GET THE APP

Contribution of Inducible Nitric Oxide Synthase to the Transforma
Journal of Leukemia

Journal of Leukemia
Open Access

ISSN: 2329-6917

+44 1300 500008

Research Article - (2015) Volume 3, Issue 4

Contribution of Inducible Nitric Oxide Synthase to the Transformation of HTLV-1 Infected CD4+ T-Cells

Hicham H. Baydoun1 and Lee Ratner1,2*
1Department of Medicine, Division of Molecular Oncology, Washington University School of Medicine, St Louis, USA
2Department of Molecular Microbiology, Washington University School of Medicine, St Louis, USA
*Corresponding Author: Lee Ratner, Department of Medicine, Division of Molecular Oncology, Washington University School of Medicine, St Louis, USA, Tel: 314-362-8836, Fax: 314-747-2120 Email:

Abstract

The Human T-cell Leukemia Virus type 1 (HTLV-1), is the first retrovirus associated with a human cancer. HTLV-1 is the causative agent of an aggressive and fatal malignancy of CD4+ T lymphocytes known as Adult T-cell Leukemia lymphoma (ATLL). Since the discovery of the virus in 1980, intensive investigations have been undertaken to determine how HTLV-1 drives the transformation process in infected cells. This is because the oncogenic features of HTLV-1 make it an excellent tool to dissect the molecular pathways involved in cancer development. More important, HTLV-1 induced leukemia is a typical inflammation-mediated malignancy with constitutive activation of the NF-kB pathway, which is also a critical determinant in many other cancers. How NF-kB contributes to the leukemogenic process is not completely defined. We recently demonstrated that the NF-kB pathway induces the expression of inducible nitric oxide synthase (iNOS) in HTLV-1 induced leukemia. iNOS enzymatically generates nitric oxide, which is an oxidative and nitrosative agent of DNA and protein. Nitric oxide was found to be associated with a large number of DNA Double Strand Breaks (DSBs) in HTLV-1 transformed cells. Here, we will review the major effects of nitric oxide on HTLV-1 induced leukemia.

Keywords: HTLV; Nitric oxide; Tax

Introduction

The Human T-cell Leukemia Virus type 1 (HTLV-1) is the etiological agent of Adult T-cell Leukemia Lymphoma (ATLL), a rare and aggressive T-cell malignancy. The transmission of the virus occurs sexually or by IV drug abuse, but the most efficient way of viral transmission is through breast-feeding from an infected mother to her baby [1,2] (Figure 1). This is because the breast epithelial cells regulate a physiological recruitment of lymphoid and myeloid cells from the circulation into the milk, while secreting nutritive molecules, antibiotic substances, growth factors, inflammatory cytokines, and chemokines [3]. As a result, breast milk allows contact between lymphoid cells which promotes cell to cell transmission of the virus, a more efficient manner of virus spread as compared to free particle infection [4,5]. Yet, for unknown reasons, only a few percent of infected individuals develop ATLL after a long period of latency [6]. Currently, there is no way to predict which infected patients will develop ATLL, and there is no effective treatment for those entering the acute phase of the disease. Of note, it is still not known whether the integration of the proviral DNA into specific loci in the human genome has a role in ATLL development [7]. Moreover, the concept of the monoclonal disease development has recently been debated as a result of deep sequencing results, which showed that multiple clones can evolve during progression of the disease [8]. It is also not understood why ATLL develops only in CD4+ T-cells, while the virus is present in almost all lymphoid and myeloid progenitors, including hematopoietic stem cells (HSC) [9,10]. However, data obtained from HTLV-1 infected humanized mouse (HIS) demonstrated that high frequency of HTLV-1 infection was found in the double positive T-cells during lymphogenesis suggesting that lymphoid progenitors constitute the niche of HTLV-1 infection. The other infected cells either represent the latent reservoirs of the virus or lack properties to support the process of transformation [11-14]. Because HTLV-1 infection has evolved mechanisms that activate CD4+ T-cells and impair the immune CTL response, the outcome of the disease largely depends on two antagonist factors, the proviral load and the efficiency of the immune response against the infected cells [6,15]. Activation of proliferation and inhibition of tumor suppressors are also two major hallmarks of oncogenic events occurring during the long period of latent infection. However, the accumulation of genetic defects is believed to be a driving force for transformation [16]. How and when these genetic defects accumulate is still under intense investigation.

leukemia-hematopoietic-cells

Figure 1: Characteristics of HTLV-1 associated malignancy. Number of infected individuals and their world endemic distribution as well as the mode of the virus transmission and the prevalence of the disease. Although HTLV-1 infect other hematopoietic cells, ATLL is a clonal expansion of CD4+, CD3+ and CD25+ T-cells with abnormalities in cell division as seen with DAPI staining of ATLL cells.

Over the last decade, there is increasing evidence that inflammation is a hidden force that drives many malignancies [17-20]. More than 90% of cancers are associated with some forms of chronic inflammation, which are induced by infections, obesity, use of tobacco, and exposure to different mutagenic agents. Most likely, inflammation is related to cancer through processes that involve genotoxicity, aberrant tissue repair, proliferative responses, invasion and metastasis. In contrast to somatic inherited mutations, inflammation may induce random mutations, which will contribute to additional cooperating events for the initiation and maintenance of the transformation process [21,22]. HTLV-1 induced leukemia is a typical inflammation-mediated malignancy with constitutive activation of the NF-kB pathway [23], which is also a critical determinant in many other cancers [24-29]. The NF-kB pathway activates the expression of a large number of genes involved in immunity and the inflammatory response, apoptosis, proliferation, differentiation, and survival [30,31]. Cytokines and chemokines are the first effectors of the inflammatory response [32]. They contribute to the proliferation of pre-neoplastic cells, but little is known about the downstream effectors that are involved in this process, and whether a specific connection exists between NF-kB activation and accumulation of genetic defects [25].

The downstream effectors pose another level of complexity during the inflammatory response. Many of them act as double edged swords with beneficial and detrimental effects, depending on the physiological environment and magnitude of expression [33]. They are mainly involved in fighting infection and inflammatory conditions. However in pre-neoplastic cells in which the apoptotic machinery has been compromised, the inflammatory downstream effectors may induce constitutive damage that drives the transformation process [34].

Inducible nitric oxide synthase (iNOS) is one of the most common downstream inflammatory effectors. It was found to be overexpressed in chronic inflammatory diseases as well as in various types of cancer [35-45]. iNOS is an enzyme catalyzing the production of nitric oxide (NO), which is an important regulatory molecule in both inflammation and cancer development [46-48]. NO is the precursor of the highly reactive nitrogen species peroxynitrite (ONOO-), an obligatory factor of oxidative and nitrosative modifications of DNA and proteins [49-52] (Figure 2). It has recently been shown that selective inhibitors of iNOS, that reduce the release of nitric oxide in vivo, inhibited the progression of tumorigenesis in several cancers models [40,53-57]. Moreover, the inhibition of lipopolysaccharide (LPS) induced NF-kB activation inhibits iNOS expression and NO production, and inhibited inflammation-mediated tumorigenesis in mouse models [58-65].

leukemia-nitric-oxide-synthase

Figure 2: Nitric oxide induces a cell damage response as a defense mechanism. Inducible nitric oxide synthase (iNOS) as well as oxidases are expressed in response to inflammatory responses, which are induced by stimuli such as toxicity, hypoxia, infection, inflammation, radiation, chemical stress and obesity. iNOS and oxidases produce respectively nitric oxide, NO and superoxide, O2-. The reaction between O2- and NO generates a much more stable and highly reactive molecule, peroxynitrite ONOO-. ONOO- is an obligatory factor for oxidative and nitrosative modification of DNA and proteins.

In contrast to other constitutively expressed isoforms, iNOS is only expressed in response to inflammatory cytokines such as TNF-α and IL-1β and transcriptional activators, such NF-kB [48,66-69]. In chronic inflammation models and inflammation-related tumorigenesis, iNOS may be persistently stimulated by cytokines and NF-kB activation in the tumor microenvironment. iNOS/NO signaling can also induce cyclo-oxygenase-2 (COX-2), which is another link between inflammation and cancer [59,70]. In view of the diverse effects of iNOS-produced NO, it is important to determine how cells regulate their iNOS/NO system. Nevertheless, the progressive alterations of DNA and protein modifications are probably the major outcome. In the present review, we will focus on these two functions of NO in the context of HTLV-1 induced leukemia, a representative human malignancy for which the etiological agent is clearly identified.

Expression of Nitric Oxide Synthase

Three isoforms constitute the family of nitric oxide synthase (NOS) that catalyze the production of nitric oxide (NO) from L-arginine [69,71,72]. Neuronal nitric oxide synthase (nNOS or NOS1) and endothelial nitric oxide synthase (eNOS or NOS3) are constitutively expressed at steady state in the corresponding tissues, and are involved in neurotransmission and vasodilation, respectively [69]. In contrast, inducible nitric oxide synthase is expressed de novo, in response to inflammatory mediators, and its expression varies depending on the physiological environment and the magnitude of the inflammatory response [73,74]. While nNOS and eNOS catalyze low levels of NO synthesis in a Ca2+ dependent manner, iNOS generates high levels of NO independent of Ca2+ [75]. iNOS expression has also been detected in a wide array of cells and tissues, including pulmonary and colonic epithelium, and hepatocytes, but the immune cells, mainly macrophages and neutrophils, are considered as the major sources of iNOS synthesis [36,76-86]. They generate large amounts of NO in the extrinsic environment, with a primary microbiocidal activity. iNOS can also be induced and expressed in virally-infected lymphocytes, specifically T cells, but little is known about its intrinsic effects in activated regulatory CD4+ T-cells[87-91].

While all nitric oxide synthetases enzymatically catalyze NO production, they only share 50% amino acid sequence similarity, and are located on different chromosomes [69,76]. Nitric oxide synthetases also differ by the mechanisms regulating their expression. Despite the fact that post-transcriptional, co-translational, and post-translational regulation play roles in NOS expression, the predominant regulatory mechanism is transcriptional regulation [71,72,80,92,93]. The mechanism of transcriptional regulation of human iNOS is much more complex than those mediating the expression of constitutively expressed nNOS and eNOS, as well as the expression of murine iNOS [94,95].

For almost two decade, murine macrophages were used as a main research tool for iNOS investigations. A combination of lipopolysaccharides (LPS) that activate the Toll-like receptor, and interferon-γ (IFN-γ) that activates the interferon type II response through JAK1/STAT1 pathway was sufficient to induce the expression of murine iNOS [96-101]. However, human iNOS involves a complex mechanism of transcriptional regulation that requires a mixture of cytokines and transcription factors. There is considerable evidence to suggest that many signaling pathways are involved in human iNOS expression that include IKK-IκB-NF-kB, JAK/STAT, PI3K-Akt, and MAPKs as well as the ubiquitin-proteasome degradation pathways. NF-kB, AP1, STAT1α, IRF-1, Oct1, C/EBPβ, ATF-2 and cAMP responsive element are specific transcription factors that have been described to interact and activate the human iNOS promoter [72,94,100,102-117]. iNOS expression can also stimulated in hypoxic conditions, and HIF-1α is one of the transcription factors that participate in the induction of iNOS expression [118,119].

Regulation of the expression of inducible nitric oxide synthase in HTLV-1 infected cells

Significant questions have been addressed to understand the effect of iNOS/NO on tumor biology. High expression of iNOS has been targeted by selective inhibitors in many animal cancer models, including colon, breast, prostate, bladder, skin, esophageal, and head and neck cancers, but the mechanisms that involve iNOS/NO in tumorigenesis are yet to be determined [120-135]. It is very important to clearly delineate the mechanisms of iNOS expression and NO production and their effects in the tumor microenvironment before designing prevention and therapeutic strategies that target iNOS/NO signaling. This is because iNOS/NO signaling has an extrinsic effect, with anti-infection and anti-tumoral actions mainly generated by macrophages and neutrophils, while its intrinsic effect has a tumor promoting activity within the infected/inflamed cells. Therefore, it is more relevant to explore iNOS/NO signaling in a virally-induced tumor like HTLV-1 induced leukemia, in which the infected cells represent the appropriate recipient for investigating the intrinsic effect of iNOS/NO.

HTLV-1 induced leukemogenesis in CD4+ T-cells is mediated by the expression of the viral oncoprotein Tax [6,16,23,136-140]. Among its many oncogenic functions, Tax induces a potent inflammatory response through activation of the NF-kB pathway. Because Tax has an intermittent expression during the early stages of infection and it is rarely measureable during the acute phase of the disease, it is still not clear whether Tax exerts different functions between these two phases of virally-induced tumorigenesis [6,138,141]. In fact, Tax follows the approach of “hit and run”, in which it promotes the oncogenic events during the infection course and it hides to prevent its elimination by the immune CTL response. Although Tax induces irreversible events that lead to the transformation of HTLV-1 infected lymphocytes, its expression might still be required to maintain the ATLL malignancy at later stages [138]. Most likely, Tax drives the initial events of immortalization in HTLV-1 infected cells through the expression of inflammatory mediators by NF-kB activation. In fact, Tax transgenic mouse model develop an ATLL like disease, and express high levels of TNFα, IL-1β, IL-6 and IFNγ, which participate in iNOS expression[140,142]. Moreover, we and others have demonstrated high levels of iNOS expression in HTLV-1 infected cells and this activation required the NF-kB pathway [87,143,144].

We recently reported a comprehensive study, in which we investigated the mechanism of iNOS expression in HTLV-1 infected T-cells (Figure 3). By using a human iNOS promoter reporter and small interfering RNA, we demonstrated that an activation of the classical NF-kB pathway was sufficient to activate iNOS expression by HTLV-1 Tax, and this activation involved a combination of both NF-kB and JAK/STAT pathways [87]. Moreover, we reported an increase of the active phosphorylated form of the signal transducer and activator of transcription, p-STAT1, and the interferon regulator factor 1, IRF-1 in HTLV-1 and Tax expressing cells. We also demonstrated that JAK1/STAT1 activation by type II interferon, but not type I interferon, is required to induce iNOS expression. In addition to NF-kB transcription factors, p-STAT1 and IRF-1 have previously been reported to activate iNOS expression in chronic inflammation models [145-147]. Interestingly, the inhibition of the JAK/STAT pathway by ruxolitinib, a JAK1/JAK2 inhibitor reduced iNOS expression and NO production. Although JAK/STAT activation is intended to induce an interferon-dependent, antiviral response, its role in HTLV-1 transformation was not well defined [148]. Interestingly, a recent study showed that a combination of ruxolitinib and navitoclax, a Bcl-2/Bcl-xl inhibitor, dramatically lowered tumor burden and prolonged survival in an ATL murine model [149]. This combination strongly blocked ex vivo proliferation of ATL patients’ PBMCs, indicating that JAK/STAT pathway promotes the proliferation of HTLV-1 infected cells.

leukemia-post-translational-modification

Figure 3: S-nitrosylation of Key proliferative proteins – In addition to the well-characterized NO function as a signal transducer, Snitrosylation, which is a covalent addition of NO to the thiol group of cysteine, has emerged as a major post-translational modification of proteins. Over the past decade, the number of substrates modified by s-nitrosylation has considerably increased to include the small GTPase Ras, the protein kinase AKT, and the phosphotase and tensin homolog PTEN, which are all examples of proteins involved in proliferation, cell survival and tumorigenesis. Snitrosylation was found to be a major player in their respective functions.

Whether iNOS/NO signaling induces and/or maintains the immortalization process in HTLV-1 infected cells is not clear. But, the establishment in vitro of newly infected PBMCs with an HTLV-1 infected cell line showed that new infection was associated with an increase of iNOS expression, suggesting that de novo expression of iNOS is induced with productive infection. The high level of expression of iNOS, which is detected in HTLV-1 transformed cell lines and in ATLL patients’ samples, also suggest that iNOS/NO signaling is required to maintain the phenotype of HTLV-1 induced leukemia [87].

Origin of Genomic Instability in HTLV-1 Infected Cells

Genetic instability is defined by increased rates of DNA damage and by the inability to maintain a faithful integrity of the genome within the infected cells. Although DNA damage is continuously created by multiple sources, cells have evolved mechanisms that include tumors suppressors, cell cycle checkpoints, and DNA repair pathways to control genomic integrity. Consequently, DNA damage is repaired, or if it is left unrepaired, the cell will activate programmed cell death.

In HTLV-1 infected cells, the origin of genetic instability was associated with defects in different DNA repair mechanisms. i) Amplification of centrosomes, a common cause for improper distribution of chromosomes and aneuploidy, are two hallmarks of malignant cells. ii) Inactivation of cell division checkpoints, such Anaphase-promoting complex (APC) and mitotic spindle checkpoint (MSC) proteins such MAD1 and MAD2 lead to chromosome missegregation and accumulation of multinucleated cells, is a common phenotype of ATL cells. iii) Induction of DNA double strand breaks (DSBs) and inhibition of DNA damage repair pathways are also responsible for genetic defects in HTLV-1 infected cells. While other HTLV-1 viral proteins have been involved in the regulation of genomic instability, the Tax oncoprotein was described as the main modulator of these defects (Figure 4). DSBs are the most detrimental forms of DNA damage because they pose problems for replication, transcription, and chromosome segregation, and are often the origin of mutations found in malignant cells. DSBs were found increased in HTLV-1 and Tax expressing cells, suggesting that Tax is able to induce DSBs and inhibit DNA damage repair response. The majority of these DSBs were recently attributed to the high levels of iNOS-produced NO in HTLV-1 infected cells. In fact, the inhibition of iNOS activity by the selective inhibitor 1400W (N-[[3-(amino methyl) phenyl] methyl]-ethanimidamide, dihydrochloride) reduced the number of DSBs in HTLV-1 infected by 60%, as observed by immunofluorescence staining of γ-H2AX foci. However, the genetic depletion of iNOS expression by a specific shRNA and complete inhibition of NO production eliminated the majority of DSBs from HTLV-1 infected cells. Under the same conditions, the active phosphorylated forms of ATM, ATR, 53BP1, Chk2 and H2AX of the DNA damage response were abolished. Similar results were obtained in the same study by comet alkaline assays that functionally test the DNA single and double strand breaks at the level of single cells. In the presence of an iNOS inhibitor or by genetic depletion of iNOS expression, the comet tails were reduced, suggesting an attenuation of DNA damage in HTLV-1 infected cells.

leukemia-Tax-Mediated-Genetic

Figure 4: Model of Tax-Mediated Genetic Defects - HTLV-1 induced leukemogenesis is mediated by the expression of the viral oncoprotein Tax. Among its many oncogenic functions, Tax induces DNA Double Strand Breaks and inhibits DNA repair responses, which are two mechanisms at the origin of genetic defects. DSBs, which are the most detrimental forms of DNA damage, are found to be induced in HTLV-1 and Tax expressing cells by iNOS-produced NO and this induction required an activation of NF-kB and JAK/ STAT, p-STAT1, IRF-1 signaling pathways.

iNOS/NO and DNA oxidation/nitration

Initial mutations in pre-neoplastic cells are believed to be induced by DNA oxidative and nitrosative molecules, such as reactive oxygen species and reactive nitrogen intermediates. NADPH oxidase (nicotinamide adenine dinucleotide phosphate-oxidase) produces the superoxide O2-, and iNOS utilizes L-arginine to generate NO. Both superoxide and nitric oxide are unstable molecules. However, the reaction between O2- and NO generates a much more stable and highly reactive molecule, peroxynitrite ONOO-. ONOO- is an obligatory factor for oxidative and nitrosative modification of DNA and proteins. DNA oxidative/nitrative damage is seen as modification of deoxyribonucleic acid, and it mostly occurs on guanine (G) because of its high oxidation/nitration potential relatively to cytidine, thymidine, and adenine (Figure 4). The hydroxyl at the C8 position of guanine is oxidized to generate 7,8-dihydro-8-hydroxyguanine (8-OHG), which forms 8-oxo-dihydroguanine (8-oxo-dG) or the ring-opened 2,6-diamino-5-formamido-4-hydroxy-pyrimidine (FapyG), two of the most abundant oxidative DNA adducts [150-152] (Figure 5). In order to be repaired, the oxidized guanine, 8 hydroxyguanine (8-oxo-dG), or the nitrated guanine (8-nitro-dG) must be removed by a specific DNA glycosylase and repaired as a single strand mutation by base or nucleotide excision repair (BER or NER) (for an excellent review, see ref [150]). If it is not repaired, the modified guanine will have preference to pair with an adenine during the synthesis of DNA. However, modifications on adjacent guanines of both strands, which is followed by excision of the modified nucleotides often creates double strand DNA breaks (DSBs) that will be repaired by one of the two most common DNA repair pathways, homologous recombination (HR) or the non-homologous end joining (NHEJ) [153-155]. If the type of DNA damage occurs during the DNA replication in the S-phase of the cell cycle, the DNA will be faithfully repaired by the HR DNA repair pathway. If DNA damage occurs at other phases of the cell cycle (G1 or G2), the DNA repair will be directed by the error-prone NHEJ pathway and will create permanent deletions in the genome [156,157]. Thus, this type of DNA damage that constitutes a danger to the affected cell is totally arbitrary [156].

leukemia-deoxyribonucleic-ribonucleic

Figure 5: NO and its derivative products oxidize and nitrate DNA and cause DNA modifications that are at the origin of DNA double strand breaks. Although all deoxyribonucleic or ribonucleic acids can be oxidized or nitrated, guanine (G) has the highest oxidation/ nitration potential. Two of the most abundant oxidative DNA adducts are the 7,8-dihydro-8-hydroxyguanine (8-oxo-dG), and the 8-nitro-2'-deoxyguanosine (8 -nitro -dG ), which are generated by oxidation and nitration of the hydroxyl at the C8 position of guanine, respectively.

In chronic inflammatory models, we can imagine a scenario in which continuous mutations by oxidation/nitration are randomly created until a combination of selectively stable mutations have initiated the immortalization process [157]. Another set of mutations are probably required to switch on the transformation process. Consequently, the incidence of a given malignancy is random and it depends on the frequency of mutations incurred during an oxidative exposition. Here, it is important to analyze the mutation signature of different cancer models to understand this phenomenon, in which lung carcinoma and melanoma are incurring the highest rate of mutations because they are exposed to the maximum oxidation and ionization factors [158]. In fact, the oxidative/nitrosative damage is a continuous process that does not stop once transformation of a clone of cells is switched on. This is perceived by the presence of heterogeneous clones in cancers [159] in vivo, and by continuous expression of oxidative agents in transformed cell lines in vitro [87].

iNOS/NO and S-nitrosylation of Proteins

NO is also a major source for protein S-nitrosylation, a covalent modification of cysteine thiol. The list of proteins modified by S-nitrosylation is currently gaining more attention [160] because emerging data has shown a role of S-nitrosylation in multiple pathways important for tumorigenesis [62,161-165]. NO, specifically generated by iNOS, is extremely important for S-nitrosylation, and its resultant signaling pathways. Experimental data elucidated a correlation between iNOS-mediated nitrosylation and an aggressive tumor phenotype for breast cancer, lung cancer, colon cancer and prostate cancer [166]. The small GTPase Ras, the protein kinase AKT, and the phosphotase and tensin homolog PTEN are all examples of proteins involved in proliferation, cell survival and tumorigenesis. S-nitrosylation was found to be a major player in their respective functions.

Modification of a single conserved cysteine residue in the small GTPase Ras (Cys118 in human H-Ras) was one of the earliest described targets of S-nitrosylation [166]. This modification stimulates guanine nucleotide exchange and downstream pathways, including activation of mitogen-activated protein kinase signaling (MAPK). Recent findings showed that iNOS expression promotes tumorigenesis in ER-negative breast cancer by a mechanism in which NO induces S-nitrosylation of wild-type Ras, leading to phosphorylation and activation of the transcription factor Ets-1 through the Ras/MEK/ERK pathway [166]. Interestingly, the Ras protein has been shown to be involved in the survival of HTLV-1 infected cells [167,168]. Akt is another multifunctional regulatory protein that is also involved in cellular metabolism, proliferation and survival. Recent discoveries showed that Akt kinase activity was augmented when it was nitrosylated by iNOS at Cysteine 224 [169,170]. Why iNOS differs from other NOS isoforms is still unclear. However, it is likely to be linked to a specific stimulus that only targets iNOS activation. PTEN, one of the main phosphotases regulating Akt dephosphorylation, is selectively S-nitrosylated by low concentrations of NO at a speci?c cysteine residue (Cys-83). S-nitrosylation of PTEN inhibits its activity and stimulates Akt activation [170] (Figure 3).

The dysfunction of Akt and PTEN was extensively studied in HTLV-1 infected cells [137,171-173]. Whether S-nitrosylation of these proteins has an effect on HTLV-1 induced leukemogenesis was not investigated. In the light of the new published data, it is important to characterize the S-nitrosylation modification of Ras, Akt and PTEN in those cells, and to functionally test the influence of S-nitrosylation on the signaling pathway of Ras, Akt and PTEN. Inhibition of NO production by inhibitors, or by shRNA, the use of Tax mutants defective in iNOS activation, or the use of Ras C118S, Akt C224S or PTEN C83S mutants should be investigated to determine the role of S-nitrosylation on the oncogenic activities of these proteins in HTLV-1 infected cells.

iNOS/NO as a Marker for Diagnosis and Treatment

Nitric oxide is an important cellular signaling molecule involved in many physiological and pathological processes. The NO synthesized in the endothelial and neuronal tissues induces vasodilation and neurotransmission actions, respectively. However, NO generated by iNOS in the immune cells has a central role in fighting infections, and it can be extremely mutagenic in chronically infected cells. The genetic alterations induced by NO are important requirements for induction of malignancy. iNOS-produced NO seems to play a critical role in cancer development because it was detected in various cancers and inhibitors targeting iNOS in animal models dramatically reduced tumorigenesis. Thus, iNOS/NO signaling can be considered as a novel and potential therapeutic target and iNOS/NO measurement can serve as useful assays in providing diagnostic for potential malignancies. Further experimentation on iNOS/NO signaling is required in order to develop new strategies for cancer prevention and treatment.

References

  1. Gessain A, Cassar O (2012) Epidemiological Aspects and World Distribution of HTLV-1 Infection. Front Microbiol 3: 388.
  2. European Centre for Disease Prevention and Control (2015) Geographical distribution of areas with a high prevalence of HTLV-1 infection. ECDC Technical Report.
  3. Michie CA, Gilmour J (2001) Breast feeding and the risks of viral transmission. Arch Dis Child 84: 381-382.
  4. Pais-Correia AM, Sachse M, Guadagnini S, Robbiati V, Lasserre R, et al. (2010) Biofilm-like extracellular viral assemblies mediate HTLV-1 cell-to-cell transmission at virological synapses. Nat Med 16: 83-89.
  5. Pique C, Jones KS (2012) Pathways of cell-cell transmission of HTLV-1. Front Microbiol 3: 378.
  6. Bangham CR, Ratner L (2015) How does HTLV-1 cause adult T-cell leukaemia/lymphoma (ATL) Curr Opin Virol 14: 93-100.
  7. Cook LB, Melamed A, Niederer H, Valganon M, Laydon D, et al. (2014) The role of HTLV-1 clonality, proviral structure, and genomic integration site in adult T-cell leukemia/lymphoma. Blood 123: 3925-3931.
  8. Melamed A, Laydon DJ, Gillet NA, Tanaka Y, Taylor GP, et al. (2013) Genome-wide determinants of proviral targeting, clonal abundance and expression in natural HTLV-1 infection. PLoS Pathog 9: e1003271.
  9. Feuer G, Fraser JK, Zack JA, Lee F, Feuer R, et al. (1996) Human T-cell leukemia virus infection of human hematopoietic progenitor cells: maintenance of virus infection during differentiation in vitro and in vivo. J Virol 70: 4038-4044.
  10. Matsuoka M, Jeang KT (2007) Human T-cell leukaemia virus type 1 (HTLV-1) infectivity and cellular transformation. Nat Rev Cancer 7: 270-280.
  11. Banerjee P, Tripp A, Lairmore MD, Crawford L, Sieburg M, et al. (2010) Adult T-cell leukemia/lymphoma development in HTLV-1-infected humanized SCID mice. Blood 115: 2640-2648.
  12. Villaudy J, Wencker M, Gadot N, Gillet NA, Scoazec JY, et al. (2011) HTLV-1 propels thymic human T cell development in "human immune system" Rag2⁻/⁻ gamma c⁻/⁻ mice. PLoS Pathog 7: e1002231.
  13. Nagai Y, Kawahara M, Hishizawa M, Shimazu Y, Sugino N, et al. (2015) T memory stem cells are the hierarchical apex of adult T-cell leukemia. Blood 125: 3527-3535.
  14. Gattinoni L (2015) The dark side of T memory stem cells. Blood 125: 3519-3520.
  15. Rowan AG, Bangham CR (2012) Is There a Role for HTLV-1-Specific CTL in Adult T-Cell Leukemia/Lymphoma? Leuk Res Treatment 2012: 391953.
  16. Yasunaga J, Matsuoka M (2011) Molecular mechanisms of HTLV-1 infection and pathogenesis. Int J Hematol 94: 435-442.
  17. Colotta F, Allavena P, Sica A, Garlanda C, Mantovani A (2009) Cancer-related inflammation, the seventh hallmark of cancer: links to genetic instability. Carcinogenesis 30: 1073-1081.
  18. Lu H, Ouyang W, Huang C (2006) Inflammation, a key event in cancer development. Mol Cancer Res 4: 221-233.
  19. Grivennikov SI, Greten FR, Karin M (2010) Immunity, inflammation, and cancer. Cell 140: 883-899.
  20. Crusz SM, Balkwill FR (2015) Inflammation and cancer: advances and new agents. Nat Rev Clin Oncol 12: 584-596.
  21. Elinav E, Nowarski R, Thaiss CA, Hu B, Jin C, et al. (2013) Inflammation-induced cancer: crosstalk between tumours, immune cells and microorganisms. Nat Rev Cancer 13: 759-771.
  22. Trinchieri G (2012) Cancer and inflammation: an old intuition with rapidly evolving new concepts. Annu Rev Immunol 30: 677-706.
  23. Rauch DA, Ratner L (2011) Targeting HTLV-1 activation of NFκB in mouse models and ATLL patients. Viruses 3: 886-900.
  24. Sethi G, Sung B, Aggarwal BB (2008) Nuclear factor-kappaB activation: from bench to bedside. Exp Biol Med (Maywood) 233: 21-31.
  25. Shalapour S, Karin M (2015) Immunity, inflammation, and cancer: an eternal fight between good and evil. J Clin Invest 125: 3347-3355.
  26. DiDonato JA, Mercurio F, Karin M (2012) NF-κB and the link between inflammation and cancer. Immunol Rev 246: 379-400.
  27. Ben-Neriah Y, Karin M (2011) Inflammation meets cancer, with NF-κB as the matchmaker. Nat Immunol 12: 715-723.
  28. Terzić J, Grivennikov S, Karin E, Karin M (2010) Inflammation and colon cancer. Gastroenterology 138: 2101-2114.
  29. Gukovsky I, Li N, Todoric J, Gukovskaya A, Karin M (2013) Inflammation, autophagy, and obesity: common features in the pathogenesis of pancreatitis and pancreatic cancer. Gastroenterology 144: 1199-1209.
  30. Hayden MS, Ghosh S (2012) NF-κB, the first quarter-century: remarkable progress and outstanding questions. Genes Dev 26: 203-234.
  31. Napetschnig J, Wu H (2013) Molecular basis of NF-κB signaling. Annu Rev Biophys 42: 443-468.
  32. Scheiermann C, Kunisaki Y, Frenette PS (2013) Circadian control of the immune system. Nat Rev Immunol 13: 190-198.
  33. Shachar I, Karin N (2013) The dual roles of inflammatory cytokines and chemokines in the regulation of autoimmune diseases and their clinical implications. J Leukoc Biol 93: 51-61.
  34. Grimm EA, Sikora AG, Ekmekcioglu S (2013) Molecular Pathways: Inflammation-associated nitric-oxide production as a cancer-supporting redox mechanism and a potential therapeutic target. Clin Cancer Res 19: 5557-5563.
  35. Muntané J, la Mata MD (2010) Nitric oxide and cancer. World J Hepatol 2: 337-344.
  36. Jabłońska E, Puzewska W, Marcińczyk M, Grabowska Z, Jabłoński J (2005) iNOS expression and NO production by neutrophils in cancer patients. Arch Immunol Ther Exp (Warsz) 53: 175-179.
  37. Loibl S, Buck A, Strank C, von Minckwitz G, Roller M, et al. (2005) The role of early expression of inducible nitric oxide synthase in human breast cancer. Eur J Cancer 41: 265-271.
  38. Masri FA, Comhair SA, Koeck T, Xu W, Janocha A, et al. (2005) Abnormalities in nitric oxide and its derivatives in lung cancer. Am J Respir Crit Care Med 172: 597-605.
  39. Lechner M, Lirk P, Rieder J (2005) Inducible nitric oxide synthase (iNOS) in tumor biology: the two sides of the same coin. Semin Cancer Biol 15: 277-289.
  40. Gochman E, Mahajna J, Shenzer P, Dahan A, Blatt A, et al. (2012) The expression of iNOS and nitrotyrosine in colitis and colon cancer in humans. Acta Histochem 114: 827-835.
  41. Burke AJ, Sullivan FJ, Giles FJ, Glynn SA (2013) The yin and yang of nitric oxide in cancer progression. Carcinogenesis 34: 503-512.
  42. Zhang W, He XJ, Ma YY, Wang HJ, Xia YJ, et al. (2011) Inducible nitric oxide synthase expression correlates with angiogenesis, lymphangiogenesis, and poor prognosis in gastric cancer patients. Hum Pathol 42: 1275-1282.
  43. Rafiei A, Hosseini V, Janbabai G, Fazli B, Ajami A, et al. (2012) Inducible nitric oxide synthetase genotype and Helicobacter pylori infection affect gastric cancer risk. World J Gastroenterol 18: 4917-4924.
  44. Sandes EO, Lodillinsky C, Langle Y, Belgorosky D, Marino L, et al. (2012) Inducible nitric oxide synthase and PPARγ are involved in bladder cancer progression. J Urol 188: 967-973.
  45. Huang FY, Chan AO, Rashid A, Wong DK, Cho CH, et al. (2012) Helicobacter pylori induces promoter methylation of E-cadherin via interleukin-1β activation of nitric oxide production in gastric cancer cells. Cancer 118: 4969-4980.
  46. Bian K, Murad F (2014) What is next in nitric oxide research? From cardiovascular system to cancer biology. Nitric Oxide 43: 3-7.
  47. Srinivas P, Wink D, Mohanakumar KP, Pillai MR (2014) The legacy of nitric oxide: impact on disease biology. Nitric Oxide 43: 1-2.
  48. Kim YI, Park SW, Kang IJ, Shin MK, Lee MH (2015) Activin suppresses LPS-induced Toll-like receptor, cytokine and inducible nitric oxide synthase expression in normal human melanocytes by inhibiting NF-kappa B and MAPK pathway activation. International Journal of Molecular Medicine 36: 1165-1172.
  49. Goldstein S, Squadrito GL, Pryor WA, Czapski G (1996) Direct and indirect oxidations by peroxynitrite, neither involving the hydroxyl radical. Free Radic Biol Med 21: 965-974.
  50. Kissner R, Nauser T, Bugnon P, Lye PG, Koppenol WH (1998) Formation and properties of peroxynitrite as studied by laser flash photolysis, high-pressure stopped-flow technique, and pulse radiolysis. Chem Res Toxicol 11: 557.
  51. Mungrue IN, Husain M, Stewart DJ (2002) The role of NOS in heart failure: lessons from murine genetic models. Heart Fail Rev 7: 407-422.
  52. Mungrue IN, Gros R, You X, Pirani A, Azad A, et al. (2002) Cardiomyocyte overexpression of iNOS in mice results in peroxynitrite generation, heart block, and sudden death. J Clin Invest 109: 735-743.
  53. Janakiram NB, Rao CV (2012) iNOS-selective inhibitors for cancer prevention: promise and progress. Future Med Chem 4: 2193-2204.
  54. Janakiram NB, Rao CV (2012) Chemoprevention of Colon Cancer by iNOS-Selective Inhibitors. For Immunopathol Dis Therap 3: 155-167.
  55. Anttila MA, Voutilainen K, Merivalo S, Saarikoski S, Kosma VM (2007) Prognostic significance of iNOS in epithelial ovarian cancer. Gynecol Oncol 105: 97-103.
  56. Harada K, Supriatno, Kawaguchi S, Tomitaro O, Yoshida H, et al. (2004) Overexpression of iNOS gene suppresses the tumorigenicity and metastasis of oral cancer cells. In vivo 18: 449-455.
  57. McCarthy HO, Zholobenko AV, Wang Y, Canine B, Robson T, et al. (2011) Evaluation of a multi-functional nanocarrier for targeted breast cancer iNOS gene therapy. Int J Pharm 405: 196-202.
  58. Singh S, Gupta AK (2011) Nitric oxide: role in tumour biology and iNOS/NO-based anticancer therapies. Cancer Chemother Pharmacol 67: 1211-1224.
  59. Kang J, Zhang Y, Cao X, Fan J, Li G, et al. (2012) Lycorine inhibits lipopolysaccharide-induced iNOS and COX-2 up-regulation in RAW264.7 cells through suppressing P38 and STATs activation and increases the survival rate of mice after LPS challenge. Int Immunopharmacol 12: 249-256.
  60. Oh WJ, Jung U, Eom HS, Shin HJ, Park HR (2013) Inhibition of lipopolysaccharide-induced proinflammatory responses by Buddleja officinalis extract in BV-2 microglial cells via negative regulation of NF-kB and ERK1/2 signaling. Molecules 18: 9195-9206.
  61. Li DY, Xue MY, Geng ZR, Chen PY (2012) The suppressive effects of Bursopentine (BP5) on oxidative stress and NF-ĸB activation in lipopolysaccharide-activated murine peritoneal macrophages. Cell Physiol Biochem 29: 9-20.
  62. Kelleher ZT, Potts EN, Brahmajothi MV, Foster MW, Auten RL, et al. (2011) NOS2 regulation of LPS-induced airway inflammation via S-nitrosylation of NF-{kappa}B p65. Am J Physiol Lung Cell Mol Physiol 301: 327-333.
  63. Kim SJ, Ha MS, Choi EY, Choi JI, Choi IS (2005) Nitric oxide production and inducible nitric oxide synthase expression induced by Prevotella nigrescens lipopolysaccharide. FEMS Immunol Med Microbiol 43: 51-58.
  64. Harmey JH, Bucana CD, Lu W, Byrne AM, McDonnell S, et al. (2002) Lipopolysaccharide-induced metastatic growth is associated with increased angiogenesis, vascular permeability and tumor cell invasion. Int J Cancer 101: 415-422.
  65. Pidgeon GP, Harmey JH, Kay E, Da Costa M, Redmond HP, et al. (1999) The role of endotoxin/lipopolysaccharide in surgically induced tumour growth in a murine model of metastatic disease. Br J Cancer 81: 1311-1317.
  66. Machado-Silva W, Alfinito-Kreis R, Carvalho LSF, Quinaglia-e-Silva JC, Almeida OLR, et al. (2015) Endothelial nitric oxide synthase genotypes modulate peripheral vasodilatory properties after myocardial infarction. Gene 568: 165-169.
  67. Gambaryan S, Tsikas D (2015) A review and discussion of platelet nitric oxide and nitric oxide synthase: do blood platelets produce nitric oxide from L-arginine or nitrite? Amino Acids 47: 1779-1793.
  68. Shen JQ, Yang QL, Xue Y, Cheng XB, Jiang ZH, et al. (2015) Inducible nitric oxide synthase response and associated cytokine gene expression in the spleen of mice infected with Clonorchis sinensis. Parasitol Res 114: 1661-1670.
  69. Förstermann U, Sessa WC (2012) Nitric oxide synthases: regulation and function. Eur Heart J 33: 829-837, 837a-837d.
  70. Liu Q, Inoue H, Mahendran R (2008) Transcriptional regulation of the COX-2 expression by nitric oxide in colon cancer cell lines. Oncol Rep 19: 269-274.
  71. Pautz A, Art J, Hahn S, Nowag S, Voss C, et al. (2010) Regulation of the expression of inducible nitric oxide synthase. Nitric Oxide 23: 75-93.
  72. Kleinert H, Pautz A, Linker K, Schwarz PM (2004) Regulation of the expression of inducible nitric oxide synthase. Eur J Pharmacol 500: 255-266.
  73. Maier E, Mittermeir M, Ess S, Neuper T, Schmiedlechner A, et al. (2015) Prerequisites for Functional Interleukin 31 Signaling and Its Feedback Regulation by Suppressor of Cytokine Signaling 3 (SOCS3). J Biol Chem 290: 24747-24759.
  74. Lechner M, Rieder J, Tilg H (2007) Helicobacter pylori infection, iNOS, and gastric cancer: the impact of another possible link. J Surg Oncol 95: 271-272.
  75. Rabender CS, Alam A, Sundaresan G, Cardnell RJ, Yakovlev VA, et al. (2015) The Role of Nitric Oxide Synthase Uncoupling in Tumor Progression. Mol Cancer Res 13: 1034-1043.
  76. Campbell MG, Smith BC, Potter CS, Carragher B, Marletta MA (2014) Molecular architecture of mammalian nitric oxide synthases. Proc Natl Acad Sci U S A 111: E3614-3623.
  77. Bogdan C (2015) Nitric oxide synthase in innate and adaptive immunity: an update. Trends Immunol 36: 161-178.
  78. Rahat MA, Hemmerlein B (2013) Macrophage-tumor cell interactions regulate the function of nitric oxide. Front Physiol 4: 144.
  79. Boscá L, Zeini M, Través PG, Hortelano S (2005) Nitric oxide and cell viability in inflammatory cells: a role for NO in macrophage function and fate. Toxicology 208: 249-258.
  80. Schmidt N, Pautz A, Art J, Rauschkolb P, Jung M, et al. (2010) Transcriptional and post-transcriptional regulation of iNOS expression in human chondrocytes. Biochem Pharmacol 79: 722-732.
  81. Ratajczak-Wrona W, Jablonska E, Garley M, Jablonski J, Radziwon P, et al. (2013) Role of AP-1 family proteins in regulation of inducible nitric oxide synthase (iNOS) in human neutrophils. J Immunotoxicol 10: 32-39.
  82. Shelton JL, Wang L, Cepinskas G, Sandig M, Scott JA, et al. (2007) Inducible NO synthase (iNOS) in human neutrophils but not pulmonary microvascular endothelial cells (PMVEC) mediates septic protein leak in vitro. Microvasc Res 74: 23-31.
  83. Patel JD, Krupka T, Anderson JM (2007) iNOS-mediated generation of reactive oxygen and nitrogen species by biomaterial-adherent neutrophils. J Biomed Mater Res A 80: 381-390.
  84. Nordone SK, Gookin JL (2010) Lymphocytes and not IFN-gamma mediate expression of iNOS by intestinal epithelium in murine cryptosporidiosis. Parasitol Res 106: 1507-1511.
  85. Hettich C, Wilker S, Mentlein R, Lucius R, Roider J, et al. (2014) The retinal pigment epithelium (RPE) induces FasL and reduces iNOS and Cox2 in primary monocytes. Graefes Arch Clin Exp Ophthalmol 252: 1747-1754.
  86. Jin C, Guo J, Qiu X, Ma K, Xiang M, et al. (2014) IGF-1 induces iNOS expression via the p38 MAPK signal pathway in the anti-apoptotic process in pulmonary artery smooth muscle cells during PAH. J Recept Signal Transduct Res 34: 325-331.
  87. Baydoun HH, Cherian MA, Green P, Ratner L (2015) Inducible nitric oxide synthase mediates DNA double strand breaks in Human T-Cell Leukemia Virus Type 1-induced leukemia/lymphoma. Retrovirology 12: 71.
  88. Jayaraman P, Alfarano MG, Svider PF, Parikh F, Lu G, et al. (2014) iNOS expression in CD4+ T cells limits Treg induction by repressing TGFÎ: combined iNOS inhibition and Treg depletion unmask endogenous antitumor immunity. Clin Cancer Res 20: 6439-6451.
  89. Jianjun Yang, Zhang R, Lu G, Shen Y, Peng L, et al. (2013) T cell–derived inducible nitric oxide synthase switches off Th17 cell differentiation. J Exp Med 210: 1447-1462.
  90. Isayama K, Murao Y, Saito F, Hirakawa A, Nakatani T (2012) Effects of hypertonic saline on CD4+CD25+Foxp3+ regulatory T cells after hemorrhagic shock in relation to iNOS and cytokines. J Surg Res 172: 137-145.
  91. Chong SZ, Wong KL, Lin G, Yang CM, Wong SC, et al. (2011) Human CD8⁺ T cells drive Th1 responses through the differentiation of TNF/iNOS-producing dendritic cells. Eur J Immunol 41: 1639-1651.
  92. Yi B, Ozerova M, Zhang GX, Yan GJ, Huang SD, et al. (2015) Post-Transcriptional Regulation of Endothelial Nitric Oxide Synthase Expression by Polypyrimidine Tract-Binding Protein 1. Arteriosclerosis Thrombosis and Vascular Biology 35: 2153-2160.
  93. Korhonen R, Linker K, Pautz A, Förstermann U, Moilanen E, et al. (2007) Post-transcriptional regulation of human inducible nitric-oxide synthase expression by the Jun N-terminal kinase. Mol Pharmacol 71: 1427-1434.
  94. Lamas S, Lowenstein CJ, Michel T (2007) Nitric oxide signaling comes of age: 20 years and thriving. Cardiovasc Res 75: 207-209.
  95. Lorsbach RB, Murphy WJ, Lowenstein CJ, Snyder SH, Russell SW (1993) Expression of the nitric oxide synthase gene in mouse macrophages activated for tumor cell killing. Molecular basis for the synergy between interferon-gamma and lipopolysaccharide. J Biol Chem 268: 1908-1913.
  96. Du Q, Park KS, Guo Z, He P, Nagashima M, et al. (2006) Regulation of human nitric oxide synthase 2 expression by Wnt beta-catenin signaling. Cancer Res 66: 7024-7031.
  97. Taylor BS, Geller DA (2000) Molecular regulation of the human inducible nitric oxide synthase (iNOS) gene. Shock 13: 413-424.
  98. Geller DA, Billiar TR (1998) Molecular biology of nitric oxide synthases. Cancer Metastasis Rev 17: 7-23.
  99. Geller DA, Lowenstein CJ, Shapiro RA, Nussler AK, Di Silvio M, et al. (1993) Molecular cloning and expression of inducible nitric oxide synthase from human hepatocytes. Proc Natl Acad Sci U S A 90: 3491-3495.
  100. Xie QW, Whisnant R, Nathan C (1993) Promoter of the mouse gene encoding calcium-independent nitric oxide synthase confers inducibility by interferon gamma and bacterial lipopolysaccharide. J Exp Med 177: 1779-1784.
  101. Xie QW, Cho HJ, Calaycay J, Mumford RA, Swiderek KM, et al. (1992) Cloning and characterization of inducible nitric oxide synthase from mouse macrophages. Science 256: 225-228.
  102. Wolf G (1997) Nitric oxide and nitric oxide synthase: biology, pathology, localization. Histol Histopathol 12: 251-261.
  103. Cohen J, Evans TJ, Spink J (1998) Cytokine regulation of inducible nitric oxide synthase in vascular smooth muscle cells. Prog Clin Biol Res 397: 169-177.
  104. Zhang X, Laubach VE, Alley EW, Edwards KA, Sherman PA, et al. (1996) Transcriptional basis for hyporesponsiveness of the human inducible nitric oxide synthase gene to lipopolysaccharide/interferon-gamma. J Leukoc Biol 59: 575-585.
  105. Alley EW, Murphy WJ, Russell SW (1995) A classical enhancer element responsive to both lipopolysaccharide and interferon-gamma augments induction of the iNOS gene in mouse macrophages. Gene 158: 247-251.
  106. Kristof AS, Marks-Konczalik J, Moss J (2001) Mitogen-activated protein kinases mediate activator protein-1-dependent human inducible nitric-oxide synthase promoter activation. J Biol Chem 276: 8445-8452.
  107. Chu SC, Marks-Konczalik J, Wu HP, Banks TC, Moss J (1998) Analysis of the cytokine-stimulated human inducible nitric oxide synthase (iNOS) gene: characterization of differences between human and mouse iNOS promoters. Biochem Biophys Res Commun 248: 871-878.
  108. Mashmoushi AK, Oates JC (2015) Lipopolysaccharide induces inducible nitric oxide synthase-dependent podocyte dysfunction via a hypoxia-inducible factor 1alpha and cell division control protein 42 and Ras-related C3 botulinum toxin substrate 1 pathway. Free Radic Biol Med 84: 185-195.
  109. Taylor BS, Alarcon LH, Billiar TR (1998) Inducible nitric oxide synthase in the liver: regulation and function. Biochemistry (Mosc) 63: 766-781.
  110. Ziesché E, Bachmann M, Kleinert H, Pfeilschifter J, Mühl H (2007) The interleukin-22/STAT3 pathway potentiates expression of inducible nitric-oxide synthase in human colon carcinoma cells. J Biol Chem 282: 16006-16015.
  111. Dawn B, Xuan YT, Guo Y, Rezazadeh A, Stein AB, et al. (2004) IL-6 plays an obligatory role in late preconditioning via JAK-STAT signaling and upregulation of iNOS and COX-2. Cardiovasc Res 64: 61-71.
  112. Shao L, Guo Z, Geller DA (2007) Transcriptional suppression of cytokine-induced iNOS gene expression by IL-13 through IRF-1/ISRE signaling. Biochem Biophys Res Commun 362: 582-586.
  113. Guo Z, Shao L, Du Q, Park KS, Geller DA (2007) Identification of a classic cytokine-induced enhancer upstream in the human iNOS promoter. FASEB J 21: 535-542.
  114. Saha RN, Jana M, Pahan K (2007) MAPK p38 regulates transcriptional activity of NF-kappaB in primary human astrocytes via acetylation of p65. J Immunol 179: 7101-7109.
  115. Ganster RW, Taylor BS, Shao L, Geller DA (2001) Complex regulation of human inducible nitric oxide synthase gene transcription by Stat 1 and NF-kappa B. Proc Natl Acad Sci U S A 98: 8638-8643.
  116. Harbrecht BG, Taylor BS, Xu Z, Ramalakshmi S, Ganster RW, et al. (2001) cAMP inhibits inducible nitric oxide synthase expression and NF-kappaB-binding activity in cultured rat hepatocytes. J Surg Res 99: 258-264.
  117. Vila-del Sol V, Díaz-Muñoz MD, Fresno M (2007) Requirement of tumor necrosis factor alpha and nuclear factor-kappaB in the induction by IFN-gamma of inducible nitric oxide synthase in macrophages. J Leukoc Biol 81: 272-283.
  118. Bhattacharya I, Dominguez AP, Dragert K, Humar R, Haas E, er al. (2015) Hypoxia potentiates tumor necrosis factor-alpha induced expression of inducible nitric oxide synthase and cyclooxygenase-2 in white and brown adipocytes. Biochemical and Biophysical Research Communications 461: 287-292.
  119. Kojima M, Morisaki T, Tsukahara Y, Uchiyama A, Matsunari Y, et al. (1999) Nitric oxide synthase expression and nitric oxide production in human colon carcinoma tissue. J Surg Oncol 70: 222-229.
  120. Vakkala M, Kahlos K, Lakari E, Pääkkö P, Kinnula V, et al. (2000) Inducible nitric oxide synthase expression, apoptosis, and angiogenesis in in situ and invasive breast carcinomas. Clin Cancer Res 6: 2408-2416.
  121. De Paepe B, Verstraeten VM, De Potter CR, Bullock GR (2002) Increased angiotensin II type-2 receptor density in hyperplasia, DCIS and invasive carcinoma of the breast is paralleled with increased iNOS expression. Histochem Cell Biol 117: 13-19.
  122. Aaltoma SH, Lipponen PK, Kosma VM (2001) Inducible nitric oxide synthase (iNOS) expression and its prognostic value in prostate cancer. Anticancer Res 21: 3101-3106.
  123. Swana HS, Smith SD, Perrotta PL, Saito N, Wheeler MA, et al. (1999) Inducible nitric oxide synthase with transitional cell carcinoma of the bladder. J Urol 161: 630-634.
  124. He H, Feng YS, Zang LH, Liu WW, Ding LQ, et al. (2014) Nitric oxide induces apoptosis and autophagy; autophagy down-regulates NO synthesis in physalin A-treated A375-S2 human melanoma cells. Food Chem Toxicol 71: 128-135.
  125. Shigyo H, Nonaka S, Katada A, Bandoh N, Ogino T, et al. (2007) Inducible nitric oxide synthase expression in various laryngeal lesions in relation to carcinogenesis, angiogenesis, and patients' prognosis. Acta Otolaryngol 127: 970-979.
  126. Igawa S, Hayashi I, Tanaka N, Hiruma H, Majima M, et al. (2004) Nitric oxide generated by iNOS reduces deformability of Lewis lung carcinoma cells. Cancer Sci 95: 342-347.
  127. Brennan PA, Dennis S, Poller D, Quintero M, Puxeddu R, et al. (2008) Inducible nitric oxide synthase: correlation with extracapsular spread and enhancement of tumor cell invasion in head and neck squamous cell carcinoma. Head Neck 30: 208-214.
  128. Brennan PA, Sharma S, Bowden JR, Umar T (2003) Expression of inducible nitric oxide synthase in bone metastases. Eur J Surg Oncol 29: 619-623.
  129. Brennan PA, Thomas GJ, Langdon JD (2003) The role of nitric oxide in oral diseases. Arch Oral Biol 48: 93-100.
  130. Brennan PA (2003) Nitric oxide and squamous carcinoma. J Oral Maxillofac Surg 61: 277.
  131. Brennan PA, Umar T, Wilson AW, Mellor TK (2002) Expression of type 2 nitric oxide synthase and vascular endothelial growth factor in oral dysplasia. J Oral Maxillofac Surg 60: 1455-1460.
  132. Brennan PA, Umar T, Smith GI, McCauley P, Peters WJ, et al. (2002) Expression of type 2 nitric oxide synthase and p53 in Warthin's tumour of the parotid. J Oral Pathol Med 31: 458-462.
  133. Brennan PA, Palacios-Callender M, Umar T, Tant S, Langdon JD (2002) Expression of type 2 nitric oxide synthase and p21 in oral squamous cell carcinoma. Int J Oral Maxillofac Surg 31: 200-205.
  134. Brennan PA, Umar T, Smith GI, Lo CH, Tant S (2002) Expression of nitric oxide synthase-2 in cutaneous squamous cell carcinoma of the head and neck. Br J Oral Maxillofac Surg 40: 191-194.
  135. Brennan PA, Moncada S (2002) From pollutant gas to biological messenger: the diverse actions of nitric oxide in cancer. Ann R Coll Surg Engl 84: 75-78.
  136. Matsuoka M, Jeang KT (2011) Human T-cell leukemia virus type 1 (HTLV-1) and leukemic transformation: viral infectivity, Tax, HBZ and therapy. Oncogene 30: 1379-1389.
  137. Bellon M, Nicot C (2008) Central role of PI3K in transcriptional activation of hTERT in HTLV-I-infected cells. Blood 112: 2946-2955.
  138. Bellon M, Baydoun HH, Yao Y, Nicot C (2010) HTLV-I Tax-dependent and -independent events associated with immortalization of human primary T lymphocytes. Blood 115: 2441-2448.
  139. Giam CZ, Jeang KT (2007) HTLV-1 Tax and adult T-cell leukemia. Front Biosci 12: 1496-1507.
  140. Grossman WJ, Ratner L (1996) Transgenic mouse models for HTLV-I infection. J Acquir Immune Defic Syndr Hum Retrovirol 13 Suppl 1: S162-169.
  141. Yamagishi M, Nakano K, Miyake A, Yamochi T, Kagami Y, et al. (2012) Polycomb-mediated loss of miR-31 activates NIK-dependent NF-κB pathway in adult T cell leukemia and other cancers. Cancer Cell 21: 121-135.
  142. Grossman WJ, Kimata JT, Wong FH, Zutter M, Ley TJ, et al. (1995) Development of leukemia in mice transgenic for the tax gene of human T-cell leukemia virus type I. Proc Natl Acad Sci USA 92: 1057-1061.
  143. Mori N, Nunokawa Y, Yamada Y, Ikeda S, Tomonaga M, et al. (1999) 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 94: 2862-2870.
  144. Sonoki T, Matsuzaki H, Nagasaki A, Hata H, Yoshida M, et al. (1999) 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 13: 713-718.
  145. Stempelj M, Kedinger M, Augenlicht L, Klampfer L (2007) Essential role of the JAK/STAT1 signaling pathway in the expression of inducible nitric-oxide synthase in intestinal epithelial cells and its regulation by butyrate. J Biol Chem 282: 9797-9804.
  146. Do H, Pyo S, Sohn EH (2010) Suppression of iNOS expression by fucoidan is mediated by regulation of p38 MAPK, JAK/STAT, AP-1 and IRF-, and depends on up-regulation of scavenger receptor B1 expression in TNF-alpha- and IFN-gamma-stimulated C6 glioma cells. J Nutr Biochem 21: 671-679.
  147. Bachmaier K, Neu N, Pummerer C, Duncan GS, Mak TW, et al. (1997) iNOS expression and nitrotyrosine formation in the myocardium in response to inflammation is controlled by the interferon regulatory transcription factor 1. Circulation 96: 585-591.
  148. Zhang M, Mathews Griner LA, Ju W, Duveau DY, Guha R, et al. (2015) Selective targeting of JAK/STAT signaling is potentiated by Bcl-xL blockade in IL-2-dependent adult T-cell leukemia. Proc Natl Acad Sci U S A 112: 12480-12485.
  149. Brooks SC, Adhikary S, Rubinson EH, Eichman BF (2013) Recent advances in the structural mechanisms of DNA glycosylases. Biochim Biophys Acta 1834: 247-271.
  150. Evans MD, Dizdaroglu M, Cooke MS (2004) Oxidative DNA damage and disease: induction, repair and significance. Mutat Res 567: 1-61.
  151. Burrows CJ, Muller JG (1998) Oxidative Nucleobase Modifications Leading to Strand Scission. Chem Rev 98: 1109-1152.
  152. Chapman JR, Taylor MR, Boulton SJ (2012) Playing the end game: DNA double-strand break repair pathway choice. Mol Cell 47: 497-510.
  153. Brandsma I, Gent DC (2012) Pathway choice in DNA double strand break repair: observations of a balancing act. Genome Integr 3: 9.
  154. Mateos-Gomez PA, Gong F, Nair N, Miller KM, Lazzerini-Denchi E, et al. (2015) Mammalian polymerase θ promotes alternative NHEJ and suppresses recombination. Nature 518: 254-257.
  155. Bielas JH, Loeb KR, Rubin BP, True LD, Loeb LA (2006) Human cancers express a mutator phenotype. Proc Natl Acad Sci USA 103: 18238-18242.
  156. Balkwill FR, Mantovani A (2012) Cancer-related inflammation: common themes and therapeutic opportunities. Semin Cancer Biol 22: 33-40.
  157. Alexandrov LB, Nik-Zainal S, Wedge DC, Aparicio SA, Behjati S, et al. (2013) Signatures of mutational processes in human cancer. Nature 500: 415-421.
  158. Meacham CE, Morrison SJ (2013) Tumour heterogeneity and cancer cell plasticity. Nature 501: 328-337.
  159. Hess DT, Stamler JS (2012) Regulation by S-nitrosylation of protein post-translational modification. J Biol Chem 287: 4411-4418.
  160. Hess DT, Matsumoto A, Kim SO, Marshall HE, Stamler JS (2005) Protein S-nitrosylation: purview and parameters. Nat Rev Mol Cell Biol 6: 150-166.
  161. Foster MW, Hess DT, Stamler JS (2009) Protein S-nitrosylation in health and disease: a current perspective. Trends Mol Med 15: 391-404.
  162. Guan W, Sha J, Chen X, Xing Y, Yan J, et al. (2012) S-Nitrosylation of mitogen activated protein kinase phosphatase-1 suppresses radiation-induced apoptosis. Cancer Lett 314: 137-146.
  163. Marshall HE, Gow A (2012) Regulation of cellular processes by S-nitrosylation. Preface. Biochim Biophys Acta 1820: 673-674.
  164. Sha Y, Marshall HE (2012) S-nitrosylation in the regulation of gene transcription. Biochim Biophys Acta 1820: 701-711.
  165. Marshall HE, Foster MW (2012) S-nitrosylation of Ras in breast cancer. Breast Cancer Res 14: 113.
  166. Vajente N, Trevisan R, Saggioro D (2009) HTLV-1 Tax protein cooperates with Ras in protecting cells from apoptosis. Apoptosis 14: 153-163.
  167. Stoppa G, Rumiato E, Saggioro D (2012) Ras signaling contributes to survival of human T-cell leukemia/lymphoma virus type 1 (HTLV-1) Tax-positive T-cells. Apoptosis 17: 219-228.
  168. Wu M, Katta A, Gadde MK, Liu H, Kakarla SK, et al. (2009) Aging-associated dysfunction of Akt/protein kinase B: S-nitrosylation and acetaminophen intervention. PLoS One 4: e6430.
  169. Numajiri N, Takasawa K, Nishiya T, Tanaka H, Ohno K, et al. (2011) On-off system for PI3-kinase-Akt signaling through S-nitrosylation of phosphatase with sequence homology to tensin (PTEN). Proc Natl Acad Sci U S A 108: 10349-10354.
  170. Jeong SJ, Dasgupta A, Jung KJ, Um JH, Burke A, et al. (2008) PI3K/AKT inhibition induces caspase-dependent apoptosis in HTLV-1-transformed cells. Virology 370: 264-272.
  171. Saito K, Saito M, Taniura N, Okuwa T, Ohara Y (2010) Activation of the PI3K-Akt pathway by human T cell leukemia virus type 1 (HTLV-1) oncoprotein Tax increases Bcl3 expression, which is associated with enhanced growth of HTLV-1-infected T cells. Virology 403: 173-180.
  172. Jeong SJ, Pise-Masison CA, Radonovich MF, Park HU, Brady JN (2005) Activated AKT regulates NF-kappaB activation, p53 inhibition and cell survival in HTLV-1-transformed cells. Oncogene 24: 6719-6728.
Citation: Hicham HB, Lee R (2015) Contribution of Inducible Nitric Oxide Synthase to the Transformation of HTLV-1 Infected CD4+ T-Cells. J Leuk 3:199.

Copyright: © 2015 Hicham HB, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Top