax tug-of-war: Cellular senescence and death or cellular transformation. Pathogens 2024, 13, 87. [Google Scholar] [CrossRef]
  33. Sodroski, J.G.; Rosen, C.A.; Haseltine, W.A. Trans-acting transcriptional activation of the long terminal repeat of human T lymphotropic viruses in infected cells. Science 1984, 225, 381–385. [Google Scholar] [CrossRef]
  34. Fujisawa, J.; Seiki, M.; Kiyokawa, T.; Yoshida, M. Functional activation of the long terminal repeat of human T-cell leukemia virus type I by a trans-acting factor. Proc. Natl. Acad. Sci. USA 1985, 82, 2277–2281. [Google Scholar] [CrossRef]
  35. Qu, Z.; Xiao, G. Human T-cell lymphotropic virus: A model of NF-κB-associated tumorigenesis. Viruses 2011, 3, 714–749. [Google Scholar] [CrossRef]
  36. Currer, R.; Van Duyne, R.; Jaworski, E.; Guendel, I.; Sampey, G.; Das, R.; Narayanan, A.; Kashanchi, F. HTLV tax: A fascinating multifunctional co-regulator of viral and cellular pathways. Front. Microbiol. 2012, 3, 406. [Google Scholar] [CrossRef]
  37. Fochi, S.; Mutascio, S.; Bertazzoni, U.; Zipeto, D.; Romanelli, M.G. HTLV Deregulation of the NF-κB pathway: An update on Tax and antisense proteins role. Front. Microbiol. 2018, 9, 285. [Google Scholar] [CrossRef] [PubMed]
  38. Grassmann, R.; Berchtold, S.; Radant, I.; Alt, M.; Fleckenstein, B.; Sodroski, J.G.; Haseltine, W.A.; Ramstedt, U. Role of human T-cell leukemia virus type 1 X region proteins in immortalization of primary human lymphocytes in culture. J. Virol. 1992, 66, 4570–4575. [Google Scholar] [CrossRef] [PubMed]
  39. Akagi, T.; Ono, H.; Shimotohno, K. Characterization of T cells immortalized by Tax1 of human T-cell leukemia virus type 1. Blood 1995, 86, 4243–4249. [Google Scholar] [CrossRef] [PubMed]
  40. Akagi, T.; Ono, H.; Nyunoya, H.; Shimotohno, K. Characterization of peripheral blood T-lymphocytes transduced with HTLV-I Tax mutants with different trans-activating phenotypes. Oncogene 1997, 14, 2071–2078. [Google Scholar] [CrossRef]
  41. Robek, M.D.; Ratner, L. Immortalization of CD4+ and CD8+ T lymphocytes by human T-cell leukemia virus type 1 Tax mutants expressed in a functional molecular clone. J. Virol. 1999, 73, 4856–4865. [Google Scholar] [CrossRef]
  42. Robek, M.D.; Ratner, L. Immortalization of T lymphocytes by human T-cell leukemia virus type 1 is independent of the tax-CBP/p300 interaction. J. Virol. 2000, 74, 11988–11992. [Google Scholar] [CrossRef]
  43. Ratner, L.; Portis, T.; Robek, M.; Harding, J.; Grossman, W. Studies of the immortalizing activity of HTLV type 1 Tax, using an infectious molecular clone and transgenic mice. AIDS Res. Hum. Retroviruses 2000, 16, 1647–1651. [Google Scholar] [CrossRef] [PubMed]
  44. Takeda, S.; Maeda, M.; Morikawa, S.; Taniguchi, Y.; Yasunaga, J.; Nosaka, K.; Tanaka, Y.; Matsuoka, M. Genetic and epigenetic inactivation of tax gene in adult T-cell leukemia cells. Int. J. Cancer 2004, 109, 559–567. [Google Scholar] [CrossRef] [PubMed]
  45. Zhi, H.; Yang, L.; Kuo, Y.L.; Ho, Y.K.; Shih, H.M.; Giam, C.Z. NF-κB hyper-activation by HTLV-1 tax induces cellular senescence, but can be alleviated by the viral anti-sense protein HBZ. PLoS Pathog. 2011, 7, e1002025. [Google Scholar] [CrossRef] [PubMed]
  46. Ho, Y.K.; Zhi, H.; DeBiaso, D.; Philip, S.; Shih, H.M.; Giam, C.Z. HTLV-1 tax-induced rapid senescence is driven by the transcriptional activity of NF-κB and depends on chronically activated IKKα and p65/RelA. J. Virol. 2012, 86, 9474–9483. [Google Scholar] [CrossRef] [PubMed]
  47. Gaudray, G.; Gachon, F.; Basbous, J.; Biard-Piechaczyk, M.; Devaux, C.; Mesnard, J.M. The complementary strand of the human T-cell leukemia virus type 1 RNA genome encodes a bZIP transcription factor that down-regulates viral transcription. J. Virol. 2002, 76, 12813–12822. [Google Scholar] [CrossRef]
  48. Lemasson, I.; Lewis, M.R.; Polakowski, N.; Hivin, P.; Cavanagh, M.H.; Thebault, S.; Barbeau, B.; Nyborg, J.K.; Mesnard, J.M. Human T-cell leukemia virus type 1 (HTLV-1) bZIP protein interacts with the cellular transcription factor CREB to inhibit HTLV-1 transcription. J. Virol. 2007, 81, 1543–1553. [Google Scholar] [CrossRef]
  49. Zhao, T.; Yasunaga, J.; Satou, Y.; Nakao, M.; Takahashi, M.; Fujii, M.; Matsuoka, M. Human T-cell leukemia virus type 1 bZIP factor selectively suppresses the classical pathway of NF-κB. Blood 2009, 113, 2755–2764. [Google Scholar] [CrossRef]
  50. Mesnard, J.M.; Barbeau, B.; Cesaire, R.; Peloponese, J.M. Roles of HTLV-1 basic Zip factor (HBZ) in viral chronicity and leukemic transformation. Potential new therapeutic approaches to prevent and treat HTLV-1-related diseases. Viruses 2015, 7, 6490–6505. [Google Scholar] [CrossRef]
  51. Matsuoka, M.; Mesnard, J.M. HTLV-1 bZIP factor: The key viral gene for pathogenesis. Retrovirology 2020, 17, 2. [Google Scholar] [CrossRef]
  52. Giam, C.Z. HTLV-1 replication and adult T cell leukemia development. Recent. Results Cancer Res. 2021, 217, 209–243. [Google Scholar] [CrossRef]
  53. Zhang, W.; Nisbet, J.W.; Albrecht, B.; Ding, W.; Kashanchi, F.; Bartoe, J.T.; Lairmore, M.D. Human T-lymphotropic virus type 1 p30II regulates gene transcription by binding CREB binding protein/p300. J. Virol. 2001, 75, 9885–9895. [Google Scholar] [CrossRef]
  54. Nicot, C.; Dundr, M.; Johnson, J.M.; Fullen, J.R.; Alonzo, N.; Fukumoto, R.; Princler, G.L.; Derse, D.; Misteli, T.; Franchini, G. HTLV-1-encoded p30II is a post-transcriptional negative regulator of viral replication. Nat. Med. 2004, 10, 197–201. [Google Scholar] [CrossRef]
  55. Green, P.L. HTLV-1 p30II: Selective repressor of gene expression. Retrovirology 2004, 1, 40. [Google Scholar] [CrossRef] [PubMed]
  56. Andresen, V.; Pise-Masison, C.A.; Sinha-Datta, U.; Bellon, M.; Valeri, V.; Washington Parks, R.; Cecchinato, V.; Fukumoto, R.; Nicot, C.; Franchini, G. Suppression of HTLV-1 replication by Tax-mediated rerouting of the p13 viral protein to nuclear speckles. Blood 2011, 118, 1549–1559. [Google Scholar] [CrossRef] [PubMed]
  57. Harrod, R. Silencers of HTLV-1 and HTLV-2: The pX-encoded latency-maintenance factors. Retrovirology 2019, 16, 25. [Google Scholar] [CrossRef] [PubMed]
  58. Moles, R.; Sarkis, S.; Galli, V.; Omsland, M.; Purcell, D.F.J.; Yurick, D.; Khoury, G.; Pise-Masison, C.A.; Franchini, G. p30 protein: A critical regulator of HTLV-1 viral latency and host immunity. Retrovirology 2019, 16, 42. [Google Scholar] [CrossRef]
  59. Kataoka, K.; Nagata, Y.; Kitanaka, A.; Shiraishi, Y.; Shimamura, T.; Yasunaga, J.; Totoki, Y.; Chiba, K.; Sato-Otsubo, A.; Nagae, G.; et al. Integrated molecular analysis of adult T cell leukemia/lymphoma. Nat. Genet. 2015, 47, 1304–1315. [Google Scholar] [CrossRef]
  60. Kogure, Y.; Kataoka, K. Genetic alterations in adult T-cell leukemia/lymphoma. Cancer Sci. 2017, 108, 1719–1725. [Google Scholar] [CrossRef]
  61. Sakihama, S.; Morichika, K.; Saito, R.; Miyara, M.; Miyagi, T.; Hayashi, M.; Uchihara, J.; Tomoyose, T.; Ohshiro, K.; Nakayama, S.; et al. Genetic profile of adult T-cell leukemia/lymphoma in Okinawa: Association with prognosis, ethnicity, and HTLV-1 strains. Cancer Sci. 2021, 112, 1300–1309. [Google Scholar] [CrossRef]
  62. Chlichlia, K.; Khazaie, K. HTLV-1 Tax: Linking transformation, DNA damage and apoptotic T-cell death. Chem. Biol. Interact. 2010, 188, 359–365. [Google Scholar] [CrossRef]
  63. Nicot, C. HTLV-I Tax-mediated inactivation of cell cycle checkpoints and DNA repair pathways contribute to cellular transformation: “A Random Mutagenesis Model”. J. Cancer Sci. 2015, 2. [Google Scholar] [CrossRef]
  64. Boxus, M.; Twizere, J.C.; Legros, S.; Dewulf, J.F.; Kettmann, R.; Willems, L. The HTLV-1 Tax interactome. Retrovirology 2008, 5, 76. [Google Scholar] [CrossRef]
  65. Azran, I.; Schavinsky-Khrapunsky, Y.; Aboud, M. Role of Tax protein in human T-cell leukemia virus type-I leukemogenicity. Retrovirology 2004, 1, 20. [Google Scholar] [CrossRef]
  66. Tanaka, A.; Takahashi, C.; Yamaoka, S.; Nosaka, T.; Maki, M.; Hatanaka, M. Oncogenic transformation by the tax gene of human T-cell leukemia virus type I in vitro. Proc. Natl. Acad. Sci. USA 1990, 87, 1071–1075. [Google Scholar] [CrossRef] [PubMed]
  67. Smith, M.R.; Greene, W.C. Type I human T cell leukemia virus tax protein transforms rat fibroblasts through the cyclic adenosine monophosphate response element binding protein/activating transcription factor pathway. J. Clin. Investig. 1991, 88, 1038–1042. [Google Scholar] [CrossRef]
  68. Matsumoto, K.; Shibata, H.; Fujisawa, J.I.; Inoue, H.; Hakura, A.; Tsukahara, T.; Fujii, M. Human T-cell leukemia virus type 1 Tax protein transforms rat fibroblasts via two distinct pathways. J. Virol. 1997, 71, 4445–4451. [Google Scholar] [CrossRef]
  69. Neuveut, C.; Low, K.G.; Maldarelli, F.; Schmitt, I.; Majone, F.; Grassmann, R.; Jeang, K.T. Human T-cell leukemia virus type 1 Tax and cell cycle progression: Role of cyclin D-cdk and p110Rb. Mol. Cell Biol. 1998, 18, 3620–3632. [Google Scholar] [CrossRef] [PubMed]
  70. Schmitt, I.; Rosin, O.; Rohwer, P.; Gossen, M.; Grassmann, R. Stimulation of cyclin-dependent kinase activity and G1- to S-phase transition in human lymphocytes by the human T-cell leukemia/lymphotropic virus type 1 Tax protein. J. Virol. 1998, 72, 633–640. [Google Scholar] [CrossRef] [PubMed]
  71. Ohtani, K.; Iwanaga, R.; Arai, M.; Huang, Y.; Matsumura, Y.; Nakamura, M. Cell type-specific E2F activation and cell cycle progression induced by the oncogene product Tax of human T-cell leukemia virus type I. J. Biol. Chem. 2000, 275, 11154–11163. [Google Scholar] [CrossRef] [PubMed]
  72. Nerenberg, M.; Hinrichs, S.H.; Reynolds, R.K.; Khoury, G.; Jay, G. The tat gene of human T-lymphotropic virus type 1 induces mesenchymal tumors in transgenic mice. Science 1987, 237, 1324–1329. [Google Scholar] [CrossRef]
  73. Hinrichs, S.H.; Nerenberg, M.; Reynolds, R.K.; Khoury, G.; Jay, G. A transgenic mouse model for human neurofibromatosis. Science 1987, 237, 1340–1343. [Google Scholar] [CrossRef]
  74. Hall, A.P.; Irvine, J.; Blyth, K.; Cameron, E.R.; Onions, D.E.; Campbell, M.E. Tumours derived from HTLV-I tax transgenic mice are characterized by enhanced levels of apoptosis and oncogene expression. J. Pathol. 1998, 186, 209–214. [Google Scholar] [CrossRef]
  75. Grossman, W.J.; Kimata, J.T.; Wong, F.H.; Zutter, M.; Ley, T.J.; Ratner, L. Development of leukemia in mice transgenic for the tax gene of human T-cell leukemia virus type I. Proc. Natl. Acad. Sci. USA 1995, 92, 1057–1061. [Google Scholar] [CrossRef]
  76. Hasegawa, H.; Sawa, H.; Lewis, M.J.; Orba, Y.; Sheehy, N.; Yamamoto, Y.; Ichinohe, T.; Tsunetsugu-Yokota, Y.; Katano, H.; Takahashi, H.; et al. Thymus-derived leukemia-lymphoma in mice transgenic for the Tax gene of human T-lymphotropic virus type I. Nat. Med. 2006, 12, 466–472. [Google Scholar] [CrossRef] [PubMed]
  77. Rauch, D.; Gross, S.; Harding, J.; Niewiesk, S.; Lairmore, M.; Piwnica-Worms, D.; Ratner, L. Imaging spontaneous tumorigenesis: Inflammation precedes development of peripheral NK tumors. Blood 2009, 113, 1493–1500. [Google Scholar] [CrossRef]
  78. Ohsugi, T.; Kumasaka, T.; Okada, S.; Urano, T. The Tax protein of HTLV-1 promotes oncogenesis in not only immature T cells but also mature T cells. Nat. Med. 2007, 13, 527–528. [Google Scholar] [CrossRef] [PubMed]
  79. Niewiesk, S. Animals models of human T cell leukemia virus type I leukemogenesis. ILAR J. 2016, 57, 3–11. [Google Scholar] [CrossRef]
  80. Hasegawa, H.; Sano, K.; Ainai, A.; Suzuki, T. Application of HTLV-1 tax transgenic mice for therapeutic intervention. Adv. Biol. Regul. 2018, 68, 10–12. [Google Scholar] [CrossRef]
  81. El Hajj, H.; El-Sabban, M.; Hasegawa, H.; Zaatari, G.; Ablain, J.; Saab, S.T.; Janin, A.; Mahfouz, R.; Nasr, R.; Kfoury, Y.; et al. Therapy-induced selective loss of leukemia-initiating activity in murine adult T cell leukemia. J. Exp. Med. 2010, 207, 2785–2792. [Google Scholar] [CrossRef]
  82. Mahgoub, M.; Yasunaga, J.I.; Iwami, S.; Nakaoka, S.; Koizumi, Y.; Shimura, K.; Matsuoka, M. Sporadic on/off switching of HTLV-1 Tax expression is crucial to maintain the whole population of virus-induced leukemic cells. Proc. Natl. Acad. Sci. USA 2018, 115, E1269–E1278. [Google Scholar] [CrossRef] [PubMed]
  83. Hleihel, R.; Skayneh, H.; de The, H.; Hermine, O.; Bazarbachi, A. Primary cells from patients with adult T cell leukemia/lymphoma depend on HTLV-1 Tax expression for NF-κB activation and survival. Blood Cancer J. 2023, 13, 67. [Google Scholar] [CrossRef] [PubMed]
  84. Jegado, B.; Kashanchi, F.; Dutartre, H.; Mahieux, R. STLV-1 as a model for studying HTLV-1 infection. Retrovirology 2019, 16, 41. [Google Scholar] [CrossRef]
  85. Hussein, O.; Mahgoub, M.; Shichijo, T.; Nakagawa, S.; Tanabe, J.; Akari, H.; Miura, T.; Matsuoka, M.; Yasunaga, J.I. Evolution of primate T-cell leukemia virus type 1 accessory genes and functional divergence of its antisense proteins. PLoS Pathog. 2025, 21, e1013158. [Google Scholar] [CrossRef]
  86. Miura, M.; Yasunaga, J.; Tanabe, J.; Sugata, K.; Zhao, T.; Ma, G.; Miyazato, P.; Ohshima, K.; Kaneko, A.; Watanabe, A.; et al. Characterization of simian T-cell leukemia virus type 1 in naturally infected Japanese macaques as a model of HTLV-1 infection. Retrovirology 2013, 10, 118. [Google Scholar] [CrossRef]
  87. Suehiro, Y.; Hasegawa, A.; Iino, T.; Sasada, A.; Watanabe, N.; Matsuoka, M.; Takamori, A.; Tanosaki, R.; Utsunomiya, A.; Choi, I.; et al. Clinical outcomes of a novel therapeutic vaccine with Tax peptide-pulsed dendritic cells for adult T cell leukaemia/lymphoma in a pilot study. Br. J. Haematol. 2015, 169, 356–367. [Google Scholar] [CrossRef]
  88. Kannagi, M.; Hasegawa, A.; Nagano, Y.; Iino, T.; Okamura, J.; Suehiro, Y. Maintenance of long remission in adult T-cell leukemia by Tax-targeted vaccine: A hope for disease-preventive therapy. Cancer Sci. 2019, 110, 849–857. [Google Scholar] [CrossRef]
  89. Iino, T.; Hasegawa, A.; Matsutani, T.; Akashi, K.; Kannagi, M.; Suehiro, Y. Elimination of residual adult T-cell leukaemia clones by Tax-targeted dendritic cell vaccine. EJHaem 2025, 6, e1072. [Google Scholar] [CrossRef] [PubMed]
  90. Cann, A.J.; Rosenblatt, J.D.; Wachsman, W.; Shah, N.P.; Chen, I.S. Identification of the gene responsible for human T-cell leukaemia virus transcriptional regulation. Nature 1985, 318, 571–574. [Google Scholar] [CrossRef]
  91. Chen, I.S.; Slamon, D.J.; Rosenblatt, J.D.; Shah, N.P.; Quan, S.G.; Wachsman, W. The x gene is essential for HTLV replication. Science 1985, 229, 54–58. [Google Scholar] [CrossRef] [PubMed]
  92. Fujisawa, J.; Seiki, M.; Sato, M.; Yoshida, M. A transcriptional enhancer sequence of HTLV-I is responsible for trans-activation mediated by p40 chi HTLV-I. EMBO J. 1986, 5, 713–718. [Google Scholar] [CrossRef] [PubMed]
  93. Shimotohno, K.; Takano, M.; Teruuchi, T.; Miwa, M. Requirement of multiple copies of a 21-nucleotide sequence in the U3 regions of human T-cell leukemia virus type I and type II long terminal repeats for trans-acting activation of transcription. Proc. Natl. Acad. Sci. USA 1986, 83, 8112–8116. [Google Scholar] [CrossRef]
  94. Zhao, L.J.; Giam, C.Z. Interaction of the human T-cell lymphotrophic virus type I (HTLV-I) transcriptional activator Tax with cellular factors that bind specifically to the 21-base-pair repeats in the HTLV-I enhancer. Proc. Natl. Acad. Sci. USA 1991, 88, 11445–11449. [Google Scholar] [CrossRef]
  95. Zhao, L.J.; Giam, C.Z. Human T-cell lymphotropic virus type I (HTLV-I) transcriptional activator, Tax, enhances CREB binding to HTLV-I 21-base-pair repeats by protein-protein interaction. Proc. Natl. Acad. Sci. USA 1992, 89, 7070–7074. [Google Scholar] [CrossRef]
  96. Jeang, K.T.; Boros, I.; Brady, J.; Radonovich, M.; Khoury, G. Characterization of cellular factors that interact with the human T-cell leukemia virus type I p40x-responsive 21-base-pair sequence. J. Virol. 1988, 62, 4499–4509. [Google Scholar] [CrossRef]
  97. Suzuki, T.; Fujisawa, J.I.; Toita, M.; Yoshida, M. The trans-activator tax of human T-cell leukemia virus type 1 (HTLV-1) interacts with cAMP-responsive element (CRE) binding and CRE modulator proteins that bind to the 21-base-pair enhancer of HTLV-1. Proc. Natl. Acad. Sci. USA 1993, 90, 610–614. [Google Scholar] [CrossRef]
  98. Franklin, A.A.; Kubik, M.F.; Uittenbogaard, M.N.; Brauweiler, A.; Utaisincharoen, P.; Matthews, M.A.; Dynan, W.S.; Hoeffler, J.P.; Nyborg, J.K. Transactivation by the human T-cell leukemia virus Tax protein is mediated through enhanced binding of activating transcription factor-2 (ATF-2) ATF-2 response and cAMP element-binding protein (CREB). J. Biol. Chem. 1993, 268, 21225–21231. [Google Scholar] [CrossRef]
  99. Gachon, F.; Peleraux, A.; Thebault, S.; Dick, J.; Lemasson, I.; Devaux, C.; Mesnard, J.M. CREB-2, a cellular CRE-dependent transcription repressor, functions in association with Tax as an activator of the human T-cell leukemia virus type 1 promoter. J. Virol. 1998, 72, 8332–8337. [Google Scholar] [CrossRef] [PubMed]
  100. Gachon, F.; Thebault, S.; Peleraux, A.; Devaux, C.; Mesnard, J.M. Molecular interactions involved in the transactivation of the human T-cell leukemia virus type 1 promoter mediated by Tax and CREB-2 (ATF-4). Mol. Cell Biol. 2000, 20, 3470–3481. [Google Scholar] [CrossRef]
  101. Baranger, A.M.; Palmer, C.R.; Hamm, M.K.; Giebler, H.A.; Brauweiler, A.; Nyborg, J.K.; Schepartz, A. Mechanism of DNA-binding enhancement by the human T-cell leukaemia virus transactivator Tax. Nature 1995, 376, 606–608. [Google Scholar] [CrossRef]
  102. Adya, N.; Giam, C.Z. Distinct regions in human T-cell lymphotropic virus type I tax mediate interactions with activator protein CREB and basal transcription factors. J. Virol. 1995, 69, 1834–1841. [Google Scholar] [CrossRef]
  103. Yin, M.J.; Gaynor, R.B. Complex formation between CREB and Tax enhances the binding affinity of CREB for the human T-cell leukemia virus type 1 21-base-pair repeats. Mol. Cell Biol. 1996, 16, 3156–3168. [Google Scholar] [CrossRef]
  104. Yin, M.J.; Gaynor, R.B. HTLV-1 21 bp repeat sequences facilitate stable association between Tax and CREB to increase CREB binding affinity. J. Mol. Biol. 1996, 264, 20–31. [Google Scholar] [CrossRef] [PubMed]
  105. Lenzmeier, B.A.; Giebler, H.A.; Nyborg, J.K. Human T-cell leukemia virus type 1 Tax requires direct access to DNA for recruitment of CREB binding protein to the viral promoter. Mol. Cell Biol. 1998, 18, 721–731. [Google Scholar] [CrossRef] [PubMed]
  106. Youn, H.G.; Matsumoto, J.; Tanaka, Y.; Shimotohno, K. SR-related protein TAXREB803/SRL300 is an important cellular factor for the transactivational function of human T-cell lymphotropic virus type 1 Tax. J. Virol. 2003, 77, 10015–10027. [Google Scholar] [CrossRef] [PubMed]
  107. Kwok, R.P.; Laurance, M.E.; Lundblad, J.R.; Goldman, P.S.; Shih, H.; Connor, L.M.; Marriott, S.J.; Goodman, R.H. Control of cAMP-regulated enhancers by the viral transactivator Tax through CREB and the co-activator CBP. Nature 1996, 380, 642–646. [Google Scholar] [CrossRef] [PubMed]
  108. Giebler, H.A.; Loring, J.E.; van Orden, K.; Colgin, M.A.; Garrus, J.E.; Escudero, K.W.; Brauweiler, A.; Nyborg, J.K. Anchoring of CREB binding protein to the human T-cell leukemia virus type 1 promoter: A molecular mechanism of Tax transactivation. Mol. Cell Biol. 1997, 17, 5156–5164. [Google Scholar] [CrossRef]
  109. Kashanchi, F.; Duvall, J.F.; Kwok, R.P.; Lundblad, J.R.; Goodman, R.H.; Brady, J.N. The coactivator CBP stimulates human T-cell lymphotrophic virus type I Tax transactivation in vitro. J. Biol. Chem. 1998, 273, 34646–34652. [Google Scholar] [CrossRef]
  110. Geiger, T.R.; Sharma, N.; Kim, Y.M.; Nyborg, J.K. The human T-cell leukemia virus type 1 tax protein confers CBP/p300 recruitment and transcriptional activation properties to phosphorylated CREB. Mol. Cell Biol. 2008, 28, 1383–1392. [Google Scholar] [CrossRef]
  111. Harrod, R.; Tang, Y.; Nicot, C.; Lu, H.S.; Vassilev, A.; Nakatani, Y.; Giam, C.Z. An exposed KID-like domain in human T-cell lymphotropic virus type 1 Tax is responsible for the recruitment of coactivators CBP/p300. Mol. Cell Biol. 1998, 18, 5052–5061. [Google Scholar] [CrossRef]
  112. Vendel, A.C.; McBryant, S.J.; Lumb, K.J. KIX-mediated assembly of the CBP-CREB-HTLV-1 tax coactivator-activator complex. Biochemistry 2003, 42, 12481–12487. [Google Scholar] [CrossRef] [PubMed]
  113. Lemasson, I.; Nyborg, J.K. Human T-cell leukemia virus type I tax repression of p73β is mediated through competition for the C/H1 domain of CBP. J. Biol. Chem. 2001, 276, 15720–15727. [Google Scholar] [CrossRef] [PubMed]
  114. Scoggin, K.E.; Ulloa, A.; Nyborg, J.K. The oncoprotein Tax binds the SRC-1-interacting domain of CBP/p300 to mediate transcriptional activation. Mol. Cell Biol. 2001, 21, 5520–5530. [Google Scholar] [CrossRef] [PubMed]
  115. Ramirez, J.A.; Nyborg, J.K. Molecular characterization of HTLV-1 Tax interaction with the KIX domain of CBP/p300. J. Mol. Biol. 2007, 372, 958–969. [Google Scholar] [CrossRef]
  116. Siu, Y.T.; Chin, K.T.; Siu, K.L.; Yee Wai Choy, E.; Jeang, K.T.; Jin, D.Y. TORC1 and TORC2 coactivators are required for tax activation of the human T-cell leukemia virus type 1 long terminal repeats. J. Virol. 2006, 80, 7052–7059. [Google Scholar] [CrossRef]
  117. Kim, Y.M.; Geiger, T.R.; Egan, D.I.; Sharma, N.; Nyborg, J.K. The HTLV-1 tax protein cooperates with phosphorylated CREB, TORC2 and p300 to activate CRE-dependent cyclin D1 transcription. Oncogene 2010, 29, 2142–2152. [Google Scholar] [CrossRef]
  118. Kim, Y.M.; Ramirez, J.A.; Mick, J.E.; Giebler, H.A.; Yan, J.P.; Nyborg, J.K. Molecular characterization of the Tax-containing HTLV-1 enhancer complex reveals a prominent role for CREB phosphorylation in Tax transactivation. J. Biol. Chem. 2007, 282, 18750–18757. [Google Scholar] [CrossRef]
  119. Jiang, H.; Lu, H.; Schiltz, R.L.; Pise-Masison, C.A.; Ogryzko, V.V.; Nakatani, Y.; Brady, J.N. PCAF interacts with tax and stimulates tax transactivation in a histone acetyltransferase-independent manner. Mol. Cell Biol. 1999, 19, 8136–8145. [Google Scholar] [CrossRef]
  120. Lu, H.; Pise-Masison, C.A.; Fletcher, T.M.; Schiltz, R.L.; Nagaich, A.K.; Radonovich, M.; Hager, G.; Cole, P.A.; Brady, J.N. Acetylation of nucleosomal histones by p300 facilitates transcription from tax-responsive human T-cell leukemia virus type 1 chromatin template. Mol. Cell Biol. 2002, 22, 4450–4462. [Google Scholar] [CrossRef]
  121. Sharma, N.; Nyborg, J.K. The coactivators CBP/p300 and the histone chaperone NAP1 promote transcription-independent nucleosome eviction at the HTLV-1 promoter. Proc. Natl. Acad. Sci. USA 2008, 105, 7959–7963. [Google Scholar] [CrossRef]
  122. Easley, R.; Carpio, L.; Guendel, I.; Klase, Z.; Choi, S.; Kehn-Hall, K.; Brady, J.N.; Kashanchi, F. Human T-lymphotropic virus type 1 transcription and chromatin-remodeling complexes. J. Virol. 2010, 84, 4755–4768. [Google Scholar] [CrossRef] [PubMed]
  123. Lu, H.; Pise-Masison, C.A.; Linton, R.; Park, H.U.; Schiltz, R.L.; Sartorelli, V.; Brady, J.N. Tax relieves transcriptional repression by promoting histone deacetylase 1 release from the human T-cell leukemia virus type 1 long terminal repeat. J. Virol. 2004, 78, 6735–6743. [Google Scholar] [CrossRef] [PubMed]
  124. Jain, P.; Lavorgna, A.; Sehgal, M.; Gao, L.; Ginwala, R.; Sagar, D.; Harhaj, E.W.; Khan, Z.K. Myocyte enhancer factor (MEF)-2 plays essential roles in T-cell transformation associated with HTLV-1 infection by stabilizing complex between Tax and CREB. Retrovirology 2015, 12, 23. [Google Scholar] [CrossRef]
  125. Hayden, M.S.; Ghosh, S. Shared principles in NF-κB signaling. Cell 2008, 132, 344–362. [Google Scholar] [CrossRef]
  126. Bonizzi, G.; Karin, M. The two NF-κB activation pathways and their role in innate and adaptive immunity. Trends Immunol. 2004, 25, 280–288. [Google Scholar] [CrossRef]
  127. Sun, S.C. The non-canonical NF-κB pathway in immunity and inflammation. Nat. Rev. Immunol. 2017, 17, 545–558. [Google Scholar] [CrossRef] [PubMed]
  128. Hinz, M.; Scheidereit, C. The IκB kinase complex in NF-κB regulation and beyond. EMBO Rep. 2014, 15, 46–61. [Google Scholar] [CrossRef]
  129. DiDonato, J.A.; Mercurio, F.; Karin, M. NF-κB and the link between inflammation and cancer. Immunol. Rev. 2012, 246, 379–400. [Google Scholar] [CrossRef]
  130. Hoesel, B.; Schmid, J.A. The complexity of NF-κB signaling in inflammation and cancer. Mol. Cancer 2013, 12, 86. [Google Scholar] [CrossRef]
  131. Pires, B.R.B.; Silva, R.; Ferreira, G.M.; Abdelhay, E. NF-kappaB: Two Sides of the Same Coin. Genes 2018, 9, 24. [Google Scholar] [CrossRef] [PubMed]
  132. Xiao, G.; Harhaj, E.W.; Sun, S.C. NF-κB-inducing kinase regulates the processing of NF-κB2 p100. Mol. Cell 2001, 7, 401–409. [Google Scholar] [CrossRef]
  133. Thu, Y.M.; Richmond, A. NF-κB inducing kinase: A key regulator in the immune system and in cancer. Cytokine Growth Factor. Rev. 2010, 21, 213–226. [Google Scholar] [CrossRef] [PubMed]
  134. Sun, S.C.; Elwood, J.; Beraud, C.; Greene, W.C. Human T-cell leukemia virus type I Tax activation of NF-κB/Rel involves phosphorylation and degradation of IκBα and RelA (p65)-mediated induction of the c-rel gene. Mol. Cell Biol. 1994, 14, 7377–7384. [Google Scholar] [CrossRef]
  135. Maggirwar, S.B.; Harhaj, E.; Sun, S.C. Activation of NF-κB/Rel by Tax involves degradation of IκBα and is blocked by a proteasome inhibitor. Oncogene 1995, 11, 993–998. [Google Scholar]
  136. Uhlik, M.; Good, L.; Xiao, G.; Harhaj, E.W.; Zandi, E.; Karin, M.; Sun, S.C. NF-κB-inducing kinase and IκB kinase participate in human T-cell leukemia virus I Tax-mediated NF-κB activation. J. Biol. Chem. 1998, 273, 21132–21136. [Google Scholar] [CrossRef]
  137. Suzuki, T.; Hirai, H.; Murakami, T.; Yoshida, M. Tax protein of HTLV-1 destabilizes the complexes of NF-κB and IκB-α and induces nuclear translocation of NF-κB for transcriptional activation. Oncogene 1995, 10, 1199–1207. [Google Scholar] [PubMed]
  138. Petropoulos, L.; Hiscott, J. Association between HTLV-1 Tax and IκBα is dependent on the IκBα phosphorylation state. Virology 1998, 252, 189–199. [Google Scholar] [CrossRef]
  139. Good, L.; Sun, S.C. Persistent activation of NF-κB/Rel by human T-cell leukemia virus type 1 tax involves degradation of IκBβ. J. Virol. 1996, 70, 2730–2735. [Google Scholar] [CrossRef]
  140. Hirai, H.; Suzuki, T.; Fujisawa, J.; Inoue, J.; Yoshida, M. Tax protein of human T-cell leukemia virus type I binds to the ankyrin motifs of inhibitory factor κB and induces nuclear translocation of transcription factor NF-κB proteins for transcriptional activation. Proc. Natl. Acad. Sci. USA 1994, 91, 3584–3588. [Google Scholar] [CrossRef]
  141. Hirai, H.; Fujisawa, J.; Suzuki, T.; Ueda, K.; Muramatsu, M.; Tsuboi, A.; Arai, N.; Yoshida, M. Transcriptional activator Tax of HTLV-1 binds to the NF-κB precursor p105. Oncogene 1992, 7, 1737–1742. [Google Scholar] [PubMed]
  142. Watanabe, M.; Muramatsu, M.; Hirai, H.; Suzuki, T.; Fujisawa, J.; Yoshida, M.; Arai, K.; Arai, N. HTLV-I encoded Tax in association with NF-κB precursor p105 enhances nuclear localization of NF-κB p50 and p65 in transfected cells. Oncogene 1993, 8, 2949–2958. [Google Scholar] [PubMed]
  143. Rousset, R.; Desbois, C.; Bantignies, F.; Jalinot, P. Effects on NF-κB1/p105 processing of the interaction between the HTLV-1 transactivator Tax and the proteasome. Nature 1996, 381, 328–331. [Google Scholar] [CrossRef] [PubMed]
  144. Lanoix, J.; Lacoste, J.; Pepin, N.; Rice, N.; Hiscott, J. Overproduction of NFKB2 (lyt-10) and c-Rel: A mechanism for HTLV-I Tax-mediated trans-activation via the NF-κB signalling pathway. Oncogene 1994, 9, 841–852. [Google Scholar]
  145. Murakami, T.; Hirai, H.; Suzuki, T.; Fujisawa, J.; Yoshida, M. HTLV-1 Tax enhances NF-κB2 expression and binds to the products p52 and p100, but does not suppress the inhibitory function of p100. Virology 1995, 206, 1066–1074. [Google Scholar] [CrossRef]
  146. Suzuki, T.; Hirai, H.; Fujisawa, J.; Fujita, T.; Yoshida, M. A trans-activator Tax of human T-cell leukemia virus type 1 binds to NF-κB p50 and serum response factor (SRF) and associates with enhancer DNAs of the NF-κB site and CArG box. Oncogene 1993, 8, 2391–2397. [Google Scholar]
  147. Suzuki, T.; Hirai, H.; Yoshida, M. Tax protein of HTLV-1 interacts with the Rel homology domain of NF-κB p65 and c-Rel proteins bound to the NF-κB binding site and activates transcription. Oncogene 1994, 9, 3099–3105. [Google Scholar]
  148. Bex, F.; McDowall, A.; Burny, A.; Gaynor, R. The human T-cell leukemia virus type 1 transactivator protein Tax colocalizes in unique nuclear structures with NF-κB proteins. J. Virol. 1997, 71, 3484–3497. [Google Scholar] [CrossRef]
  149. Bex, F.; Yin, M.J.; Burny, A.; Gaynor, R.B. Differential transcriptional activation by human T-cell leukemia virus type 1 Tax mutants is mediated by distinct interactions with CREB binding protein and p300. Mol. Cell Biol. 1998, 18, 2392–2405. [Google Scholar] [CrossRef]
  150. Yamaoka, S.; Courtois, G.; Bessia, C.; Whiteside, S.T.; Weil, R.; Agou, F.; Kirk, H.E.; Kay, R.J.; Israel, A. Complementation cloning of NEMO, a component of the IkappaB kinase complex essential for NF-kappaB activation. Cell 1998, 93, 1231–1240. [Google Scholar] [CrossRef]
  151. Geleziunas, R.; Ferrell, S.; Lin, X.; Mu, Y.; Cunningham, E.T., Jr.; Grant, M.; Connelly, M.A.; Hambor, J.E.; Marcu, K.B.; Greene, W.C. Human T-cell leukemia virus type 1 Tax induction of NF-κB involves activation of the IκB kinase α (IKKα) and IKKβ cellular kinases. Mol. Cell Biol. 1998, 18, 5157–5165. [Google Scholar] [CrossRef]
  152. Harhaj, E.W.; Sun, S.C. IKK γ serves as a docking subunit of the IκB kinase (IKK) and mediates interaction of IKK with the human T-cell leukemia virus Tax protein. J. Biol. Chem. 1999, 274, 22911–22914. [Google Scholar] [CrossRef]
  153. Chu, Z.L.; Shin, Y.A.; Yang, J.M.; DiDonato, J.A.; Ballard, D.W. IKKγ mediates the interaction of cellular IκB kinases with the tax transforming protein of human T cell leukemia virus type 1. J. Biol. Chem. 1999, 274, 15297–15300. [Google Scholar] [CrossRef]
  154. Jin, D.Y.; Giordano, V.; Kibler, K.V.; Nakano, H.; Jeang, K.T. Role of adapter function in oncoprotein-mediated activation of NF-κB. Human T-cell leukemia virus type I Tax interacts directly with IκB kinase γ. J. Biol. Chem. 1999, 274, 17402–17405. [Google Scholar] [CrossRef]
  155. Harhaj, E.W.; Good, L.; Xiao, G.; Uhlik, M.; Cvijic, M.E.; Rivera-Walsh, I.; Sun, S.C. Somatic mutagenesis studies of NF-κB signaling in human T cells: Evidence for an essential role of IKK γ in NF-κB activation by T-cell costimulatory signals and HTLV-I Tax protein. Oncogene 2000, 19, 1448–1456. [Google Scholar] [CrossRef]
  156. Sun, S.C.; Harhaj, E.W.; Xiao, G.; Good, L. Activation of I-κB kinase by the HTLV type 1 Tax protein: Mechanistic insights into the adaptor function of IKKγ. AIDS Res. Hum. Retroviruses 2000, 16, 1591–1596. [Google Scholar] [CrossRef]
  157. O’Mahony, A.M.; Montano, M.; Van Beneden, K.; Chen, L.F.; Greene, W.C. Human T-cell lymphotropic virus type 1 tax induction of biologically active NF-κB requires IκB kinase-1-mediated phosphorylation of RelA/p65. J. Biol. Chem. 2004, 279, 18137–18145. [Google Scholar] [CrossRef]
  158. Wu, X.; Sun, S.C. Retroviral oncoprotein Tax deregulates NF-κB by activating Tak1 and mediating the physical association of Tak1-IKK. EMBO Rep. 2007, 8, 510–515. [Google Scholar] [CrossRef] [PubMed]
  159. Yu, Q.; Minoda, Y.; Yoshida, R.; Yoshida, H.; Iha, H.; Kobayashi, T.; Yoshimura, A.; Takaesu, G. HTLV-1 Tax-mediated TAK1 activation involves TAB2 adapter protein. Biochem. Biophys. Res. Commun. 2008, 365, 189–194. [Google Scholar] [CrossRef] [PubMed]
  160. Yin, M.J.; Christerson, L.B.; Yamamoto, Y.; Kwak, Y.T.; Xu, S.; Mercurio, F.; Barbosa, M.; Cobb, M.H.; Gaynor, R.B. HTLV-I Tax protein binds to MEKK1 to stimulate IκB kinase activity and NF-κB activation. Cell 1998, 93, 875–884. [Google Scholar] [CrossRef] [PubMed]
  161. Sebald, A.; Mattioli, I.; Schmitz, M.L. T cell receptor-induced lipid raft recruitment of the IκB kinase complex is necessary and sufficient for NF-κB activation occurring in the cytosol. Eur. J. Immunol. 2005, 35, 318–325. [Google Scholar] [CrossRef] [PubMed]
  162. Harhaj, N.S.; Sun, S.C.; Harhaj, E.W. Activation of NF-κB by the human T cell leukemia virus type I Tax oncoprotein is associated with ubiquitin-dependent relocalization of IκB kinase. J. Biol. Chem. 2007, 282, 4185–4192. [Google Scholar] [CrossRef] [PubMed]
  163. Huang, J.; Ren, T.; Guan, H.; Jiang, Y.; Cheng, H. HTLV-1 Tax is a critical lipid raft modulator that hijacks IκB kinases to the microdomains for persistent activation of NF-κB. J. Biol. Chem. 2009, 284, 6208–6217. [Google Scholar] [CrossRef]
  164. Journo, C.; Filipe, J.; About, F.; Chevalier, S.A.; Afonso, P.V.; Brady, J.N.; Flynn, D.; Tangy, F.; Israel, A.; Vidalain, P.O.; et al. NRP/Optineurin Cooperates with TAX1BP1 to potentiate the activation of NF-κB by human T-lymphotropic virus type 1 tax protein. PLoS Pathog. 2009, 5, e1000521. [Google Scholar] [CrossRef]
  165. Zhang, H.; Chen, L.; Cai, S.H.; Cheng, H. Identification of TBK1 and IKKε, the non-canonical IκB kinases, as crucial pro-survival factors in HTLV-1-transformed T lymphocytes. Leuk. Res. 2016, 46, 37–44. [Google Scholar] [CrossRef]
  166. Shibata, Y.; Tanaka, Y.; Gohda, J.; Inoue, J. Activation of the IκB kinase complex by HTLV-1 Tax requires cytosolic factors involved in Tax-induced polyubiquitination. J. Biochem. 2011, 150, 679–686. [Google Scholar] [CrossRef]
  167. Shembade, N.; Harhaj, E.W. Role of post-translational modifications of HTLV-1 Tax in NF-κB activation. World J. Biol. Chem. 2010, 1, 13–20. [Google Scholar] [CrossRef]
  168. Mohanty, S.; Han, T.; Choi, Y.B.; Lavorgna, A.; Zhang, J.; Harhaj, E.W. The E3/E4 ubiquitin conjugation factor UBE4B interacts with and ubiquitinates the HTLV-1 Tax oncoprotein to promote NF-κB activation. PLoS Pathog. 2020, 16, e1008504. [Google Scholar] [CrossRef] [PubMed]
  169. Ho, Y.K.; Zhi, H.; Bowlin, T.; Dorjbal, B.; Philip, S.; Zahoor, M.A.; Shih, H.M.; Semmes, O.J.; Schaefer, B.; Glover, J.N.; et al. HTLV-1 Tax stimulates ubiquitin E3 ligase, Ring Finger Protein 8, to assemble Lysine 63-linked polyubiquitin chains for TAK1 and IKK activation. PLoS Pathog. 2015, 11, e1005102. [Google Scholar] [CrossRef]
  170. Shibata, Y.; Tokunaga, F.; Goto, E.; Komatsu, G.; Gohda, J.; Saeki, Y.; Tanaka, K.; Takahashi, H.; Sawasaki, T.; Inoue, S.; et al. HTLV-1 Tax induces formation of the active macromolecular IKK complex by generating Lys63- and Met1-linked hybrid polyubiquitin chains. PLoS Pathog. 2017, 13, e1006162. [Google Scholar] [CrossRef]
  171. Wang, C.; Long, W.; Peng, C.; Hu, L.; Zhang, Q.; Wu, A.; Zhang, X.; Duan, X.; Wong, C.C.; Tanaka, Y.; et al. HTLV-1 Tax functions as a ubiquitin E3 ligase for direct IKK activation via synthesis of mixed-linkage polyubiquitin chains. PLoS Pathog. 2016, 12, e1005584. [Google Scholar] [CrossRef]
  172. Beraud, C.; Sun, S.C.; Ganchi, P.; Ballard, D.W.; Greene, W.C. Human T-cell leukemia virus type I Tax associates with and is negatively regulated by the NF-κB2 p100 gene product: Implications for viral latency. Mol. Cell Biol. 1994, 14, 1374–1382. [Google Scholar] [CrossRef]
  173. Xiao, G.; Cvijic, M.E.; Fong, A.; Harhaj, E.W.; Uhlik, M.T.; Waterfield, M.; Sun, S.C. Retroviral oncoprotein Tax induces processing of NF-κB2/p100 in T cells: Evidence for the involvement of IKKα. EMBO J. 2001, 20, 6805–6815. [Google Scholar] [CrossRef]
  174. Xiao, G.; Fong, A.; Sun, S.C. Induction of p100 processing by NF-κB-inducing kinase involves docking IκB kinase α (IKK α) to p100 and IKKα-mediated phosphorylation. J. Biol. Chem. 2004, 279, 30099–30105. [Google Scholar] [CrossRef]
  175. Fong, A.; Sun, S.C. Genetic evidence for the essential role of β-transducin repeat-containing protein in the inducible processing of NF-κB2/p100. J. Biol. Chem. 2002, 277, 22111–22114. [Google Scholar] [CrossRef]
  176. Qu, Z.; Qing, G.; Rabson, A.; Xiao, G. Tax deregulation of NF-κB2 p100 processing involves both β-TrCP-dependent and -independent mechanisms. J. Biol. Chem. 2004, 279, 44563–44572. [Google Scholar] [CrossRef] [PubMed]
  177. Shuh, M.; Derse, D. Ternary complex factors and cofactors are essential for human T-cell leukemia virus type 1 tax transactivation of the serum response element. J. Virol. 2000, 74, 11394–11397. [Google Scholar] [CrossRef] [PubMed]
  178. Winter, H.Y.; Marriott, S.J. Human T-cell leukemia virus type 1 Tax enhances serum response factor DNA binding and alters site selection. J. Virol. 2007, 81, 6089–6098. [Google Scholar] [CrossRef] [PubMed]
  179. Ohtani, K.; Nakamura, M.; Saito, S.; Nagata, K.; Sugamura, K.; Hinuma, Y. Electroporation: Application to human lymphoid cell lines for stable introduction of a transactivator gene of human T-cell leukemia virus type I. Nucleic Acids Res. 1989, 17, 1589–1604. [Google Scholar] [CrossRef]
  180. Nagata, K.; Ohtani, K.; Nakamura, M.; Sugamura, K. Activation of endogenous c-fos proto-oncogene expression by human T-cell leukemia virus type I-encoded p40tax protein in the human T-cell line, Jurkat. J. Virol. 1989, 63, 3220–3226. [Google Scholar] [CrossRef]
  181. Miura, S.; Ohtani, K.; Numata, N.; Niki, M.; Ohbo, K.; Ina, Y.; Gojobori, T.; Tanaka, Y.; Tozawa, H.; Nakamura, M.; et al. Molecular cloning and characterization of a novel glycoprotein, gp34, that is specifically induced by the human T-cell leukemia virus type I transactivator p40tax. Mol. Cell Biol. 1991, 11, 1313–1325. [Google Scholar] [CrossRef]
  182. Nakashima, K.; Kawakami, A.; Hida, A.; Yamasaki, S.; Nakamura, H.; Kamachi, M.; Miyashita, T.; Tanaka, F.; Izumi, Y.; Tamai, M.; et al. Protection of mitochondrial perturbation by human T-lymphotropic virus type 1 tax through induction of Bcl-xL expression. J. Lab. Clin. Med. 2003, 142, 341–347. [Google Scholar] [CrossRef]
  183. Kawakami, A.; Nakashima, T.; Sakai, H.; Urayama, S.; Yamasaki, S.; Hida, A.; Tsuboi, M.; Nakamura, H.; Ida, H.; Migita, K.; et al. Inhibition of caspase cascade by HTLV-I tax through induction of NF-κB nuclear translocation. Blood 1999, 94, 3847–3854. [Google Scholar] [CrossRef]
  184. Baba, M.; Imai, T.; Yoshida, T.; Yoshie, O. Constitutive expression of various chemokine genes in human T-cell lines infected with human T-cell leukemia virus type 1: Role of the viral transactivator Tax. Int. J. Cancer 1996, 66, 124–129. [Google Scholar] [CrossRef]
  185. Iwata, S.; Souta-Kuribara, A.; Yamakawa, A.; Sasaki, T.; Shimizu, T.; Hosono, O.; Kawasaki, H.; Tanaka, H.; Dang, N.H.; Watanabe, T.; et al. HTLV-I Tax induces and associates with Crk-associated substrate lymphocyte type (Cas-L). Oncogene 2005, 24, 1262–1271. [Google Scholar] [CrossRef] [PubMed]
  186. Jin, Z.; Nagakubo, D.; Shirakawa, A.K.; Nakayama, T.; Shigeta, A.; Hieshima, K.; Yamada, Y.; Yoshie, O. CXCR7 is inducible by HTLV-1 Tax and promotes growth and survival of HTLV-1-infected T cells. Int. J. Cancer 2009, 125, 2229–2235. [Google Scholar] [CrossRef]
  187. Lemasson, I.; Robert-Hebmann, V.; Hamaia, S.; Duc Dodon, M.; Gazzolo, L.; Devaux, C. Transrepression of lck gene expression by human T-cell leukemia virus type 1-encoded p40tax. J. Virol. 1997, 71, 1975–1983. [Google Scholar] [CrossRef]
  188. Yamashita, I.; Katamine, S.; Moriuchi, R.; Nakamura, Y.; Miyamoto, T.; Eguchi, K.; Nagataki, S. Transactivation of the human interleukin-6 gene by human T-lymphotropic virus type 1 Tax protein. Blood 1994, 84, 1573–1578. [Google Scholar] [CrossRef]
  189. Higashimura, N.; Takasawa, N.; Tanaka, Y.; Nakamura, M.; Sugamura, K. Induction of OX40, a receptor of gp34, on T cells by trans-acting transcriptional activator, Tax, of human T-cell leukemia virus type I. Jpn. J. Cancer Res. 1996, 87, 227–231. [Google Scholar] [CrossRef]
  190. Chen, X.; Zachar, V.; Zdravkovic, M.; Guo, M.; Ebbesen, P.; Liu, X. Role of the Fas/Fas ligand pathway in apoptotic cell death induced by the human T cell lymphotropic virus type I Tax transactivator. J. Gen. Virol. 1997, 78 Pt 12, 3277–3285. [Google Scholar] [CrossRef] [PubMed]
  191. Tanaka, Y.; Fukudome, K.; Hayashi, M.; Takagi, S.; Yoshie, O. Induction of ICAM-1 and LFA-3 by Tax1 of human T-cell leukemia virus type 1 and mechanism of down-regulation of ICAM-1 or LFA-1 in adult-T-cell-leukemia cell lines. Int. J. Cancer 1995, 60, 554–561. [Google Scholar] [CrossRef]
  192. Mori, N.; Murakami, S.; Oda, S.; Eto, S. Human T-cell leukemia virus type I tax induces intracellular adhesion molecule-1 expression in T cells. Blood 1994, 84, 350–351. [Google Scholar] [CrossRef]
  193. Tanaka, Y.; Mizuguchi, M.; Takahashi, Y.; Fujii, H.; Tanaka, R.; Fukushima, T.; Tomoyose, T.; Ansari, A.A.; Nakamura, M. Human T-cell leukemia virus type-I Tax induces the expression of CD83 on T cells. Retrovirology 2015, 12, 56. [Google Scholar] [CrossRef]
  194. Naito, T.; Ushirogawa, H.; Fukushima, T.; Tanaka, Y.; Saito, M. EOS, an Ikaros family zinc finger transcription factor, interacts with the HTLV-1 oncoprotein Tax and is downregulated in peripheral blood mononuclear cells of HTLV-1-infected individuals, irrespective of clinical statuses. Virol. J. 2019, 16, 160. [Google Scholar] [CrossRef]
  195. Sasada, T.; Nakamura, H.; Masutani, H.; Ueda, S.; Sono, H.; Takabayashi, A.; Yodoi, J. Thioredoxin-mediated redox control of human T cell lymphotropic virus type I (HTLV-I) gene expression. Mol. Immunol. 2002, 38, 723–732. [Google Scholar] [CrossRef]
  196. Liu, X.; Chen, X.; Zachar, V.; Chang, C.; Ebbesen, P. Transcriptional activation of human TR3/nur77 gene expression by human T-lymphotropic virus type I Tax protein through two AP-1-like elements. J. Gen. Virol. 1999, 80 Pt 12, 3073–3081. [Google Scholar] [CrossRef] [PubMed]
  197. Beck, Z.; Bacsi, A.; Liu, X.; Ebbesen, P.; Andirko, I.; Csoma, E.; Konya, J.; Nagy, E.; Toth, F.D. Differential patterns of human cytomegalovirus gene expression in various T-cell lines carrying human T-cell leukemia-lymphoma virus type I: Role of Tax-activated cellular transcription factors. J. Med. Virol. 2003, 71, 94–104. [Google Scholar] [CrossRef] [PubMed]
  198. Nagakubo, D.; Jin, Z.; Hieshima, K.; Nakayama, T.; Shirakawa, A.K.; Tanaka, Y.; Hasegawa, H.; Hayashi, T.; Tsukasaki, K.; Yamada, Y.; et al. Expression of CCR9 in HTLV-1+ T cells and ATL cells expressing Tax. Int. J. Cancer 2007, 120, 1591–1597. [Google Scholar] [CrossRef] [PubMed]
  199. Ng, P.W.; Iha, H.; Iwanaga, Y.; Bittner, M.; Chen, Y.; Jiang, Y.; Gooden, G.; Trent, J.M.; Meltzer, P.; Jeang, K.T.; et al. Genome-wide expression changes induced by HTLV-1 Tax: Evidence for MLK-3 mixed lineage kinase involvement in Tax-mediated NF-κB activation. Oncogene 2001, 20, 4484–4496. [Google Scholar] [CrossRef]
  200. Kasai, T.; Jeang, K.T. Two discrete events, human T-cell leukemia virus type I Tax oncoprotein expression and a separate stress stimulus, are required for induction of apoptosis in T-cells. Retrovirology 2004, 1, 7. [Google Scholar] [CrossRef]
  201. Chen, X.; Zachar, V.; Chang, C.; Ebbesen, P.; Liu, X. Differential expression of Nur77 family members in human T-lymphotropic virus type 1-infected cells: Transactivation of the TR3/nur77 gene by Tax protein. J. Virol. 1998, 72, 6902–6906. [Google Scholar] [CrossRef]
  202. Chen, L.; Ma, S.; Li, B.; Fink, T.; Zachar, V.; Takahashi, M.; Cuttichia, J.; Tsui, L.C.; Ebbesen, P.; Liu, X. Transcriptional activation of immediate-early gene ETR101 by human T-cell leukaemia virus type I Tax. J. Gen. Virol. 2003, 84, 3203–3214. [Google Scholar] [CrossRef] [PubMed]
  203. Yamagata, T.; Mitani, K.; Ueno, H.; Kanda, Y.; Yazaki, Y.; Hirai, H. Triple synergism of human T-lymphotropic virus type 1-encoded tax, GATA-binding protein, and AP-1 is required for constitutive expression of the interleukin-5 gene in adult T-cell leukemia cells. Mol. Cell Biol. 1997, 17, 4272–4281. [Google Scholar] [CrossRef]
  204. Valentin, H.; Lemasson, I.; Hamaia, S.; Casse, H.; Konig, S.; Devaux, C.; Gazzolo, L. Transcriptional activation of the vascular cell adhesion molecule-1 gene in T lymphocytes expressing human T-cell leukemia virus type 1 Tax protein. J. Virol. 1997, 71, 8522–8530. [Google Scholar] [CrossRef] [PubMed]
  205. Mori, N.; Mukaida, N.; Ballard, D.W.; Matsushima, K.; Yamamoto, N. Human T-cell leukemia virus type I Tax transactivates human interleukin 8 gene through acting concurrently on AP-1 and nuclear factor-κB-like sites. Cancer Res. 1998, 58, 3993–4000. [Google Scholar]
  206. Fujii, M.; Niki, T.; Mori, T.; Matsuda, T.; Matsui, M.; Nomura, N.; Seiki, M. HTLV-1 Tax induces expression of various immediate early serum responsive genes. Oncogene 1991, 6, 1023–1029. [Google Scholar]
  207. Fufa, T.D.; Byun, J.S.; Wakano, C.; Fernandez, A.G.; Pise-Masison, C.A.; Gardner, K. The Tax oncogene enhances ELL incorporation into p300 and P-TEFb containing protein complexes to activate transcription. Biochem. Biophys. Res. Commun. 2015, 465, 5–11. [Google Scholar] [CrossRef]
  208. Siekevitz, M.; Feinberg, M.B.; Holbrook, N.; Wong-Staal, F.; Greene, W.C. Activation of interleukin 2 and interleukin 2 receptor (Tac) promoter expression by the trans-activator (tat) gene product of human T-cell leukemia virus, type I. Proc. Natl. Acad. Sci. USA 1987, 84, 5389–5393. [Google Scholar] [CrossRef] [PubMed]
  209. Wano, Y.; Feinberg, M.; Hosking, J.B.; Bogerd, H.; Greene, W.C. Stable expression of the tax gene of type I human T-cell leukemia virus in human T cells activates specific cellular genes involved in growth. Proc. Natl. Acad. Sci. USA 1988, 85, 9733–9737. [Google Scholar] [CrossRef]
  210. Hoyos, B.; Ballard, D.W.; Bohnlein, E.; Siekevitz, M.; Greene, W.C. Kappa B-specific DNA binding proteins: Role in the regulation of human interleukin-2 gene expression. Science 1989, 244, 457–460. [Google Scholar] [CrossRef]
  211. McGuire, K.L.; Curtiss, V.E.; Larson, E.L.; Haseltine, W.A. Influence of human T-cell leukemia virus type I tax and rex on interleukin-2 gene expression. J. Virol. 1993, 67, 1590–1599. [Google Scholar] [CrossRef]
  212. Good, L.; Maggirwar, S.B.; Sun, S.C. Activation of the IL-2 gene promoter by HTLV-I tax involves induction of NF-AT complexes bound to the CD28-responsive element. EMBO J. 1996, 15, 3744–3750. [Google Scholar] [CrossRef]
  213. Curtiss, V.E.; Smilde, R.; McGuire, K.L. Requirements for interleukin 2 promoter transactivation by the Tax protein of human T-cell leukemia virus type 1. Mol. Cell Biol. 1996, 16, 3567–3575. [Google Scholar] [CrossRef]
  214. Cross, S.L.; Feinberg, M.B.; Wolf, J.B.; Holbrook, N.J.; Wong-Staal, F.; Leonard, W.J. Regulation of the human interleukin-2 receptor α chain promoter: Activation of a nonfunctional promoter by the transactivator gene of HTLV-I. Cell 1987, 49, 47–56. [Google Scholar] [CrossRef] [PubMed]
  215. Leung, K.; Nabel, G.J. HTLV-1 transactivator induces interleukin-2 receptor expression through an NF-κB-like factor. Nature 1988, 333, 776–778. [Google Scholar] [CrossRef]
  216. Ruben, S.; Poteat, H.; Tan, T.H.; Kawakami, K.; Roeder, R.; Haseltine, W.; Rosen, C.A. Cellular transcription factors and regulation of IL-2 receptor gene expression by HTLV-I tax gene product. Science 1988, 241, 89–92. [Google Scholar] [CrossRef] [PubMed]
  217. Ballard, D.W.; Bohnlein, E.; Lowenthal, J.W.; Wano, Y.; Franza, B.R.; Greene, W.C. HTLV-I tax induces cellular proteins that activate the κB element in the IL-2 receptor α gene. Science 1988, 241, 1652–1655. [Google Scholar] [CrossRef]
  218. Green, J.E.; Begley, C.G.; Wagner, D.K.; Waldmann, T.A.; Jay, G. trans activation of granulocyte-macrophage colony-stimulating factor and the interleukin-2 receptor in transgenic mice carrying the human T-lymphotropic virus type 1 tax gene. Mol. Cell Biol. 1989, 9, 4731–4737. [Google Scholar] [CrossRef]
  219. Marriott, S.J.; Trinh, D.; Brady, J.N. Activation of interleukin-2 receptor α expression by extracellular HTLV-I Tax1 protein: A potential role in HTLV-I pathogenesis. Oncogene 1992, 7, 1749–1755. [Google Scholar]
  220. Ashrafi, F.; Nassiri, M.; Javadmanesh, A.; Rahimi, H.; Rezaee, S.A. Epigenetics evaluation of the oncogenic mechanisms of two closely related bovine and human deltaretroviruses: A system biology study. Microb. Pathog. 2020, 139, 103845. [Google Scholar] [CrossRef] [PubMed]
  221. Ross, S.H.; Cantrell, D.A. Signaling and Function of Interleukin-2 in T Lymphocytes. Annu. Rev. Immunol. 2018, 36, 411–433. [Google Scholar] [CrossRef]
  222. Ohbo, K.; Takasawa, N.; Ishii, N.; Tanaka, N.; Nakamura, M.; Sugamura, K. Functional analysis of the human interleukin 2 receptor γ chain gene promoter. J. Biol. Chem. 1995, 270, 7479–7486. [Google Scholar] [CrossRef] [PubMed]
  223. Nakamura, M.; Takasawa, N.; Ohbo, K.; Higashimura, N.; Ohtani, K.; Tanaka, Y.; Sugamura, K. HTLV-I Tax trans-activation and cell growth signaling. Leukemia 1997, 11 (Suppl. S3), 7–9. [Google Scholar] [PubMed]
  224. Maeda, M.; Tanabe-Shibuya, J.; Miyazato, P.; Masutani, H.; Yasunaga, J.I.; Usami, K.; Shimizu, A.; Matsuoka, M. IL-2/IL-2 receptor pathway plays a crucial role in the growth and malignant transformation of HTLV-1-infected T cells to develop adult T-cell leukemia. Front. Microbiol. 2020, 11, 356. [Google Scholar] [CrossRef]
  225. Satoh, M.; Toma, H.; Sugahara, K.; Etoh, K.; Shiroma, Y.; Kiyuna, S.; Takara, M.; Matsuoka, M.; Yamaguchi, K.; Nakada, K.; et al. Involvement of IL-2/IL-2R system activation by parasite antigen in polyclonal expansion of CD4+25+ HTLV-1-infected T-cells in human carriers of both HTLV-1 and S. stercoralis. Oncogene 2002, 21, 2466–2475. [Google Scholar] [CrossRef]
  226. Trejo, S.R.; Fahl, W.E.; Ratner, L. The tax protein of human T-cell leukemia virus type 1 mediates the transactivation of the c-sis/platelet-derived growth factor-B promoter through interactions with the zinc finger transcription factors Sp1 and NGFI-A/Egr-1. J. Biol. Chem. 1997, 272, 27411–27421. [Google Scholar] [CrossRef]
  227. Fung, M.M.; Chu, Y.L.; Fink, J.L.; Wallace, A.; McGuire, K.L. IL-2- and STAT5-regulated cytokine gene expression in cells expressing the Tax protein of HTLV-1. Oncogene 2005, 24, 4624–4633. [Google Scholar] [CrossRef]
  228. Chen, J.; Petrus, M.; Bryant, B.R.; Phuc Nguyen, V.; Stamer, M.; Goldman, C.K.; Bamford, R.; Morris, J.C.; Janik, J.E.; Waldmann, T.A. Induction of the IL-9 gene by HTLV-I Tax stimulates the spontaneous proliferation of primary adult T-cell leukemia cells by a paracrine mechanism. Blood 2008, 111, 5163–5172. [Google Scholar] [CrossRef] [PubMed]
  229. Azimi, N.; Brown, K.; Bamford, R.N.; Tagaya, Y.; Siebenlist, U.; Waldmann, T.A. Human T cell lymphotropic virus type I Tax protein trans-activates interleukin 15 gene transcription through an NF-κB site. Proc. Natl. Acad. Sci. USA 1998, 95, 2452–2457. [Google Scholar] [CrossRef]
  230. Azimi, N.; Mariner, J.; Jacobson, S.; Waldmann, T.A. How does interleukin 15 contribute to the pathogenesis of HTLV type 1-associated myelopathy/tropical spastic paraparesis? AIDS Res. Hum. Retroviruses 2000, 16, 1717–1722. [Google Scholar] [CrossRef]
  231. Mariner, J.M.; Lantz, V.; Waldmann, T.A.; Azimi, N. Human T cell lymphotropic virus type I Tax activates IL-15Rα gene expression through an NF-κB site. J. Immunol. 2001, 166, 2602–2609. [Google Scholar] [CrossRef] [PubMed]
  232. Mizuguchi, M.; Asao, H.; Hara, T.; Higuchi, M.; Fujii, M.; Nakamura, M. Transcriptional activation of the interleukin-21 gene and its receptor gene by human T-cell leukemia virus type 1 Tax in human T-cells. J. Biol. Chem. 2009, 284, 25501–25511. [Google Scholar] [CrossRef]
  233. Rajaei, T.; Farajifard, H.; Rafatpanah, H.; Bustani, R.; Valizadeh, N.; Rajaei, B.; Rezaee, S.A. Role of IL-21 in HTLV-1 infections with emphasis on HTLV-1-associated myelopathy/tropical spastic paraparesis (HAM/TSP). Med. Microbiol. Immunol. 2017, 206, 195–201. [Google Scholar] [CrossRef]
  234. Kim, S.J.; Kehrl, J.H.; Burton, J.; Tendler, C.L.; Jeang, K.T.; Danielpour, D.; Thevenin, C.; Kim, K.Y.; Sporn, M.B.; Roberts, A.B. Transactivation of the transforming growth factor β1 (TGF-β1) gene by human T lymphotropic virus type 1 tax: A potential mechanism for the increased production of TGF-β1 in adult T cell leukemia. J. Exp. Med. 1990, 172, 121–129. [Google Scholar] [CrossRef]
  235. Kim, S.J.; Winokur, T.S.; Lee, H.D.; Danielpour, D.; Kim, K.Y.; Geiser, A.G.; Chen, L.S.; Sporn, M.B.; Roberts, A.B.; Jay, G. Overexpression of transforming growth factor-β in transgenic mice carrying the human T-cell lymphotropic virus type I tax gene. Mol. Cell Biol. 1991, 11, 5222–5228. [Google Scholar] [CrossRef]
  236. Baum, P.R.; Gayle, R.B., 3rd; Ramsdell, F.; Srinivasan, S.; Sorensen, R.A.; Watson, M.L.; Seldin, M.F.; Baker, E.; Sutherland, G.R.; Clifford, K.N.; et al. Molecular characterization of murine and human OX40/OX40 ligand systems: Identification of a human OX40 ligand as the HTLV-1-regulated protein gp34. EMBO J. 1994, 13, 3992–4001. [Google Scholar] [CrossRef]
  237. Pankow, R.; Durkop, H.; Latza, U.; Krause, H.; Kunzendorf, U.; Pohl, T.; Bulfone-Paus, S. The HTLV-I tax protein transcriptionally modulates OX40 antigen expression. J. Immunol. 2000, 165, 263–270. [Google Scholar] [CrossRef]
  238. Saito, M.; Tanaka, R.; Arishima, S.; Matsuzaki, T.; Ishihara, S.; Tokashiki, T.; Ohya, Y.; Takashima, H.; Umehara, F.; Izumo, S.; et al. Increased expression of OX40 is associated with progressive disease in patients with HTLV-1-associated myelopathy/tropical spastic paraparesis. Retrovirology 2013, 10, 51. [Google Scholar] [CrossRef]
  239. Ohtani, K.; Tsujimoto, A.; Tsukahara, T.; Numata, N.; Miura, S.; Sugamura, K.; Nakamura, M. Molecular mechanisms of promoter regulation of the gp34 gene that is trans-activated by an oncoprotein Tax of human T cell leukemia virus type I. J. Biol. Chem. 1998, 273, 14119–14129. [Google Scholar] [CrossRef] [PubMed]
  240. Motai, Y.; Takahashi, M.; Takachi, T.; Higuchi, M.; Hara, T.; Mizuguchi, M.; Aoyagi, Y.; Terai, S.; Tanaka, Y.; Fujii, M. Human T-cell leukemia virus type 1 (HTLV-1) Tax1 oncoprotein but not HTLV-2 Tax2 induces the expression of OX40 ligand by interacting with p52/p100 and RelB. Virus Genes 2016, 52, 4–13. [Google Scholar] [CrossRef] [PubMed]
  241. Harhaj, E.W.; Harhaj, N.S.; Grant, C.; Mostoller, K.; Alefantis, T.; Sun, S.C.; Wigdahl, B. Human T cell leukemia virus type I Tax activates CD40 gene expression via the NF-κB pathway. Virology 2005, 333, 145–158. [Google Scholar] [CrossRef]
  242. Harhaj, N.S.; Janic, B.; Ramos, J.C.; Harrington, W.J., Jr.; Harhaj, E.W. Deregulated expression of CD40 ligand in HTLV-I infection: Distinct mechanisms of downregulation in HTLV-I-transformed cell lines and ATL patients. Virology 2007, 362, 99–108. [Google Scholar] [CrossRef]
  243. El-Sabban, M.E.; Merhi, R.A.; Haidar, H.A.; Arnulf, B.; Khoury, H.; Basbous, J.; Nijmeh, J.; de The, H.; Hermine, O.; Bazarbachi, A. Human T-cell lymphotropic virus type 1-transformed cells induce angiogenesis and establish functional gap junctions with endothelial cells. Blood 2002, 99, 3383–3389. [Google Scholar] [CrossRef]
  244. Watanabe, T.; Yamaguchi, K.; Takatsuki, K.; Osame, M.; Yoshida, M. Constitutive expression of parathyroid hormone-related protein gene in human T cell leukemia virus type 1 (HTLV-1) carriers and adult T cell leukemia patients that can be trans-activated by HTLV-1 tax gene. J. Exp. Med. 1990, 172, 759–765. [Google Scholar] [CrossRef]
  245. Mori, N.; Ejima, E.; Prager, D. Transactivation of parathyroid hormone-related protein gene expression by human T-cell leukemia virus type I tax. Eur. J. Haematol. 1996, 56, 116–117. [Google Scholar] [CrossRef]
  246. Dittmer, J.; Gitlin, S.D.; Reid, R.L.; Brady, J.N. Transactivation of the P2 promoter of parathyroid hormone-related protein by human T-cell lymphotropic virus type I Tax1: Evidence for the involvement of transcription factor Ets1. J. Virol. 1993, 67, 6087–6095. [Google Scholar] [CrossRef]
  247. Ejima, E.; Rosenblatt, J.D.; Massari, M.; Quan, E.; Stephens, D.; Rosen, C.A.; Prager, D. Cell-type-specific transactivation of the parathyroid hormone-related protein gene promoter by the human T-cell leukemia virus type I (HTLV-I) tax and HTLV-II tax proteins. Blood 1993, 81, 1017–1024. [Google Scholar] [CrossRef] [PubMed]
  248. Dittmer, J.; Pise-Masison, C.A.; Clemens, K.E.; Choi, K.S.; Brady, J.N. Interaction of human T-cell lymphotropic virus type I Tax, Ets1, and Sp1 in transactivation of the PTHrP P2 promoter. J. Biol. Chem. 1997, 272, 4953–4958. [Google Scholar] [CrossRef] [PubMed]
  249. Nadella, M.V.; Dirksen, W.P.; Nadella, K.S.; Shu, S.; Cheng, A.S.; Morgenstern, J.A.; Richard, V.; Fernandez, S.A.; Huang, T.H.; Guttridge, D.; et al. Transcriptional regulation of parathyroid hormone-related protein promoter P2 by NF-κB in adult T-cell leukemia/lymphoma. Leukemia 2007, 21, 1752–1762. [Google Scholar] [CrossRef] [PubMed]
  250. Nadella, M.V.; Shu, S.T.; Dirksen, W.P.; Thudi, N.K.; Nadella, K.S.; Fernandez, S.A.; Lairmore, M.D.; Green, P.L.; Rosol, T.J. Expression of parathyroid hormone-related protein during immortalization of human peripheral blood mononuclear cells by HTLV-1: Implications for transformation. Retrovirology 2008, 5, 46. [Google Scholar] [CrossRef]
  251. Yoshie, O.; Fujisawa, R.; Nakayama, T.; Harasawa, H.; Tago, H.; Izawa, D.; Hieshima, K.; Tatsumi, Y.; Matsushima, K.; Hasegawa, H.; et al. Frequent expression of CCR4 in adult T-cell leukemia and human T-cell leukemia virus type 1-transformed T cells. Blood 2002, 99, 1505–1511. [Google Scholar] [CrossRef]
  252. Yoshie, O. Expression of CCR4 in adult T-cell leukemia. Leuk. Lymphoma 2005, 46, 185–190. [Google Scholar] [CrossRef]
  253. Nakayama, T.; Hieshima, K.; Arao, T.; Jin, Z.; Nagakubo, D.; Shirakawa, A.K.; Yamada, Y.; Fujii, M.; Oiso, N.; Kawada, A.; et al. Aberrant expression of Fra-2 promotes CCR4 expression and cell proliferation in adult T-cell leukemia. Oncogene 2008, 27, 3221–3232. [Google Scholar] [CrossRef] [PubMed]
  254. Sawada, M.; Suzumura, A.; Kondo, N.; Marunouchi, T. Induction of cytokines in glial cells by trans activator of human T-cell lymphotropic virus type I. FEBS Lett. 1992, 313, 47–50. [Google Scholar] [CrossRef] [PubMed]
  255. Mori, N.; Prager, D. Transactivation of the interleukin-1alpha promoter by human T-cell leukemia virus type I and type II Tax proteins. Blood 1996, 87, 3410–3417. [Google Scholar] [CrossRef]
  256. Banerjee, P.; Rochford, R.; Antel, J.; Canute, G.; Wrzesinski, S.; Sieburg, M.; Feuer, G. Proinflammatory cytokine gene induction by human T-cell leukemia virus type 1 (HTLV-1) and HTLV-2 Tax in primary human glial cells. J. Virol. 2007, 81, 1690–1700. [Google Scholar] [CrossRef]
  257. Nakayama, Y.; Ishikawa, C.; Tamaki, K.; Senba, M.; Fujita, J.; Mori, N. Interleukin-1α produced by human T-cell leukaemia virus type I-infected T cells induces intercellular adhesion molecule-1 expression on lung epithelial cells. J. Med. Microbiol. 2011, 60, 1750–1761. [Google Scholar] [CrossRef]
  258. Rauch, D.A.; Harding, J.C.; Ratner, L. IL-15 deficient tax mice reveal a role for IL-1α in tumor immunity. PLoS ONE 2014, 9, e85028. [Google Scholar] [CrossRef] [PubMed]
  259. Tsukada, J.; Misago, M.; Serino, Y.; Ogawa, R.; Murakami, S.; Nakanishi, M.; Tonai, S.; Kominato, Y.; Morimoto, I.; Auron, P.E.; et al. Human T-cell leukemia virus type I Tax transactivates the promoter of human prointerleukin-1β gene through association with two transcription factors, nuclear factor-interleukin-6 and Spi-1. Blood 1997, 90, 3142–3153. [Google Scholar] [CrossRef]
  260. Li-Weber, M.; Giaisi, M.; Chlichlia, K.; Khazaie, K.; Krammer, P.H. Human T cell leukemia virus type I Tax enhances IL-4 gene expression in T cells. Eur. J. Immunol. 2001, 31, 2623–2632. [Google Scholar] [CrossRef]
  261. Blumenthal, S.G.; Aichele, G.; Wirth, T.; Czernilofsky, A.P.; Nordheim, A.; Dittmer, J. Regulation of the human interleukin-5 promoter by Ets transcription factors. Ets1 and Ets2, but not Elf-1, cooperate with GATA3 and HTLV-I Tax1. J. Biol. Chem. 1999, 274, 12910–12916. [Google Scholar] [CrossRef]
  262. Muraoka, O.; Kaisho, T.; Tanabe, M.; Hirano, T. Transcriptional activation of the interleukin-6 gene by HTLV-1 p40tax through an NF-κB-like binding site. Immunol. Lett. 1993, 37, 159–165. [Google Scholar] [CrossRef] [PubMed]
  263. Mori, N.; Shirakawa, F.; Shimizu, H.; Murakami, S.; Oda, S.; Yamamoto, K.; Eto, S. Transcriptional regulation of the human interleukin-6 gene promoter in human T-cell leukemia virus type I-infected T-cell lines: Evidence for the involvement of NF-κB. Blood 1994, 84, 2904–2911. [Google Scholar] [CrossRef]
  264. Mori, N.; Shirakawa, F.; Abe, M.; Kamo, Y.; Koyama, Y.; Murakami, S.; Shimizu, H.; Yamamoto, K.; Oda, S.; Eto, S. Human T-cell leukemia virus type I tax transactivates the interleukin-6 gene in human rheumatoid synovial cells. J. Rheumatol. 1995, 22, 2049–2054. [Google Scholar]
  265. Horiuchi, S.; Yamamoto, N.; Dewan, M.Z.; Takahashi, Y.; Yamashita, A.; Yoshida, T.; Nowell, M.A.; Richards, P.J.; Jones, S.A.; Yamamoto, N. Human T-cell leukemia virus type-I Tax induces expression of interleukin-6 receptor (IL-6R): Shedding of soluble IL-6R and activation of STAT3 signaling. Int. J. Cancer 2006, 119, 823–830. [Google Scholar] [CrossRef]
  266. Refaat, A.; Zhou, Y.; Suzuki, S.; Takasaki, I.; Koizumi, K.; Yamaoka, S.; Tabuchi, Y.; Saiki, I.; Sakurai, H. Distinct roles of transforming growth factor-β-activated kinase 1 (TAK1)-c-Rel and interferon regulatory factor 4 (IRF4) pathways in human T cell lymphotropic virus 1-transformed T helper 17 cells producing interleukin-9. J. Biol. Chem. 2011, 286, 21092–21099. [Google Scholar] [CrossRef]
  267. Mori, N.; Murakami, S.; Oda, S.; Prager, D.; Eto, S. Production of interleukin 8 in adult T-cell leukemia cells: Possible transactivation of the interleukin 8 gene by human T-cell leukemia virus type I tax. Cancer Res. 1995, 55, 3592–3597. [Google Scholar] [PubMed]
  268. Mori, N.; Gill, P.S.; Mougdil, T.; Murakami, S.; Eto, S.; Prager, D. Interleukin-10 gene expression in adult T-cell leukemia. Blood 1996, 88, 1035–1045. [Google Scholar] [CrossRef] [PubMed]
  269. Mori, N.; Prager, D. Interleukin-10 gene expression and adult T-cell leukemia. Leuk. Lymphoma 1998, 29, 239–248. [Google Scholar] [CrossRef]
  270. Ahuja, J.; Lepoutre, V.; Wigdahl, B.; Khan, Z.K.; Jain, P. Induction of pro-inflammatory cytokines by human T-cell leukemia virus type-1 Tax protein as determined by multiplexed cytokine protein array analyses of human dendritic cells. Biomed. Pharmacother. 2007, 61, 201–208. [Google Scholar] [CrossRef]
  271. Chung, H.K.; Young, H.A.; Goon, P.K.; Heidecker, G.; Princler, G.L.; Shimozato, O.; Taylor, G.P.; Bangham, C.R.; Derse, D. Activation of interleukin-13 expression in T cells from HTLV-1-infected individuals and in chronically infected cell lines. Blood 2003, 102, 4130–4136. [Google Scholar] [CrossRef]
  272. Waldele, K.; Schneider, G.; Ruckes, T.; Grassmann, R. Interleukin-13 overexpression by tax transactivation: A potential autocrine stimulus in human T-cell leukemia virus-infected lymphocytes. J. Virol. 2004, 78, 6081–6090. [Google Scholar] [CrossRef] [PubMed]
  273. Silbermann, K.; Schneider, G.; Grassmann, R. Stimulation of interleukin-13 expression by human T-cell leukemia virus type 1 oncoprotein Tax via a dually active promoter element responsive to NF-κB and NFAT. J. Gen. Virol. 2008, 89, 2788–2798. [Google Scholar] [CrossRef]
  274. Dodon, M.D.; Li, Z.; Hamaia, S.; Gazzolo, L. Tax protein of human T-cell leukaemia virus type 1 induces interleukin 17 gene expression in T cells. J. Gen. Virol. 2004, 85, 1921–1932. [Google Scholar] [CrossRef] [PubMed]
  275. Lavorgna, A.; Matsuoka, M.; Harhaj, E.W. A critical role for IL-17RB signaling in HTLV-1 tax-induced NF-κB activation and T-cell transformation. PLoS Pathog. 2014, 10, e1004418. [Google Scholar] [CrossRef]
  276. Araya, N.; Sato, T.; Ando, H.; Tomaru, U.; Yoshida, M.; Coler-Reilly, A.; Yagishita, N.; Yamauchi, J.; Hasegawa, A.; Kannagi, M.; et al. HTLV-1 induces a Th1-like state in CD4+CCR4+ T cells. J. Clin. Investig. 2014, 124, 3431–3442. [Google Scholar] [CrossRef]
  277. Mendez, E.; Kawanishi, T.; Clemens, K.; Siomi, H.; Soldan, S.S.; Calabresi, P.; Brady, J.; Jacobson, S. Astrocyte-specific expression of human T-cell lymphotropic virus type 1 (HTLV-1) Tax: Induction of tumor necrosis factor α and susceptibility to lysis by CD8+ HTLV-1-specific cytotoxic T cells. J. Virol. 1997, 71, 9143–9149. [Google Scholar] [CrossRef]
  278. Cowan, E.P.; Alexander, R.K.; Daniel, S.; Kashanchi, F.; Brady, J.N. Induction of tumor necrosis factor α in human neuronal cells by extracellular human T-cell lymphotropic virus type 1 Tax. J. Virol. 1997, 71, 6982–6989. [Google Scholar] [CrossRef]
  279. Tzagarakis-Foster, C.; Geleziunas, R.; Lomri, A.; An, J.; Leitman, D.C. Estradiol represses human T-cell leukemia virus type 1 Tax activation of tumor necrosis factor-α gene transcription. J. Biol. Chem. 2002, 277, 44772–44777. [Google Scholar] [CrossRef] [PubMed]
  280. Paul, N.L.; Millet, I.; Ruddle, N.H. The lymphotoxin promoter is stimulated by HTLV-I tax activation of NF-κB in human T-cell lines. Cytokine 1993, 5, 372–378. [Google Scholar] [CrossRef]
  281. Tschachler, E.; Bohnlein, E.; Felzmann, S.; Reitz, M.S., Jr. Human T-lymphotropic virus type I tax regulates the expression of the human lymphotoxin gene. Blood 1993, 81, 95–100. [Google Scholar] [CrossRef]
  282. Nimer, S.D.; Gasson, J.C.; Hu, K.; Smalberg, I.; Williams, J.L.; Chen, I.S.; Rosenblatt, J.D. Activation of the GM-CSF promoter by HTLV-I and -II tax proteins. Oncogene 1989, 4, 671–676. [Google Scholar]
  283. Himes, S.R.; Coles, L.S.; Katsikeros, R.; Lang, R.K.; Shannon, M.F. HTLV-1 tax activation of the GM-CSF and G-CSF promoters requires the interaction of NF-κB with other transcription factor families. Oncogene 1993, 8, 3189–3197. [Google Scholar]
  284. Miyazato, A.; Kawakami, K.; Iwakura, Y.; Saito, A. Chemokine synthesis and cellular inflammatory changes in lungs of mice bearing p40tax of human T-lymphotropic virus type 1. Clin. Exp. Immunol. 2000, 120, 113–124. [Google Scholar] [CrossRef] [PubMed]
  285. Mori, N.; Ueda, A.; Ikeda, S.; Yamasaki, Y.; Yamada, Y.; Tomonaga, M.; Morikawa, S.; Geleziunas, R.; Yoshimura, T.; Yamamoto, N. Human T-cell leukemia virus type I tax activates transcription of the human monocyte chemoattractant protein-1 gene through two nuclear factor-κB sites. Cancer Res. 2000, 60, 4939–4945. [Google Scholar]
  286. Saito, M.; Sejima, H.; Naito, T.; Ushirogawa, H.; Matsuzaki, T.; Matsuura, E.; Tanaka, Y.; Nakamura, T.; Takashima, H. The CC chemokine ligand (CCL) 1, upregulated by the viral transactivator Tax, can be downregulated by minocycline: Possible implications for long-term treatment of HTLV-1-associated myelopathy/tropical spastic paraparesis. Virol. J. 2017, 14, 234. [Google Scholar] [CrossRef]
  287. Sharma, V.; May, C.C. Human T-cell lymphotrophic virus type-I tax gene induces secretion of human macrophage inflammatory protein-1α. Biochem. Biophys. Res. Commun. 1999, 262, 429–432. [Google Scholar] [CrossRef]
  288. Napolitano, M.; Modi, W.S.; Cevario, S.J.; Gnarra, J.R.; Seuanez, H.N.; Leonard, W.J. The gene encoding the Act-2 cytokine. Genomic structure, HTLV-I/Tax responsiveness of 5′ upstream sequences, and chromosomal localization. J. Biol. Chem. 1991, 266, 17531–17536. [Google Scholar] [CrossRef]
  289. Sharma, V.; Lorey, S.L. Autocrine role of macrophage inflammatory protein-1 β in human T-cell lymphotropic virus type-I tax-transfected Jurkat T-cells. Biochem. Biophys. Res. Commun. 2001, 287, 910–913. [Google Scholar] [CrossRef]
  290. Mori, N.; Krensky, A.M.; Ohshima, K.; Tomita, M.; Matsuda, T.; Ohta, T.; Yamada, Y.; Tomonaga, M.; Ikeda, S.; Yamamoto, N. Elevated expression of CCL5/RANTES in adult T-cell leukemia cells: Possible transactivation of the CCL5 gene by human T-cell leukemia virus type I tax. Int. J. Cancer 2004, 111, 548–557. [Google Scholar] [CrossRef]
  291. Imaizumi, Y.; Sugita, S.; Yamamoto, K.; Imanishi, D.; Kohno, T.; Tomonaga, M.; Matsuyama, T. Human T cell leukemia virus type-I Tax activates human macrophage inflammatory protein-3α/CCL20 gene transcription via the NF-κB pathway. Int. Immunol. 2002, 14, 147–155. [Google Scholar] [CrossRef]
  292. Hieshima, K.; Nagakubo, D.; Nakayama, T.; Shirakawa, A.K.; Jin, Z.; Yoshie, O. Tax-inducible production of CC chemokine ligand 22 by human T cell leukemia virus type 1 (HTLV-1)-infected T cells promotes preferential transmission of HTLV-1 to CCR4-expressing CD4+ T cells. J. Immunol. 2008, 180, 931–939. [Google Scholar] [CrossRef]
  293. Naito, T.; Yasunaga, J.I.; Mitobe, Y.; Shirai, K.; Sejima, H.; Ushirogawa, H.; Tanaka, Y.; Nakamura, T.; Hanada, K.; Fujii, M.; et al. Distinct gene expression signatures induced by viral transactivators of different HTLV-1 subgroups that confer a different risk of HAM/TSP. Retrovirology 2018, 15, 72. [Google Scholar] [CrossRef]
  294. Souza, F.D.S.; Freitas, N.L.; Gomes, Y.C.P.; Torres, R.C.; Echevarria-Lima, J.; da Silva-Filho, I.L.; Leite, A.; de Lima, M.; da Silva, M.T.T.; Araujo, A.Q.C.; et al. Following the clues: Usefulness of biomarkers of neuroinflammation and neurodegeneration in the investigation of HTLV-1-associated myelopathy progression. Front. Immunol. 2021, 12, 737941. [Google Scholar] [CrossRef]
  295. Arai, M.; Ohashi, T.; Tsukahara, T.; Murakami, T.; Hori, T.; Uchiyama, T.; Yamamoto, N.; Kannagi, M.; Fujii, M. Human T-cell leukemia virus type 1 Tax protein induces the expression of lymphocyte chemoattractant SDF-1/PBSF. Virology 1998, 241, 298–303. [Google Scholar] [CrossRef]
  296. Moriuchi, M.; Moriuchi, H.; Fauci, A.S. HTLV type I Tax activation of the CXCR4 promoter by association with nuclear respiratory factor 1. AIDS Res. Hum. Retroviruses 1999, 15, 821–827. [Google Scholar] [CrossRef]
  297. Araya, N.; Yamagishi, M.; Nakashima, M.; Asahara, N.; Kiyohara, K.; Aratani, S.; Yagishita, N.; Horibe, E.; Ishizaki, I.; Watanabe, T.; et al. Virus-induced RGMa expression drives neurodegeneration in HTLV-1-associated myelopathy. JCI Insight 2025, 10, e184530. [Google Scholar] [CrossRef]
  298. Valentin, H.; Hamaia, S.; Konig, S.; Gazzolo, L. Vascular cell adhesion molecule-1 induced by human T-cell leukaemia virus type 1 Tax protein in T-cells stimulates proliferation of human T-lymphocytes. J. Gen. Virol. 2001, 82, 831–835. [Google Scholar] [CrossRef]
  299. Murakami, Y. Involvement of a cell adhesion molecule, TSLC1/IGSF4, in human oncogenesis. Cancer Sci. 2005, 96, 543–552. [Google Scholar] [CrossRef]
  300. Hartsough, E.J.; Weiss, M.B.; Heilman, S.A.; Purwin, T.J.; Kugel, C.H., 3rd; Rosenbaum, S.R.; Erkes, D.A.; Tiago, M.; HooKim, K.; Chervoneva, I.; et al. CADM1 is a TWIST1-regulated suppressor of invasion and survival. Cell Death Dis. 2019, 10, 281. [Google Scholar] [CrossRef]
  301. Nakahata, S.; Morishita, K. CADM1/TSLC1 is a novel cell surface marker for adult T-cell leukemia/lymphoma. J. Clin. Exp. Hematop. 2012, 52, 17–22. [Google Scholar] [CrossRef]
  302. Zhang, J.; Yamada, O.; Matsushita, Y.; Chagan-Yasutan, H.; Hattori, T. Transactivation of human osteopontin promoter by human T-cell leukemia virus type 1-encoded Tax protein. Leuk. Res. 2010, 34, 763–768. [Google Scholar] [CrossRef]
  303. Zhang, J.; Yamada, O.; Kida, S.; Matsushita, Y.; Yamaoka, S.; Chagan-Yasutan, H.; Hattori, T. Identification of CD44 as a downstream target of noncanonical NF-κB pathway activated by human T-cell leukemia virus type 1-encoded Tax protein. Virology 2011, 413, 244–252. [Google Scholar] [CrossRef]
  304. Chagan-Yasutan, H.; Tsukasaki, K.; Takahashi, Y.; Oguma, S.; Harigae, H.; Ishii, N.; Zhang, J.; Fukumoto, M.; Hattori, T. Involvement of osteopontin and its signaling molecule CD44 in clinicopathological features of adult T cell leukemia. Leuk. Res. 2011, 35, 1484–1490. [Google Scholar] [CrossRef]
  305. Wu, M.; Wong, H.Y.; Lin, J.L.; Moliner, A.; Schwarz, H. Induction of CD137 expression by viral genes reduces T cell costimulation. J. Cell Physiol. 2019, 234, 21076–21088. [Google Scholar] [CrossRef]
  306. Kitajima, I.; Kawahara, K.; Hanyu, N.; Shin, H.; Tokioka, T.; Soejima, Y.; Tsutsui, J.; Ozawa, M.; Shimayama, T.; Maruyama, I. Enhanced E-cadherin expression and increased calcium-dependent cell-cell adhesion in human T-cell leukemia virus type I Tax-expressing PC12 cells. J. Cell Sci. 1996, 109 Pt 3, 609–617. [Google Scholar] [CrossRef]
  307. Pichler, K.; Kattan, T.; Gentzsch, J.; Kress, A.K.; Taylor, G.P.; Bangham, C.R.; Grassmann, R. Strong induction of 4-1BB, a growth and survival promoting costimulatory receptor, in HTLV-1-infected cultured and patients’ T cells by the viral Tax oncoprotein. Blood 2008, 111, 4741–4751. [Google Scholar] [CrossRef]
  308. Masamoto, I.; Yoshimitsu, M.; Kuroki, A.; Horai, S.; Ezinne, C.C.; Kozako, T.; Hachiman, M.; Kamada, Y.; Baba, M.; Arima, N. Clinical significance of CD70 expression on T cells in human T-lymphotropic virus type-1 carriers and adult T cell leukemia/lymphoma patients. Leuk. Lymphoma 2016, 57, 685–691. [Google Scholar] [CrossRef]
  309. Ishikawa, C.; Kawakami, H.; Uchihara, J.N.; Senba, M.; Mori, N. CD69 overexpression by human T-cell leukemia virus type 1 Tax transactivation. Biochim. Biophys. Acta 2013, 1833, 1542–1552. [Google Scholar] [CrossRef]
  310. Sawada, M.; Suzumura, A.; Yoshida, M.; Marunouchi, T. Human T-cell leukemia virus type I trans activator induces class I major histocompatibility complex antigen expression in glial cells. J. Virol. 1990, 64, 4002–4006. [Google Scholar] [CrossRef]
  311. Green, J.E. trans activation of nerve growth factor in transgenic mice containing the human T-cell lymphotropic virus type I tax gene. Mol. Cell Biol. 1991, 11, 4635–4641. [Google Scholar] [CrossRef]
  312. Joshi, J.B.; Dave, H.P. Transactivation of the proenkephalin gene promoter by the Tax1 protein of human T-cell lymphotropic virus type I. Proc. Natl. Acad. Sci. USA 1992, 89, 1006–1010. [Google Scholar] [CrossRef]
  313. Low, K.G.; Chu, H.M.; Schwartz, P.M.; Daniels, G.M.; Melner, M.H.; Comb, M.J. Novel interactions between human T-cell leukemia virus type I Tax and activating transcription factor 3 at a cyclic AMP-responsive element. Mol. Cell Biol. 1994, 14, 4958–4974. [Google Scholar] [CrossRef]
  314. Fu, W.; Shah, S.R.; Jiang, H.; Hilt, D.C.; Dave, H.P.; Joshi, J.B. Transactivation of proenkephalin gene by HTLV-1 tax1 protein in glial cells: Involvement of Fos/Jun complex at an AP-1 element in the proenkephalin gene promoter. J. Neurovirol 1997, 3, 16–27. [Google Scholar] [CrossRef]
  315. Alexandre, C.; Verrier, B. Four regulatory elements in the human c-fos promoter mediate transactivation by HTLV-1 Tax protein. Oncogene 1991, 6, 543–551. [Google Scholar]
  316. Iwakura, Y.; Tosu, M.; Yoshida, E.; Saijo, S.; Nakayama-Yamada, J.; Itagaki, K.; Asano, M.; Siomi, H.; Hatanaka, M.; Takeda, T.; et al. Augmentation of c-fos and c-jun expression in transgenic mice carrying the human T-cell leukemia virus type-I tax gene. Virus Genes. 1995, 9, 161–170. [Google Scholar] [CrossRef] [PubMed]
  317. Sakamoto, K.M.; Nimer, S.D.; Rosenblatt, J.D.; Gasson, J.C. HTLV-I and HTLV-II tax trans-activate the human EGR-1 promoter through different cis-acting sequences. Oncogene 1992, 7, 2125–2130. [Google Scholar]
  318. Fujii, M.; Tsuchiya, H.; Chuhjo, T.; Akizawa, T.; Seiki, M. Interaction of HTLV-1 Tax1 with p67SRF causes the aberrant induction of cellular immediate early genes through CArG boxes. Genes. Dev. 1992, 6, 2066–2076. [Google Scholar] [CrossRef]
  319. Li, B.; Fink, T.; Ebbesen, P.; Liu, X.D.; Zachar, V. Expression of butyrate response factor 1 in HTLV-1-transformed cells and its transactivation by tax protein. Arch. Virol. 2003, 148, 1787–1804. [Google Scholar] [CrossRef]
  320. Duyao, M.P.; Kessler, D.J.; Spicer, D.B.; Bartholomew, C.; Cleveland, J.L.; Siekevitz, M.; Sonenshein, G.E. Transactivation of the c-myc promoter by human T cell leukemia virus type 1 tax is mediated by NFκB. J. Biol. Chem. 1992, 267, 16288–16291. [Google Scholar] [CrossRef]
  321. Gazon, H.; Barbeau, B.; Mesnard, J.M.; Peloponese, J.M., Jr. Hijacking of the AP-1 signaling pathway during development of ATL. Front. Microbiol. 2017, 8, 2686. [Google Scholar] [CrossRef]
  322. Fujii, M.; Iwai, K.; Oie, M.; Fukushi, M.; Yamamoto, N.; Kannagi, M.; Mori, N. Activation of oncogenic transcription factor AP-1 in T cells infected with human T cell leukemia virus type 1. AIDS Res. Hum. Retroviruses 2000, 16, 1603–1606. [Google Scholar] [CrossRef]
  323. Mori, N.; Fujii, M.; Iwai, K.; Ikeda, S.; Yamasaki, Y.; Hata, T.; Yamada, Y.; Tanaka, Y.; Tomonaga, M.; Yamamoto, N. Constitutive activation of transcription factor AP-1 in primary adult T-cell leukemia cells. Blood 2000, 95, 3915–3921. [Google Scholar]
  324. Iwai, K.; Mori, N.; Oie, M.; Yamamoto, N.; Fujii, M. Human T-cell leukemia virus type 1 tax protein activates transcription through AP-1 site by inducing DNA binding activity in T cells. Virology 2001, 279, 38–46. [Google Scholar] [CrossRef]
  325. Huang, Q.; Niu, Z.; Han, J.; Liu, X.; Lv, Z.; Li, H.; Yuan, L.; Li, X.; Sun, S.; Wang, H.; et al. HTLV-1 Tax upregulates early growth response protein 1 through nuclear factor-κB signaling. Oncotarget 2017, 8, 51123–51133. [Google Scholar] [CrossRef] [PubMed]
  326. Arima, N.; Molitor, J.A.; Smith, M.R.; Kim, J.H.; Daitoku, Y.; Greene, W.C. Human T-cell leukemia virus type I Tax induces expression of the Rel-related family of κB enhancer-binding proteins: Evidence for a pretranslational component of regulation. J. Virol. 1991, 65, 6892–6899. [Google Scholar] [CrossRef] [PubMed]
  327. Ishikawa, C.; Senba, M.; Hashimoto, T.; Imaizumi, A.; Mori, N. Expression and significance of Pim-3 kinase in adult T-cell leukemia. Eur. J. Haematol. 2017, 99, 495–504. [Google Scholar] [CrossRef]
  328. Bellon, M.; Nicot, C. Feedback loop regulation between Pim kinases and Tax keeps human T-cell leukemia virus type 1 viral replication in check. J. Virol. 2022, 96, e0196021. [Google Scholar] [CrossRef] [PubMed]
  329. Bellon, M.; Nicot, C. Targeting Pim kinases in hematological cancers: Molecular and clinical review. Mol. Cancer 2023, 22, 18. [Google Scholar] [CrossRef]
  330. Migone, T.S.; Lin, J.X.; Cereseto, A.; Mulloy, J.C.; O’Shea, J.J.; Franchini, G.; Leonard, W.J. Constitutively activated Jak-STAT pathway in T cells transformed with HTLV-I. Science 1995, 269, 79–81. [Google Scholar] [CrossRef]
  331. Takemoto, S.; Mulloy, J.C.; Cereseto, A.; Migone, T.S.; Patel, B.K.; Matsuoka, M.; Yamaguchi, K.; Takatsuki, K.; Kamihira, S.; White, J.D.; et al. Proliferation of adult T cell leukemia/lymphoma cells is associated with the constitutive activation of JAK/STAT proteins. Proc. Natl. Acad. Sci. USA 1997, 94, 13897–13902. [Google Scholar] [CrossRef] [PubMed]
  332. Nakamura, N.; Fujii, M.; Tsukahara, T.; Arai, M.; Ohashi, T.; Wakao, H.; Kannagi, M.; Yamamoto, N. Human T-cell leukemia virus type 1 Tax protein induces the expression of STAT1 and STAT5 genes in T-cells. Oncogene 1999, 18, 2667–2675. [Google Scholar] [CrossRef]
  333. Sharma, S.; Mamane, Y.; Grandvaux, N.; Bartlett, J.; Petropoulos, L.; Lin, R.; Hiscott, J. Activation and regulation of interferon regulatory factor 4 in HTLV type 1-infected T lymphocytes. AIDS Res. Hum. Retroviruses 2000, 16, 1613–1622. [Google Scholar] [CrossRef] [PubMed]
  334. Sharma, S.; Grandvaux, N.; Mamane, Y.; Genin, P.; Azimi, N.; Waldmann, T.; Hiscott, J. Regulation of IFN regulatory factor 4 expression in human T cell leukemia virus-I-transformed T cells. J. Immunol. 2002, 169, 3120–3130. [Google Scholar] [CrossRef] [PubMed]
  335. Cherian, M.A.; Olson, S.; Sundaramoorthi, H.; Cates, K.; Cheng, X.; Harding, J.; Martens, A.; Challen, G.A.; Tyagi, M.; Ratner, L.; et al. An activating mutation of interferon regulatory factor 4 (IRF4) in adult T-cell leukemia. J. Biol. Chem. 2018, 293, 6844–6858. [Google Scholar] [CrossRef]
  336. Ishikawa, C.; Senba, M.; Barnes, B.J.; Mori, N. Constitutive expression of IRF-5 in HTLV-1-infected T cells. Int. J. Oncol. 2015, 47, 361–369. [Google Scholar] [CrossRef]
  337. Shimizu, T.; Kawakita, S.; Li, Q.H.; Fukuhara, S.; Fujisawa, J. Human T-cell leukemia virus type 1 Tax protein stimulates the interferon-responsive enhancer element via NF-κB activity. FEBS Lett. 2003, 539, 73–77. [Google Scholar] [CrossRef]
  338. Yamanashi, Y.; Mori, S.; Yoshida, M.; Kishimoto, T.; Inoue, K.; Yamamoto, T.; Toyoshima, K. Selective expression of a protein-tyrosine kinase, p56lyn, in hematopoietic cells and association with production of human T-cell lymphotropic virus type I. Proc. Natl. Acad. Sci. USA 1989, 86, 6538–6542. [Google Scholar] [CrossRef]
  339. Uchiumi, F.; Semba, K.; Yamanashi, Y.; Fujisawa, J.; Yoshida, M.; Inoue, K.; Toyoshima, K.; Yamamoto, T. Characterization of the promoter region of the src family gene lyn and its trans activation by human T-cell leukemia virus type I-encoded p40tax. Mol. Cell Biol. 1992, 12, 3784–3795. [Google Scholar] [CrossRef]
  340. Mills, G.B.; Arima, N.; May, C.; Hill, M.; Schmandt, R.; Li, J.; Miyamoto, N.G.; Greene, W.C. Neither the LCK nor the FYN kinases are obligatory for IL-2-mediated signal transduction in HTLV-I-infected human T cells. Int. Immunol. 1992, 4, 1233–1243. [Google Scholar] [CrossRef]
  341. Letafati, A.; Bahavar, A.; Tabarraei, A.; Norouzi, M.; Amiri, A.; Mozhgani, S.H. Human T-cell lymphotropic virus type 1 (HTLV-1) grip on T-cells: Investigating the viral tapestry of activation. Infect. Agent. Cancer 2024, 19, 23. [Google Scholar] [CrossRef]
  342. Zhang, Y.J.; Song, J.R.; Zhao, M.J. NR4A1 regulates cerebral ischemia-induced brain injury by regulating neuroinflammation through interaction with NF-κB/p65. Biochem. Biophys. Res. Commun. 2019, 518, 59–65. [Google Scholar] [CrossRef]
  343. Takahashi, R.; Yamagishi, M.; Nakano, K.; Yamochi, T.; Yamochi, T.; Fujikawa, D.; Nakashima, M.; Tanaka, Y.; Uchimaru, K.; Utsunomiya, A.; et al. Epigenetic deregulation of Ellis Van Creveld confers robust Hedgehog signaling in adult T-cell leukemia. Cancer Sci. 2014, 105, 1160–1169. [Google Scholar] [CrossRef]
  344. Yamazaki, S. The Nuclear NF-κB Regulator IκBξ: Updates on Its Molecular Functions and Pathophysiological Roles. Cells 2024, 13, 1467. [Google Scholar] [CrossRef] [PubMed]
  345. Kimura, R.; Senba, M.; Cutler, S.J.; Ralph, S.J.; Xiao, G.; Mori, N. Human T cell leukemia virus type I tax-induced IκB-ξ modulates tax-dependent and tax-independent gene expression in T cells. Neoplasia 2013, 15, 1110–1124. [Google Scholar] [CrossRef] [PubMed]
  346. Soltani, S.; Mozhgani, S.H.; Roohinezhad, R.; Emami, S.H.; Emami, M.H.; Solooki, S.; Fattah Hesari, M.; Doroozeh, N.; Norouzi, M. Leukemia-related signaling pathways among HTLV-1-derived adult T cell leukemia/lymphoma and asymptomatic carriers in comparison to normal group. AIDS Res. Hum. Retroviruses 2025, 41, 113–119. [Google Scholar] [CrossRef] [PubMed]
  347. Ishikawa, C.; Mori, N. FX1, a BCL6 inhibitor, reactivates BCL6 target genes and suppresses HTLV-1-infected T cells. Investig. New Drugs 2022, 40, 245–254. [Google Scholar] [CrossRef]
  348. Ishikawa, C.; Mori, N. Exportin-1 is critical for cell proliferation and survival in adult T cell leukemia. Investig. New Drugs 2022, 40, 718–727. [Google Scholar] [CrossRef]
  349. Zhong, W.; Cao, X.; Pan, G.; Niu, Q.; Feng, X.; Xu, M.; Li, M.; Huang, Y.; Yi, Q.; Yan, D. ORP4L is a prerequisite for the induction of T-cell leukemogenesis associated with human T-cell leukemia virus 1. Blood 2022, 139, 1052–1065. [Google Scholar] [CrossRef]
  350. Bellon, M.; Nicot, C. Increased H19/miR-675 expression in adult T-cell leukemia is associated with a unique Notch signature pathway. Int. J. Mol. Sci. 2024, 25, 5130. [Google Scholar] [CrossRef]
  351. Bellon, M.; Moles, R.; Chaib-Mezrag, H.; Pancewicz, J.; Nicot, C. JAG1 overexpression contributes to Notch1 signaling and the migration of HTLV-1-transformed ATL cells. J. Hematol. Oncol. 2018, 11, 119. [Google Scholar] [CrossRef] [PubMed]
  352. Pancewicz, J.; Taylor, J.M.; Datta, A.; Baydoun, H.H.; Waldmann, T.A.; Hermine, O.; Nicot, C. Notch signaling contributes to proliferation and tumor formation of human T-cell leukemia virus type 1-associated adult T-cell leukemia. Proc. Natl. Acad. Sci. USA 2010, 107, 16619–16624. [Google Scholar] [CrossRef]
  353. Yeh, C.H.; Bellon, M.; Pancewicz-Wojtkiewicz, J.; Nicot, C. Oncogenic mutations in the FBXW7 gene of adult T-cell leukemia patients. Proc. Natl. Acad. Sci. USA 2016, 113, 6731–6736. [Google Scholar] [CrossRef] [PubMed]
  354. Bellon, M.; Yeh, C.h.; Bai, X.T.; Nicot, C. The HTLV-I oncoprotein Tax inactivates the tumor suppressor FBXW7. J. Virol. 2024, 98, e0040524. [Google Scholar] [CrossRef]
  355. Cheng, W.; Zheng, T.; Wang, Y.; Cai, K.; Wu, W.; Zhao, T.; Xu, R. Activation of Notch1 signaling by HTLV-1 Tax promotes proliferation of adult T-cell leukemia cells. Biochem. Biophys. Res. Commun. 2019, 512, 598–603. [Google Scholar] [CrossRef]
  356. Miyake-Nishijima, R.; Iwata, S.; Saijo, S.; Kobayashi, H.; Kobayashi, S.; Souta-Kuribara, A.; Hosono, O.; Kawasaki, H.; Tanaka, H.; Ikeda, E.; et al. Role of Crk-associated substrate lymphocyte type in the pathophysiology of rheumatoid arthritis in tax transgenic mice and in humans. Arthritis Rheum. 2003, 48, 1890–1900. [Google Scholar] [CrossRef]
  357. Katayose, T.; Iwata, S.; Oyaizu, N.; Hosono, O.; Yamada, T.; Dang, N.H.; Hatano, R.; Tanaka, H.; Ohnuma, K.; Morimoto, C. The role of Cas-L/NEDD9 as a regulator of collagen-induced arthritis in a murine model. Biochem. Biophys. Res. Commun. 2015, 460, 1069–1075. [Google Scholar] [CrossRef]
  358. Ohashi, M.; Sakurai, M.; Higuchi, M.; Mori, N.; Fukushi, M.; Oie, M.; Coffey, R.J.; Yoshiura, K.; Tanaka, Y.; Uchiyama, M.; et al. Human T-cell leukemia virus type 1 Tax oncoprotein induces and interacts with a multi-PDZ domain protein, MAGI-3. Virology 2004, 320, 52–62. [Google Scholar] [CrossRef]
  359. Shoji, T.; Higuchi, M.; Kondo, R.; Takahashi, M.; Oie, M.; Tanaka, Y.; Aoyagi, Y.; Fujii, M. Identification of a novel motif responsible for the distinctive transforming activity of human T-cell leukemia virus (HTLV) type 1 Tax1 protein from HTLV-2 Tax2. Retrovirology 2009, 6, 83. [Google Scholar] [CrossRef]
  360. Hirata, A.; Higuchi, M.; Niinuma, A.; Ohashi, M.; Fukushi, M.; Oie, M.; Akiyama, T.; Tanaka, Y.; Gejyo, F.; Fujii, M. PDZ domain-binding motif of human T-cell leukemia virus type 1 Tax oncoprotein augments the transforming activity in a rat fibroblast cell line. Virology 2004, 318, 327–336. [Google Scholar] [CrossRef] [PubMed]
  361. Arpin-Andre, C.; Mesnard, J.M. The PDZ domain-binding motif of the human T cell leukemia virus type 1 tax protein induces mislocalization of the tumor suppressor hScrib in T cells. J. Biol. Chem. 2007, 282, 33132–33141. [Google Scholar] [CrossRef]
  362. Makokha, G.N.; Takahashi, M.; Higuchi, M.; Saito, S.; Tanaka, Y.; Fujii, M. Human T-cell leukemia virus type 1 Tax protein interacts with and mislocalizes the PDZ domain protein MAGI-1. Cancer Sci. 2013, 104, 313–320. [Google Scholar] [CrossRef]
  363. Tsubata, C.; Higuchi, M.; Takahashi, M.; Oie, M.; Tanaka, Y.; Gejyo, F.; Fujii, M. PDZ domain-binding motif of human T-cell leukemia virus type 1 Tax oncoprotein is essential for the interleukin 2 independent growth induction of a T-cell line. Retrovirology 2005, 2, 46. [Google Scholar] [CrossRef]
  364. Peres, E.; Blin, J.; Ricci, E.P.; Artesi, M.; Hahaut, V.; Van den Broeke, A.; Corbin, A.; Gazzolo, L.; Ratner, L.; Jalinot, P.; et al. PDZ domain-binding motif of Tax sustains T-cell proliferation in HTLV-1-infected humanized mice. PLoS Pathog. 2018, 14, e1006933. [Google Scholar] [CrossRef] [PubMed]
  365. Cherian, M.A.; Baydoun, H.H.; Al-Saleem, J.; Shkriabai, N.; Kvaratskhelia, M.; Green, P.; Ratner, L. Akt Pathway Activation by Human T-cell Leukemia Virus Type 1 Tax Oncoprotein. J. Biol. Chem. 2015, 290, 26270–26281. [Google Scholar] [CrossRef]
  366. Adachi, Y.; Kitahara-Ozawa, A.; Sugamura, K.; Lee, W.J.; Yodoi, J.; Maki, M.; Murachi, T.; Hatanaka, M. Expression of calpain II gene in human hematopoietic system cells infected with human T-cell leukemia virus type I. J. Biol. Chem. 1992, 267, 19373–19378. [Google Scholar] [CrossRef] [PubMed]
  367. Morford, L.A.; Forrest, K.; Logan, B.; Overstreet, L.K.; Goebel, J.; Brooks, W.H.; Roszman, T.L. Calpain II colocalizes with detergent-insoluble rafts on human and Jurkat T-cells. Biochem. Biophys. Res. Commun. 2002, 295, 540–546. [Google Scholar] [CrossRef] [PubMed]
  368. Malumbres, M. Cyclin-dependent kinases. Genome Biol. 2014, 15, 122. [Google Scholar] [CrossRef]
  369. Sherr, C.J. D-type cyclins. Trends Biochem. Sci. 1995, 20, 187–190. [Google Scholar] [CrossRef]
  370. Ohtani, K. Implication of transcription factor E2F in regulation of DNA replication. Front. Biosci. 1999, 4, D793–D804. [Google Scholar] [CrossRef]
  371. Zhou, Y.; Nakajima, R.; Shirasawa, M.; Fikriyanti, M.; Zhao, L.; Iwanaga, R.; Bradford, A.P.; Kurayoshi, K.; Araki, K.; Ohtani, K. Expanding Roles of the E2F-RB-p53 Pathway in Tumor Suppression. Biology 2023, 12, 1511. [Google Scholar] [CrossRef]
  372. Kent, L.N.; Leone, G. The broken cycle: E2F dysfunction in cancer. Nat. Rev. Cancer 2019, 19, 326–338. [Google Scholar] [CrossRef]
  373. Iwanaga, R.; Ohtani, K.; Hayashi, T.; Nakamura, M. Molecular mechanism of cell cycle progression induced by the oncogene product Tax of human T-cell leukemia virus type I. Oncogene 2001, 20, 2055–2067. [Google Scholar] [CrossRef] [PubMed]
  374. Santiago, F.; Clark, E.; Chong, S.; Molina, C.; Mozafari, F.; Mahieux, R.; Fujii, M.; Azimi, N.; Kashanchi, F. Transcriptional up-regulation of the cyclin D2 gene and acquisition of new cyclin-dependent kinase partners in human T-cell leukemia virus type 1-infected cells. J. Virol. 1999, 73, 9917–9927. [Google Scholar] [CrossRef] [PubMed]
  375. Mori, N.; Fujii, M.; Hinz, M.; Nakayama, K.; Yamada, Y.; Ikeda, S.; Yamasaki, Y.; Kashanchi, F.; Tanaka, Y.; Tomonaga, M.; et al. Activation of cyclin D1 and D2 promoters by human T-cell leukemia virus type I tax protein is associated with IL-2-independent growth of T cells. Int. J. Cancer 2002, 99, 378–385. [Google Scholar] [CrossRef]
  376. Lemasson, I.; Thebault, S.; Sardet, C.; Devaux, C.; Mesnard, J.M. Activation of E2F-mediated transcription by human T-cell leukemia virus type I Tax protein in a p16INK4A-negative T-cell line. J. Biol. Chem. 1998, 273, 23598–23604. [Google Scholar] [CrossRef] [PubMed]
  377. Huang, Y.; Ohtani, K.; Iwanaga, R.; Matsumura, Y.; Nakamura, M. Direct trans-activation of the human cyclin D2 gene by the oncogene product Tax of human T-cell leukemia virus type I. Oncogene 2001, 20, 1094–1102. [Google Scholar] [CrossRef]
  378. Iwanaga, R.; Ozono, E.; Fujisawa, J.; Ikeda, M.A.; Okamura, N.; Huang, Y.; Ohtani, K. Activation of the cyclin D2 and cdk6 genes through NF-κB is critical for cell-cycle progression induced by HTLV-I Tax. Oncogene 2008, 27, 5635–5642. [Google Scholar] [CrossRef]
  379. Fisher, R.P. Secrets of a double agent: CDK7 in cell-cycle control and transcription. J. Cell Sci. 2005, 118, 5171–5180. [Google Scholar] [CrossRef]
  380. Fisher, R.P.; Morgan, D.O. A novel cyclin associates with MO15/CDK7 to form the CDK-activating kinase. Cell 1994, 78, 713–724. [Google Scholar] [CrossRef]
  381. Yee, A.; Nichols, M.A.; Wu, L.; Hall, F.L.; Kobayashi, R.; Xiong, Y. Molecular cloning of CDK7-associated human MAT1, a cyclin-dependent kinase-activating kinase (CAK) assembly factor. Cancer Res. 1995, 55, 6058–6062. [Google Scholar]
  382. Akhtar, M.S.; Heidemann, M.; Tietjen, J.R.; Zhang, D.W.; Chapman, R.D.; Eick, D.; Ansari, A.Z. TFIIH kinase places bivalent marks on the carboxy-terminal domain of RNA polymerase II. Mol. Cell 2009, 34, 387–393. [Google Scholar] [CrossRef]
  383. Glover-Cutter, K.; Larochelle, S.; Erickson, B.; Zhang, C.; Shokat, K.; Fisher, R.P.; Bentley, D.L. TFIIH-associated Cdk7 kinase functions in phosphorylation of C-terminal domain Ser7 residues, promoter-proximal pausing, and termination by RNA polymerase II. Mol. Cell Biol. 2009, 29, 5455–5464. [Google Scholar] [CrossRef]
  384. Shirasawa, M.; Nakajima, R.; Zhou, Y.; Zhao, L.; Fikriyanti, M.; Iwanaga, R.; Bradford, A.P.; Kurayoshi, K.; Araki, K.; Ohtani, K. Activation of the CDK7 gene, coding for the catalytic subunit of the cyclin-dependent kinase (CDK)-activating kinase (CAK) and general transcription factor II H, by the trans-activator protein Tax of Human T-Cell Leukemia Virus yype-1. Genes 2024, 15, 1080. [Google Scholar] [CrossRef]
  385. Fujinaga, K.; Huang, F.; Peterlin, B.M. P-TEFb: The master regulator of transcription elongation. Mol. Cell 2023, 83, 393–403. [Google Scholar] [CrossRef] [PubMed]
  386. Mann, M.C.; Strobel, S.; Fleckenstein, B.; Kress, A.K. The transcription elongation factor ELL2 is specifically upregulated in HTLV-1-infected T-cells and is dependent on the viral oncoprotein Tax. Virology 2014, 464–465, 98–110. [Google Scholar] [CrossRef] [PubMed]
  387. Kohrt, S.; Strobel, S.; Mann, M.C.; Sticht, H.; Fleckenstein, B.; Thoma-Kress, A.K. Characterizing the Interaction between the HTLV-1 transactivator Tax-1 with transcription elongation factor ELL2 and its impact on viral transactivation. Int. J. Mol. Sci. 2021, 22, 13597. [Google Scholar] [CrossRef] [PubMed]
  388. Akagi, T.; Ono, H.; Shimotohno, K. Expression of cell-cycle regulatory genes in HTLV-I infected T-cell lines: Possible involvement of Tax1 in the altered expression of cyclin D2, p18Ink4 and p21Waf1/Cip1/Sdi1. Oncogene 1996, 12, 1645–1652. [Google Scholar]
  389. de La Fuente, C.; Deng, L.; Santiago, F.; Arce, L.; Wang, L.; Kashanchi, F. Gene expression array of HTLV type 1-infected T cells: Up-regulation of transcription factors and cell cycle genes. AIDS Res. Hum. Retroviruses 2000, 16, 1695–1700. [Google Scholar] [CrossRef]
  390. de La Fuente, C.; Santiago, F.; Chong, S.Y.; Deng, L.; Mayhood, T.; Fu, P.; Stein, D.; Denny, T.; Coffman, F.; Azimi, N.; et al. Overexpression of p21waf1 in human T-cell lymphotropic virus type 1-infected cells and its association with cyclin A/cdk2. J. Virol. 2000, 74, 7270–7283. [Google Scholar] [CrossRef]
  391. Chowdhury, I.H.; Farhadi, A.; Wang, X.F.; Robb, M.L.; Birx, D.L.; Kim, J.H. Human T-cell leukemia virus type 1 Tax activates cyclin-dependent kinase inhibitor p21/Waf1/Cip1 expression through a p53-independent mechanism: Inhibition of cdk2. Int. J. Cancer 2003, 107, 603–611. [Google Scholar] [CrossRef] [PubMed]
  392. Zhang, L.; Zhi, H.; Liu, M.; Kuo, Y.L.; Giam, C.Z. Induction of p21CIP1/WAF1 expression by human T-lymphotropic virus type 1 Tax requires transcriptional activation and mRNA stabilization. Retrovirology 2009, 6, 35. [Google Scholar] [CrossRef] [PubMed]
  393. LaBaer, J.; Garrett, M.D.; Stevenson, L.F.; Slingerland, J.M.; Sandhu, C.; Chou, H.S.; Fattaey, A.; Harlow, E. New functional activities for the p21 family of CDK inhibitors. Genes. Dev. 1997, 11, 847–862. [Google Scholar] [CrossRef] [PubMed]
  394. Cheng, M.; Olivier, P.; Diehl, J.A.; Fero, M.; Roussel, M.F.; Roberts, J.M.; Sherr, C.J. The p21Cip1 and p27Kip1 CDK ‘inhibitors’ are essential activators of cyclin D-dependent kinases in murine fibroblasts. EMBO J. 1999, 18, 1571–1583. [Google Scholar] [CrossRef]
  395. Kehn, K.; Deng, L.; de la Fuente, C.; Strouss, K.; Wu, K.; Maddukuri, A.; Baylor, S.; Rufner, R.; Pumfery, A.; Bottazzi, M.E.; et al. The role of cyclin D2 and p21/waf1 in human T-cell leukemia virus type 1 infected cells. Retrovirology 2004, 1, 6. [Google Scholar] [CrossRef]
  396. Watanabe, M.; Nakahata, S.; Hamasaki, M.; Saito, Y.; Kawano, Y.; Hidaka, T.; Yamashita, K.; Umeki, K.; Taki, T.; Taniwaki, M.; et al. Downregulation of CDKN1A in adult T-cell leukemia/lymphoma despite overexpression of CDKN1A in human T-lymphotropic virus 1-infected cell lines. J. Virol. 2010, 84, 6966–6977. [Google Scholar] [CrossRef]
  397. Haller, K.; Ruckes, T.; Schmitt, I.; Saul, D.; Derow, E.; Grassmann, R. Tax-dependent stimulation of G1 phase-specific cyclin-dependent kinases and increased expression of signal transduction genes characterize HTLV type 1-transformed T cells. AIDS Res. Hum. Retroviruses 2000, 16, 1683–1688. [Google Scholar] [CrossRef] [PubMed]
  398. Haller, K.; Wu, Y.; Derow, E.; Schmitt, I.; Jeang, K.T.; Grassmann, R. Physical interaction of human T-cell leukemia virus type 1 Tax with cyclin-dependent kinase 4 stimulates the phosphorylation of retinoblastoma protein. Mol. Cell Biol. 2002, 22, 3327–3338. [Google Scholar] [CrossRef]
  399. Kehn, K.; Fuente Cde, L.; Strouss, K.; Berro, R.; Jiang, H.; Brady, J.; Mahieux, R.; Pumfery, A.; Bottazzi, M.E.; Kashanchi, F. The HTLV-I Tax oncoprotein targets the retinoblastoma protein for proteasomal degradation. Oncogene 2005, 24, 525–540. [Google Scholar] [CrossRef]
  400. Suzuki, T.; Kitao, S.; Matsushime, H.; Yoshida, M. HTLV-1 Tax protein interacts with cyclin-dependent kinase inhibitor p16INK4A and counteracts its inhibitory activity towards CDK4. EMBO J. 1996, 15, 1607–1614. [Google Scholar] [CrossRef]
  401. Low, K.G.; Dorner, L.F.; Fernando, D.B.; Grossman, J.; Jeang, K.T.; Comb, M.J. Human T-cell leukemia virus type 1 Tax releases cell cycle arrest induced by p16INK4a. J. Virol. 1997, 71, 1956–1962. [Google Scholar] [CrossRef]
  402. Suzuki, T.; Narita, T.; Uchida-Toita, M.; Yoshida, M. Down-regulation of the INK4 family of cyclin-dependent kinase inhibitors by tax protein of HTLV-1 through two distinct mechanisms. Virology 1999, 259, 384–391. [Google Scholar] [CrossRef]
  403. Lemoine, F.J.; Marriott, S.J. Accelerated G1 phase progression induced by the human T cell leukemia virus type I (HTLV-I) Tax oncoprotein. J. Biol. Chem. 2001, 276, 31851–31857. [Google Scholar] [CrossRef]
  404. Liang, M.H.; Geisbert, T.; Yao, Y.; Hinrichs, S.H.; Giam, C.Z. Human T-lymphotropic virus type 1 oncoprotein tax promotes S-phase entry but blocks mitosis. J. Virol. 2002, 76, 4022–4033. [Google Scholar] [CrossRef]
  405. Liu, B.; Hong, S.; Tang, Z.; Yu, H.; Giam, C.Z. HTLV-I Tax directly binds the Cdc20-associated anaphase-promoting complex and activates it ahead of schedule. Proc. Natl. Acad. Sci. USA 2005, 102, 63–68. [Google Scholar] [CrossRef]
  406. Harhaj, E.W.; Good, L.; Xiao, G.; Sun, S.C. Gene expression profiles in HTLV-I-immortalized T cells: Deregulated expression of genes involved in apoptosis regulation. Oncogene 1999, 18, 1341–1349. [Google Scholar] [CrossRef] [PubMed]
  407. Tsukahara, T.; Kannagi, M.; Ohashi, T.; Kato, H.; Arai, M.; Nunez, G.; Iwanaga, Y.; Yamamoto, N.; Ohtani, K.; Nakamura, M.; et al. Induction of Bcl-xL expression by human T-cell leukemia virus type 1 Tax through NF-κB in apoptosis-resistant T-cell transfectants with Tax. J. Virol. 1999, 73, 7981–7987. [Google Scholar] [CrossRef]
  408. Nicot, C.; Mahieux, R.; Takemoto, S.; Franchini, G. Bcl-XL is up-regulated by HTLV-I and HTLV-II in vitro and in ex vivo ATLL samples. Blood 2000, 96, 275–281. [Google Scholar] [CrossRef] [PubMed]
  409. Zhang, M.; Mathews Griner, L.A.; Ju, W.; Duveau, D.Y.; Guha, R.; Petrus, M.N.; Wen, B.; Maeda, M.; Shinn, P.; Ferrer, M.; et al. Selective targeting of JAK/STAT signaling is potentiated by Bcl-xL blockade in IL-2-dependent adult T-cell leukemia. Proc. Natl. Acad. Sci. USA 2015, 112, 12480–12485. [Google Scholar] [CrossRef]
  410. Macaire, H.; Riquet, A.; Moncollin, V.; Biemont-Trescol, M.C.; Duc Dodon, M.; Hermine, O.; Debaud, A.L.; Mahieux, R.; Mesnard, J.M.; Pierre, M.; et al. Tax protein-induced expression of antiapoptotic Bfl-1 protein contributes to survival of human T-cell leukemia virus type 1 (HTLV-1)-infected T-cells. J. Biol. Chem. 2012, 287, 21357–21370. [Google Scholar] [CrossRef] [PubMed]
  411. Waldele, K.; Silbermann, K.; Schneider, G.; Ruckes, T.; Cullen, B.R.; Grassmann, R. Requirement of the human T-cell leukemia virus (HTLV-1) tax-stimulated HIAP-1 gene for the survival of transformed lymphocytes. Blood 2006, 107, 4491–4499. [Google Scholar] [CrossRef]
  412. Zane, L.; Sibon, D.; Legras, C.; Lachuer, J.; Wierinckx, A.; Mehlen, P.; Delfau-Larue, M.H.; Gessain, A.; Gout, O.; Pinatel, C.; et al. Clonal expansion of HTLV-1 positive CD8+ cells relies on cIAP-2 but not on c-FLIP expression. Virology 2010, 407, 341–351. [Google Scholar] [CrossRef] [PubMed]
  413. Mori, N.; Yamada, Y.; Hata, T.; Ikeda, S.; Yamasaki, Y.; Tomonaga, M.; Yamamoto, N. Expression of survivin in HTLV-I-infected T-cell lines and primary ATL cells. Biochem. Biophys. Res. Commun. 2001, 282, 1110–1113. [Google Scholar] [CrossRef]
  414. Kamihira, S.; Yamada, Y.; Hirakata, Y.; Tomonaga, M.; Sugahara, K.; Hayashi, T.; Dateki, N.; Harasawa, H.; Nakayama, K. Aberrant expression of caspase cascade regulatory genes in adult T-cell leukaemia: Survivin is an important determinant for prognosis. Br. J. Haematol. 2001, 114, 63–69. [Google Scholar] [CrossRef]
  415. Kawakami, H.; Tomita, M.; Matsuda, T.; Ohta, T.; Tanaka, Y.; Fujii, M.; Hatano, M.; Tokuhisa, T.; Mori, N. Transcriptional activation of survivin through the NF-κB pathway by human T-cell leukemia virus type I tax. Int. J. Cancer 2005, 115, 967–974. [Google Scholar] [CrossRef] [PubMed]
  416. Krueger, A.; Fas, S.C.; Giaisi, M.; Bleumink, M.; Merling, A.; Stumpf, C.; Baumann, S.; Holtkotte, D.; Bosch, V.; Krammer, P.H.; et al. HTLV-1 Tax protects against CD95-mediated apoptosis by induction of the cellular FLICE-inhibitory protein (c-FLIP). Blood 2006, 107, 3933–3939. [Google Scholar] [CrossRef]
  417. Okamoto, K.; Fujisawa, J.; Reth, M.; Yonehara, S. Human T-cell leukemia virus type-I oncoprotein Tax inhibits Fas-mediated apoptosis by inducing cellular FLIP through activation of NF-κB. Genes. Cells 2006, 11, 177–191. [Google Scholar] [CrossRef]
  418. Wang, W.; Zhou, J.; Shi, J.; Zhang, Y.; Liu, S.; Liu, Y.; Zheng, D. Human T-cell leukemia virus type 1 Tax-deregulated autophagy pathway and c-FLIP expression contribute to resistance against death receptor-mediated apoptosis. J. Virol. 2014, 88, 2786–2798. [Google Scholar] [CrossRef]
  419. Daniel, S.; Arvelo, M.B.; Patel, V.I.; Longo, C.R.; Shrikhande, G.; Shukri, T.; Mahiou, J.; Sun, D.W.; Mottley, C.; Grey, S.T.; et al. A20 protects endothelial cells from TNF-, Fas-, and NK-mediated cell death by inhibiting caspase 8 activation. Blood 2004, 104, 2376–2384. [Google Scholar] [CrossRef]
  420. Laherty, C.D.; Perkins, N.D.; Dixit, V.M. Human T cell leukemia virus type I Tax and phorbol 12-myristate 13-acetate induce expression of the A20 zinc finger protein by distinct mechanisms involving nuclear factor κB. J. Biol. Chem. 1993, 268, 5032–5039. [Google Scholar] [CrossRef] [PubMed]
  421. Saitoh, Y.; Hamano, A.; Mochida, K.; Kakeya, A.; Uno, M.; Tsuruyama, E.; Ichikawa, H.; Tokunaga, F.; Utsunomiya, A.; Watanabe, T.; et al. A20 targets caspase-8 and FADD to protect HTLV-I-infected cells. Leukemia 2016, 30, 716–727. [Google Scholar] [CrossRef]
  422. Choi, Y.B.; Harhaj, E.W. HTLV-1 tax stabilizes MCL-1 via TRAF6-dependent K63-linked polyubiquitination to promote cell survival and transformation. PLoS Pathog. 2014, 10, e1004458. [Google Scholar] [CrossRef]
  423. Chuang, S.E.; Doong, S.L.; Lin, M.T.; Cheng, A.L. Tax of the human T-lymphotropic virus type I transactivates promoter of the MDR-1 gene. Biochem. Biophys. Res. Commun. 1997, 238, 482–486. [Google Scholar] [CrossRef] [PubMed]
  424. Lau, A.; Gant, T.W.; Cann, A.J. The mechanism of trans-activation of the MDR1 gene by human T-cell leukemia virus. Biochem. Biophys. Res. Commun. 1998, 249, 397–404. [Google Scholar] [CrossRef]
  425. Lau, A.; Nightingale, S.; Taylor, G.P.; Gant, T.W.; Cann, A.J. Enhanced MDR1 gene expression in human T-cell leukemia virus-I-infected patients offers new prospects for therapy. Blood 1998, 91, 2467–2474. [Google Scholar] [CrossRef] [PubMed]
  426. Sakaki, Y.; Terashi, K.; Yamaguchi, A.; Kawamata, N.; Tokito, Y.; Mori, H.; Umehara, M.; Yoshiyama, T.; Ohtsubo, H.; Arimura, K.; et al. Human T-cell lymphotropic virus type I Tax activates lung resistance-related protein expression in leukemic clones established from an adult T-cell leukemia patient. Exp. Hematol. 2002, 30, 340–345. [Google Scholar] [CrossRef]
  427. Chlichlia, K.; Busslinger, M.; Peter, M.E.; Walczak, H.; Krammer, P.H.; Schirrmacher, V.; Khazaie, K. ICE-proteases mediate HTLV-I Tax-induced apoptotic T-cell death. Oncogene 1997, 14, 2265–2272. [Google Scholar] [CrossRef]
  428. Rivera-Walsh, I.; Waterfield, M.; Xiao, G.; Fong, A.; Sun, S.C. NF-κB signaling pathway governs TRAIL gene expression and human T-cell leukemia virus-I Tax-induced T-cell death. J. Biol. Chem. 2001, 276, 40385–40388. [Google Scholar] [CrossRef]
  429. Gabet, A.S.; Mortreux, F.; Charneau, P.; Riou, P.; Duc-Dodon, M.; Wu, Y.; Jeang, K.T.; Wattel, E. Inactivation of hTERT transcription by Tax. Oncogene 2003, 22, 3734–3741. [Google Scholar] [CrossRef] [PubMed]
  430. Sinha-Datta, U.; Horikawa, I.; Michishita, E.; Datta, A.; Sigler-Nicot, J.C.; Brown, M.; Kazanji, M.; Barrett, J.C.; Nicot, C. Transcriptional activation of hTERT through the NF-κB pathway in HTLV-I-transformed cells. Blood 2004, 104, 2523–2531. [Google Scholar] [CrossRef]
  431. Matsumura-Arioka, Y.; Ohtani, K.; Hara, T.; Iwanaga, R.; Nakamura, M. Identification of two distinct elements mediating activation of telomerase (hTERT) gene expression in association with cell growth in human T cells. Int. Immunol. 2005, 17, 207–215. [Google Scholar] [CrossRef]
  432. Hara, T.; Matsumura-Arioka, Y.; Ohtani, K.; Nakamura, M. Role of human T-cell leukemia virus type I Tax in expression of the human telomerase reverse transcriptase (hTERT) gene in human T-cells. Cancer Sci. 2008, 99, 1155–1163. [Google Scholar] [CrossRef] [PubMed]
  433. Zane, L.; Sibon, D.; Capraro, V.; Galia, P.; Karam, M.; Delfau-Larue, M.H.; Gilson, E.; Gessain, A.; Gout, O.; Hermine, O.; et al. HTLV-1 positive and negative T cells cloned from infected individuals display telomerase and telomere genes deregulation that predominate in activated but untransformed CD4+ T cells. Int. J. Cancer 2012, 131, 821–833. [Google Scholar] [CrossRef]
  434. Bellon, M.; Yuan, Y.; Nicot, C. Transcription independent stimulation of telomerase enzymatic activity by HTLV-I Tax through stimulation of IKK. J. Cancer Sci. 2021, 8. [Google Scholar] [CrossRef]
  435. Miyake, H.; Suzuki, T.; Hirai, H.; Yoshida, M. Trans-activator Tax of human T-cell leukemia virus type 1 enhances mutation frequency of the cellular genome. Virology 1999, 253, 155–161. [Google Scholar] [CrossRef]
  436. Philpott, S.M.; Buehring, G.C. Defective DNA repair in cells with human T-cell leukemia/bovine leukemia viruses: Role of tax gene. J. Natl. Cancer Inst. 1999, 91, 933–942. [Google Scholar] [CrossRef]
  437. Lemoine, F.J.; Marriott, S.J. Genomic instability driven by the human T-cell leukemia virus type I (HTLV-I) oncoprotein, Tax. Oncogene 2002, 21, 7230–7234. [Google Scholar] [CrossRef] [PubMed]
  438. Baydoun, H.H.; Bai, X.T.; Shelton, S.; Nicot, C. HTLV-I tax increases genetic instability by inducing DNA double strand breaks during DNA replication and switching repair to NHEJ. PLoS ONE 2012, 7, e42226. [Google Scholar] [CrossRef] [PubMed]
  439. Ressler, S.; Morris, G.F.; Marriott, S.J. Human T-cell leukemia virus type 1 Tax transactivates the human proliferating cell nuclear antigen promoter. J. Virol. 1997, 71, 1181–1190. [Google Scholar] [CrossRef]
  440. Kao, S.Y.; Marriott, S.J. Disruption of nucleotide excision repair by the human T-cell leukemia virus type 1 Tax protein. J. Virol. 1999, 73, 4299–4304. [Google Scholar] [CrossRef]
  441. Lemoine, F.J.; Kao, S.Y.; Marriott, S.J. Suppression of DNA repair by HTLV type 1 Tax correlates with Tax trans-activation of proliferating cell nuclear antigen gene expression. AIDS Res. Hum. Retroviruses 2000, 16, 1623–1627. [Google Scholar] [CrossRef]
  442. Edwards, D.C.; Marriott, S.J. Human T-cell leukemia virus type 1 Tax relieves repression of proliferating cell nuclear antigen gene expression. J. Virol. 2008, 82, 11714–11722. [Google Scholar] [CrossRef] [PubMed]
  443. Mori, N.; Nunokawa, Y.; Yamada, Y.; Ikeda, S.; Tomonaga, M.; Yamamoto, N. Expression of human inducible nitric oxide synthase gene in T-cell lines infected with human T-cell leukemia virus type-I and primary adult T-cell leukemia cells. Blood 1999, 94, 2862–2870. [Google Scholar] [CrossRef] [PubMed]
  444. Sonoki, T.; Matsuzaki, H.; Nagasaki, A.; Hata, H.; Yoshida, M.; Matsuoka, M.; Kuribayashi, N.; Kimura, T.; Harada, N.; Takatsuki, K.; et al. Detection of inducible nitric oxide synthase (iNOS) mRNA by RT-PCR in ATL patients and HTLV-I infected cell lines: Clinical features and apoptosis by NOS inhibitor. Leukemia 1999, 13, 713–718. [Google Scholar] [CrossRef] [PubMed]
  445. Goto, H.; Nakamura, T.; Shirabe, S.; Ueki, Y.; Nishiura, Y.; Furuya, T.; Tsujino, A.; Nakane, S.; Eguchi, K.; Nagataki, S. Up-regulation of iNOS mRNA expression and increased production of NO in human monoblast cell line, U937 transfected by HTLV-I tax gene. Immunobiology 1997, 197, 513–521. [Google Scholar] [CrossRef]
  446. Chaib-Mezrag, H.; Lemacon, D.; Fontaine, H.; Bellon, M.; Bai, X.T.; Drac, M.; Coquelle, A.; Nicot, C. Tax impairs DNA replication forks and increases DNA breaks in specific oncogenic genome regions. Mol. Cancer 2014, 13, 205. [Google Scholar] [CrossRef]
  447. Baydoun, H.H.; Cherian, M.A.; Green, P.; Ratner, L. Inducible nitric oxide synthase mediates DNA double strand breaks in human T-cell leukemia virus type 1-induced leukemia/lymphoma. Retrovirology 2015, 12, 71. [Google Scholar] [CrossRef]
  448. Kinjo, T.; Ham-Terhune, J.; Peloponese, J.M., Jr.; Jeang, K.T. Induction of reactive oxygen species by human T-cell leukemia virus type 1 tax correlates with DNA damage and expression of cellular senescence marker. J. Virol. 2010, 84, 5431–5437. [Google Scholar] [CrossRef]
  449. Takahashi, M.; Higuchi, M.; Makokha, G.N.; Matsuki, H.; Yoshita, M.; Tanaka, Y.; Fujii, M. HTLV-1 Tax oncoprotein stimulates ROS production and apoptosis in T cells by interacting with USP10. Blood 2013, 122, 715–725. [Google Scholar] [CrossRef]
  450. Marriott, S.J.; Semmes, O.J. Impact of HTLV-I Tax on cell cycle progression and the cellular DNA damage repair response. Oncogene 2005, 24, 5986–5995. [Google Scholar] [CrossRef]
  451. He, Y.; Pasupala, N.; Zhi, H.; Dorjbal, B.; Hussain, I.; Shih, H.M.; Bhattacharyya, S.; Biswas, R.; Miljkovic, M.; Semmes, O.J.; et al. NF-κB-induced R-loop accumulation and DNA damage select for nucleotide excision repair deficiencies in adult T cell leukemia. Proc. Natl. Acad. Sci. USA 2021, 118, e2005568118. [Google Scholar] [CrossRef] [PubMed]
  452. Giam, C.Z.; Pasupala, N. NF-κB-induced R-loops and genomic instability in HTLV-1-infected and adult T-cell leukemia cells. Viruses 2022, 14, 877. [Google Scholar] [CrossRef]
  453. Dayaram, T.; Lemoine, F.J.; Donehower, L.A.; Marriott, S.J. Activation of WIP1 phosphatase by HTLV-1 Tax mitigates the cellular response to DNA damage. PLoS ONE 2013, 8, e55989. [Google Scholar] [CrossRef]
  454. Zane, L.; Yasunaga, J.; Mitagami, Y.; Yedavalli, V.; Tang, S.W.; Chen, C.Y.; Ratner, L.; Lu, X.; Jeang, K.T. Wip1 and p53 contribute to HTLV-1 Tax-induced tumorigenesis. Retrovirology 2012, 9, 114. [Google Scholar] [CrossRef] [PubMed]
  455. Haoudi, A.; Daniels, R.C.; Wong, E.; Kupfer, G.; Semmes, O.J. Human T-cell leukemia virus-I tax oncoprotein functionally targets a subnuclear complex involved in cellular DNA damage-response. J. Biol. Chem. 2003, 278, 37736–37744. [Google Scholar] [CrossRef] [PubMed]
  456. Chandhasin, C.; Ducu, R.I.; Berkovich, E.; Kastan, M.B.; Marriott, S.J. Human T-cell leukemia virus type 1 tax attenuates the ATM-mediated cellular DNA damage response. J. Virol. 2008, 82, 6952–6961. [Google Scholar] [CrossRef]
  457. Park, H.U.; Jeong, J.H.; Chung, J.H.; Brady, J.N. Human T-cell leukemia virus type 1 Tax interacts with Chk1 and attenuates DNA-damage induced G2 arrest mediated by Chk1. Oncogene 2004, 23, 4966–4974. [Google Scholar] [CrossRef]
  458. Park, H.U.; Jeong, S.J.; Jeong, J.H.; Chung, J.H.; Brady, J.N. Human T-cell leukemia virus type 1 Tax attenuates γ-irradiation-induced apoptosis through physical interaction with Chk2. Oncogene 2006, 25, 438–447. [Google Scholar] [CrossRef]
  459. Boxus, M.; Twizere, J.C.; Legros, S.; Kettmann, R.; Willems, L. Interaction of HTLV-1 Tax with minichromosome maintenance proteins accelerates the replication timing program. Blood 2012, 119, 151–160. [Google Scholar] [CrossRef]
  460. Suzuki, T.; Uchida-Toita, M.; Andoh, T.; Yoshida, M. HTLV-1 tax oncoprotein binds to DNA topoisomerase I and inhibits its catalytic activity. Virology 2000, 270, 291–298. [Google Scholar] [CrossRef]
  461. Yoshida, M.; Suzuki, T. HTLV type 1 Tax oncoprotein binds to DNA topoisomerase I and inhibits its catalytic activity. AIDS Res. Hum. Retroviruses 2000, 16, 1639–1645. [Google Scholar] [CrossRef] [PubMed]
  462. Zhi, H.; Guo, X.; Ho, Y.K.; Pasupala, N.; Engstrom, H.A.A.; Semmes, O.J.; Giam, C.Z. RNF8 dysregulation and down-regulation during HTLV-1 infection promote genomic instability in adult T-cell leukemia. PLoS Pathog. 2020, 16, e1008618. [Google Scholar] [CrossRef] [PubMed]
  463. Jonas, S.; Izaurralde, E. Towards a molecular understanding of microRNA-mediated gene silencing. Nat. Rev. Genet. 2015, 16, 421–433. [Google Scholar] [CrossRef]
  464. Kobayashi, H.; Tomari, Y. RISC assembly: Coordination between small RNAs and Argonaute proteins. Biochim. Biophys. Acta 2016, 1859, 71–81. [Google Scholar] [CrossRef]
  465. Michlewski, G.; Caceres, J.F. Post-transcriptional control of miRNA biogenesis. RNA 2019, 25, 1–16. [Google Scholar] [CrossRef]
  466. Treiber, T.; Treiber, N.; Meister, G. Regulation of microRNA biogenesis and its crosstalk with other cellular pathways. Nat Rev Mol. Cell Biol. 2019, 20, 5–20. [Google Scholar] [CrossRef] [PubMed]
  467. Ruggero, K.; Corradin, A.; Zanovello, P.; Amadori, A.; Bronte, V.; Ciminale, V.; D’Agostino, D.M. Role of microRNAs in HTLV-1 infection and transformation. Mol. Asp. Med. 2010, 31, 367–382. [Google Scholar] [CrossRef]
  468. Hermeking, H. The miR-34 family in cancer and apoptosis. Cell Death Differ. 2010, 17, 193–199. [Google Scholar] [CrossRef]
  469. Misso, G.; Di Martino, M.T.; De Rosa, G.; Farooqi, A.A.; Lombardi, A.; Campani, V.; Zarone, M.R.; Gulla, A.; Tagliaferri, P.; Tassone, P.; et al. Mir-34: A new weapon against cancer? Mol. Ther. Nucleic Acids 2014, 3, e194. [Google Scholar] [CrossRef]
  470. Schmid, G.; Notaro, S.; Reimer, D.; Abdel-Azim, S.; Duggan-Peer, M.; Holly, J.; Fiegl, H.; Rossler, J.; Wiedemair, A.; Concin, N.; et al. Expression and promotor hypermethylation of miR-34a in the various histological subtypes of ovarian cancer. BMC Cancer 2016, 16, 102. [Google Scholar] [CrossRef]
  471. Sharma, V.K.; Raimondi, V.; Ruggero, K.; Pise-Masison, C.A.; Cavallari, I.; Silic-Benussi, M.; Ciminale, V.; D’Agostino, D.M. Expression of miR-34a in T-cells infected by human T-lymphotropic virus 1. Front. Microbiol. 2018, 9, 832. [Google Scholar] [CrossRef]
  472. Pichler, K.; Schneider, G.; Grassmann, R. MicroRNA miR-146a and further oncogenesis-related cellular microRNAs are dysregulated in HTLV-1-transformed T lymphocytes. Retrovirology 2008, 5, 100. [Google Scholar] [CrossRef]
  473. Tomita, M.; Tanaka, Y.; Mori, N. MicroRNA miR-146a is induced by HTLV-1 tax and increases the growth of HTLV-1-infected T-cells. Int. J. Cancer 2012, 130, 2300–2309. [Google Scholar] [CrossRef] [PubMed]
  474. Liu, F.; Kong, X.; Lv, L.; Gao, J. MiR-155 targets TP53INP1 to regulate liver cancer stem cell acquisition and self-renewal. FEBS Lett. 2015, 589, 500–506. [Google Scholar] [CrossRef] [PubMed]
  475. Yeung, M.L.; Yasunaga, J.; Bennasser, Y.; Dusetti, N.; Harris, D.; Ahmad, N.; Matsuoka, M.; Jeang, K.T. Roles for microRNAs, miR-93 and miR-130b, and tumor protein 53-induced nuclear protein 1 tumor suppressor in cell growth dysregulation by human T-cell lymphotrophic virus 1. Cancer Res. 2008, 68, 8976–8985. [Google Scholar] [CrossRef]
  476. Van Duyne, R.; Guendel, I.; Klase, Z.; Narayanan, A.; Coley, W.; Jaworski, E.; Roman, J.; Popratiloff, A.; Mahieux, R.; Kehn-Hall, K.; et al. Localization and sub-cellular shuttling of HTLV-1 tax with the miRNA machinery. PLoS ONE 2012, 7, e40662. [Google Scholar] [CrossRef]
  477. Yamagishi, M.; Fujikawa, D.; Watanabe, T.; Uchimaru, K. HTLV-1-mediated epigenetic pathway to adult T-cell leukemia-lymphoma. Front. Microbiol. 2018, 9, 1686. [Google Scholar] [CrossRef] [PubMed]
  478. Ratner, L. Epigenetic regulation of human T-cell leukemia virus gene expression. Microorganisms 2021, 10, 84. [Google Scholar] [CrossRef]
  479. Watanabe, T.; Yamashita, S.; Ureshino, H.; Kamachi, K.; Kurahashi, Y.; Fukuda-Kurahashi, Y.; Yoshida, N.; Hattori, N.; Nakamura, H.; Sato, A.; et al. Targeting aberrant DNA hypermethylation as a driver of ATL leukemogenesis by using the new oral demethylating agent OR-2100. Blood 2020, 136, 871–884. [Google Scholar] [CrossRef]
  480. Nosaka, K.; Maeda, M.; Tamiya, S.; Sakai, T.; Mitsuya, H.; Matsuoka, M. Increasing methylation of the CDKN2A gene is associated with the progression of adult T-cell leukemia. Cancer Res. 2000, 60, 1043–1048. [Google Scholar]
  481. Taniguchi, A.; Nemoto, Y.; Yokoyama, A.; Kotani, N.; Imai, S.; Shuin, T.; Daibata, M. Promoter methylation of the bone morphogenetic protein-6 gene in association with adult T-cell leukemia. Int. J. Cancer 2008, 123, 1824–1831. [Google Scholar] [CrossRef]
  482. Yasunaga, J.; Taniguchi, Y.; Nosaka, K.; Yoshida, M.; Satou, Y.; Sakai, T.; Mitsuya, H.; Matsuoka, M. Identification of aberrantly methylated genes in association with adult T-cell leukemia. Cancer Res. 2004, 64, 6002–6009. [Google Scholar] [CrossRef] [PubMed]
  483. Tang, H.M.; Gao, W.W.; Chan, C.P.; Cheng, Y.; Deng, J.J.; Yuen, K.S.; Iha, H.; Jin, D.Y. SIRT1 Suppresses Human T-Cell Leukemia Virus Type 1 Transcription. J. Virol. 2015, 89, 8623–8631. [Google Scholar] [CrossRef] [PubMed]
  484. Kamoi, K.; Yamamoto, K.; Misawa, A.; Miyake, A.; Ishida, T.; Tanaka, Y.; Mochizuki, M.; Watanabe, T. SUV39H1 interacts with HTLV-1 Tax and abrogates Tax transactivation of HTLV-1 LTR. Retrovirology 2006, 3, 5. [Google Scholar] [CrossRef]
  485. Yamamoto, K.; Ishida, T.; Nakano, K.; Yamagishi, M.; Yamochi, T.; Tanaka, Y.; Furukawa, Y.; Nakamura, Y.; Watanabe, T. SMYD3 interacts with HTLV-1 Tax and regulates subcellular localization of Tax. Cancer Sci. 2011, 102, 260–266. [Google Scholar] [CrossRef]
  486. Satou, Y.; Miyazato, P.; Ishihara, K.; Yaguchi, H.; Melamed, A.; Miura, M.; Fukuda, A.; Nosaka, K.; Watanabe, T.; Rowan, A.G.; et al. The retrovirus HTLV-1 inserts an ectopic CTCF-binding site into the human genome. Proc. Natl. Acad. Sci. USA 2016, 113, 3054–3059. [Google Scholar] [CrossRef] [PubMed]
  487. Cheng, X.; Joseph, A.; Castro, V.; Chen-Liaw, A.; Skidmore, Z.; Ueno, T.; Fujisawa, J.I.; Rauch, D.A.; Challen, G.A.; Martinez, M.P.; et al. Epigenomic regulation of human T-cell leukemia virus by chromatin-insulator CTCF. PLoS Pathog. 2021, 17, e1009577. [Google Scholar] [CrossRef]
  488. Melamed, A.; Yaguchi, H.; Miura, M.; Witkover, A.; Fitzgerald, T.W.; Birney, E.; Bangham, C.R. The human leukemia virus HTLV-1 alters the structure and transcription of host chromatin in cis. Elife 2018, 7, e36245. [Google Scholar] [CrossRef]
  489. Yaguchi, H.; Melamed, A.; Ramanayake, S.; Kiik, H.; Witkover, A.; Bangham, C.R.M. The impact of HTLV-1 expression on the 3D structure and expression of host chromatin. PLoS Pathog. 2024, 20, e1011716. [Google Scholar] [CrossRef]
  490. Ego, T.; Tanaka, Y.; Shimotohno, K. Interaction of HTLV-1 Tax and methyl-CpG-binding domain 2 positively regulates the gene expression from the hypermethylated LTR. Oncogene 2005, 24, 1914–1923. [Google Scholar] [CrossRef]
  491. Rank, G.; Cerruti, L.; Simpson, R.J.; Moritz, R.L.; Jane, S.M.; Zhao, Q. Identification of a PRMT5-dependent repressor complex linked to silencing of human fetal globin gene expression. Blood 2010, 116, 1585–1592. [Google Scholar] [CrossRef] [PubMed]
  492. Panfil, A.R.; Al-Saleem, J.; Howard, C.M.; Mates, J.M.; Kwiek, J.J.; Baiocchi, R.A.; Green, P.L. PRMT5 is upregulated in HTLV-1-mediated T-cell transformation and selective inhibition alters viral gene expression and infected cell survival. Viruses 2015, 8, 7. [Google Scholar] [CrossRef] [PubMed]
  493. Ernzen, K.; Melvin, C.; Yu, L.; Phelps, C.; Niewiesk, S.; Green, P.L.; Panfil, A.R. The PRMT5 inhibitor EPZ015666 is effective against HTLV-1-transformed T-cell lines in vitro and in vivo. Front. Microbiol. 2023, 14, 1101544. [Google Scholar] [CrossRef]
  494. Fujikawa, D.; Nakagawa, S.; Hori, M.; Kurokawa, N.; Soejima, A.; Nakano, K.; Yamochi, T.; Nakashima, M.; Kobayashi, S.; Tanaka, Y.; et al. Polycomb-dependent epigenetic landscape in adult T-cell leukemia. Blood 2016, 127, 1790–1802. [Google Scholar] [CrossRef]
  495. Zhang, X.; Yi, K.; Wang, B.; Chu, K.; Liu, J.; Zhang, J.; Fang, J.; Zhao, T. EZH2 activates HTLV-1 bZIP factor-mediated TGF-β signaling in adult T-cell leukemia. J. Med. Virol. 2024, 96, e70025. [Google Scholar] [CrossRef]
  496. Ichikawa, T.; Nakahata, S.; Fujii, M.; Iha, H.; Shimoda, K.; Morishita, K. The regulation of NDRG2 expression during ATLL development after HTLV-1 infection. Biochim. Biophys. Acta Mol. Basis Dis. 2019, 1865, 2633–2646. [Google Scholar] [CrossRef]
  497. Jeong, S.J.; Lu, H.; Cho, W.K.; Park, H.U.; Pise-Masison, C.; Brady, J.N. Coactivator-associated arginine methyltransferase 1 enhances transcriptional activity of the human T-cell lymphotropic virus type 1 long terminal repeat through direct interaction with Tax. J. Virol. 2006, 80, 10036–10044. [Google Scholar] [CrossRef]
  498. Hayati, R.F.; Nakajima, R.; Zhou, Y.; Shirasawa, M.; Zhao, L.; Fikriyanti, M.; Iwanaga, R.; Bradford, A.P.; Kurayoshi, K.; Araki, K.; et al. Trans-activation of the coactivator-associated arginine methyltransferase 1 (Carm1) gene by the oncogene product Tax of human T-cell leukemia virus type 1. Genes 2024, 15, 698. [Google Scholar] [CrossRef]
  499. Furukawa, K.; Akagi, T.; Nagata, Y.; Yamada, Y.; Shimotohno, K.; Cheung, N.K.; Shiku, H.; Furukawa, K. GD2 ganglioside on human T-lymphotropic virus type I-infected T cells: Possible activation of β-1,4-N-acetylgalactosaminyltransferase gene by p40tax. Proc. Natl. Acad. Sci. USA 1993, 90, 1972–1976. [Google Scholar] [CrossRef] [PubMed]
  500. Lu, J.; Holmgren, A. The thioredoxin antioxidant system. Free Radic. Biol. Med. 2014, 66, 75–87. [Google Scholar] [CrossRef]
  501. Yodoi, J.; Tursz, T. ADF, a growth-promoting factor derived from adult T cell leukemia and homologous to thioredoxin: Involvement in lymphocyte immortalization by HTLV-I and EBV. Adv. Cancer Res. 1991, 57, 381–411. [Google Scholar] [CrossRef] [PubMed]
  502. Rosen, A.; Lundman, P.; Carlsson, M.; Bhavani, K.; Srinivasa, B.R.; Kjellstrom, G.; Nilsson, K.; Holmgren, A. A CD4+ T cell line-secreted factor, growth promoting for normal and leukemic B cells, identified as thioredoxin. Int. Immunol. 1995, 7, 625–633. [Google Scholar] [CrossRef] [PubMed]
  503. Wakasugi, N.; Tagaya, Y.; Wakasugi, H.; Mitsui, A.; Maeda, M.; Yodoi, J.; Tursz, T. Adult T-cell leukemia-derived factor/thioredoxin, produced by both human T-lymphotropic virus type I- and Epstein-Barr virus-transformed lymphocytes, acts as an autocrine growth factor and synergizes with interleukin 1 and interleukin 2. Proc. Natl. Acad. Sci. USA 1990, 87, 8282–8286. [Google Scholar] [CrossRef] [PubMed]
  504. Los, M.; Khazaie, K.; Schulze-Osthoff, K.; Baeuerle, P.A.; Schirrmacher, V.; Chlichlia, K. Human T cell leukemia virus-I (HTLV-I) Tax-mediated apoptosis in activated T cells requires an enhanced intracellular prooxidant state. J. Immunol. 1998, 161, 3050–3055. [Google Scholar] [CrossRef]
  505. Masutani, H.; Hirota, K.; Sasada, T.; Ueda-Taniguchi, Y.; Taniguchi, Y.; Sono, H.; Yodoi, J. Transactivation of an inducible anti-oxidative stress protein, human thioredoxin by HTLV-I Tax. Immunol. Lett. 1996, 54, 67–71. [Google Scholar] [CrossRef]
  506. Ishikawa, C.; Mori, N. Pivotal role of dihydroorotate dehydrogenase as a therapeutic target in adult T-cell leukemia. Eur. J. Haematol. 2024, 113, 99–109. [Google Scholar] [CrossRef]
  507. Fukudome, K.; Furuse, M.; Fukuhara, N.; Orita, S.; Imai, T.; Takagi, S.; Nagira, M.; Hinuma, Y.; Yoshie, O. Strong induction of ICAM-1 in human T cells transformed by human T-cell-leukemia virus type 1 and depression of ICAM-1 or LFA-1 in adult T-cell-leukemia-derived cell lines. Int. J. Cancer 1992, 52, 418–427. [Google Scholar] [CrossRef]
  508. Tanaka, Y.; Hayashi, M.; Takagi, S.; Yoshie, O. Differential transactivation of the intercellular adhesion molecule 1 gene promoter by Tax1 and Tax2 of human T-cell leukemia viruses. J. Virol. 1996, 70, 8508–8517. [Google Scholar] [CrossRef]
  509. Owen, S.M.; Rudolph, D.L.; Dezzutti, C.S.; Shibata, N.; Naik, S.; Caughman, S.W.; Lal, R.B. Transcriptional activation of the intercellular adhesion molecule 1 (CD54) gene by human T lymphotropic virus types I and II Tax is mediated through a palindromic response element. AIDS Res. Hum. Retroviruses 1997, 13, 1429–1437. [Google Scholar] [CrossRef]
  510. Nejmeddine, M.; Negi, V.S.; Mukherjee, S.; Tanaka, Y.; Orth, K.; Taylor, G.P.; Bangham, C.R. HTLV-1-Tax and ICAM-1 act on T-cell signal pathways to polarize the microtubule-organizing center at the virological synapse. Blood 2009, 114, 1016–1025. [Google Scholar] [CrossRef]
  511. Hiyoshi, M.; Takahashi, N.; Eltalkhawy, Y.M.; Noyori, O.; Lotfi, S.; Panaampon, J.; Okada, S.; Tanaka, Y.; Ueno, T.; Fujisawa, J.I.; et al. M-Sec induced by HTLV-1 mediates an efficient viral transmission. PLoS Pathog. 2021, 17, e1010126. [Google Scholar] [CrossRef]
  512. Percher, F.; Curis, C.; Peres, E.; Artesi, M.; Rosewick, N.; Jeannin, P.; Gessain, A.; Gout, O.; Mahieux, R.; Ceccaldi, P.E.; et al. HTLV-1-induced leukotriene B4 secretion by T cells promotes T cell recruitment and virus propagation. Nat. Commun. 2017, 8, 15890. [Google Scholar] [CrossRef]
  513. Moriuchi, M.; Moriuchi, H. A milk protein lactoferrin enhances human T cell leukemia virus type I and suppresses HIV-1 infection. J. Immunol. 2001, 166, 4231–4236. [Google Scholar] [CrossRef]
  514. Moriuchi, M.; Moriuchi, H. Induction of lactoferrin gene expression in myeloid or mammary gland cells by human T-cell leukemia virus type 1 (HTLV-1) tax: Implications for milk-borne transmission of HTLV-1. J. Virol. 2006, 80, 7118–7126. [Google Scholar] [CrossRef] [PubMed]
  515. Gauthier, S.; Pelletier, I.; Ouellet, M.; Vargas, A.; Tremblay, M.J.; Sato, S.; Barbeau, B. Induction of galectin-1 expression by HTLV-I Tax and its impact on HTLV-I infectivity. Retrovirology 2008, 5, 105. [Google Scholar] [CrossRef] [PubMed]
  516. Hsu, D.K.; Hammes, S.R.; Kuwabara, I.; Greene, W.C.; Liu, F.T. Human T lymphotropic virus-I infection of human T lymphocytes induces expression of the β-galactoside-binding lectin, galectin-3. Am. J. Pathol. 1996, 148, 1661–1670. [Google Scholar] [PubMed]
  517. Albrecht, B.; Collins, N.D.; Burniston, M.T.; Nisbet, J.W.; Ratner, L.; Green, P.L.; Lairmore, M.D. Human T-lymphotropic virus type 1 open reading frame I p12I is required for efficient viral infectivity in primary lymphocytes. J. Virol. 2000, 74, 9828–9835. [Google Scholar] [CrossRef]
  518. Curis, C.; Percher, F.; Jeannin, P.; Montange, T.; Chevalier, S.A.; Seilhean, D.; Cartier, L.; Couraud, P.O.; Gout, O.; Gessain, A.; et al. Human T-lymphotropic virus type 1-induced overexpression of activated leukocyte cell adhesion molecule (ALCAM) facilitates trafficking of infected lymphocytes through the blood-brain barrier. J. Virol. 2016, 90, 7303–7312. [Google Scholar] [CrossRef]
  519. Lilienbaum, A.; Duc Dodon, M.; Alexandre, C.; Gazzolo, L.; Paulin, D. Effect of human T-cell leukemia virus type I tax protein on activation of the human vimentin gene. J. Virol. 1990, 64, 256–263. [Google Scholar] [CrossRef]
  520. Lilienbaum, A.; Paulin, D. Activation of the human vimentin gene by the Tax human T-cell leukemia virus. I. Mechanisms of regulation by the NF-κB transcription factor. J. Biol. Chem. 1993, 268, 2180–2188. [Google Scholar] [CrossRef]
  521. Hiraiwa, N.; Hiraiwa, M.; Kannagi, R. Human T-cell leukemia virus-1 encoded Tax protein transactivates α 1-->3 fucosyltransferase Fuc-T VII, which synthesizes sialyl Lewis X, a selectin ligand expressed on adult T-cell leukemia cells. Biochem. Biophys. Res. Commun. 1997, 231, 183–186. [Google Scholar] [CrossRef]
  522. Hiraiwa, N.; Yabuta, T.; Yoritomi, K.; Hiraiwa, M.; Tanaka, Y.; Suzuki, T.; Yoshida, M.; Kannagi, R. Transactivation of the fucosyltransferase VII gene by human T-cell leukemia virus type 1 Tax through a variant cAMP-responsive element. Blood 2003, 101, 3615–3621. [Google Scholar] [CrossRef]
  523. Nakachi, S.; Nakazato, T.; Ishikawa, C.; Kimura, R.; Mann, D.A.; Senba, M.; Masuzaki, H.; Mori, N. Human T-cell leukemia virus type 1 tax transactivates the matrix metalloproteinase 7 gene via JunD/AP-1 signaling. Biochim. Biophys. Acta 2011, 1813, 731–741. [Google Scholar] [CrossRef]
  524. Mori, N.; Sato, H.; Hayashibara, T.; Senba, M.; Hayashi, T.; Yamada, Y.; Kamihira, S.; Ikeda, S.; Yamasaki, Y.; Morikawa, S.; et al. Human T-cell leukemia virus type I Tax transactivates the matrix metalloproteinase-9 gene: Potential role in mediating adult T-cell leukemia invasiveness. Blood 2002, 99, 1341–1349. [Google Scholar] [CrossRef]
  525. Uchijima, M.; Sato, H.; Fujii, M.; Seiki, M. Tax proteins of human T-cell leukemia virus type 1 and 2 induce expression of the gene encoding erythroid-potentiating activity (tissue inhibitor of metalloproteinases-1, TIMP-1). J. Biol. Chem. 1994, 269, 14946–14950. [Google Scholar] [CrossRef]
  526. Kress, A.K.; Kalmer, M.; Rowan, A.G.; Grassmann, R.; Fleckenstein, B. The tumor marker Fascin is strongly induced by the Tax oncoprotein of HTLV-1 through NF-κB signals. Blood 2011, 117, 3609–3612. [Google Scholar] [CrossRef]
  527. Mohr, C.F.; Gross, C.; Bros, M.; Reske-Kunz, A.B.; Biesinger, B.; Thoma-Kress, A.K. Regulation of the tumor marker Fascin by the viral oncoprotein Tax of human T-cell leukemia virus type 1 (HTLV-1) depends on promoter activation and on a promoter-independent mechanism. Virology 2015, 485, 481–491. [Google Scholar] [CrossRef] [PubMed]
  528. Vincent, P.; Collette, Y.; Marignier, R.; Vuaillat, C.; Rogemond, V.; Davoust, N.; Malcus, C.; Cavagna, S.; Gessain, A.; Machuca-Gayet, I.; et al. A role for the neuronal protein collapsin response mediator protein 2 in T lymphocyte polarization and migration. J. Immunol. 2005, 175, 7650–7660. [Google Scholar] [CrossRef] [PubMed]
  529. Varrin-Doyer, M.; Nicolle, A.; Marignier, R.; Cavagna, S.; Benetollo, C.; Wattel, E.; Giraudon, P. Human T lymphotropic virus type 1 increases T lymphocyte migration by recruiting the cytoskeleton organizer CRMP2. J. Immunol. 2012, 188, 1222–1233. [Google Scholar] [CrossRef] [PubMed]
  530. Yi, T.; Lee, B.H.; Park, R.W.; Kim, I.S. Transactivation of fibronectin promoter by HTLV-I Tax through NF-κB pathway. Biochem. Biophys. Res. Commun. 2000, 276, 579–586. [Google Scholar] [CrossRef]
  531. Sherr, C.J.; McCormick, F. The RB and p53 pathways in cancer. Cancer Cell 2002, 2, 103–112. [Google Scholar] [CrossRef]
  532. Pise-Masison, C.A.; Choi, K.S.; Radonovich, M.; Dittmer, J.; Kim, S.J.; Brady, J.N. Inhibition of p53 transactivation function by the human T-cell lymphotropic virus type 1 Tax protein. J. Virol. 1998, 72, 1165–1170. [Google Scholar] [CrossRef]
  533. Suzuki, T.; Uchida-Toita, M.; Yoshida, M. Tax protein of HTLV-1 inhibits CBP/p300-mediated transcription by interfering with recruitment of CBP/p300 onto DNA element of E-box or p53 binding site. Oncogene 1999, 18, 4137–4143. [Google Scholar] [CrossRef]
  534. Ariumi, Y.; Kaida, A.; Lin, J.Y.; Hirota, M.; Masui, O.; Yamaoka, S.; Taya, Y.; Shimotohno, K. HTLV-1 tax oncoprotein represses the p53-mediated trans-activation function through coactivator CBP sequestration. Oncogene 2000, 19, 1491–1499. [Google Scholar] [CrossRef] [PubMed]
  535. Pise-Masison, C.A.; Mahieux, R.; Jiang, H.; Ashcroft, M.; Radonovich, M.; Duvall, J.; Guillerm, C.; Brady, J.N. Inactivation of p53 by human T-cell lymphotropic virus type 1 Tax requires activation of the NF-κB pathway and is dependent on p53 phosphorylation. Mol. Cell Biol. 2000, 20, 3377–3386. [Google Scholar] [CrossRef] [PubMed]
  536. Pise-Masison, C.A.; Mahieux, R.; Radonovich, M.; Jiang, H.; Brady, J.N. Human T-lymphotropic virus type I Tax protein utilizes distinct pathways for p53 inhibition that are cell type-dependent. J. Biol. Chem. 2001, 276, 200–205. [Google Scholar] [CrossRef] [PubMed]
  537. Jeong, S.J.; Radonovich, M.; Brady, J.N.; Pise-Masison, C.A. HTLV-I Tax induces a novel interaction between p65/RelA and p53 that results in inhibition of p53 transcriptional activity. Blood 2004, 104, 1490–1497. [Google Scholar] [CrossRef]
  538. Jeong, S.J.; Pise-Masison, C.A.; Radonovich, M.F.; Park, H.U.; Brady, J.N. A novel NF-κB pathway involving IKKβ and p65/RelA Ser-536 phosphorylation results in p53 Inhibition in the absence of NF-κB transcriptional activity. J. Biol. Chem. 2005, 280, 10326–10332. [Google Scholar] [CrossRef]
  539. Jeong, S.J.; Pise-Masison, C.A.; Radonovich, M.F.; Park, H.U.; Brady, J.N. Activated AKT regulates NF-κB activation, p53 inhibition and cell survival in HTLV-1-transformed cells. Oncogene 2005, 24, 6719–6728. [Google Scholar] [CrossRef]
  540. Brauweiler, A.; Garrus, J.E.; Reed, J.C.; Nyborg, J.K. Repression of bax gene expression by the HTLV-1 Tax protein: Implications for suppression of apoptosis in virally infected cells. Virology 1997, 231, 135–140. [Google Scholar] [CrossRef]
  541. Muhleisen, A.; Giaisi, M.; Kohler, R.; Krammer, P.H.; Li-Weber, M. Tax contributes apoptosis resistance to HTLV-1-infected T cells via suppression of Bid and Bim expression. Cell Death Dis. 2014, 5, e1575. [Google Scholar] [CrossRef] [PubMed]
  542. Olagnier, D.; Sze, A.; Bel Hadj, S.; Chiang, C.; Steel, C.; Han, X.; Routy, J.P.; Lin, R.; Hiscott, J.; van Grevenynghe, J. HTLV-1 Tax-mediated inhibition of FOXO3a activity is critical for the persistence of terminally differentiated CD4+ T cells. PLoS Pathog. 2014, 10, e1004575. [Google Scholar] [CrossRef]
  543. Bopp, T.; Becker, C.; Klein, M.; Klein-Hessling, S.; Palmetshofer, A.; Serfling, E.; Heib, V.; Becker, M.; Kubach, J.; Schmitt, S.; et al. Cyclic adenosine monophosphate is a key component of regulatory T cell-mediated suppression. J. Exp. Med. 2007, 204, 1303–1310. [Google Scholar] [CrossRef]
  544. Bielenberg, M.; Kurelic, R.; Frantz, S.; Nikolaev, V.O. A mini-review: Phosphodiesterases in charge to balance intracellular cAMP during T-cell activation. Front. Immunol. 2024, 15, 1365484. [Google Scholar] [CrossRef]
  545. Kress, A.K.; Schneider, G.; Pichler, K.; Kalmer, M.; Fleckenstein, B.; Grassmann, R. Elevated cyclic AMP levels in T lymphocytes transformed by human T-cell lymphotropic virus type 1. J. Virol. 2010, 84, 8732–8742. [Google Scholar] [CrossRef]
  546. Weil, R.; Levraud, J.P.; Dodon, M.D.; Bessia, C.; Hazan, U.; Kourilsky, P.; Israel, A. Altered expression of tyrosine kinases of the Src and Syk families in human T-cell leukemia virus type 1-infected T-cell lines. J. Virol. 1999, 73, 3709–3717. [Google Scholar] [CrossRef]
  547. Cheng, J.; Kydd, A.R.; Nakase, K.; Noonan, K.M.; Murakami, A.; Tao, H.; Dwyer, M.; Xu, C.; Zhu, Q.; Marasco, W.A. Negative regulation of the SH2-homology containing protein-tyrosine phosphatase-1 (SHP-1) P2 promoter by the HTLV-1 Tax oncoprotein. Blood 2007, 110, 2110–2120. [Google Scholar] [CrossRef]
  548. Nakase, K.; Cheng, J.; Zhu, Q.; Marasco, W.A. Mechanisms of SHP-1 P2 promoter regulation in hematopoietic cells and its silencing in HTLV-1-transformed T cells. J. Leukoc. Biol. 2009, 85, 165–174. [Google Scholar] [CrossRef]
  549. Feigenbaum, L.; Fujita, K.; Collins, F.S.; Jay, G. Repression of the NF1 gene by Tax may explain the development of neurofibromas in human T-lymphotropic virus type 1 transgenic mice. J. Virol. 1996, 70, 3280–3285. [Google Scholar] [CrossRef] [PubMed]
  550. Kibler, K.V.; Jeang, K.T. CREB/ATF-dependent repression of cyclin A by human T-cell leukemia virus type 1 Tax protein. J. Virol. 2001, 75, 2161–2173. [Google Scholar] [CrossRef] [PubMed]
  551. Beauvois, A.; Gazon, H.; Chauhan, P.S.; Jamakhani, M.; Jacques, J.R.; Thiry, M.; Dejardin, E.; Valentin, E.D.; Twizere, J.C.; Peloponese, J.M.; et al. The helicase-like transcription factor redirects the autophagic flux and restricts human T cell leukemia virus type 1 infection. Proc. Natl. Acad. Sci. USA 2023, 120, e2216127120. [Google Scholar] [CrossRef]
  552. Suzuki, S.; Zhou, Y.; Refaat, A.; Takasaki, I.; Koizumi, K.; Yamaoka, S.; Tabuchi, Y.; Saiki, I.; Sakurai, H. Human T cell lymphotropic virus 1 manipulates interferon regulatory signals by controlling the TAK1-IRF3 and IRF4 pathways. J. Biol. Chem. 2010, 285, 4441–4446. [Google Scholar] [CrossRef]
  553. Diani, E.; Avesani, F.; Bergamo, E.; Cremonese, G.; Bertazzoni, U.; Romanelli, M.G. HTLV-1 Tax protein recruitment into IKKε and TBK1 kinase complexes enhances IFN-I expression. Virology 2015, 476, 92–99. [Google Scholar] [CrossRef]
  554. Yuen, C.K.; Chan, C.P.; Fung, S.Y.; Wang, P.H.; Wong, W.M.; Tang, H.V.; Yuen, K.S.; Chan, C.P.; Jin, D.Y.; Kok, K.H. Suppression of type I interferon production by human T-cell leukemia virus type 1 oncoprotein Tax through inhibition of IRF3 phosphorylation. J. Virol. 2016, 90, 3902–3912. [Google Scholar] [CrossRef]
  555. Wang, D.; Guo, M.X.; Hu, H.M.; Zhao, Z.Z.; Qiu, H.L.; Shao, H.J.; Zhu, C.G.; Xue, L.; Shi, Y.B.; Li, W.X. Human T-cell leukemia virus type 1 oncoprotein tax represses ZNF268 expression through the cAMP-responsive element-binding protein/activating transcription factor pathway. J. Biol. Chem. 2008, 283, 16299–16308. [Google Scholar] [CrossRef]
  556. Yamagata, T.; Nishida, J.; Tanaka, S.; Sakai, R.; Mitani, K.; Yoshida, M.; Taniguchi, T.; Yazaki, Y.; Hirai, H. A novel interferon regulatory factor family transcription factor, ICSAT/Pip/LSIRF, that negatively regulates the activity of interferon-regulated genes. Mol. Cell Biol. 1996, 16, 1283–1294. [Google Scholar] [CrossRef] [PubMed]
  557. Takachi, T.; Takahashi, M.; Takahashi-Yoshita, M.; Higuchi, M.; Obata, M.; Mishima, Y.; Okuda, S.; Tanaka, Y.; Matsuoka, M.; Saitoh, A.; et al. Human T-cell leukemia virus type 1 Tax oncoprotein represses the expression of the BCL11B tumor suppressor in T-cells. Cancer Sci. 2015, 106, 461–465. [Google Scholar] [CrossRef] [PubMed]
  558. Gilli, S.C.; Salles, T.S.; Saad, S.T. Decreased GATA3 mRNA expression in human T-cell lymphotropic virus type 1 (HTLV-1) infection. Scand. J. Infect. Dis. 2000, 32, 161–163. [Google Scholar] [CrossRef] [PubMed]
  559. Gilli, S.C.; Salles, T.S.; Saad, S.T. Regulation of the GATA3 promoter by human T-cell lymphotropic virus type I Tax protein. J. Cell Biochem. 2004, 93, 1178–1187. [Google Scholar] [CrossRef]
  560. Riou, P.; Bex, F.; Gazzolo, L. The human T cell leukemia/lymphotropic virus type 1 Tax protein represses MyoD-dependent transcription by inhibiting MyoD-binding to the KIX domain of p300. A potential mechanism for Tax-mediated repression of the transcriptional activity of basic helix-loop-helix factors. J. Biol. Chem. 2000, 275, 10551–10560. [Google Scholar] [CrossRef]
  561. Colgin, M.A.; Nyborg, J.K. The human T-cell leukemia virus type 1 oncoprotein Tax inhibits the transcriptional activity of c-Myb through competition for the CREB binding protein. J. Virol. 1998, 72, 9396–9399. [Google Scholar] [CrossRef] [PubMed]
  562. Nicot, C.; Mahieux, R.; Opavsky, R.; Cereseto, A.; Wolff, L.; Brady, J.N.; Franchini, G. HTLV-I Tax transrepresses the human c-Myb promoter independently of its interaction with CBP or p300. Oncogene 2000, 19, 2155–2164. [Google Scholar] [CrossRef]
  563. Nicot, C.; Opavsky, R.; Mahieux, R.; Johnson, J.M.; Brady, J.N.; Wolff, L.; Franchini, G. Tax oncoprotein trans-represses endogenous B-myb promoter activity in human T cells. AIDS Res. Hum. Retroviruses 2000, 16, 1629–1632. [Google Scholar] [CrossRef]
  564. Nicot, C.; Mahieux, R.; Pise-Masison, C.; Brady, J.; Gessain, A.; Yamaoka, S.; Franchini, G. Human T-cell lymphotropic virus type 1 Tax represses c-Myb-dependent transcription through activation of the NF-κB pathway and modulation of coactivator usage. Mol. Cell Biol. 2001, 21, 7391–7402. [Google Scholar] [CrossRef]
  565. Sobol, R.W.; Horton, J.K.; Kuhn, R.; Gu, H.; Singhal, R.K.; Prasad, R.; Rajewsky, K.; Wilson, S.H. Requirement of mammalian DNA polymerase-β in base-excision repair. Nature 1996, 379, 183–186. [Google Scholar] [CrossRef]
  566. Jeang, K.T.; Widen, S.G.; Semmes, O.J.t.; Wilson, S.H. HTLV-I trans-activator protein, tax, is a trans-repressor of the human β-polymerase gene. Science 1990, 247, 1082–1084. [Google Scholar] [CrossRef] [PubMed]
  567. Uittenbogaard, M.N.; Armstrong, A.P.; Chiaramello, A.; Nyborg, J.K. Human T-cell leukemia virus type I Tax protein represses gene expression through the basic helix-loop-helix family of transcription factors. J. Biol. Chem. 1994, 269, 22466–22469. [Google Scholar] [CrossRef] [PubMed]
  568. Wencker, M.; Sausse, C.; Derse, D.; Gazzolo, L.; Duc Dodon, M. Human T-cell leukemia virus type 1 Tax protein down-regulates pre-T-cell receptor α gene transcription in human immature thymocytes. J. Virol. 2007, 81, 301–308. [Google Scholar] [CrossRef]
  569. Rahman, S.; Quann, K.; Pandya, D.; Singh, S.; Khan, Z.K.; Jain, P. HTLV-1 Tax mediated downregulation of miRNAs associated with chromatin remodeling factors in T cells with stably integrated viral promoter. PLoS ONE 2012, 7, e34490. [Google Scholar] [CrossRef]
  570. Manel, N.; Kim, F.J.; Kinet, S.; Taylor, N.; Sitbon, M.; Battini, J.L. The ubiquitous glucose transporter GLUT-1 is a receptor for HTLV. Cell 2003, 115, 449–459. [Google Scholar] [CrossRef]
  571. Ghez, D.; Lepelletier, Y.; Lambert, S.; Fourneau, J.M.; Blot, V.; Janvier, S.; Arnulf, B.; van Endert, P.M.; Heveker, N.; Pique, C.; et al. Neuropilin-1 is involved in human T-cell lymphotropic virus type 1 entry. J. Virol. 2006, 80, 6844–6854. [Google Scholar] [CrossRef]
  572. Lambert, S.; Bouttier, M.; Vassy, R.; Seigneuret, M.; Petrow-Sadowski, C.; Janvier, S.; Heveker, N.; Ruscetti, F.W.; Perret, G.; Jones, K.S.; et al. HTLV-1 uses HSPG and neuropilin-1 for entry by molecular mimicry of VEGF165. Blood 2009, 113, 5176–5185. [Google Scholar] [CrossRef] [PubMed]
  573. Jones, K.S.; Petrow-Sadowski, C.; Bertolette, D.C.; Huang, Y.; Ruscetti, F.W. Heparan sulfate proteoglycans mediate attachment and entry of human T-cell leukemia virus type 1 virions into CD4+ T cells. J. Virol. 2005, 79, 12692–12702. [Google Scholar] [CrossRef] [PubMed]
  574. Pinon, J.D.; Klasse, P.J.; Jassal, S.R.; Welson, S.; Weber, J.; Brighty, D.W.; Sattentau, Q.J. Human T-cell leukemia virus type 1 envelope glycoprotein gp46 interacts with cell surface heparan sulfate proteoglycans. J. Virol. 2003, 77, 9922–9930. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mechanisms of oncogenesis induced by animal retroviruses. (A) Introduction of a viral oncogene (v-onc). The virus possesses an activated oncogene derived from a cellular proto-oncogene and introduces the activated oncogene into the host cell genome upon infection. (B) Promoter insertion into or near a host proto-oncogene. When the virus genome is inserted into or near a host cell proto-oncogene, it enhances the expression of that gene due to promoter activity of the viral LTR.
Figure 1. Mechanisms of oncogenesis induced by animal retroviruses. (A) Introduction of a viral oncogene (v-onc). The virus possesses an activated oncogene derived from a cellular proto-oncogene and introduces the activated oncogene into the host cell genome upon infection. (B) Promoter insertion into or near a host proto-oncogene. When the virus genome is inserted into or near a host cell proto-oncogene, it enhances the expression of that gene due to promoter activity of the viral LTR.
Genes 16 01221 g001
Figure 2. Organization of the HTLV-1 genome. In addition to common structures of retroviral genomes such as LTRs, gag, pro, pol and env, the HTLV-1 genome contains a pX region, which lacks homology to the mammalian genome and codes for several regulatory proteins, such as Tax and Rex. An additional regulatory protein HBZ is encoded in the anti-sense strand.
Figure 2. Organization of the HTLV-1 genome. In addition to common structures of retroviral genomes such as LTRs, gag, pro, pol and env, the HTLV-1 genome contains a pX region, which lacks homology to the mammalian genome and codes for several regulatory proteins, such as Tax and Rex. An additional regulatory protein HBZ is encoded in the anti-sense strand.
Genes 16 01221 g002
Figure 3. The mechanism of oncogenesis exploited by HTLV-1. The trans-acting transcriptional activation mechanism. Tax encoded in the pX region functions as a transcriptional activator, which activates cellular transcription factors (TFs) to induce the expression of growth-promoting genes to facilitate the growth of host T cells.
Figure 3. The mechanism of oncogenesis exploited by HTLV-1. The trans-acting transcriptional activation mechanism. Tax encoded in the pX region functions as a transcriptional activator, which activates cellular transcription factors (TFs) to induce the expression of growth-promoting genes to facilitate the growth of host T cells.
Genes 16 01221 g003
Figure 4. Tax activates HTLV-1 LTR primarily via CREB. Tax indirectly binds to CRE of 21 bp-motifs in HTLV-1 LTR via CREB, while recognizing GC-rich sequences flanking CRE. Tax recruits transcriptional coactivators such as CBP/p300 and P/CAF to activate transcription. KID: kinase-inducible domain (KID)-like domain, KIX: kinase-inducible domain (KID) interacting domain, and Ac: acetylation.
Figure 4. Tax activates HTLV-1 LTR primarily via CREB. Tax indirectly binds to CRE of 21 bp-motifs in HTLV-1 LTR via CREB, while recognizing GC-rich sequences flanking CRE. Tax recruits transcriptional coactivators such as CBP/p300 and P/CAF to activate transcription. KID: kinase-inducible domain (KID)-like domain, KIX: kinase-inducible domain (KID) interacting domain, and Ac: acetylation.
Genes 16 01221 g004
Figure 5. The mechanisms of activation of the canonical and non-canonical NF-κB pathways by Tax. (A) Tax activates IKK complex through association with IKKγ (NEMO) and TAK1. Tax also binds to RNF8 and LUBAC, generating polyubiquitin chains to facilitate association between TAK1 and IKK. Activated IKK complex phosphorylates IκB targeting it for proteasome-mediated degradation, thereby releasing NF-κB. Tax associates with IκB and releases NF-κB. Tax also associates with RelA and c-Rel to enhance transcription. (B) Tax associates with and activates IKKα to phosphorylate p100, triggering limited proteolysis and generating p52. Tax also associates with and activates NIK to activate IKKα.
Figure 5. The mechanisms of activation of the canonical and non-canonical NF-κB pathways by Tax. (A) Tax activates IKK complex through association with IKKγ (NEMO) and TAK1. Tax also binds to RNF8 and LUBAC, generating polyubiquitin chains to facilitate association between TAK1 and IKK. Activated IKK complex phosphorylates IκB targeting it for proteasome-mediated degradation, thereby releasing NF-κB. Tax associates with IκB and releases NF-κB. Tax also associates with RelA and c-Rel to enhance transcription. (B) Tax associates with and activates IKKα to phosphorylate p100, triggering limited proteolysis and generating p52. Tax also associates with and activates NIK to activate IKKα.
Genes 16 01221 g005
Figure 6. Representatives of host cell target genes regulated by Tax. The expected roles of the gene products involved in early transformation of HTLV-1-infected cells are depicted. Solid arrows indicate activation, dashed arrows indicate pathways, and ⊥ indicates inhibition.
Figure 6. Representatives of host cell target genes regulated by Tax. The expected roles of the gene products involved in early transformation of HTLV-1-infected cells are depicted. Solid arrows indicate activation, dashed arrows indicate pathways, and ⊥ indicates inhibition.
Genes 16 01221 g006
Table 1. Cytokines, their receptors, and cell surface molecules induced by Tax.
Table 1. Cytokines, their receptors, and cell surface molecules induced by Tax.
GeneGene Description References
Interleukins and those receptors
IL-1αInterleukin-1α[254,255,256,257,258]
IL-1βInterleukin-1β[256,259]
IL-2Interleukin-2[208,209,210,211,212,213]
IL-2RαInterleukin-2 receptor α
(CD25, Tac antigen)
[208,209,214,215,216,217,218,219]
IL-2RγInterleukin-2 receptor γ
(CD132, common γ chain)
[222]
IL-4Interleukin-4[260]
IL-5Interleukin-5[203,227,261]
IL-6Interleukin-6[188,254,256,262,263,264,266]
IL-6RInterleukin-6 receptor[265]
IL-8Interleukin-8[205,267]
IL-9Interleukin-9[227,266]
IL-10Interleukin-10[83,268]
IL-12Interleukin-12[270]
IL-13Interleukin-13[227,271,272,273]
IL-15Interleukin-15[229]
IL-15RInterleukin-15 receptor[231]
IL-17Interleukin-17[274]
IL17RInterleukin-17 receptor[275]
IL-21Interleukin-21[232]
IL-21RInterleukin-21 receptor[232]
Chemokines and those receptors
CCL1C-C motif chemokine ligand 1[286]
CCL2
(MCP-1)
C-C motif chemokine ligand 2 (monocyte chemoattractant protein-1)[270,284,285]
CCL3
(MIP-1α)
C-C motif chemokine ligand 3
(macrophage inflammatory protein 1α)
[184,284,285,287]
CCL4
(MIP-1β)
C-C motif chemokine ligand 4
(macrophage inflammatory protein 1β)
[288,289]
CCL5
(RANTES)
C-C motif chemokine ligand 5
(regulated on activation, normal T cell expressed and secreted)
[284,290]
CCL7
(MCP-3)
C-C motif chemokine ligand 7
(monocyte chemoattractant protein-3)
[270]
CCR9C-C chemokine receptor type 9[198]
CCL11C-C motif chemokine ligand 11
(Eotaxin)
[270]
CCL20
(MIP-3α)
C-C motif chemokine ligand 20
(macrophage inflammatory protein 3α)
[291]
CCL22C-C motif chemokine ligand 22[292]
CXCR4C-X-C chemokine receptor type 4[296]
CXCR7C-X-C chemokine receptor type 7[186]
CXCL10
(IP-10)
C-X-C motif chemokine ligand 10
(interferon γ-induced protein 10)
[184,284,293]
CXCL12
(SDF-1)
(PBSF)
C-X-C motif chemokine ligand 12
(stromal cell-derived factor 1)
(pre-B cell growth-stimulating factor)
[295]
XCL1Chemokine, C motif, and ligand 1[184]
Growth factors
c-sisPlatelet-derived growth factor subunit B[226]
TGF-β1Transforming growth factor β1[234,235]
VEGFVascular endothelial growth factor[243]
GM-CSF
(CSF2)
Granulocyte–macrophage colony-stimulating factor
(colony-stimulating factor 2)
[209,218,282,283]
G-CSFGranulocyte colony-stimulating factor[283]
NGFNerve growth factor[311]
Costimulatory molecules
gp34Tumor necrosis factor ligand superfamily member 4
(CD252)
[181,239,240]
OX40Cluster of differentiation 134[189,237]
CD40Cluster of differentiation 40[241]
CD40LCD40 ligand[242]
Tumor necrosis factor-related
TNF-αTumor necrosis factor α[256,270,277,278,279]
TNF-βTumor necrosis factor β
(Lymphotoxin)
[256,280,281]
TNFRSF9/4–1BB/CD137/ILATumor necrosis factor receptor superfamily member 9[307]
CD70Tumor necrosis factor ligand superfamily member 7[308]
Cell adhesion molecules
VCAM-1Vascular cell adhesion molecule 1 (CD106)[204,298]
E-cadherin [306]
Others
PTHrPParathyroid hormone-related peptide[245,246,247,248]
IFN-γInterferon γ[270,271]
OPNOsteopontin[302]
CD44Cluster of differentiation 44[303]
LFA-3
(CD58)
Lymphocyte function-associated antigen, type 3
(cluster of differentiation 58)
[191]
CD69Cluster of differentiation 69[309]
CD83Cluster of differentiation 83[193]
CD137Cluster of differentiation 137[305]
MHC class IMajor histocompatibility complex, class I[310]
proenkephalin [312,313,314]
RGMaRepulsive guidance molecule A[297]
Table 2. Proto-oncogenes, transcription factors, and growth signal transducers induced by Tax.
Table 2. Proto-oncogenes, transcription factors, and growth signal transducers induced by Tax.
GeneGene DescriptionReferences
Proto-oncogenes
c-Fosv-Fos FBJ murine osteosarcoma viral oncogene homolog[180,315,316]
c-Junv-jun avian sarcoma virus 17 oncogene homolog[206,316]
c-Mycv-Myc avian myelocytomatosis viral oncogene homolog[320]
c-Relv-Rel avian reticuloendotheliosis viral oncogene homolog[134,326]
Transcription factors
Fra-1Fos-related antigen 1[206]
JunD [206]
Egr-1Early growth response 1[206,317,318]
Egr-2Early growth response 2[206,318]
ETR101Immediate-early response gene 2 (IER2)[202]
BRF1
(ZFP36L1)
Butyrate response factor 1
(Zinc finger protein 36-like 1)
[319]
p105Nuclear factor kappa-B, subunit 1 (NFKB1)[326]
STAT1Signal transducer and activator of transcription 1[332]
STAT5Signal transducer and activator of transcription 5[332]
IRF-4Interferon regulatory factor 4[333]
IRF-5Interferon regulatory factor 5[336]
Nur77Nuclear receptor sub-family 4, group A, member 1 (NR4A1)[196,201]
BCL6B-cell/CLL lymphoma 6[347]
Signal transducers
Pim-1Serine/threonine protein kinase PIM1[327]
Pim-3Serine/threonine protein kinase PIM3[327,328]
LynLyn protooncogene, Src family tyrosine kinase[339]
JAG1Jagged 1[351]
Cas-LCrk-associated substrate lymphocyte type[185]
MAGI3Membrane-associated guanylate kinase, WW and PDZ domains-containing 3[358]
Others
EVCEllis Van Creveld[343]
IκB-ζInhibitor of NF-κB-ζ[345]
Table 3. Cell cycle regulators induced by Tax.
Table 3. Cell cycle regulators induced by Tax.
GeneGene DescriptionReferences
CCND1Cyclin D1[117,376]
CCND2Cyclin D2[374,375,377,378]
E2F1E2F transcription factor 1[373,376]
CDK6Cyclin-dependent kinase 6[378]
CDK7Cyclin-dependent kinase 7[384]
ELL2Elongation factor, RNA polymerase II, 2[386]
CDKN1ACyclin-dependent kinase inhibitor 1A
CDK-interacting protein 1 (CIP1)
Wildtype p53-activated fragment 1 (WAF1)
p21
[389,391,392]
Table 4. Genes related to apoptosis and cell survival induced by Tax.
Table 4. Genes related to apoptosis and cell survival induced by Tax.
Gene Gene Description References
Bcl-xLBcl2-related protein, long isoform[407,408]
BFL1Bcl2-related protein A1[410]
BIRC3Baculoviral IAP repeat-containing 3 (cIAP2, HIAP-1)[411,412]
XIAPX-chromosome-linked inhibitor of apoptosis[183]
SurvivinBaculoviral IAP repeat-containing 5 (BIRC5)[415]
c-FLIPCellular FLICE-inhibitory protein[416,417,418]
A20Tumor necrosis factor-α-induced protein 3 (TNFAIP3)[420]
MDR1Multidrug resistance protein 1
ATP-binding cassette sub-family B member 1 (ABCB1)Ppermeability glycoprotein (P-glycoprotein, P-gp)
[423,424,425]
LRPLung resistance-related protein[426]
FasLFas ligand[190]
TRAILTNF-related apoptosis-inducing ligand[428]
Table 5. Genes involved in the suppression of DNA repair and induction of genome instability induced by Tax.
Table 5. Genes involved in the suppression of DNA repair and induction of genome instability induced by Tax.
GeneGene DescriptionReferences
PCNAProliferating cell nuclear antigen[439,440,441]
iNOSInducible nitric oxide synthase[443,445]
WIP1Wildtype p53-induced phosphatase 1
Protein phosphatase, magnesium/manganese-dependent, 1D (PPM1D)
[453]
Table 6. MicroRNAs (miRNAs) induced by Tax.
Table 6. MicroRNAs (miRNAs) induced by Tax.
GeneTarget GenesReference
miR-146a [472]
miR-155Tumor protein p53-inducible nuclear protein 1 (TP53INP1)[473]
Table 7. Epigenetic genes induced by Tax.
Table 7. Epigenetic genes induced by Tax.
GeneGene DescriptionReferences
EZH2Enhancer of zeste homolog 2[494,495]
CARM1Coactivator-associated arginine methyltransferase 1
Protein arginine methyltransferase 4 (PRMT4)
[498]
Table 8. Genes involved in cellular metabolism induced by Tax.
Table 8. Genes involved in cellular metabolism induced by Tax.
GeneGene DescriptionReference
GM2/GD2 synthaseβ-1,4-N-acetylgalactosaminyltransferase[499]
TRXThioredoxin[505]
DHODHDihydroorotate dehydrogenase[506]
Table 9. Genes involved in viral transmission induced by Tax.
Table 9. Genes involved in viral transmission induced by Tax.
GeneGene DescriptionReferences
ICAM-1Intercellular adhesion molecule 1 (CD54)[191,192,508,509]
TNFAIP2TNF-α-induced protein 2 (M-Sec)[511]
PLA2G4CPhospholipase A2, group IVC[512]
LactoferrinLactotransferrin (LTF)[514]
Galectin-1Lectin, galactoside-binding, soluble, 1 (LGALS1)[515]
Galectin-3Lectin, galactoside-binding, soluble, 3 (LGALS3)[516]
Table 10. Genes involved in invasion and infiltration induced by Tax.
Table 10. Genes involved in invasion and infiltration induced by Tax.
GeneGene DescriptionReferences
ALCAMActivated cell adhesion molecule[518]
Vimentin [519,520]
FucT VIIFucosyltransferase 7[521,522]
MMP-7Matrix metalloproteinase 7[523]
MMP-9Matrix metalloproteinase 9[524]
TIMP-1Tissue inhibitors of matrix metalloproteinases-1[525]
FascinActin-bundling protein[526,527]
CRMP2Collapsin response mediator protein 2[529]
FNFibronectin[530]
Table 11. Genes downregulated by Tax.
Table 11. Genes downregulated by Tax.
Gene Gene Description References
Pro-apoptotic genes
BaxBcl-2-associated X protein[540]
BimBcl-2-interacting mediator of cell death[541,542]
BidBH3-interacting domain death agonist[541]
PUMAp53-upregulated modulator of apoptosis
Bcl2-binding component 3 (BBC3)
[542]
Signal transducers
PDE3BPhosphodiesterase 3B[545]
LckLymphocyte-specific protein-tyrosine kinase[187]
Zap-70Zeta-chain-associated protein kinase
Syk-related tyrosine kinase (SRK)
[546]
SHP-1Src homology-2-containing protein-tyrosine phosphatase 1
Tyrosine-protein phosphatase non-receptor type 6 (PTPN6)
[547]
NF1Neurofibromatosis type I[549]
Cell cycle regulators
p18Ink4cCyclin-dependent kinase inhibitor 2C (CDKN2C)[402]
CCNA2Cyclin A[550]
Transcription factors
HLTFHelicase-like transcription factor[551]
ZNF268Zinc finger protein 268[555]
BCL11BB-cell CLL/Lymphoma 11B[557]
c-Mybv-Myb avian myeloblastosis viral oncogene homolog[562,563,564]
B-Mybv-Myb avian myeloblastosis viral oncogene homolog-like 2[563]
Others
Pol βDNA polymerase β[566,567]
Type I IFNType I Interferon[554]
pTCRαPre-T-cell receptor α[568]
Target genes
miR-149p300 and P/CAF[569]
miR-873p300 and P/CAF[569]
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

Shirasawa, M.; Nakajima, R.; Zhou, Y.; Fikriyanti, M.; Iwanaga, R.; Bradford, A.P.; Kurayoshi, K.; Araki, K.; Ohtani, K. Transcriptional Activation Mechanisms and Target Genes of the Oncogene Product Tax of Human T-Cell Leukemia Virus Type 1. Genes 2025, 16, 1221. https://doi.org/10.3390/genes16101221

AMA Style

Shirasawa M, Nakajima R, Zhou Y, Fikriyanti M, Iwanaga R, Bradford AP, Kurayoshi K, Araki K, Ohtani K. Transcriptional Activation Mechanisms and Target Genes of the Oncogene Product Tax of Human T-Cell Leukemia Virus Type 1. Genes. 2025; 16(10):1221. https://doi.org/10.3390/genes16101221

Chicago/Turabian Style

Shirasawa, Mashiro, Rinka Nakajima, Yaxuan Zhou, Mariana Fikriyanti, Ritsuko Iwanaga, Andrew P. Bradford, Kenta Kurayoshi, Keigo Araki, and Kiyoshi Ohtani. 2025. "Transcriptional Activation Mechanisms and Target Genes of the Oncogene Product Tax of Human T-Cell Leukemia Virus Type 1" Genes 16, no. 10: 1221. https://doi.org/10.3390/genes16101221

APA Style

Shirasawa, M., Nakajima, R., Zhou, Y., Fikriyanti, M., Iwanaga, R., Bradford, A. P., Kurayoshi, K., Araki, K., & Ohtani, K. (2025). Transcriptional Activation Mechanisms and Target Genes of the Oncogene Product Tax of Human T-Cell Leukemia Virus Type 1. Genes, 16(10), 1221. https://doi.org/10.3390/genes16101221

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