Next Article in Journal
Intersociety Position Statement on the Prevention of Ophthalmia Neonatorum in Italy
Next Article in Special Issue
Current Insights into the Maturation of Epstein–Barr Virus Particles
Previous Article in Journal
A Novel View of the Diversity of Anoxygenic Phototrophic Bacteria Inhabiting the Chemocline of Meromictic Karst Lakes
Previous Article in Special Issue
The Research Progress on Immortalization of Human B Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Assessing the Clinical Implications of Epstein-Barr Virus Infection and Their Application to the Diagnosis and Treatment of Nasopharyngeal Carcinoma

by
Tomokazu Yoshizaki
*,†,
Satoru Kondo
,
Hirotomo Dochi
,
Eiji Kobayashi
,
Harue Mizokami
,
Shigetaka Komura
and
Kazuhira Endo
Division of Otolaryngology-Head and Neck Surgery, Graduate School of Medical Science, Kanazawa University, 13-1 Takaramachi, Kanazawa 920-8641, Japan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Microorganisms 2024, 12(1), 14; https://doi.org/10.3390/microorganisms12010014 (registering DOI)
Submission received: 12 October 2023 / Revised: 12 December 2023 / Accepted: 16 December 2023 / Published: 20 December 2023
(This article belongs to the Special Issue Epstein–Barr Virus Infection and Associated Diseases 2.0)

Abstract

:
Reports about the oncogenic mechanisms underlying nasopharyngeal carcinoma (NPC) have been accumulating since the discovery of Epstein-Barr virus (EBV) in NPC cells. EBV is the primary causative agent of NPC. EBV–host and tumor–immune system interactions underlie the unique representative pathology of NPC, which is an undifferentiated cancer cell with extensive lymphocyte infiltration. Recent advances in the understanding of immune evasion and checkpoints have changed the treatment of NPC in clinical settings. The main EBV genes involved in NPC are LMP1, which is the primary EBV oncogene, and BZLF1, which induces the lytic phase of EBV. These two multifunctional genes affect host cell behavior, including the tumor–immune microenvironment and EBV behavior. Latent infections, elevated concentrations of the anti-EBV antibody and plasma EBV DNA have been used as biomarkers of EBV-associated NPC. The massive infiltration of lymphocytes in the stroma suggests the immunogenic characteristics of NPC as a virus-infected tumor and, at the same time, also indicates the presence of a sophisticated immunosuppressive system within NPC tumors. In fact, immune checkpoint inhibitors have shown promise in improving the prognosis of NPC patients with recurrent and metastatic disease. However, patients with advanced NPC still require invasive treatments. Therefore, there is a pressing need to develop an effective screening system for early-stage detection of NPC in patients. Various modalities, such as nasopharyngeal cytology, cell-free DNA methylation, and deep learning-assisted nasopharyngeal endoscopy for screening and diagnosis, have been introduced. Each modality has its advantages and disadvantages. A reciprocal combination of these modalities will improve screening and early diagnosis of NPC.

1. Introduction

Human gamma herpesviruses, including the Epstein-Barr virus (EBV; human herpesvirus 4; HHV-4) and Kaposi’s sarcoma-associated herpesvirus (KSHV; human herpesvirus 8; HHV-8), have oncogenic properties. EBV and KSHV are lymphotropic viruses, but they also infect epithelial cells, and their infected cells may develop carcinomas [1]. EBV was the first human oncovirus discovered in Burkitt’s lymphoma (BL) in 1964 [2]. Kenyan patients with nasopharyngeal carcinoma (NPC) whose sera were selected as controls in a BL study showed more antibody precipitation than those with BL [3]. Four years later, EBV was identified in undifferentiated NPC cells [4]. The definite association of EBV with NPC was confirmed by the ubiquitous identification of EBV DNA in the tumor cells of NPC tissues in endemic and non-endemic areas for NPC [5]. Based on outstanding papers published in the past half-century, we introduce new insights into the relevance of the unique clinicopathological characteristics of NPC associated with EBV infection.

1.1. Histopathology of NPC and EBV Infection

EBV infection is closely associated with pathological characteristics of NPC. The World Health Organization (WHO) classification system, which is based on the grade of differentiation of tumors, is generally accepted for the pathological classification of NPC. WHO grades I, II, and III represent keratinizing, non-keratinizing-differentiated, and non-keratinizing-undifferentiated NPC, respectively. WHO II and III are considered to indicate EBV-associated NPC [6]. However, there has been a debate on the relevance of EBV in WHO I NPC diagnosis. A previous prevalent opinion was that WHO I NPC was originally an EBV-driven tumor. Throughout the tumor progression, EBV escaped or was eliminated from tumor cells. Thus, a small amount of EBV DNA was detected in WHO type I tumors. However, a recent prevalent opinion based on serological and histological studies is that WHO I NPC is a squamous cell carcinoma similar to general head and neck carcinoma.
EBV-associated WHO II and III NPCs are characterized by the prominent infiltration of lymphocytes in the tumor-surrounding area and the so-called lymphoepithelioma. This difference in histological features has been investigated in various directions regarding the association between EBV and the clinical features of NPC.
WHO I NPC accounts for less than 20% of NPC cases globally. This rate is relatively low in endemic areas, such as Southeast Asia and southern China. In other words, the prevalence of non-keratinizing (WHO II, III) NPC is higher in endemic areas (>95%) and is predominantly associated with EBV infection. However, patients with NPC present elevated IgG and IgA concentrations in response to the viral capsid antigen (VCA) and early antigen (EA) of EBV regardless of endemic or non-endemic area [7]. For WHO I NPC, initial studies reported similar pathology and EBV serologic profiles similar to those of other head and neck carcinomas [8,9], whereas other studies have suggested that all types of NPC result in elevated concentrations of EBV antigens [10]. Recent high-resolution analyses, such as duplex multiplex assays for EBV IgA and IgG antibodies, have shown that very fine adjustment of sample sera is mandatory for the precise quantification of antibody titers [11,12]. Presumably, the mixed review of the serological association of WHO I NPC is attributable to technical problems. However, it is generally accepted that WHO I represents NPC that is unrelated to EBV, and WHO II and III represent EBV-associated NPC. The clinical characteristics of WHO I NPC differ from those of WHO II and III NPCs. A multicenter prospective trial revealed that patients with WHO I NPC had no distant metastatic recurrence, and all relapsed sites were locoregional areas. This pattern is similar to the patterns of conventional head and neck cancers. In contrast, patients with WHO II and III NPCs had significantly higher rates of distant metastatic recurrence. These results indicate that EBV contributes to the high metastatic properties of WHO II and III NPCs [13] (Table 1).
Recent trends of human papillomavirus (HPV) infections in head and neck carcinomas, especially oropharyngeal carcinomas, have been incorporated into NPC research. HPV was detected in two of the 58 NPC tissues at our institute. The HPV-positive cases were classified as WHO II, and EBERs were detected using in situ hybridization. One case had HPV18, and the others had HPV16 and 18 [14].
Reports have been inconsistent, likely because of the limited number of patients and the ethnic and geographic differences among the study populations. Several studies have detected HPV in NPC, with some demonstrating a dichotomy between EBV and HPV infections predominantly in non-endemic regions [15,16,17]. Others have reported cases of EBV and HPV co-infection, predominantly in patients from endemic regions [18,19,20]. Similar to EBV, a serological test has been developed for HPV and applied to screening for HPV-associated diseases [21].

1.2. EBV Gene Expression in NPC

The default program for EBV infection in nasopharyngeal epithelial cells is lytic infection. Therefore, switching the mode of infection from lytic to latent and maintaining latent infection are essential steps in NPC pathogenesis. EBV latency programs can be categorized into three types. Latency affects various diseases and infected cell types [22,23,24]. Two EBV-encoded genes, EBER and EBNA1, are expressed in all three types of latencies. Type I latency is characteristic of Burkitt’s lymphoma, and it is defined as the expression of a minimal number of latent EBV genes, which are EBER and EBNA1. For type II latency, LMP1 and LMP2A are expressed in addition to EBER and EBNA1. Type II latency is observed in NPC, Hodgkin’s disease, and T/NK-cell lymphoma. Type III latency is recognized in lymphoproliferative diseases in immunocompromised hosts such as organ recipients or HIV patients [25,26]. The remarkable expression of BART-miRNAs in NPC and EBV-associated gastric cancer strongly suggests an important carcinogenic role for BART-miRNAs in EBV-associated epithelial malignancies [27].
NPC is an EBV type II latency malignant tumor that expresses three EBV-encoded proteins (EBNA1, LMP1, and LMP2) and two EBV-encoded transcripts, EBERs and BARTs [22,23,25]. Among the EBV genes, we have mainly focused on the EBV primary oncogene LMP1 as it promotes all aspects of “the hallmark of cancer” advocated by Hanahan and Weinberg [28,29].

1.3. Effect of LMP1 Expression in NPC

LMP1 expression is generally observed in moderate-to-severe dysplastic lesions in the EBV-infected nasopharyngeal epithelium or preinvasive NPC. Thus, LMP1 plays an important role in driving EBV-infected premalignant cells into the early stages of NPC [30]. The comprehensive study on LMP1 revealed its oncogenic properties by activating multiple cellular signal pathways such as NF-κB, AP-1, and PI3 kinase [31]. Among them, the activation of the NF-κB pathway is a major characteristic of LMP1. Several steps of carcinogenesis, including anti-apoptosis, anti-cellular stresses, cell metabolism, acquisition of cancer stemness, and immune attacks, are mediated by the NF-κB-associated pathway [32,33,34,35,36,37,38,39]. Moreover, the contribution of LMP1 to the highly metastatic features of NPC, represented by the downregulation of cell adhesion, upregulation of stromal destruction, and angiogenesis, is also attributable to this NF-κB signaling [40,41,42,43,44] (Figure 1).
Interestingly, the mutually exclusive relationship of LMP1 expression with somatic mutations of negative regulators that are located upstream of the NF-κB pathway in NPC has been reported from the endemic region. These findings indicate a special role of NF-κB activation via constitutive activation through somatic mutations dominated by CYLD, TRAF3, NFKBI, and NLRC5, or LMP1 in the pathogenesis of NPC [45,46]. These somatic alterations of NF-κB regulators may supplant the need for LMP1-mediated NF-κB activation in tumor cells during tumor progression [46].

2. Effect of BZLF1 Expression in NPC

Two EBV infection modes, the latent and lytic phases, are believed to be tightly regulated by specific gene expression patterns; however, evidence of the simultaneous expression of latent and lytic genes within the same cell has been accumulated. Among lytic EBV gene products, we focus on BZLF1 as it is the most studied and well-known initiator of the EBV replication switch.
BZLF1, identified in studies on the defective EBV genome, is a trigger gene that facilitates the progression of EBV infection from the latent to the lytic phase [47,48]. The induction of BZLF1 in latently infected cells usually activates the EBV lytic cascade. However, this lytic cascade terminates without producing mature EBV virions; this is called the abortive lytic infection [49]. Thus, the BZLF1 protein ZEBRA can affect host cells during abortive lytic infection [49].
The roles of ZEBRA in EBV infection, the lytic cycle, and oncogenesis have been extensively studied. ZEBRA is an AP-1 mimicking transcriptional factor. Thus, ZEBRA transactivates EBV genes, as well as various host cellular genes related to proliferation, inflammation, angiogenesis, and metastasis. ZEBRA downregulates apoptosis-related genes, MHC class II genes, and interferon regulatory factors, eventually acquiring resistance to cell death and evasion during immune surveillance [38]. ZEBRA also interacts with various cellular proteins, including NF-κB, and changes their functions [50,51,52]. Thus, the effects of ZEBRA are complex. The fate of EBV-infected cells during abortive lytic infection requires further investigation.
Both LMP1 and ZEBRA expression have been identified in NPC tumor tissues [48,52]. However, ZEBRA is still clinically valuable. The expression of the protein in NPC tissues and elevation of anti-ZEBRA IgG concentrations in sera are biomarkers for poor prognosis in patients with NPC, as well as the diagnosis of early-stage NPC [50,53,54,55]. In addition, elevated concentrations of IgA antibodies against the EBV VCA are high-risk markers for NPC development [56]. Studies on EBV strains isolated from NPC and other EBV-associated diseases suggest the existence of aberrant strains predisposed to spontaneous lytic replication and specific EBV-associated diseases [57,58,59,60,61].

2.1. Genetic and Epigenetic Alteration in NPC Host Genome

LMP1 and BZLF1 are key EBV genes that drive EBV-infected epithelial cells toward NPC. However, the expression of EBV genes alone is insufficient for the development of clinical NPC. The accumulation of multiple irreversible somatic genetic alterations, in addition to EBV infection, is required for the development of NPC [22]. To date, there have been several reports of frequently identified genetic changes in patients with EBV-associated NPC. The frequent identification of allelic loss at chromosomal loci 3p21.3 and 9p21 in both EBV-positive and -negative pre-invasive lesions suggests that these genetic changes have occurred before EBV infection [22,62,63]. These genetic alterations imply inactivation of tumor suppressor genes. Among the genes, RAS association domain family 1A (RASSF1A) and cyclin-dependent kinase inhibitor 2A (p16/CDKN2A) are candidate genes involved in the initiation of NPC [22,64], as inactivation of RASSF1A and p16 may provide a growth advantage for the clonal expansion of EBV-infected premalignant epithelial cells.
Recent genome-wide analyses of invasive and pre-invasive NPC have elucidated common somatic alterations in the NPC host genome. Focal loss of alleles on chromosomes 3p, 9p, 11q, 13q, 14q, and 16q and amplification of chromosome 11q13, where the cyclin D1 loci are located, are frequently observed [64,65,66]. Among these genes, the inactivation of p16/CDKN2A by homozygous deletion and promoter hypermethylation has been consistently found in almost all NPC samples examined [67], suggesting that the expression of the p16 gene product is commonly downregulated in NPC.
Epigenetic changes can be induced by EBV infection, and they eventually play a role in tumorigenesis [68]. CpG hypermethylation is a common mechanism involved in the inactivation of tumor suppressor genes in human cancers and has been observed in EBV-associated malignant tumors, including Burkitt’s lymphoma, NPC, and EBV-associated gastric cancer [69,70]. A specific EBV epigenetic signature has been more thoroughly studied in EBV-associated gastric cancer, whereas that in EBV-associated NPC is slightly less defined. Nonetheless, promoter methylation of multiple genes at chromosomes 3p21.3 (including RASSF1A [71]), 6p22.1-21.3 (including the MHC locus [72]), and 9p21 (including CDKN2A (p16), CDKN2B (p15) [73], and others [74]) has been identified. To date, the relationship between latent EBV infection and the methylation profile observed in NPC has not been fully elucidated.
ZEBRA acts as a pioneer factor. Considering that ZEBRA is a homologue of c-Fos and binds to the AP-1 site, it is reasonable to speculate that ZEBRA expression in NPC tumor cells could modulate gene expression in both heterochromatin and euchromatin regions. Further studies are needed to explore the broad effects of ZEBRA expression in NPC.

2.2. Potential Risk Factors Other Than EBV

In addition to host genetics and EBV infection, other potential risk factors identified by epidemiological studies include the family history of nasopharyngeal carcinoma, active and passive tobacco smoking, consumption of preserved foods and alcohol, and oral hygiene [75,76,77,78,79]. In addition to these traditional carcinogenic factors, the role of microbiota in the pathogenesis of NPC has been gradually elucidated.
Microorganisms within the gut and other niches may contribute to carcinogenesis, influence cancer immunosurveillance, and respond to immunotherapy. Thus, targeting the gut and tumor microbiota in cancer is an issue of interest [78]. The association of EBV has been firmly established as a microbial risk factor for NPC. However, the crucial role of the nasopharynx as a niche of the upper respiratory tract microbiome and the clinical implications of the microbiota in NPC tissues remain largely obscure.
Qiao et al. conducted a transcription study on 12 paired NPC tissues with either high or low bacterial loads. Proliferative genes in cell cycle pathways and metastasis-associated pathways were upregulated in tumors with a high bacterial load. Conversely, both T- and B-cell-mediated immune responses and interferon pathways, as well as O-linked glycosylation, are activated in tumors with a low bacterial load [80]. Subsequently, they conducted a multicenter cohort study that included 802 patients with NPC. They confirmed the existence of microbiota mainly originating from the nasopharynx within NPC tissues. Moreover, the intratumoral bacterial load was negatively associated with tumor-infiltrating immune cells and positively associated with poor prognosis in patients with NPC [81]. The results also showed that the levels of expression of various immune-related genes, such as CXCL13, were negatively associated with the load of intratumoral bacteria, particularly Porphyromonas. Butyrate produced by Porphyromonas gingivalis contributes to the regulation of histone acetylation, which reactivates EBV, implying the existence of crosstalk between the bacteria and EBV [82].

2.3. Prognostic Relevance of Cell-Free EBV DNA

The UICC/AJCC staging system has been evolving, and the eighth system is of clinical use [83]. However, this diagnostic system is anatomically based and has limitations in predicting the prognosis or treatment benefits because it is affected by other clinical factors and molecular biomarkers. Several researchers have attempted to incorporate these factors into the system to establish better predictive markers of clinical outcomes.
Among the candidate factors, the combination of pretreatment plasma EBV DNA and clinicopathological variables is a reliable system for arriving at a more accurate prognosis for patients with NPC [84]. Other biomarkers, such as somatic and EBV DNA methylation and miRNA and mRNA expression, have also shown prognostic value and potential clinical applications in patients with NPC [85,86]. Genome-wide screening techniques have revealed the prognostic value of gene expression-based signature analysis and may be beneficial in selecting patients for more intensive treatments [85,86]. However, the diagnostic and prognostic value of circulating EBV DNA is commonly accepted, and quantification kits for circulating EBV, including serum, plasma, and whole blood samples, are now commercially available. These findings have promoted the use of circulating EBV DNA in NPC treatment.
The successful incorporation of pretreatment plasma EBV DNA into the eighth edition has been reported [87,88]. Furthermore, both pretreatment and EBV DNA responses after induction chemotherapy or post-treatment are useful predictors of clinical outcomes [89,90]. These observations have suggested that EBV DNA is a promising liquid biopsy-based biomarker for risk-stratified treatment adaptation.
Longitudinal EBV DNA quantification during and after NPC treatment is now accepted as a valuable method of monitoring NPC. Despite the well-established value of EBV DNA in NPC monitoring, international standardization of assay methods, such as variations in the number of target repeats in the EBV genome, has been a huge obstacle for prospective large-scale trials, even when applying the same assay system [91,92]. In 2016, at the National Cancer Institute EBV testing harmonization workshop for nasopharyngeal carcinoma, NPC and laboratory medicine experts presented the limitations of current assays for EBV DNA quantification and discussed methods for improving combination assays in the future [93]. Studies on the application of EBV DNA as a biomarker for NPC staging are ongoing.

3. Immune Microenvironment of NPC and EBV

As mentioned above, EBV-associated NPC is characterized by the prominent infiltration of lymphocytes in the tumor-surrounding area and the so-called lymphoepithelioma. Based on the unique pathological features of NPC, the relationship between the immune system and NPC tumor cells has been intensively investigated.
The nasopharynx contains a unique lymphoid tissue called the nasopharynx-associated lymphoid tissue (NALT), which differs from other lymphoid tissues involved in organogenesis. NALT plays an important role in the mucosal immune system [94]. The secretion of anti-EBV IgA from the NALT into the airway lumen and serum is a marker for NPC screening and diagnosis. New insights into the tumor microenvironment of NPC have been growing, including the tumor–immune microenvironment (TIME). The tumor microenvironment is mainly composed of heterogeneous cellular components such as tumor cells, fibroblasts, endothelial cells, and leukocytes. It also contains various mediators, such as cytokines and exosomes, which influence the behavior of NPC tumors and, eventually, the prognosis of NPC patients. The TIME consists of immunostimulant and immunosuppressive components. Thus, the TIME component also serves as a prognostic biomarker and a potential target for novel therapies [95].
Tumor-infiltrating lymphocytes (TIL) in NPC tissue contain various immune cells, and the proportion of the immune population dynamically changes with tumor progression and treatment. However, TIL levels usually reflect the treatment outcomes of patients. Abundant intratumoral and stromal TILs are predictive markers of favorable outcomes in patients with NPC [96].

3.1. Immune Evasion Mechanism in NPC

Malignant tumor cells begin the construction of the TIME once they are recognized as targets of immune cells. Subsequently, the tumor cells escape the immune attack. Generally, viral proteins are presented as antigens combined with MHC class I molecules to recruit CD 8-positive cytotoxic T cells (CTL) to the virus-infected tumor cells. Three EBV genes, EBNA1, LMP1, and LMP2, are expressed in NPC. However, EBV-encoded gene products expressed in NPC cells must escape from immune cells, mainly from CTL, for the development of NPC.
In the EBV latent infection program, the most efficient strategy for immune escape is to keep the level of expression of the EBV antigen per cell as low as possible. Maintaining low antigen expression leads to low CTL epitope presentation in EBV-infected cells. Crotzer et al. reported that the presentation of PRIYDLIEL-like epitopes in EBNA3C is less than one epitope per cell [97], resulting in very low or no detection of EBV-transformed B-cells by CTL clones with high affinity to the peptide [98].
In addition to the low levels of latent gene expression, EBNA1 contains a Gly-Ara repeat domain that suppresses the translation of EBNA1 and prevents the processing of EBNA1, resulting in the inhibition of antigen presentation by MHC class I [99].
Upon activation, LMP1 localizes and aggregates within lipid rafts on the cell membrane [100]. LMP1 mutants that lose their aggregating properties at lipid rafts are recognized by HLA epitope-sensitized CTL, whereas wild-type and LMP mutants that retain their aggregating properties at lipid rafts are not. These results suggest that the aggregation of LMP1 in lipid rafts protects LMP1 from being processed and presented with MHC class I, which allows LMP1 expression in NPC cells to escape immune surveillance [101].
In contrast, lytic infection involves the expression of more than 60 viral gene products with high copy numbers per cell [102]. For the transmission of EBV from carriers who have established T-cell immunity to the EBV gene to other individuals, cells with EBV lytic infection must escape T-cell recognition long enough to produce mature EBV particles. Even in cells under abortive lytic infection, the expression of immunogenic ZEBRA should affect the tumor–immune cell interactions.
BZLF1 suppresses the transcription of inflammatory factors TNF-α and IFN-γ and prevents their response during the EBV lytic infection. This implies that the EBV lytic cycle employs a distinct strategy to evade the antiviral inflammatory response [103].

3.2. Superantigen Induction by EBV

One theory states that despite these mechanisms of immune evasion, some EBV gene expression leads to lymphocyte infiltration. This is one theory. The other theory is the induction of superantigens by EBV genes. Superantigens are microbial pathogen-derived proteins that induce strong T-cell responses. HERV-K18 was the first human endogenous provirus to originate from a retrovirus with superantigen activity [104]. HERV-K18 preferentially activates human TCRBV13 T cells [105]. Essentially, it exists in a dormant state. However, the expression of its env gene was observed in latently infected EBV cells. Sutkowski showed that both LMP1 and LMP2A are able to transactivate the HERV-K18 superantigen. Both LMP1 and LMP2A are expressed in NPC cells [106]. HERV-K18 may induce heavy T-cell infiltration, and the secretion of cytokines by the interaction of superantigen-expressing tumor cells with activated T-cells can contribute to the highly metastatic features of NPC [107,108].

3.3. Modulation of Immune Checkpoint in NPC

Generally, the number of intratumoral and stromal TILs is associated with a favorable prognosis in patients with NPC [96]. The programmed death-1/programmed death-ligand 1 (PD-1/PD-L1) axis plays an important role in T-cell tolerance and immune escape of tumor cells. PD-L1 expression is higher in EBV-positive than in EBV-negative NPC cell lines. It can be due to the abundant IFN-γ induction by EBV infection. In addition, LMP1 and IFN-γ pathways cooperate to regulate PD-L1 [44]. LMP1 knockdown suppresses PD-L1 expression in EBV-positive cell lines. IFN-γ upregulates PD-L1 expression independently and synergistically with LMP1 in NPC tissue [109].
PD-1 is expressed on TIL, while PD-L1 is expressed in tumor and stromal cells. The interaction between PD-1 and PD-L1 suppresses CTL attacks on tumor cells [109]. Thus, the co-expression of PD-1 and PD-L1 in the tumor microenvironment usually predicts the recurrence and metastasis of NPC after initial therapy. Antibodies that block this pathway have been clinically used against various cancers, including NPC [95,110]. Adoptive T-cell therapy targeting the EBV gene products is a potential therapeutic strategy for NPC. PD-L1 expression in cancer cells can inhibit the effector functions of adoptively transferred EBV-specific T-cells. The combination of EBV-specific adoptive T-cell and PD-L1 blockade therapies has been reported to be more effective [111].
Several immunotherapies have been explored for NPC; however, these trials have yielded inconsistent conclusions, probably due to different immune microenvironments [109,110,112,113,114,115]. For example, LMP1 promotes myeloid-derived suppressor cell (MDSC) expansion in the tumor microenvironment by promoting extra-mitochondrial glycolysis in malignant cells. In addition to RAGE, LMP1 promotes the expression of multiple glycolytic genes, such as GLUT1. This metabolic reprogramming induces the NOD-like receptor family protein 3 inflammasome and activates the arachidonic cascade, which in turn activates various cytokines such as IL-1β, IL-6, and GM-CSF. These changes in the tumor environment result in NPC-derived MDSC induction [116].

3.4. Challenge for EBV Vaccine Development

The close association of EBV infection with NPC suggests that controlling EBV infection and targeting EBV-infected cells are effective strategies for preventing NPC development and managing developed NPC. However, despite the urgent need for prophylactic or therapeutic solutions, there has been no licensed EBV vaccine thus far [117]. The major obstacles to developing a prophylactic vaccine include the global prevalence of EBV and the fact that infection typically occurs during early childhood to adolescence, resulting in a limited number of control cohorts without EBV infection. In addition, the complexity of the EBV replication system and its infectivity to T lymphocytes and NK cells present challenges in eliminating all EBV target cells [117]. With recent advancements in messenger RNA vaccines, exemplified by their successful application in the COVID-19 pandemic, various viruses, including EBV, have been considered candidates for this technique [118]. Moderna has announced the initiation of a phase 1 study for its mRNA Epstein-Barr virus (EBV) vaccine, code-named mRNA-1189. mRNA-1189 comprises five mRNAs encoding envelope glycoproteins (gp320, gH, gL, gp42, and gp220) that bind to target cell surface receptors, playing key roles in initiating EBV infection of target cells. The administration of mRNA-1189 aims to induce a broad immune response that could prevent EBV infection in various types of cells, ultimately reducing the symptoms of infectious mononucleosis. According to the manufacturer, preclinical testing of mRNA-1189 in mice and nonhuman primates demonstrated high and durable levels of antigen-specific antibodies against B cell and epithelial cell infection [119]. The announced phase 1 clinical trial (NCT05164094) will assess the safety and tolerability of three different doses of mRNA-1189 in healthy adults aged 18 to 30. The humoral immune response will be evaluated up to day 197 [120]. (https://clinicaltrials.gov/study/NCT05164094 (accessed on 11 October 2023)).
mRNA-1189, which encodes additional latent EBV genes, may serve as a candidate for therapeutic vaccines. Certainly, mRNA technology is poised to bring about significant progress in the future of vaccinology.

4. Early Diagnosis of NPC

The prognosis of patients with advanced-stage disease, including NPC, has not improved despite various treatment modalities and the systematic clinical trials that have been developed. Currently, patients with early-stage NPC are treated with radiotherapy alone, whereas those with an advanced stage are treated with a combination of chemotherapy and radiotherapy. Patients with advanced-stage NPC have worse survival rates and quality of life than those with early-stage disease [9,117]. Thus, early detection and timely treatment of NPC is an effective strategy to improve prognosis and quality of life and reduce the medical burden of patients with NPC [118,119]. Nasopharyngeal endoscopy, EBV serology, and plasma EBV DNA have been introduced and assessed by several researchers for the early detection of NPC. Here, we discuss the advantages and disadvantages of these modalities (Table 2).
Nasopharyngeal endoscopy with pathological examination of suspected lesions is the gold standard for the diagnosis of NPC [120]. However, it is primarily used for individuals with suspected NPC. Therefore, it is not suitable for early screening, especially large-scale screening. In addition, current laboratory blood tests (e.g., anti-EBV antibody screening and EBV DNA testing) can provide high specificity and sensitivity, but they still have low positive predictive values (PPVs) and produce more false-positive results, leading to repeated nasal endoscopic examinations, biopsies, and follow-up. To improve the early diagnosis rate of NPC, He et al. developed a deep-learning-based NPC detection model based on a large-scale video frame dataset for nasopharyngeal endoscopy, which enabled more accurate biopsy site selection during endoscopy [121].
There are currently three common clinical EBV antibody detection targets: VCA-IgA, EBNA1-IgA, and EA-IgA [122]. The responses to these antibodies are markedly heterogeneous, and the response to a single antibody is not sufficient for NPC screening. Moreover, the traditional immunoenzyme labeling method is complicated, and sometimes the results are reproducible. Therefore, it is gradually being replaced by ELISA. With the development of new techniques for the detection of dual and triple antibodies, the sensitivity and specificity of the serological diagnosis of NPC have significantly improved [123]. In recent years, researchers have made efforts to improve the sensitivity, specificity, and PPV for detecting early-stage NPC. However, an acceptable PPV (<10%) has not been achieved.
The IgA serological test is the most generally accepted NPC screening method, but the quantification of plasma EBV DNA originating from the tumor is a more sensitive biomarker for screening, as well as for predicting and detecting recurrent NPC disease. Compared with serological screening, plasma EBV DNA detection is more sensitive and specific. The PPV of plasma EBV DNA is also better than that of antibody testing but still lower than 20%. In addition, the cost of EBV DNA testing is high. EBV DNA testing requires advanced equipment and standardization among laboratories that use this method, making it more difficult to scale out EBV DNA testing in populations from all high-risk areas [123].
In a recent review by Yuan et al., EBV detection in nasopharyngeal swabs as an additional test for EBV serology in high-risk individuals reduced the number of candidates for close follow-up. This method achieved a 40% improvement relative to serological examination alone [124]. However, the PPV remained low. The hypermethylation status of gene promoters in nasopharyngeal swabs can also be used for NPC screening. However, the sample sizes of the relevant studies were not sufficiently large. Large-scale studies are necessary to assess its effectiveness for NPC screening [125,126] (Table 2).
The anti-BNLF2b total antibody (P85-Ab), a novel biomarker, was validated through a large-scale prospective screening program and compared with the two standard antibodies, EBNA1-IgA and VCA-IgA.
Forty-seven patients with NPC (38 with early-stage disease) were identified from a cohort of 24,852 prospectively eligible participants. P85-Ab showed higher sensitivity (97.9% vs. 72.3%; ratio, 1.4 [95% CI, 1.1 to 1.6]), specificity (98.3% vs. 97.0%; ratio, 1.01 [95% CI, 1.01 to 1.02]), and PPV (10.0% vs. 4.3%; ratio, 2.3 [95% CI, 1.8 to 2.8]) than the combination of EBNA1-IgA and VCA-IgA. The combination of anti-BNLF2b total antibody with anti-EBNA1-IgA and anti-VCA-IgA improved the PPV to 44.6% (95% CI, 33.8 to 55.9), with a sensitivity of 70.2% (95% CI, 56.0 to 81.4). Serological screening using these three antibodies may provide a promising novel biomarker [127].

5. Conclusions

Clinicopathological features of NPC are mostly attributable to EBV infection and, eventually, virus–host and tumor–immune interaction. The infection program of EBV in NPC is basically a type II latent infection; however, EBV genes involved in the lytic infection program are also expressed. Among both program genes, latent gene LMP1 and lytic gene BZLF1 are the main players in organizing the clinicopathological characteristics of NPC. The elevation of EBV antibodies and cell-free DNA in both patients and future candidates for NPC represents crucial features. These features provide valuable insights into the development of more sensitive and cost-effective screening and diagnostic methods. The unique TME characterized by an intense infiltration of immune cells also offers clues for potential therapeutic targets in NPC. In fact, specific inhibitors aimed at immune evasion mechanisms, such as immune checkpoint inhibitors, are now in clinical use. Early detection and the development of novel drugs for advanced NPC constitute effective strategies to enhance the prognosis of patients with NPC. Efforts to clarify the mechanism of EBV-mediated NPC pathogenesis and apply the findings to NPC treatment are ongoing. With the rapid technological advances, further evidence is expected to be revealed.

Author Contributions

Conceptualization: T.Y. and S.K. (Satoru Kondo); writing—original draft preparation: T.Y., H.D., H.M. and E.K.; writing—review and editing: S.K. (Shigetaka Komura) and K.E.; funding acquisition: T.Y., S.K. (Satoru Kondo) and E.K. All authors have read and approved the manuscript.

Funding

This work was supported by JSPS KAKENHI (Grant-in-Aid for Scientific Research, Japan Society For The Promotion of Science) (grant numbers: 23H03053, 21K19557, and 19K09864).

Acknowledgments

The authors thank Asuka Nanbo for providing us the opportunity to contribute to this special issue, and Naohiro Wakisaka (Kanazawa Medical Center), Yosuke Nakanishi, and Nobuyuki Hirai (Division of Otolaryngology, Kanazawa University Hospital) for helpful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wen, K.W.; Wang, L.; Menke, J.R.; Damania, B. Cancers associated with human gammaherpesviruses. FEBS J. 2022, 289, 7631–7669. [Google Scholar] [CrossRef] [PubMed]
  2. Epstein, M.A.; Achong, B.G.; Barr, Y.M. Virus particles in cultured lymphoblasts from Burkitt’s lymphoma. Lancet 1964, 7335, 702–703. [Google Scholar] [CrossRef] [PubMed]
  3. Old, L.J.; Boyse, E.A.; Oettgen, H.F.; Harven, E.D.; Geering, G.; Williamson, B.; Clifford, P. Precipitating antibody in human serum to an antigen present in cultured Burkitt’s lymphoma cells. Proc. Natl. Acad. Sci. USA 1966, 56, 1699–1704. [Google Scholar] [CrossRef]
  4. Zur Hausen, H.; Schulte-Holthausen, I.; Klein, G.; Henle, W.; Henle, G.; Clifford, P.; Santesson, L. EBV-DNA in biopsies of Burkitt’s tumors and anaplastic carcinomas of the nasopharynx. Nature 1970, 228, 1056–1058. [Google Scholar] [CrossRef] [PubMed]
  5. Wolf, H.; zur Hausen, H.; Becker, V. EB viral genomes in epithelial nasopharyngeal carcinoma cells. Nat. New Biol. 1973, 244, 245–247. [Google Scholar] [CrossRef]
  6. Shanmugaratnam, K. Histological typing of nasopharyngeal carcinoma. IARC Sci. Publ. 1978, 20, 3–12. [Google Scholar]
  7. Yoshizaki, T. EBV-related serological biomarkers for nasopharyngeal cancer remain a hot topic. Ann. Nasopharynx Cancer 2018, 2, 5–7. [Google Scholar] [CrossRef]
  8. Neel, H.B., 3rd; Pearson, G.R.; Weiland, L.H.; Taylor, W.F.; Goepfert, H.H.; Pilch, B.Z.; Goodman, M.; Lanier, A.P.; Huang, A.T.; Hyams, V.J.; et al. Application of Epstein-Barr virus serology to the diagnosis and staging of North American patients with nasopharyngeal carcinoma. Otolaryngol. Head Neck Surg. 1983, 91, 255–262. [Google Scholar] [CrossRef]
  9. Chen, Y.P.; Chan, A.T.C.; Le, Q.T.; Blanchard, P.; Sun, Y.; Ma, J. Nasopharyngeal carcinoma. Lancet 2019, 394, 64–80. [Google Scholar] [CrossRef]
  10. Sam, C.K.; Prasad, U.; Pathmanathan, R. Serological markers in the diagnosis of histopathological types of nasopharyngeal carcinoma. Eur. J. Surg. Oncol. 1989, 15, 357–360. [Google Scholar]
  11. Schieber, J.; Pring, M.; Ness, A.; Liu, Z.; Hsu, W.L.; Brenner, N.; Butt, J.; Waterboer, T.; Simon, J. Development of a Duplex Serological Multiplex Assay for the Simultaneous Detection of Epstein-Barr Virus IgA and IgG Antibodies in Nasopharyngeal Carcinoma Patients. Cancers 2023, 15, 2578. [Google Scholar] [CrossRef] [PubMed]
  12. Lam, W.K.J.; King, A.D.; Miller, J.A.; Liu, Z.; Yu, K.J.; Chua, M.L.K.; Ma, B.B.Y.; Chen, M.Y.; Pinsky, B.A.; Lou, P.J.; et al. Recommendations for Epstein-Barr virus-based screening for nasopharyngeal cancer in high- and intermediate-risk regions. J. Natl. Cancer Inst. 2023, 115, 355–364. [Google Scholar] [CrossRef] [PubMed]
  13. Fuwa, N.; Kodaira, T.; Daimon, T.; Yoshizaki, T. The long-term outcomes of alternating chemoradiotherapy for locoregionally advanced nasopharyngeal carcinoma: A multiinstitutional phase II study. Cancer Med. 2016, 4, 1186–1195. [Google Scholar] [CrossRef] [PubMed]
  14. Kano, M.; Kondo, S.; Wakisaka, N.; Moriyama-Kita, M.; Nakanishi, Y.; Endo, K.; Murono, S.; Nakamura, H.; Yoshizaki, T. The influence of human papillomavirus on nasopharyngeal carcinoma in Japan. Auris Nasus Larynx 2017, 44, 327–332. [Google Scholar] [CrossRef] [PubMed]
  15. Lin, Z.; Khong, B.; Kwok, S.; Cao, H.; West, R.B.; Le, Q.T.; Kong, C.S. Human papillomavirus 16 detected in nasopharyngeal carcinomas in white Americans but not in endemic Southern Chinese patients. Head Neck 2014, 36, 709–714. [Google Scholar] [CrossRef] [PubMed]
  16. Dogan, S.; Hedberg, M.L.; Ferris, R.L.; Rath, T.J.; Assaad, A.M.; Chiosea, S.I. Human papillomavirus and Epstein-Barr virus in nasopharyngeal carcinoma in a low-incidence population. Head Neck 2014, 36, 511–516. [Google Scholar] [CrossRef] [PubMed]
  17. Robinson, M.; Suh, Y.E.; Paleri, V.; Devlin, D.; Ayaz, B.; Pertl, L.; Thavaraj, S. Oncogenic human papillomavirus-associated nasopharyngeal carcinoma: An observational study of correlation with ethnicity, histological subtype and outcome in a UK population. Infect. Agents Cancer 2013, 8, 30. [Google Scholar] [CrossRef] [PubMed]
  18. Laantri, N.; Attaleb, M.; Kandil, M.; Naji, F.; Mouttaki, T.; Dardari, R.; Belghmi, K.; Benchakroun, N.; El Mzibri, M.; Khyatti, M. Human papillomavirus detection in moroccan patients with nasopharyngeal carcinoma. Infect. Agents Cancer 2011, 6, 3. [Google Scholar] [CrossRef]
  19. Mirzamani, N.; Salehian, S.; Farhadi, M.; Amin Tehran, E. Detection of EBV and HPV in nasopharyngeal carcinoma by in situ hybridization Exp. Mol. Pathol. 2006, 81, 231–234. [Google Scholar] [CrossRef]
  20. Rassekh, C.H.; Rady, P.L.; Arany, I.; Tyring, S.K.; Knudsen, S.; Calhoun, K.H.; Seikaly, H.; Bailey, B.J. Combined Epstein-Barr virus and human papillomavirus infection in nasopharyngeal carcinoma. Laryngoscope 1998, 108, 362–367. [Google Scholar] [CrossRef]
  21. Robbins, H.A.; Ferreiro-Iglesias, A.; Waterboer, T.; Brenner, N.; Nygard, M.; Bender, N.; Schroeder, L.; Hildesheim, A.; Pawlita, M.; D’Souza, G.; et al. Absolute Risk of Oropharyngeal Cancer after an HPV16-E6 Serology Test and Potential Implications for Screening: Results from the Human Papillomavirus Cancer Cohort Consortium. J. Clin. Oncol. 2022, 40, 3613–3622. [Google Scholar] [CrossRef] [PubMed]
  22. Tsao, S.W.; Tsang, C.M.; Lo, K.W. Epstein-Barr virus infection and nasopharyngeal carcinoma. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2017, 372, 20160270. [Google Scholar] [CrossRef] [PubMed]
  23. Young, L.S.; Yap, L.F.; Murray, P.G. Epstein-Barr virus: More than 50 years old and still providing surprises. Nat. Rev. Cancer 2016, 16, 789–802. [Google Scholar] [CrossRef] [PubMed]
  24. Jha, H.C.; Pei, Y.; Robertson, E.S. Epstein-Barr virus: Diseases linked to infection and transformation. Front. Microbiol. 2016, 7, 1602. [Google Scholar] [CrossRef] [PubMed]
  25. Bechtel, D.; Kurth, J.; Unkel, C.; Kuppers, R. Transformation of BCR-deficient germinal-center B cells by EBV supports a major role of the virus in the pathogenesis of Hodgkin and posttransplantation lymphomas. Blood 2005, 106, 4345–4350. [Google Scholar] [CrossRef] [PubMed]
  26. Rowe, M.; Rowe, D.T.; Gregory, C.D.; Young, L.S.; Farrell, P.J.; Rupani, H.; Rickinson, A.B. Differences in B cell growth phenotype reflect novel patterns of Epstein-Barr virus latent gene expression in Burkitt’s lymphoma cells. EMBO J. 1987, 6, 2743–2751. [Google Scholar] [CrossRef]
  27. Kang, D.; Skalsky, R.L.; Cullen, B.R. EBV BART microRNAs target multiple pro-apoptotic cellular genes to promote epithelial cell survival. PLoS Pathog. 2015, 11, e1004979. [Google Scholar] [CrossRef]
  28. Hanahan, D.; Weinberg, R.A. The Hallmarks of Cancer. Cell 2000, 100, 57–70. [Google Scholar] [CrossRef]
  29. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef]
  30. Pathmanathan, R.; Prasad, U.; Sadler, R.; Flynn, K.; Raab-Traub, N. Clonal proliferations of cells infected with Epstein-Barr virus in preinvasive lesions related to nasopharyngeal carcinoma. N. Engl. J. Med. 1995, 333, 693–698. [Google Scholar] [CrossRef]
  31. Mosialos, G.; Birkenbach, M.; Yalamanchili, R.; VanArsdale, T.; Ware, C.; Kieff, E. The Epstein-Barr virus transforming protein LMP1 engages signaling proteins for the tumor necrosis factor receptor family. Cell 1995, 80, 389–399. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, M.T.; Chang, Y.T.; Chen, S.C.; Chuang, Y.C.; Chen, Y.R.; Lin, C.S.; Chen, J.Y. Epstein-Barr virus latent membrane protein 1 represses p53-mediated DNA repair and transcriptional activity. Oncogene 2005, 24, 2635–2646. [Google Scholar] [CrossRef] [PubMed]
  33. Henderson, S.; Rowe, M.; Gregory, C.; Croom-Carter, D.; Wang, F.; Longnecker, R.; Kieff, E.; Rickinson, A. Induction of bcl-2 expression by Epstein-Barr virus latent membrane protein 1 protects infected B cells from programmed cell death. Cell 1991, 65, 1107–1115. [Google Scholar] [CrossRef] [PubMed]
  34. Luo, X.; Hong, L.; Cheng, C.; Li, N.; Zhao, X.; Shi, F.; Liu, J.; Fan, J.; Zhou, J.; Bode, A.M.; et al. DNMT1 mediates metabolic reprogramming induced by Epstein-Barr virus latent membrane protein 1 and reversed by grifolin in nasopharyngeal carcinoma. Cell Death Dis. 2018, 9, 619. [Google Scholar] [CrossRef] [PubMed]
  35. Tsuji, A.; Wakisaka, N.; Kondo, S.; Murono, S.; Furukawa, M.; Yoshizaki, T. Induction of receptor for advanced glycation end products by EBV latent membrane protein 1 and its correlation with angiogenesis and cervical lymph node metastasis in nasopharyngeal carcinoma. Clin. Cancer Res. 2008, 14, 5368–5375. [Google Scholar] [CrossRef]
  36. Kondo, S.; Wakisaka, N.; Muramatsu, M.; Zen, Y.; Endo, K.; Murono, S.; Sugimoto, H.; Yamaoka, S.; Pagano, J.S.; Yoshizaki, T. Epstein-Barr virus latent membrane protein 1 induces cancer stem/progenitor-like cells in nasopharyngeal epithelial cell lines. J. Virol. 2011, 85, 11255–11264. [Google Scholar] [CrossRef]
  37. Horikawa, T.; Yang, J.; Kondo, S.; Yoshizaki, T.; Joab, I.; Furukawa, M.; Pagano, J.S. Twist and epithelial-mesenchymal transition are induced by the EBV oncoprotein latent membrane protein 1 and are associated with metastatic nasopharyngeal carcinoma. Cancer Res. 2007, 67, 1970–1978. [Google Scholar] [CrossRef]
  38. Ressing, M.E.; van Gent, M.; Gram, A.M.; Hooykaas, M.J.; Piersma, S.J.; Wiertz, E.J. Immune evasion by Epstein-Barr virus. Curr. Top. Microbiol. Immunol. 2015, 391, 355–381. [Google Scholar] [CrossRef]
  39. Kase, K.; Kondo, S.; Wakisaka, N.; Dochi, H.; Mizokami, H.; Kobayashi, E.; Kano, M.; Komori, T.; Hirai, N.; Ueno, T.; et al. Epstein-Barr Virus LMP1 Induces Soluble PD-L1 in Nasopharyngeal Carcinoma. Microorganisms 2021, 9, 603. [Google Scholar] [CrossRef]
  40. Fåhraeus, R.; Chen, W.; Trivedi, P.; Klein, G.; Obrink, B. Decreased expression of E-cadherin and increased invasive capacity in EBV-LMP-transfected human epithelial and murine adenocarcinoma cells. Int. J. Cancer 1992, 52, 834–838. [Google Scholar] [CrossRef]
  41. Yoshizaki, T.; Sato, H.; Furukawa, M.; Pagano, J. The expression of matrix metalloproteinase 9 is enhanced by Epstein-Barr virus latent membrane protein 1. Proc. Natl. Acad. Sci. USA 1998, 95, 3621–3626. [Google Scholar] [CrossRef] [PubMed]
  42. Yoshizaki, T.; Horikawa, T.; Ren, Q.; Wakisaka, N.; Takeshita, H.; Sheen, T.S.; Lee, S.Y.; Sato, H.; Furukawa, M. Induction of interleukin-8 by Epstein-Barr virus latent membrane pretein-1 and its correlation with angiogenesis in nasopharyngeal carcinoma. Clin. Cancer Res. 2001, 7, 1946–1951. [Google Scholar] [PubMed]
  43. Murono, S.; Inoue, H.; Tanabe, T.; Joab, I.; Yoshizaki, T.; Furukawa, M.; Pagano, S. Induction of cyclooxygenase-2 by Epstein-Barr virus latent membrane protein 1 is involved in vascular endothelial growth factor production in nasopharyngeal carcinoma cells. Proc. Natl. Acad. Sci. USA 2001, 98, 6905–6910. [Google Scholar] [CrossRef] [PubMed]
  44. Yoshizaki, T.; Kondo, S.; Endo, K.; Nakanishi, Y.; Aga, M.; Kobayashi, E.; Hirai, N.; Sugimoto, H.; Hatano, M.; Ueno, T.; et al. Modulation of the tumor microenvironment by Epstein-Barr virus latent membrane protein 1 in nasopharyngeal carcinoma. Cancer Sci. 2018, 109, 272–278. [Google Scholar] [CrossRef] [PubMed]
  45. Zheng, H.; Dai, W.; Cheung, A.K.; Ko, J.M.; Kan, R.; Wong, B.W.; Leong, M.M.; Deng, M.; Kwok, T.C.; Chan, J.Y.; et al. Whole-exome sequencing identifies multiple loss-of-function mutations of NF-kappaB pathway regulators in nasopharyngeal carcinoma. Proc. Natl. Acad. Sci. USA 2016, 113, 11283–11288. [Google Scholar] [CrossRef]
  46. Li, Y.Y.; Chung, G.T.; Lui, V.W.; To, K.F.; Ma, B.B.; Chow, C.; Woo, J.K.; Yip, K.Y.; Seo, J.; Hui, E.P.; et al. Exome and genome sequencing of nasopharynx cancer identifies NF-kappaB pathway activating mutations. Nat. Commun. 2017, 8, 14121. [Google Scholar] [CrossRef] [PubMed]
  47. Countryman, J.; Miller, G. Activation of expression of latent Epstein-Barr herpesvirus after gene transfer with a small cloned subfragment of heterogeneous viral DNA. Proc. Natl. Acad. Sci. USA 1985, 82, 4085–4089. [Google Scholar] [CrossRef]
  48. Ramayanti, O.; Juwana, H.; Verkuijlen, S.A.; Adham, M.; Pegtel, M.D.; Greijer, A.E.; Middeldorp, J.M. Epstein-Barr virus mRNA profiles and viral DNA methylation status in nasopharyngeal brushings from nasopharyngeal carcinoma patients reflect tumor origin. Int. J. Cancer 2017, 140, 149–162. [Google Scholar] [CrossRef]
  49. Germini, D.; Sall, F.B.; Shmakova, A.; Wiels, J.; Dokudovskaya, S.; Drouet, E.; Vassetzky, Y. Oncogenic Properties of the EBV ZEBRA Protein. Cancers 2020, 12, 1479. [Google Scholar] [CrossRef]
  50. Yoshizaki, T.; Sato, H.; Murono, S.; Pagano, J.S.; Furukawa, M. Matrix metalloproteinase 9 is induced by the Epstein-Barr virus BZLF1 transactivator. Clin. Exp. Metastasis 1999, 17, 431–436. [Google Scholar] [CrossRef]
  51. Morrison, T.E.; Mauser, A.; Wong, A.; Ting, J.P.; Kenney, S.C. Inhibition of IFN-gamma signaling by an Epstein-Barr virus immediate-early protein. Immunity 2001, 15, 787–799. [Google Scholar] [CrossRef]
  52. Rosemarie, Q.; Sugden, B. Epstein-Barr Virus: How Its Lytic Phase Contributes to Oncogenesis. Microorganisms 2020, 8, 1824. [Google Scholar] [CrossRef]
  53. Yoshizaki, T.; Miwa, H.; Takeshita, H.; Sato, H.; Furukawa, M. Elevation of antibody against Epstein-Barr virus genes BRLF1 and BZLF1 in nasopharyngeal carcinoma. J. Cancer Res. Clin. Oncol. 2000, 126, 69–73. [Google Scholar]
  54. Zhang, G.; Li, Z.; Zhou, Q. Utility of Serum EB Virus Zta Antibody in the Diagnostic of Nasopharyngeal Carcinoma: Evidences from 2126 Cases and 15,644 Controls. Front. Oncol. 2019, 9, 1391. [Google Scholar] [CrossRef]
  55. Dochi, H.; Kondo, S.; Murata, T.; Fukuyo, M.; Nanbo, A.; Wakae, K.; Jiang, W.P.; Hamabe-Horiike, T.; Tanaka, M.; Nishiuchi, T.; et al. Estrogen induces the expression of EBV lytic protein ZEBRA, a marker of poor prognosis in nasopharyngeal carcinoma. Cancer Sci. 2022, 113, 2862–2877. [Google Scholar] [CrossRef]
  56. Chien, Y.C.; Chen, J.Y.; Liu, M.Y.; Yang, H.I.; Hsu, M.M.; Chen, C.J.; Yang, C.S. Serologic markers of Epstein-Barr virus infection and nasopharyngeal carcinoma in Taiwanese men. N. Engl. J. Med. 2001, 345, 1877–1882. [Google Scholar] [CrossRef]
  57. Tsai, M.H.; Lin, X.; Shumilov, A.; Bernhardt, K.; Feederle, R.; Poirey, R.; Kopp-Schneider, A.; Pereira, B.; Almeida, R.; Delecluse, H.J. The biological properties of different Epstein-Barr virus strains explain their association with various types of cancers. Oncotarget 2017, 8, 10238–10254. [Google Scholar] [CrossRef]
  58. Okuno, Y.; Murata, T.; Sato, Y.; Muramatsu, H.; Ito, Y.; Watanabe, T.; Okuno, T.; Murakami, N.; Yoshida, K.; Sawada, A.; et al. Defective Epstein-Barr virus in chronic active infection and haematological malignancy. Nat. Microbiol. 2019, 4, 404–413. [Google Scholar] [CrossRef]
  59. Kondo, S.; Okuno, Y.; Murata, T.; Dochi, H.; Wakisaka, N.; Mizokami, H.; Moriyama-Kita, M.; Kobayashi, E.; Kano, M.; Komori, T.; et al. EBV genome variations enhance clinicopathological features of nasopharyngeal carcinoma in a non-endemic region. Cancer Sci. 2022, 113, 2446–2456. [Google Scholar] [CrossRef] [PubMed]
  60. Xu, M.; Yao, Y.; Chen, H.; Zhang, S.; Cao, S.M.; Zhang, Z.; Luo, B.; Liu, Z.; Li, Z.; Xiang, T.; et al. Genome sequencing analysis identifies Epstein-Barr virus subtypes associated with high risk of nasopharyngeal carcinoma. Nat. Genet. 2019, 51, 1131–1136. [Google Scholar] [CrossRef] [PubMed]
  61. Kanda, T.; Yajima, M.; Ikuta, K. Epstein-Barr virus strain variation and cancer. Cancer Sci. 2019, 110, 1132–1139. [Google Scholar] [CrossRef]
  62. Chan, A.S.; To, K.F.; Lo, K.W.; Ding, M.; Li, X.; Johnson, P.; Huang, D.P. Frequent chromosome 9p losses in histologically normal nasopharyngeal epithelia from southern Chinese. Int. J. Cancer 2002, 102, 300–303. [Google Scholar] [CrossRef]
  63. Chan, A.S.; To, K.F.; Lo, K.W.; Mak, K.F.; Pak, W.; Chiu, B.; Tse, G.M.; Ding, M.; Li, X.; Lee, J.C.; et al. High frequency of chromosome 3p deletion in histologically normal nasopharyngeal epithelia from southern Chinese. Cancer Res. 2000, 60, 5365–5370. [Google Scholar]
  64. Hui, A.B.; Or, Y.Y.; Takano, H.; Tsang, R.K.; To, K.F.; Guan, X.Y.; Sham, J.S.; Hung, K.W.; Lam, C.N.; van Hasselt, C.A.; et al. Array-based comparative genomic hybridization analysis identified cyclin D1 as a target oncogene at 11q13.3 in nasopharyngeal carcinoma. Cancer Res. 2005, 65, 8125–8133. [Google Scholar] [CrossRef]
  65. Lo, K.W.; Teo, P.M.; Hui, A.B.; To, K.F.; Tsang, Y.S.; Chan, S.Y.; Mak, K.F.; Lee, J.C.; Huang, D.P. High resolution allelotype of microdissected primary nasopharyngeal carcinoma. Cancer Res. 2000, 60, 3348–3353. [Google Scholar]
  66. Lin, D.C.; Meng, X.; Hazawa, M.; Nagata, Y.; Varela, A.M.; Xu, L.; Sato, Y.; Liu, L.Z.; Ding, L.W.; Sharma, A.; et al. The genomic landscape of nasopharyngeal carcinoma. Nat. Genet. 2014, 46, 866–871. [Google Scholar] [CrossRef]
  67. Lo, K.W.; Huang, D.P.; Lau, K.M. p16 gene alterations in nasopharyngeal carcinoma. Cancer Res. 1995, 55, 2039–2043. [Google Scholar]
  68. Kaneda, A.; Matsusaka, K.; Aburatani, H.; Fukayama, M. Epstein-Barr virus infection as an epigenetic driver of tumorigenesis. Cancer Res. 2012, 72, 3445–3450. [Google Scholar] [CrossRef] [PubMed]
  69. Li, L.; Zhang, Y.; Guo, B.B.; Chan, F.K.; Tao, Q. Oncogenic induction of cellular high CpG methylation by Epstein-Barr virus in malignant epithelial cells. Chin. J. Cancer 2014, 33, 604–608. [Google Scholar] [CrossRef] [PubMed]
  70. Wang, K.; Yuen, S.T.; Xu, J.; Lee, S.P.; Yan, H.H.; Shi, S.T.; Siu, H.C.; Deng, S.; Chu, K.M.; Law, S.; et al. Whole-genome sequencing and comprehensive molecular profiling identify new driver mutations in gastric cancer. Nat. Genet. 2014, 46, 573–582. [Google Scholar] [CrossRef] [PubMed]
  71. Lo, K.W.; Kwong, J.; Hui, A.B.; Chan, S.Y.; To, K.F.; Chan, A.S.; Chow, L.S.; Teo, P.M.; Johnson, P.J.; Huang, D.P. High frequency of promoter hypermethylation of RASSF1A in nasopharyngeal carcinoma. Cancer Res. 2001, 61, 3877–3881. [Google Scholar] [PubMed]
  72. Dai, W.; Cheung, A.K.; Ko, J.M.; Cheng, Y.; Zheng, H.; Ngan, R.K.; Ng, W.T.; Lee, A.W.; Yau, C.C.; Lee, V.H.; et al. Comparative methylome analysis in solid tumors reveals aberrant methylation at chromosome 6p in nasopharyngeal carcinoma. Cancer Med. 2015, 4, 1079–1090. [Google Scholar] [CrossRef] [PubMed]
  73. Lo, K.W.; Cheung, S.T.; Leung, S.F.; van Hasselt, A.; Tsang, Y.S.; Mak, K.F.; Chung, Y.F.; Woo, J.K.; Lee, J.C.; Huang, D.P. Hypermethylation of the p16 gene in nasopharyngeal carcinoma. Cancer Res. 1996, 56, 2721–2725. [Google Scholar] [PubMed]
  74. Kwong, J.; Lo, W.; To, K.F.; Teo, P.M.; Johnson, P.J.; Huang, D.P. Promoter hypermethylation of multiple genes in nasopharyngeal carcinoma. Clin. Cancer Res. 2002, 8, 131–137. [Google Scholar]
  75. Guo, X.; Johnson, R.C.; Deng, H.; Liao, J.; Guan, L.; Nelson, G.W.; Tang, M.; Zheng, Y.; de The, G.; O’Brien, S.J.; et al. Evaluation of nonviral risk factors for nasopharyngeal carcinoma in a high-risk population of Southern China. Int. J. Cancer 2009, 124, 2942–2947. [Google Scholar] [CrossRef]
  76. Tsao, S.W.; Yip, Y.L.; Tsang, C.M.; Pang, P.S.; Lau, V.M.; Zhang, G.; Lo, K.W. Etiological factors of nasopharyngeal carcinoma. Oral Oncol. 2014, 50, 330–338. [Google Scholar] [CrossRef]
  77. Liu, Z.; Chang, E.T.; Liu, Q.; Cai, Y.; Zhang, Z.; Chen, G.; Xie, S.H.; Cao, S.M.; Shao, J.Y.; Jia, W.H.; et al. Oral hygiene and risk of nasopharyngeal carcinoma—A population-based case-control study in China. Cancer Epidemiol. Biomark. Prev. 2016, 25, 1201–1207. [Google Scholar] [CrossRef]
  78. Liu, Z.; Chang, E.T.; Liu, Q.; Cai, Y.; Zhang, Z.; Chen, G.; Huang, Q.H.; Xie, S.H.; Cao, S.M.; Shao, J.Y.; et al. Quantification of familial risk of nasopharyngeal carcinoma in a high-incidence area. Cancer 2017, 123, 2716–2725. [Google Scholar] [CrossRef]
  79. Chang, E.T.; Liu, Z.; Hildesheim, A.; Liu, Q.; Cai, Y.; Zhang, Z.; Chen, G.; Xie, S.H.; Cao, S.M.; Shao, J.Y.; et al. Active and passive smoking and risk of nasopharyngeal carcinoma: A population-based case-control study in southern China. Am. J. Epidemiol. 2017, 185, 1272–1280. [Google Scholar] [CrossRef]
  80. Park, E.M.; Chelvanambi, M.; Bhutiani, N.; Kroemer, G.; Zitvogel, L.; Wargo, J.A. Targeting the gut and tumor microbiota in cancer. Nat. Med. 2022, 28, 690–703. [Google Scholar] [CrossRef]
  81. Qiao, H.; Tan, X.R.; Li, H.; Li, J.Y.; Chen, X.Z.; Li, Y.Q.; Li, W.F.; Tang, L.L.; Zhou, G.Q.; Zhang, Y.; et al. Association of Intratumoral Microbiota with Prognosis in Patients with Nasopharyngeal Carcinoma from 2 Hospitals in China. JAMA Oncol. 2022, 8, 1301–1309. [Google Scholar] [CrossRef]
  82. Imai, K.; Inoue, H.; Tamura, M.; Cueno, M.E.; Inoue, H.; Takeichi, O.; Kusama, K.; Saito, I.; Ochiai, K. The periodontal pathogen Porphyromonas gingivalis induces the Epstein-Barr virus lytic switch transactivator ZEBRA by histone modification. Biochimie 2012, 94, 839–846. [Google Scholar] [CrossRef] [PubMed]
  83. Tang, L.L.; Chen, Y.P.; Mao, Y.P.; Wang, Z.X.; Guo, R.; Chen, L.; Tian, L.; Lin, A.H.; Li, L.; Sun, Y.; et al. Validation of the 8th edition of the uicc/ajcc staging system for nasopharyngeal carcinoma from endemic areas in the intensity-modulated radiotherapy era. J. Natl. Compr. Cancer Netw. 2017, 15, 913–919. [Google Scholar] [CrossRef] [PubMed]
  84. Tang, L.Q.; Li, C.F.; Li, J.; Chen, W.H.; Chen, Q.Y.; Yuan, L.X.; Lai, X.P.; He, Y.; Xu, Y.X.; Hu, D.P.; et al. Establishment and validation of prognostic nomograms for endemic nasopharyngeal carcinoma. J. Natl. Cancer Inst. 2015, 108, djv291. [Google Scholar] [CrossRef] [PubMed]
  85. Jiang, W.; Liu, N.; Chen, X.Z.; Sun, Y.; Li, B.; Ren, X.Y.; Qin, W.F.; Jiang, N.; Xu, Y.F.; Li, Y.Q.; et al. Genome-wide identification of a methylation gene panel as a prognostic biomarker in nasopharyngeal carcinoma. Mol. Cancer Ther. 2015, 14, 2864–2873. [Google Scholar] [CrossRef] [PubMed]
  86. Liu, N.; Chen, N.Y.; Cui, R.X.; Li, W.; Li, Y.; Wei, R.; Zhang, M.; Sun, Y.; Huang, B.; Chen, M.; et al. Prognostic value of a microRNA signature in nasopharyngeal carcinoma: A microRNA expression analysis. Lancet Oncol. 2012, 13, 633–641. [Google Scholar] [CrossRef] [PubMed]
  87. Tang, X.R.; Li, Y.Q.; Liang, S.B.; Jiang, W.; Liu, F.; Ge, W.; Tang, L.; Mao, Y.; He, M.; Yang, X.; et al. Development and validation of a gene expression-based signature to predict distant metastasis in locoregionally advanced nasopharyngeal carcinoma: A retrospective, multicentre, cohort study. Lancet Oncol. 2018, 19, 382–393. [Google Scholar] [CrossRef]
  88. Xu, C.; Chen, Y.P.; Liu, X.; Li, W.F.; Chen, L.; Mao, Y.P.; Zhang, Y.; Guo, R.; Zhou, G.Q.; Tang, L.L.; et al. Establishing and applying nomograms based on the 8th edition of the UICC/AJCC staging system to select patients with nasopharyngeal carcinoma who benefit from induction chemotherapy plus concurrent chemoradiotherapy. Oral Oncol. 2017, 69, 99–107. [Google Scholar] [CrossRef]
  89. Guo, R.; Tang, L.L.; Mao, Y.P.; Du, X.J.; Chen, L.; Zhang, Z.C.; Liu, L.Z.; Tian, L.; Luo, X.T.; Xie, Y.B.; et al. Proposed modifications and incorporation of plasma Epstein-Barr virus DNA improve the TNM staging system for Epstein-Barr virus-related nasopharyngeal carcinoma. Cancer 2019, 125, 79–89. [Google Scholar] [CrossRef]
  90. Liu, L.T.; Tang, L.Q.; Chen, Q.Y.; Zhang, L.; Guo, S.S.; Guo, L.; Mo, H.Y.; Zhao, C.; Guo, X.; Cao, K.J.; et al. The prognostic value of plasma Epstein-barr viral DNA and tumor response to neoadjuvant chemotherapy in advanced-stage nasopharyngeal carcinoma. Int. J. Radiat. Oncol. Biol. Phys. 2015, 93, 862–869. [Google Scholar] [CrossRef]
  91. Chan, A.T.C.; Hui, E.P.; Ngan, R.K.C.; Tung, S.Y.; Cheng, A.C.K.; Ng, W.T.; Lee, V.H.F.; Ma, B.B.Y.; Cheng, H.C.; Wong, F.C.S.; et al. Analysis of plasma Epstein-Barr virus DNA in nasopharyngeal cancer after chemoradiation to identify high-risk patients for adjuvant chemotherapy: A randomized controlled trial. J. Clin. Oncol. 2018, 36, 3091–3100. [Google Scholar] [CrossRef] [PubMed]
  92. Le, Q.T.; Zhang, Q.; Cao, H.; Cheng, A.J.; Pinsky, B.A.; Hong, R.L.; Chang, J.T.; Wang, C.W.; Tsao, K.C.; Lo, Y.M.D.; et al. An international collaboration to harmonize the quantitative plasma Epstein-Barr virus DNA assay for future biomarker-guided trials in nasopharyngeal carcinoma. Clin. Cancer Res. 2013, 19, 2208–2215. [Google Scholar] [CrossRef] [PubMed]
  93. Kim, K.Y.; Le, Q.T.; Yom, S.S.; Pinsky, B.A.; Bratman, S.V.; Ng, R.H.; El Mubarak, H.S.; Chan, K.C.; Sander, M.; Conley, B.A. Current state of PCR-based Epstein-Barr virus DNA testing for nasopharyngeal cancer. J. Natl. Cancer Inst. 2017, 109, djx007. [Google Scholar] [CrossRef] [PubMed]
  94. Kiyono, H.; Fukuyama, S. NALT- versus PEYER’S-patch-mediated mucosal immunity. Nat. Rev. Immunol. 2004, 4, 699–710. [Google Scholar] [CrossRef] [PubMed]
  95. Xu, L.; Zou, C.; Zhang, S.; Chu, T.S.M.; Zhang, Y.; Chen, W.; Zhao, C.; Yang, L.; Xu, Z.; Dong, S.; et al. Reshaping the systemic tumor immune environment (STIE) and tumor immune microenvironment (TIME) to enhance immunotherapy efficacy in solid tumors. J. Hematol. Oncol. 2022, 15, 87. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, Y.Q.; Chen, Y.P.; Zhang, Y.; Jiang, W.; Liu, N.; Yun, J.P.; Sun, Y.; He, Q.M.; Tang, X.R.; Wen, X.; et al. Prognostic significance of tumor-infiltrating lymphocytes in nondisseminated nasopharyngeal carcinoma: A large-scale cohort study. Int. J. Cancer 2018, 142, 2558–2566. [Google Scholar] [CrossRef] [PubMed]
  97. Crotzer, V.L.; Christian, R.E.; Brooks, J.M.; Shabanowitz, J.; Settlage, R.E.; Marto, J.A.; White, F.M.; Rickinson, A.B.; Hunt, D.F. Engelhard VH. Immunodominance among EBV-derived epitopes restricted by HLA-B27 does not correlate with epitope abundance in EBV-transformed B-lymphoblastoid cell lines. J. Immunol. 2000, 164, 6120–6129. [Google Scholar] [CrossRef]
  98. Hill, A.B.; Lee, S.P.; Haurum, J.S.; Murray, N.; Yao, Q.Y.; Rowe, M.; Signoret, N.; Rickinson, A.B.; McMichael, A.J. Class I major histocompatibility complex-restricted cytotoxic T lymphocytes specific for Epstein-Barr virus (EBV)-transformed B lymphoblastoid cell lines against which they were raised. J. Exp. Med. 1995, 181, 2221–2228. [Google Scholar] [CrossRef]
  99. Münz, C. Epstein-barr virus nuclear antigen 1: From immunologically invisible to a promising T cell target. J. Exp. Med. 2004, 199, 1301–1304. [Google Scholar] [CrossRef]
  100. Clausse, B.; Fizazi, K.; Walczak, V.; Tetaud, C.; Wiels, J.; Tursz, T.; Busson, P. High concentration of the EBV Latent Membrane Protein 1 in glycosphingolipid-rich complexes from both epithelial and lymphoid cells. Virolgy 1997, 228, 285–293. [Google Scholar] [CrossRef]
  101. Smith, C.; Wakisaka, N.; Crough, T.; Peet, J.; Yoshizaki, T.; Beagley, L.; Khanna, R. Discerning regulation of cis- and trans-presentation of CD8+ T-cell epitopes by EBV-encoded oncogene LMP-1 through self-aggregation. Blood 2009, 113, 6148–6152. [Google Scholar] [CrossRef] [PubMed]
  102. Kieff, E.; Ricknson, A. Epstein-Barr virus and its replication. In Fields Virology; Knipe, D.M., Howley, P.M., Eds.; Lippincott-Raven: Philadelphia, PA, USA, 2001; pp. 2511–2573. [Google Scholar]
  103. Li, Y.; Long, X.; Huang, L.; Yang, M.; Yuan, Y.; Wang, Y.; Delecluse, H.J.; Kuang, E. Epstein-Barr Virus BZLF1-Mediated Downregulation of Proinflammatory Factors Is Essential for Optimal Lytic Viral Replication. J. Virol. 2015, 90, 887–903. [Google Scholar] [CrossRef] [PubMed]
  104. Conrad, B.; Weissmahr, R.N.; Böni, J.; Arcari, R.; Schüpbach, J.; Mach, B. A human endogenous retroviral superantigen as candidate autoimmune gene in type I diabetes. Cell 1997, 90, 303–313. [Google Scholar] [CrossRef]
  105. Sutkowski, N.; Conrad, B.; Thorley-Lawson, D.A.; Huber, B.T. Epstein-Barr virus transactivates the human endogenous retrovirus HERV-K18 that encodes a superantigen. Immunity 2001, 15, 579–589. [Google Scholar] [CrossRef] [PubMed]
  106. Sutkowski, N.; Chen, G.; Calderon, G.; Huber, B.T. Epstein-Barr virus latent membrane protein LMP-2A is sufficient for transactivation of the human endogenous retrovirus HERV-K18 superantigen. J. Virol. 2004, 78, 7852–7860. [Google Scholar] [CrossRef] [PubMed]
  107. Agathanggelou, A.; Niedobitek, G.; Chen, R.; Nicholls, J.; Yin, W.; Young, L.S. Expression of immune regulatory molecules in Epstein-Barr virus-associated nasopharyngeal carcinomas with prominent lymphoid stroma. Evidence for a functional interaction between epithelial tumor cells and infiltrating lymphoid cells. Am. J. Pathol. 1995, 147, 1152–1160. [Google Scholar] [PubMed]
  108. Niedobitek, G.; Agathanggelou, A.; Nicholls, J.M. Epstein-Barr virus infection and the pathogenesis of nasopharyngeal carcinoma: Viral gene expression, tumour cell phenotype, and the role of the lymphoid stroma. Semin. Cancer Biol. 1996, 7, 165–174. [Google Scholar] [CrossRef]
  109. Fang, W.; Zhang, J.; Hong, S.; Zhan, J.; Chen, N.; Qin, T.; Tang, Y.; Zhang, Y.; Kang, S.; Zhou, T.; et al. EBV-driven LMP1 and IFN-γ up-regulate PD-L1 in nasopharyngeal carcinoma: Implications for oncotargeted therapy. Oncotarget 2014, 5, 12189–12202. [Google Scholar] [CrossRef]
  110. Zhang, J.; Fang, W.; Qin, T.; Yang, Y.; Hong, S.; Liang, W.; Ma, Y.; Zhao, H.; Huang, Y.; Xue, C. Co-expression of PD-1 and PD-L1 predicts poor outcome in nasopharyngeal carcinoma. Med. Oncol. 2015, 32, 86. [Google Scholar] [CrossRef]
  111. Smith, C.; McGrath, M.; Neller, M.A.; Matthews, K.K.; Crooks, P.; Le Texier, L.; Panizza, B.; Porceddu, S.; Khanna, R. Complete response to PD-1 blockade following EBV-specific T-cell therapy in metastatic nasopharyngeal carcinoma. NPJ Precis. Oncol. 2021, 5, 24. [Google Scholar] [CrossRef]
  112. Lee, V.H.; Lo, A.W.; Leung, C.Y.; Shek, W.H.; Kwong, D.L.; Lam, C.O.; Tong, C.C.; Sze, C.L.; Leung, T.W. Correlation of PD-L1 expression of tumor cells with survival outcomes after radical intensity-modulated radiation therapy for non-metastatic nasopharyngeal carcinoma. PLoS ONE 2016, 11, e0157969. [Google Scholar] [CrossRef] [PubMed]
  113. Zhu, Q.; Cai, M.Y.; Chen, C.L.; Hu, H.; Lin, H.X.; Li, M.; Weng, D.S.; Zhao, J.J.; Guo, L.; Xia, J.C. Tumor cells PD-L1 expression as a favorable prognosis factor in nasopharyngeal carcinoma patients with pre-existing intratumor-infiltrating lymphocytes. Oncoimmunology 2017, 6, e1312240. [Google Scholar] [CrossRef] [PubMed]
  114. Larbcharoensub, N.; Mahaprom, K.; Jiarpinitnun, C.; Trachu, N.; Tubthong, N.; Pattaranutaporn, P.; Sirachainan, E.; Ngamphaiboon, N. Characterization of PD-L1 and PD-1 expression and CD8+tumor-infiltrating lymphocyte in Epstein-Barr virus-associated nasopharyngeal carcinoma. Am. J. Clin. Oncol. 2018, 41, 1204–1210. [Google Scholar] [CrossRef] [PubMed]
  115. Ono, T.; Azuma, K.; Kawahara, A.; Sasada, T.; Matsuo, N.; Kakuma, T.; Kamimura, H.; Maeda, R.; Hattori, C.; On, K.; et al. Prognostic stratification of patients with nasopharyngeal carcinoma based on tumor immune microenvironment. Head Neck 2018, 40, 2007–2019. [Google Scholar] [CrossRef] [PubMed]
  116. Cai, T.T.; Ye, S.B.; Liu, Y.N.; He, J.; Chen, Q.Y.; Mai, H.Q.; Zhang, C.X.; Cui, J.; Zhang, X.S.; Busson, P.; et al. LMP1-mediated glycolysis induces myeloid-derived suppressor cell expansion in nasopharyngeal carcinoma. PLoS Pathog. 2017, 13, e1006503. [Google Scholar] [CrossRef] [PubMed]
  117. Lee, A.W.; Ng, W.T.; Chan, L.; Hung, W.M.; Chan, C.; Sze, H.C.; Chan, O.S.; Chang, A.T.; Yeung, R.M. Evolution of treatment for nasopharyngeal cancer--success and setback in the intensity-modulated radiotherapy era. Radiother. Oncol. 2014, 110, 377–384. [Google Scholar] [CrossRef]
  118. Du, Y.; Zhang, W.; Lei, F.; Yu, X.; Li, Z.; Liu, X.; Ni, Y.; Deng, L.; Ji, M. Long-term survival after nasopharyngeal carcinoma treatment in a local prefecture-level hospital in southern China. Cancer Manag. Res. 2020, 12, 1329–1338. [Google Scholar] [CrossRef]
  119. Ji, M.F.; Sheng, W.; Cheng, W.M.; Ng, M.H.; Wu, B.H.; Yu, X.; Wei, K.R.; Li, F.G.; Lian, S.F.; Wang, P.P.; et al. Incidence and mortality of nasopharyngeal carcinoma: Interim analysis of a cluster randomized controlled screening trial (PRO-NPC-001) in southern China. Ann. Oncol. 2019, 30, 1630–1637. [Google Scholar] [CrossRef]
  120. Tay, J.K.; Lim, M.Y.; Kanagalingam, J. Screening in nasopharyngeal carcinoma: Current strategies and future directions. Curr. Otorhinolaryngol. Rep. 2014, 2, 1–7. [Google Scholar] [CrossRef]
  121. He, Z.; Zhang, K.; Zhao, N.; Wang, Y.; Hou, W.; Meng, Q.; Li, C.; Chen, J.; Li, J. Deep learning for real-time detection of nasopharyngeal carcinoma during nasopharyngeal endoscopy. iScience 2023, 26, 07463. [Google Scholar] [CrossRef]
  122. Coghill, A.E.; Hsu, W.L.; Pfeiffer, R.M.; Juwana, H.; Yu, K.J.; Lou, P.J.; Wang, C.P.; Chen, J.Y.; Chen, C.J.; Middeldorp, J.M.; et al. Epstein–Barr virus serology as a potential screening marker for nasopharyngeal carcinoma among high-risk individuals from multiplex families in Taiwan. Cancer Epidemiol. Biomark. Prev. 2014, 23, 1213–1219. [Google Scholar] [CrossRef] [PubMed]
  123. Yu, X.; Chen, H.; Ji, M. Epstein–Barr virus-based nasopharyngeal carcinoma population screening. Ann. Nasopharynx Cancer 2022, 6, 3. [Google Scholar] [CrossRef]
  124. Yuan, Y.; Ye, F.; Wu, J.H.; Fu, X.Y.; Huang, Z.X.; Zhang, T. Early screening of nasopharyngeal carcinoma. Head Neck 2023, 45, 2700–2709. [Google Scholar] [CrossRef] [PubMed]
  125. Yang, X.; Dai, W.; Kwong, D.L.; Szeto, C.Y.; Wong, E.H.; Ng, W.T.; Lee, A.W.; Ngan, R.K.; Yau, C.C.; Tung, S.Y.; et al. Epigenetic markers for noninvasive early detection of nasopharyngeal carcinoma by methylation-sensitive high resolution melting. Int. J. Cancer 2015, 136, E127–E135. [Google Scholar] [CrossRef]
  126. Zhang, Z.; Sun, D.; Hutajulu, S.H.; Nawaz, I.; Nguyen Van, D.; Huang, G.; Haryana, S.M.; Middeldorp, J.M.; Ernberg, I.; Hu, L.F. Development of a noninvasive method, multiplex methylation specific PCR (MMSP), for early diagnosis of nasopharyngeal carcinoma. PLoS ONE 2012, 7, e45908. [Google Scholar] [CrossRef]
  127. Li, T.; Li, F.; Guo, X.; Hong, C.; Yu, X.; Wu, B.; Lian, S.; Song, L.; Tang, J.; Wen, S.; et al. Anti-Epstein-Barr Virus BNLF2b for Mass Screening for Nasopharyngeal Cancer. N. Engl. J. Med. 2023, 389, 808–819. [Google Scholar] [CrossRef]
Figure 1. Activation of multiple cell signaling pathways by LMP1. LMP1-mediated signal activation promotes cell survival, proliferation, angiogenesis, invasion, metastasis, metabolic reprogramming, and cytokine production. LMP1 downregulates E-cadherin expression, which ultimately affects cell adhesion. As a result, LMP1 promotes six generations of “hallmarks of cancer.” Factors downregulated by LMP1 are indicated by light blue-colored boxes.
Figure 1. Activation of multiple cell signaling pathways by LMP1. LMP1-mediated signal activation promotes cell survival, proliferation, angiogenesis, invasion, metastasis, metabolic reprogramming, and cytokine production. LMP1 downregulates E-cadherin expression, which ultimately affects cell adhesion. As a result, LMP1 promotes six generations of “hallmarks of cancer.” Factors downregulated by LMP1 are indicated by light blue-colored boxes.
Microorganisms 12 00014 g001
Table 1. Nasopharyngeal cancer histology and clinicopathological features.
Table 1. Nasopharyngeal cancer histology and clinicopathological features.
WHO IWHO IIWHO III
Differentiation statuswell differentiatedmoderately to poorly differentiatedundifferentiated
Histological category in WHO classificationkeratinizingnonkeratinizing-differentiatednonkeratinizing-undifferentiated
TIL infiltrationfair to moderateheavy
EBERs in tumor(−) or faint(+)
EBV antibodiesnot elevatedelevated
Chemoradiosensitivitymoderategood
Metastatic propertylow to moderatehigh
Epidemiology20% in non-endemic area;
<5% in endemic areas
80% in non-endemic areas;
>95% in endemic areas
EBERs; EBV-encoded small RNAs, TIL; tumor-infiltrating lymphocytes, WHO; World Health Organization.
Table 2. Comparison of modalities for NPC screening and early diagnosis.
Table 2. Comparison of modalities for NPC screening and early diagnosis.
Screening ProcedureSwab CytologySerological TestCirculating EBV-DNACell Free DNA MethylationNasopharyngeal Endoscopy
BenefitsEasy samplingMost popular methodsHigher sensitivity and specificityHigher sensitivity and specificityHigher positive predictive value of approximately 90%
Quick detectionPossible combination of multiple antibody typesHigher positive predictive value but still <20%Higher positive predictive value of approximately 90%Combination of deep learning model improve accuracy
Applicable for EBV negative NPCApplicable for EBV negative NPC
DrawbacksInfluenced by sample quality, physician’s skill, cytologist’s skillLow level of positive predictive value (10%)Influenced by laboratory skill More advanced technique requiredLarge scale screening is not possible
Not globally standardizedLack of large-scale study Higher cost
Higher cost
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

Yoshizaki, T.; Kondo, S.; Dochi, H.; Kobayashi, E.; Mizokami, H.; Komura, S.; Endo, K. Recent Advances in Assessing the Clinical Implications of Epstein-Barr Virus Infection and Their Application to the Diagnosis and Treatment of Nasopharyngeal Carcinoma. Microorganisms 2024, 12, 14. https://doi.org/10.3390/microorganisms12010014

AMA Style

Yoshizaki T, Kondo S, Dochi H, Kobayashi E, Mizokami H, Komura S, Endo K. Recent Advances in Assessing the Clinical Implications of Epstein-Barr Virus Infection and Their Application to the Diagnosis and Treatment of Nasopharyngeal Carcinoma. Microorganisms. 2024; 12(1):14. https://doi.org/10.3390/microorganisms12010014

Chicago/Turabian Style

Yoshizaki, Tomokazu, Satoru Kondo, Hirotomo Dochi, Eiji Kobayashi, Harue Mizokami, Shigetaka Komura, and Kazuhira Endo. 2024. "Recent Advances in Assessing the Clinical Implications of Epstein-Barr Virus Infection and Their Application to the Diagnosis and Treatment of Nasopharyngeal Carcinoma" Microorganisms 12, no. 1: 14. https://doi.org/10.3390/microorganisms12010014

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