Early changes in quasispecies variant after antiviral therapy for chronic hepatitis B

  • Authors:
    • Yujiao Liang
    • Yoshihiko Yano
    • Wahyu Aristyaning Putri
    • Yan Mardian
    • Rina Okada
    • Toshihito Tanahashi
    • Yoshiki Murakami
    • Yoshitake Hayashi
  • View Affiliations

  • Published online on: February 2, 2018     https://doi.org/10.3892/mmr.2018.8551
  • Pages: 5528-5537
Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Hepatitis B virus (HBV) polymerase gene is targeted by nucleos(t)ide analogues (NUC), but it is unclear how HBV quasispecies of whole genome changes during early period of NUC treatment. To understand the unknown region of drug sensitivity and treatment resistance, HBV quasispecies of whole genome during early period of NUC treatment was examined using ultra‑deep sequencing. Eleven patients with chronic HBV infection who received NUC treatment were enrolled in the current study. Viral DNA was extracted from serum samples before and early period of NUC treatment. Polymerase chain reaction analysis was subsequently performed on the DNA products. The viral quasispecies of the entire genome was analyzed by ultra‑deep sequencing. The regions and positions corresponding to the changes in the quasispecies were investigated before and early period of NUC treatment. The secondary structure changes were predicted by mutations/substitutions detected using Lasergene Protean v14.1 software. The frequency of quasispecies variants increased significantly in the polymerase domain from before to early period of NUC treatment (3.08±1.28 vs. 3.51±1.47%, P<0.008), particularly the reverse transcription (RT) domain (3.76±1.25 vs. 4.52±1.37%, P<0.012). In addition, increased variation detected from HBsAg domain showed statistically significant during NUC treatment (6.81±3.26 vs. 7.81±3.26%, P<0.040). The amino acid (aa) mutations/substitutions were detected and compared from before to early period of treatment. Interestingly, most of them were located in the RT region (RT1 motif: aa21‑aa51) and small S region in the early duration of NUC treatment. Furthermore, several mutation patterns, such as cI97L and cP130T showed alterations in the secondary structure and predicted antigenicity of HBV protein. Although the HBV whole genome can be affected by NUC treatment, RT 1 motif region and small S region are more sensitive to the early period of NUC treatment. This study suggested the initial changes of HBV quasispecies might affect the long‑term drug sensitivity and resistance to NUC treatment.

Introduction

Hepatitis B virus (HBV) infection is a public health problem that causes liver diseases such as chronic hepatitis B (CHB) and progressive liver fibrosis, and may lead to cirrhosis and hepatocellular carcinoma (HCC). Approximately 2 billion people have ever been infected with HBV worldwidely, an estimated 257 million people are chronic carriers of HBV surface antigen (HBsAg), and an estimated 650,000 people will die due to CHB every year (1,2). HBV is a small enveloped virus, with a partially double-stranded DNA genome approximately 3.2 kb in length. The genome replicates by the reverse transcription (RT) activity of polymerase. The genome has four genes, the polymerase, surface, core, and X genes with overlapping open reading frames. The most commonly used drugs to treat CHB are nucleos(t)ide analogues (NUC), which inhibit HBV DNA polymerase activity and suppress HBV replication. Drugs available to treat hepatitis B include lamivudine, telbivudine, entecavir, adefovir, and tenofovir. Although all NUC are effective inhibitors of HBV replication, they are rarely curative, and few patients achieve HBsAg clearance. Therefore, long-term NUC therapy is necessary for most patients. However, long-term treatment increases the possibility of HBV genomic changes that confer drug or multidrug resistance, and reduce the treatment efficacy in CHB patients, and could progress to severe liver disease. Drug-resistance mutations were found in approximately 70% of patients treated with lamivudine for up to 5 years, 29% of patients treated with adefovir for up to 5 years, 34% of patients treated with telbivudine for up to 3 years, 1.2% of patients treated with entecavir for up to 6 years (3), and no resistance to tenofovir for up to 7 years (4). The well-known drug resistances (e.g., rtL80I/V, rtI169T, rtV173L, rtL180M, rtA181T, rtT184A, rtV191I, rtS202G, rtM204V, rtM204I, rtV207I, rtN236T, rtM250L) have been reported during long-term NUC treatment (57). In addition, C gene deletion mutants and preS gene deletion mutants in patients during NUC therapy may promote vital immune escape resulting in severe liver disease and even hepatic failure (810). Cases of severe viral reactivation, viral recurrence, and exacerbation of hepatitis have been reported after discontinuation of NUC treatment (11).

Identification of HBV genotypic NUC resistance is mainly performed by PCR amplification and Sanger sequencing. New technologies, including ultra-deep sequencing methods, that enable high-throughput analysis of thousands of amplified regions were recently developed to increase sequencing capacity while generating clonal sequences (12). Ultra-deep sequencing method is an appealing feature for clinical virologists that are involved in the diagnosis of emerging viral infections, molecular epidemiology of viral pathogens, drug-resistance testing, and also do some basic and clinical research, especially the studies of HBV and HCV (7,1215). Sanger Sequencing method was widely used for many years and ultra-deep sequencing replaced it recently due to its advantages what is high-throughput sequencing and more sensitive and efficient in detecting minor variants, which can identify mutants that constitute as little as 1% of the total viral population (7,14,16). Detecting minor population by ultra-deep sequencing assay would be particularly useful for HBV as emerging mutations may be obscured by the ongoing production of mutated variants for a prolonged period (7). In addition, more than one major population in one site can be detected by this assay while Sanger sequencing only can detected one major population in one site. Analysis of the HBV quasispecies in early period of NUC treatment will help us to clarify the substitutions related to genomic sensitivity or resistance to NUC treatment.

So far, although NUC treatment commonly treat patients with hepatitis B infection, the cost of treatment remains economic burden to the patients. Therefore, it is important to choose the appropriate medication and monitor changes in the medication, especially in the early period of treatment. The aim of this study is to understand the unknown region of drug sensitivity and treatment resistance during early period of NUC treatment.

In this study, HBV quasispecies of whole genome during early period of NUC treatment was examined using ultra-deep sequencing. we examined the initial changes in viral quasispecies in relation to drug sensitivity and treatment resistance following NUC therapy. Viral quasispecies were determined by ultra-deep sequencing.

Materials and methods

Patients and serum samples

Eleven CHB patients without advanced liver disease who received NUC treatment were enrolled in this study. The inclusion criteria were: i) HBV DNA titer >4 log copies/ml; ii) ALT ≥31 IU/l; iii) HBsAg positive for more than 6 months without antiviral therapy; iv) receiving NUC treatment after diagnosis; and v) without advanced liver disease such as liver cirrhosis. Patients co-infected with hepatitis C virus, hepatitis D virus, or human immunodeficiency virus etc. were excluded from this study. Blood samples were collected before, early period of NUC treatment and after 6 months of NUC introduction. Serum samples were stored at −80°C before analysis. NUC treatment included adefovir (ADV) and entecavir (ETV).

The demographic and clinical data, including age, sex, results of standard test for CHB, were collected from medical record after institutional review board approval was obtained. Written informed consent was obtained from each patient before all the procedures in this study were done. This study was reviewed and approved by the Ethics Committee at Kobe University.

Serological and biochemical tests, and quantification of HBV DNA

The HBsAg titer was determined for all samples using the Lumipulse HBsAg-HQ assay on a fully automated chemiluminescent enzyme immunoassay system (Lumipulse; Fujirebio, Tokyo, Japan). The sensitivity of the Lumipulse HBsAg-HQ assay is 0.005 IU/ml and its coefficient of variation is <5.9% for samples with a low HBsAg titer. Alanine aminotransferase (ALT) levels were determined using standard procedures. The HBV DNA titer was measured using the TaqMan PCR assay (Cobas TaqMan; Roche Molecular Systems, Basel, Switzerland) with a lower limit of detection of 2.1 log copies/ml (14,17).

The Cobas TaqMan methodology uses qPCR to measure PCR product accumulation through a dual-labelled fluorogenic TaqMan probe. The fluorescent signal is generated by the 5′-nuclease activity of the DNA polymerase that digests the probe, separating a fluorescent reporter dye from a quencher dye. The probe specific for the detection of HBV is labelled with a different fluorogenic reporter dye from the QS-specific probe, while both contain the same quencher dye. The accumulation of amplification products is detected by continuously monitoring the release of the fluorescent reporter dyes during the annealing/extension step of each cycle. The PCR cycle at which the increase in normalized fluorescence of a sample exceeds background noise is called the critical threshold value (Ct). The Ct is inversely proportional to the initial copy number of the target template (18).

DNA extraction and HBV amplification

HBV DNA was extracted from 200 µl of serum using a QIAamp DNA blood Mini kit (Qiagen, Tokyo, Japan) according to the manufacturer's protocol. The entire HBV genome sequence was analyzed using overlapping amplified PCR products. The primers (forward and reverse) were as follows: W1, 5′-GATTCCTGCTCAAGGAACC-3′ (nt 529–547) and 5′-GCCTACAGCCTCCTAGTAC-3′ (nt 1770–1788); W2, 5′-ACTGGGAGGAGTTGGGGGAG-3′ (nt 1729–1748) and 5′-GCTGTAGCTCTTGTTCCCAAG-3′ (nt 2827–2847); and W3, 5′-GGACTCATAAGGTGGGAAAC-3′ (nt 2461–2480) and 5′-GACTTGGCCCCCAATACCAC-3′ (nt 748–767). The PCR reaction comprised denaturing at 95°C for 5 min, followed by 35 cycles of denaturing at 95°C for 1 min, annealing at 55°C for 1 min and elongation at 72°C for 1.5 min, and a final extension at 72°C for 10 min. The presence of an amplified product was confirmed by electrophoresis on 2% agarose gels, stained with ethidium bromide, and visualized on a UV transilluminator. All of the necessary precautions were taken to prevent cross-contamination. A negative control was included in each assay.

Deep sequencing and analysis of amplified products

Ultra-deep sequencing was used to detect variations in the entire genomes of HBV quasispecies. The concentrations of the PCR products were measured using the Qubit high-sensitivity double-stranded DNA assay (Invitrogen; Thermo Fisher Scientific, Inc., Waltham, MA, USA). Next, a library of PCR products (<500 bp) of the viral genome (50 or 0.2 ng) was prepared using the Nextera DNA sample prep kit (Illumina, Inc., San Diego, CA, USA) or Nextera XT DNA sample prep kit (Illumina, Inc.), according to the manufacturer's instructions. The PCR products were uniformly sheared to 500-bp fragments using these kits, and the PCR product libraries were mixed with 1% 8 pM PhiX as a control and run on a MiSeq sequencer (Genome Analyzer; Illumina, Inc.) for paired-end 151 bp sequencing. Finally, image-based detection of the fluorescent signals was performed using MiSeq control software (Illumina, Inc.), and the images were used to develop the sequence data in the FASTQ format. Illumina paired-end sequencing was performed to generate overlapping read pairs (ORP) among relatively short-sequencing-fragment libraries combined with relatively long reads. A quality check and data trimming were necessary before assembling the sequences with Genomics Workbench software version 6.0.1 (CLC bio, Aarhus, Denmark). The sequencing results used in this analysis had a read quality of >80% of the consensus sequence estimated from a quality score of 30 (Q30) according to the manufacturer's (Illumina, Inc.) data. ORP and Q30 filtering were used to recover sequencing errors and eliminate any false-positive variants generated by PCR errors during sequencing, to increase the confidence that the sequence reads could be used to detect the viral variants with low abundance. The assembled sequence reads were mapped against the reference HBV genome of genotype C2 (GenBank accession no. AB246344). The viral quasispecies were evaluated in terms of their prevalence in the viral population, which was done using the setting ‘read conflict’ in Genomics Workbench. The nucleotide(nt) alterations leading to an amino acid (aa) changes were defined as variants. The percentage of the variant population was determined from the proportion of nt substitutions that changed the aa coverage depth (entire sequence reads in each nt position) (14). The viral quasispecies variants detected in >1% of the total viral population were classified into a major population (>10% of the total population) and a minor population (>1% to <10% of the total population). A free DNA translation tool (available at https://www.expasy.org/) was used to compare the variant population before and early period of treatment.

Calculation of antigenicity and secondary structure

To determine whether the aa changes affect the structure of polymerase protein and the antigenicity of HBsAg, hepatitis B core antigen (HBcAg) and HBx, we used the Jameson-Wolf Antigenic Index Prediction algorithm in Lasergene Protean v14.1 software (DNUCTAR Inc., Madison, WI, USA). The Antigenic Index Prediction analysis uses information from the primary aa sequence to predict the secondary structure (Chou-Fasman and Robson-Garner) (19).

Biostatistical analysis

Statistical analyses were performed with SPSS version 24.0 (IBM Corporation, S&I, Tokyo, Japan). Shapiro-Wilk test was used for checking the normal distribution of each pair. Differences between groups were examined using paired samples t-test (nominal variables) and Wilcoxon signed rank test (non-normal) and analysis of variance. P<0.05 was considered to indicate a statistically significant difference.

Results

Clinical characteristics

The subjects' demographic, biochemical and virologic data are shown in Table I. Based on the entire genome, the HBV genotype was C2 in all eleven patients. The HBV DNA titer in samples collected from early period of NUC treatment was significantly lower than those collected before treatment (6.09±0.92 vs. 8.08±1.17 log copies/ml, respectively; P<0.000). The HBsAg titer was also lower during early period of treatment than before treatment (4.02±0.43 vs. 4.21±0.55 log10 IU/ml, respectively), but it did not show statistically significant between before and early period of treatment (P<0.091). The ALT level was significantly decreased during early period of treatment compared with before treatment (93±35 vs. 132±66 IU/l, respectively; P<0.037). A mean of 3,146,683±1,935,574 (range 898,165 to 6,576,948) bp nucleotides before treatment and 3,171,583±2,283,172 (range 1,125,580 to 8,863,180) bp nucleotides during early period of treatment were mapped into the reference consensus, and mean coverage depth of 82,839±48,562 (range 22,023 to 173,168) and 100,545±83,475 (range 31,440 to 311,408) was achieved for each nucleotide site of HBV sequences, respectively. Although we could not calculate statistically due to lack of data, the amount of HBV DNA and HBsAg level have been clearly decreased after six months of NUC treatment (Table I).

Table I.

Demographic, biochemical, and virologic characteristics of the 11 patients.

Table I.

Demographic, biochemical, and virologic characteristics of the 11 patients.

HBV DNA (logcopy/ml)HBsAg (log10 IU/ml)ALT (IU/1)



No.Age/genderAntiviral treatmentDuration (days)BeforeEarlyAfter 6 MBeforeEarlyAfter 6 MBeforeEarlyAfter 6 M
  157/FETV  437.45.83.13.343.453.53120  63  35
  230/METV  328.86.43.14.653.723.23165  90131
  339/METV1489.37.36.14.314.454.35  90132  86
  466/FETV1069.16.75.74.684.464.42  62  49  40
  528/MADF  858.76.52.33.763.633.62363  82  95
  648/FETV1277.04.54.33.373.363.35163  17  18
  775/METV  648.96.44.04.034.133.76  32  43  36
  870/METV  639.36.6ND4.754.55ND123105ND
  958/FETV  145.74.82.73.953.914.09  39  29  24
1026/MADF1237.25.13.74.784.134.00192  77  56
1167/FETV  577.66.94.64.654.424.30  99  41  69
Mean 8.086.09 4.214.02 132  66
SD 1.170.92 0.550.43   93  35
P-value 0.000 0.091 0.037

[i] B/E, before treatment/early period of treatment; ETV, entecavir; ADF, adefovir; HBsAg, hepatitis B virus surface antigen; ALT, alanine aminotransferase; SD, standard deviation; ND, no data; After 6M, after 6 months of NUC introduction.

Frequency of viral quasispecies variants for each genomic domain

The frequency in major and minor viral quasispecies for each genomic domain was compared between before and early period of NUC treatment (Fig. 1). In the major population, the variation in the polymerase region was significantly greater during early period of NUC treatment than before treatment (3.51±1.47 vs. 3.08±1.28%, respectively; P<0.008; Fig. 1A). Of note, the variation in the RT domain of polymerase gene and the overlapping small S domain of the surface gene were significantly different between before and early period of treatment (3.76±1.25 vs. 4.52±1.37%, respectively, P<0.012; 6.81±3.26 vs. 7.81±3.26%, respectively, P<0.040; Fig. 1C). In the minor populations, although there were slight variations between before and early period of treatment, these were not statistically significant (Fig. 1B and D).

Amino acid changes in viral population ongoing NUC treatment

The aa mutations/substitutions that accounted for >1% of the variations in the total viral population were examined, and the variations were compared between before and early period of treatment. The positions were analyzed if they were detected in ≥6 patients (>50% of subjects). The nt and corresponding aa mutations that differed significantly are listed in Table II. Numerous mutations were detected in the RT region, which is targeted by NUC. Fig. 2 shows a map of the polymerase domain boundaries. Interestingly, most of these mutations were located in the RT region, RT1 motif (aa21-aa51) which may affect viral replication. It was suggested that the RT1 motif in the RT region was more sensitive than other subdomains in the early period of treatment. In addition, we found significant variation in the small S region which overlaps with RT region that may affect HBsAg synthesis. In terms of the other regions, we found several substitutions in the core region and X region but only cD51A, xI127T showed statistically different during early period of treatment.

Table II.

Sites and variations of mutations according to genomic region.

Table II.

Sites and variations of mutations according to genomic region.

Variations of viral quasispecies (%), (mean ± SD)

Region (aa)

Position (nt)Polymerase (RT)Surface (small S)CoreXBefore treatmentEarly period of treatmentP-value
A162G N177S (N3S) 98.40±1.77 95.84±7.620.037
G217TV376F (V30F) 18.15±5.93 21.02±5.350.029
G217C/T L195F (L21F) 19.06±6.18 22.00±5.550.027
G220TD377Y (D31Y) 13.12±3.73 15.03±3.300.044
G220C/T L196F (L22F) 13.67±3.88 15.77±3.380.039
A221CD377A (D31A)T197P (T23P) 1.54±0.56 2.01±0.400.027
A221GD377G (D31G)T197A (T23A) 2.89±1.39 3.71±0.920.013
A221TD377V (D31V)T197S (T23S) 4.50±2.69 6.11±2.180.017
C222T T197I (T23I) 2.02±1.11 2.70±0.820.018
T228C I199T (I25T) 1.03±0.46 1.44±0.260.023
A233CH381P (H35P)T201P (T27P) 10.83±2.96 12.53±2.820.011
A235CN382H (N36H) 5.67±1.68 7.09±1.530.005
T241CT384P (T38P) 3.68±1.06 4.32±1.070.041
T247GS386A (S40A) 10.18±2.25 11.65±3.340.044
A245C/T247A/G S205R (S31R) 12.82±3.02 14.77±3.880.036
C248GS386C (S40C)L206V (L32V) 1.00±0.49 1.60±0.490.010
T249C L206P (L32P) 1.82±0.77 2.43±0.890.035
C255T S208L (S34L) 1.13±0.42 1.55±0.480.022
A272CQ394P (Q48P)N214H (N40H) 4.20±1.72 5.27±2.280.025
A280TR397W (R51W) 2.42±1.20 3.39±1.720.046
G281TR397M (R51M)G217W (G43W) 1.61±0.64 1.97±0.810.043
A293CH401P (H55P)T221P (T46P) 2.30±1.41 3.05±1.740.013
A333G H234R (H60R) 0.92±0.33 1.21±0.270.043
A337GT416A (T70A) 1.02±0.47 1.39±0.270.023
G633A R334K (R160K) 3.76±4.61 5.48±4.550.036
A1138CN683H (N337H) 79.16±40.77 75.86±39.950.005
G1613AR841K 14.8±22.7 19.91±26.600.040
T1753C I127T 11.77±25.07 16.67±26.710.036
A1965C D51A 0.93±0.41 1.33±0.210.036

[i] Values show mean mutation frequency and standard deviation (SD): The mean of the ratio of total mutant clones to total aligned coverage at each nucleotide sites were detected from eleven subjects; nt: nucleotide; aa, amino acid; RT, reverse transcription.

Changes in the predicted antigenicity of HBV protein

To determine the changes in antigenicity, we used the Jameson and Wolf antigenic index prediction model (19). We used DNAStar software to compare the antigenicity between before and early period of treatment for the aa substitutions/mutations in which the variant population occurred in >1% of the total population. The aa variations linked to changes in the secondary structure and their predicted antigenicities are summarized in Table III.

Table III.

Mutation patterns in hepatitis B virus proteins.

Table III.

Mutation patterns in hepatitis B virus proteins.

Antigenic index

Mutation

Author, yearRegionsntaaBefore treatmentEarly period of treatmentDescription(Refs.)
Caligiuri et al, 2016RTA841CN238H↑   3.42.76Drug-resistance mutation(30)
Salpini et al, 2017HBsAgT784GS210R↓−0.4−0.2HCC development(37)
Kim et al, 2004; Saha et al, 2014; Carman et al, 2014HBcAgC2288AP130T↓   1.42.24MHC II(40,51,57)
Kim et al, 2016; Kim et al, 2007; Kim et al 2008HBxG1386A   V5M↑−0.1−0.3HBeAg-negative disease(45,58,59)
Kim et al, 2016; Xie et al, 2014; Li et al, 2015 A1479C   T36P↑−0.25−0.25B cell epitope(45,60,61)
Wu et al, 2010RTC294AH55Q ↑0.160.31Drug-resistance mutation(46)
Fu et al, 2016 T1078GS317A↑   0.6−0.3Drug-resistance mutation(47)
Thedja et al, 2015HBsAgG285A   G44E↑0.15   1.9Non-synonymous mutation(48)
Lee et al, 2013; Mayaphi et al, 2013; Saha et al, 2014;Pre-coreG1896A   W28*↓  0−0.4HCC development(4954)
Liu et al, 1995; Akarca et al, 1995; Kim et al, 2012
Lee et al, 2013; Saha et al, 2014; Yeh et al, 2015; Inoue et al, 2011HBcAgA2189C   I97L↓   0.30.45MHC II(49,51,55,56)
C2289TP130L↑2.241.05

[i] ↑: Increased prevalence of mutation ongoing NUC treatment. ↓: Decreased prevalence of mutation ongoing NUC treatment. -: no previous report nt, nucleotide; aa, amino acid; RT, reverse transcription; HBsAg, hepatitis B virus surface antigen; HBcAg, hepatitis B virus core antigen; HBeAg, hepatitis B e antigen; HBx, hepatitis B virus protein X; HCC, hepatocellular carcinoma; MHC, major histocompatibility complex.

Discussion

HBV infection is one of the most significant medical problems worldwide. It can cause a variety of liver diseases, including acute hepatitis, chronic hepatitis, liver cirrhosis, and HCC. Therapies for CHB are likely to delay the progression of the severe liver disease. Patients with CHB are treated with NUC, which are effective inhibitors of HBV replication. However, few patients achieve HBsAg clearance, and most patients require long-term NUC treatment. In this study, adefovir and entecavir were selected by the doctor's decision as both of them were first line drug for CHB.

Adefovir belongs to acyclic phosphate class. The structural similarity of adefovir to its natural substrate dATP, together with its small flexible linker, allows it accessibility of adefovir to the polymerase (20). The primary mutations associated with adefovir resistance are rtA181T/V, rtI233V, and rtN236T. The secondary mutations include rtV84M, rtS85A, rtV214A, rtQ215S, rtN238T/D (57). The rtN236T mutation may disrupt the hydrogen bond between rtN236 and rtS85 and rtN236 and adefovir diphosphate, thereby decreasing the binding affinity for adefovir. Similarly, other mutations (rtP237H, rtN238T/D, rtV84M, and rtS85A) around this region may also disrupt the triphosphate binding site. The mutation rtA181V/T in domain B may bend the α helices through hydrogen bond formation, and drug resistance may thus result from an allosteric change in the catalytic site. Other mutations (rtV214A, rtQ215S) do not directly interact with the nucleotide but may cause allosteric changes affecting the polymerase function (2123).

Entecavir belongs to the cyclopentane class and has potent and selective activity against HBV. Two resistance patterns have been recognized: rtM250V + rtI169T + M204V + L180M and rt184G + rtS202I + rtM204V + rtL180M. The requirement of multiple mutations for entecavir resistance suggests a higher genetic barrier to resistance. The rt250 mutation appears to alter the binding interaction between the primer strand and the template DNA. The rt169 interacts with the dNTP and may work cooperatively with the mutation at rt250. The combination of these mutations may affect elongation of the DNA strand. The rtT184 G and rtS202I mutations interact and affect the hydrophobic core region between the α helix of the DNA template binding region and the β sheet, which encodes the 2 catalytic aspartate residues in the C domain. The substitutions at residues rt184 and rt202 are predicted to magnify the impact of the lamivudine-resistant mutation in the YMDD loop, affecting substrate discrimination upon nucleotide loading and addition to the primer terminus. In the absence of the lamivudine-resistant mutations, the rtT184 G and rtS202I mutations do not affect the susceptibility of HBV to entecavir (23).

Although these two therapy show the different of mutagenicity, the potency of ADV treatment on HBsAg reduction is not significantly different with that of ETV treatment (24).

The present study revealed significant reductions in the HBV DNA titer and ALT level during NUC treatment (Table I). However, long-term NUC therapy could lead to the emergence of drug resistance and multidrug resistance mutations in the HBV genome. From previous studies (510), the well-known resistance mutations in RT region and deletion mutations were not found in this study. The present study should provide further insight into the molecular mechanism of drug resistance in early duration of treatment.

It is perhaps unsurprising that the difference in variation between before and early duration of treatment was greatest in the RT region, a target of NUC therapy. But it is still interesting that the variation in the major population increased during early period of treatment after analyzed the subdomain of RT region. The polymerase gene contains four domains (TP, Spacer, RT, and RH domains). RT plays an important role in viral replication. Because RT lacks a proofreading ability, the error rate and high viral replication rate (1011 virions per day) contribute to the high mutation rate (10−5 per nt per generation) (25,26). The RT region is divided into seven boxes (A-F). The RT1 motif, which contains boxes G and F (aa24-36 and aa37-47, respectively), is required for Hε binding (3,26). It was suggested that mutations in the RT1 motif, such as rtV30/rtD31, may disrupt protein priming and affect viral replication (2729). In this study, it is suggested that RT1 motif region is more sensitive to NUC treatment and might be a crucial factor for inhibiting the RT activity and suppress HBV replication (Table I) in the early period of NUC treatment.

There are two types of mutations associated with NUC resistance: Primary resistance mutations, which are directly responsible for drug resistance, and secondary mutations, which promote or enhance the replication competence of resistant strains. Several well-known mutations, particularly rtM204I/V (primary resistance mutations) and rtL80V/I (secondary mutations), which occur following long-term treatment and increase viral replication, were not detected in this study. However, we did detect several mutations, notably rtH55Q and rtN238H, which are frequently generated during long-term NUC treatment and are associated with M204I/V (30).

The treatment of CHB with all the by NUCs could result in changes to HBsAg, such as rtA181T/sW172*. In vitro analysis of rtA181T/sW172* HBV has shown secretion defect of the viral particles to extracellular, resulting in intracellular retention of surface proteins and lower viral loads detected extracellularly (3133). As shown in this study, the changes in RT 1 motif of polymerase gene can affect small S gene of surface gene because it overlapped with each other. Therefore, the mutation in small S gene can affect the synthesis and secretion of HBsAg that become undetectable by current commercially available kit, resulting in ‘false’ decrease of HBsAg. In addition, major HBsAg are encoded by small S gene (the large and middle S proteins are expressed at levels of about 5–15 and 1–2% compared with small S protein) (34) and includes the regions involved in the attachment of the virus to hepatocytes as well as the main epitopes recognized by neutralizing antibodies. The aa changes in the surface gene may lead to the adaptive evolution of the viral gene, resulting in altered antigenicity (35). The polymerase gene completely overlaps the HBV surface gene. It is suggested that changes in the S region where overlapped with polymerase gene (RT1 motif) might decrease the synthesis and secretion of HBsAg. Mutations in the polymerase gene decrease replication activity, but some double mutations in the polymerase gene and the ‘a’ determinant region actually increase replication activity (36). In our study, we found no double mutations in polymerase and the ‘a’ determinant region. However, several secondary resistance mutations were detected in the MHR region, notably sG44E. In addition, the variation in S210R which has been suggested that long-term therapy is associated with reduced cell cycle activity and liver carcinogenesis (37) was detected from early period of treatment. It is showed that it might change the secondary structure and then alter the antigenicity of HBsAg in the early period of treatment.

In this study, the proportion of rtH55Q, rtN238H, rtS317A, sG44E showed an increasing trend during early period of treatment as potential drug resistance mutations in the early duration of treatment. It was considered that these drug resistance mutations may enhance viral replication activity and thus affect the clinical outcome with the persist treatment; sS210R with decreased proportion may indirectly contribute to the activity of these mutations. Therefore, it was suggested that these substitutions should be monitored since the early period of treatment.

The double A1762T and G1764A mutations in the basal core promoter region are associated with altered expression of HBeAg and viral replication (38). However, the impact of basal core promoter and precore mutations on viral replication remains controversial (39). In this study, the variations A1762T/G1764A were found in 9/11 patients (81.8%), but there were no significant differences in these variants between before and early period of treatment. The most prevalent mutation in the precore region was G1896A (W28*), which introduces a translational stop codon W28* in the HBeAg sequence, and halts protein synthesis. A recent meta-analysis indicated that the precore mutation G1896A was significantly associated with increased risk of HCC and progression of liver disease, especially in Asians (40). In the early period of treatment, W28* was detected from three subjects. In other words, NUC treatment could protect or delay the progression to severe. The variations I97L and P130T in the core regions were decreased by NUC treatment. These mutations are located in the MHC class II restricted region and are significantly associated with the pathogenesis of HCC owing to the loss of HBcAg assembly (41,42). It was also reported that the mutation P130T is strongly correlated with the mutation I97L (43,44). It was suggested that these mutations may affect the secondary structure of HBcAg and increased its immunogenicity, attenuating liver carcinogenesis. cP130L with increased proportion more than 15% during early period of treatment was detected from this study. It was inferred that cP130L might alter the secondary structure of HBcAg and lead to reduce the antigenicity of HBcAg according to predicted antigenicity (Fig. 3). Subsequently, HBcAg could not be recognized by immune system resulting in immune escape. Therefore, it was considered that cP130L might be one of immune escape variants resulting in the persistence of HBV. The identification of this hypothesis should be supported by further experiments and more data.

HBx is a multifunctional protein involved in viral replication and affects host gene expression related to HCC (45). HBx also plays important roles in the establishment and maintenance of chronic infection by altering the host's innate immunity. In this study, there were trends towards increased proportion of V5M and T36P following NUC therapy, and these mutations may result in the alteration of the secondary structure of HBx. V5M is common in HCC patients, which suggested that this mutation was a genuine HCC-specific mutation that may play a key role in the progression from chronic hepatitis or liver cirrhosis to HCC (45). By contrast, T36P, which is located in the B cell epitope region, may lead to persistent infection by providing the virus with a mechanism to evade the host's humoral immune response. In this study, increased variations in V5M or T36P were found from three patients. It is unknown whether the risk of liver cancer is increased due to the aa variations. Regular follow-up and monitoring the progression of liver diseases is important to predict the clinical outcome of patients with CHB.

This study has some limitations. Only a small number of serum samples could be analyzed because of the high cost of the technique and complexity of the analysis for the HBV whole genome. In addition, it is still unclear that resistant mutations will generate after long-term NUC treatment among these patients. These limitations will be solved by long-term follow up and further studies examining additional patients.

In summary, although the HBV whole genome can be affected by NUC treatment, RT 1 motif and small S region are more sensitive to the early period of NUC treatment. This study suggested the initial changes of HBV quasispecies might affect the long-term drug sensitivity and resistance to NUC treatment.

The whole regions of HBV were affected by NUC treatment. Several variations were detected in quasispecies at several positions of the HBV genome. RT1 motif region and small S region are susceptible in the early period of NUC treatment resulting in inhibiting RT activity and synthesis and secretion of HBsAg. The initial changes of HBV quasispecies might be related to the long-term drug sensitivity and resistance. In this study, it is suggested that B and C subdomains in RT region should not only attached importance to as drug resistance mutations found, others subdomains such as G and F also should be followed during NUC treatment especially early-period of treatment. These results will be enable us to improve the development of diagnostic tools and establish more efficient therapeutic protocols for CHB.

Acknowledgements

The authors thank all of the subjects who participated in this research. We also thank the medical staff for helping with performing the study, for their invaluable advice on technical aspects and/or data analysis, and for sample preparation. This study was funded by a Grant-in-Aid from the Ministry of Education, Culture, Sports, Science and Technology, Japan (16H05826), and a Grant-in-Aid from the Japan Initiative for Global Research Network on Infectious Disease (J-GRID) supported by The Ministry of Education, Culture, Sports, Science and Technology, Japan.

Glossary

Abbreviations

Abbreviations:

HBV

hepatitis B virus

NUC

nucleos(t)ide analogues

CHB

chronic hepatitis B

References

1 

World Health Organization (WHO), . Global Hepatitis Report, 2017. WHO; Geneva: 2017, http://apps.who.int/iris/bitstream/10665/255016/1/9789241565455-eng.pdf?ua=1April 21–2017

2 

World Health Organization, . Guidelines for the Prevention, Care and Treatment of Persons with Chronic Hepatitis B Infection, 2015. WHO, Geneva: 2015, http://apps.who.int/iris/bitstream/10665/154590/1/9789241549059_eng.pdf?ua=1&ua=1March. 2015

3 

Menéndez-Arias L, Álvarez M and Pacheco B: Nucleoside/nucleotide analog inhibitors of hepatitis B virus polymerase: Mechanism of action and resistance. Curr Opin Virol. 8:1–9. 2014. View Article : Google Scholar : PubMed/NCBI

4 

Buti M, Tsai N, Petersen J, Flisiak R, Gurel S, Krastev Z, Aguilar SR, Flaherty JF, Martins EB, Charuworn P, et al: Seven-year efficacy and safety of treatment with tenofovir disoproxil fumarate for chronic hepatitis B virus infection. Dig Dis Sci. 60:1457–1464. 2015. View Article : Google Scholar : PubMed/NCBI

5 

Lei J, Wang Y, Wang LL, Zhang SJ, Chen W, Bai ZG and Xu LY: Profile of hepatitis B virus resistance mutations against nucleoside/nucleotide analogue treatment in Chinese patients with chronic hepatitis B. Virol J. 10:3132013. View Article : Google Scholar : PubMed/NCBI

6 

Rawal RK, Konreddy AK and Chu CK: Mechanism of adefovir, tenofovir and entecavir resistance: Molecular modeling studies of how a novel anti-HBV agent (FMCA) can overcome the drug resistance. Curr Med Chem. 22:3922–3932. 2015. View Article : Google Scholar : PubMed/NCBI

7 

Margeridon-Thermet S, Shulman NS, Ahmed A, Shahriar R, Liu T, Wang C, Holmes SP, Babrzadeh F, Gharizadeh B, Hanczaruk B, et al: Ultra-deep pyrosequencing of hepatitis B virus quasispecies from nucleoside and nucleotide reverse-transcriptase inhibitor (NRTI)-treated patients and NRTI-naive patients. J Infect Dis. 199:1275–1285. 2009. View Article : Google Scholar : PubMed/NCBI

8 

Ji F, Zhou L, Ma S, Li F, Ding H and Zeng C: Dynamic changes of HBV quasispecies and deletion patterns in a chronic hepatitis B patient. J Med Virol. 81:1551–1559. 2009. View Article : Google Scholar : PubMed/NCBI

9 

Günther S, Piwon N, Jung A, Iwanska A, Schmitz H and Will H: Enhanced replication contributes to enrichment of hepatitis B virus with a deletion in the core gene. Virology. 273:286–299. 2000. View Article : Google Scholar : PubMed/NCBI

10 

Zhang D, Dong P, Zhang K, Deng L, Bach C, Chen W, Li F, Protzer U, Ding H and Zeng C: Whole genome HBV deletion profiles and the accumulation of preS deletion mutant during antiviral treatment. BMC Microbiol. 12:3072012. View Article : Google Scholar : PubMed/NCBI

11 

Tsuge M, Murakami E, Imamura M, Abe H, Miki D, Hiraga N, Takahashi S, Ochi H, Nelson HC, Ginba H, et al: Serum HBV RNA and HBeAg are useful markers for the safe discontinuation of nucleotide analogue treatments in chronic hepatitis B patients. J Gastroenterol. 48:1188–1204. 2013. View Article : Google Scholar : PubMed/NCBI

12 

Chevaliez S, Rodriguez C and Pawlotsky JM: New virologic tools for management of chronic hepatitis B and C. Gastroenterology. 142:1303–1313.e1. 2012. View Article : Google Scholar : PubMed/NCBI

13 

Barzon L, Lavezzo E, Costanzi G, Franchin E, Toppo S and Palù G: Next-generation sequencing technologies in diagnostic virology. J Clin Virol. 58:346–350. 2013. View Article : Google Scholar : PubMed/NCBI

14 

Yamani LN, Yano Y, Utsumi T, Juniastuti, Wandono H, Widjanarko D, Triantanoe A, Wasityastuti W, Liang YJ, Okada R, et al: Ultradeep sequencing for detection of quasispecies variants in the major hydrophilic region of hepatitis B virus in indonesian patients. J Clin Microbiol. 53:3165–3175. 2015. View Article : Google Scholar : PubMed/NCBI

15 

Capobianchi MR, Giombini E and Rozera G: Next-generation sequencing technology in clinical virology. Clin Microbiol Infect. 19:15–22. 2013. View Article : Google Scholar : PubMed/NCBI

16 

Zoulim F and Locarnini S: Hepatitis B virus resistance to nucleos(t)ide analogues. Gastroenterology. 137:1593–1608.e1-2. 2009. View Article : Google Scholar : PubMed/NCBI

17 

Shinkai N, Matsuura K, Sugauchi F, Watanabe T, Murakami S, Iio E, Ogawa S, Nojiri S, Joh T and Tanaka Y: Application of a newly developed high-sensitivity HBsAg chemiluminescent enzyme immunoassay for hepatitis B patients with HBsAg seroclearance. J Clin Microbiol. 51:3484–3491. 2013. View Article : Google Scholar : PubMed/NCBI

18 

Weiss J, Wu H, Farrenkopf B, Schultz T, Song G, Shah S and Siegel J: Real time TaqMan PCR detection and quantitation of HBV genotypes A-G with the use of an internal quantitation standard. J Clin Virol. 30:86–93. 2004. View Article : Google Scholar : PubMed/NCBI

19 

Jameson BA and Wolf H: The antigenic index: A novel algorithm for predicting antigenic determinants. Bioinformatics. 4:181–186. 1988. View Article : Google Scholar

20 

Westland CE, Yang H, Delaney WE IV, Gibbs CS, Miller MD, Wulfsohn M, Fry J, Brosgart CL and Xiong S; 437 and 438 Study Teams, : Week 48 resistance surveillance in two phase 3 clinical studies of adefovir dipivoxil for chronic hepatitis B. Hepatology. 38:96–103. 2003. View Article : Google Scholar : PubMed/NCBI

21 

Yadav V and Chu CK: Molecular mechanisms of adefovir sensitivity and resistance in HBV polymerase mutants: A molecular dynamics study. Bioorganic Med Chem Lett. 14:4313–4317. 2004. View Article : Google Scholar

22 

Ballesteros JA, Deupi X, Olivella M, Haaksma EE and Pardo L: Serine and threonine residues bend alpha-helices in the chi(1)=g(−) conformation. Biophys J. 79:2754–2760. 2000. View Article : Google Scholar : PubMed/NCBI

23 

Ghany M and Liang TJ: Drug targets and molecular mechanisms of drug resistance in chronic hepatitis B. Gastroenterology. 132:1574–1585. 2007. View Article : Google Scholar : PubMed/NCBI

24 

Li MR, Xi HL, Wang QH, Hou FQ, Huo N, Zhang XX, Li F and Xu XY: Kinetics and prediction of HBsAg loss during long-term therapy with nucleos(t)ide analogues of different potency in patients with chronic hepatitis B. PLoS One. 9:e984762014. View Article : Google Scholar : PubMed/NCBI

25 

Xu J, Wu B, Wang JH, Huang L, Wang DY, Zhao L, Zhao GP and Wang Y: Pre-existing mutations in reverse transcriptase of hepatitis B virus in treatment-naive Chinese patients with chronic hepatitis B. PLoS One. 10:e01174292015. View Article : Google Scholar : PubMed/NCBI

26 

Nowak MA, Bonhoeffer S, Hill AM, Boehme R, Thomas HC and McDade H: Viral dynamics in hepatitis B virus infection. Proc Natl Acad Sci USA. 93:pp. 4398–4402. 1996; View Article : Google Scholar : PubMed/NCBI

27 

Jones SA, Clark DN, Cao F, Tavis JE and Hu J: Comparative analysis of hepatitis B virus polymerase sequences required for viral RNA binding, RNA packaging, and protein priming. J Virol. 88:1564–1572. 2014. View Article : Google Scholar : PubMed/NCBI

28 

Badtke MP, Khan I, Cao F, Hu J and Tavis J: An interdomain RNA binding site on the hepadnaviral polymerase that is essential for reverse transcription. Virology. 390:130–138. 2009. View Article : Google Scholar : PubMed/NCBI

29 

Clark DN and Hu J: Unveiling the roles of HBV polymerase for new antiviral strategies. Future Virol. 10:283–295. 2015. View Article : Google Scholar : PubMed/NCBI

30 

Caligiuri P, Cerruti R, Icardi G and Bruzzone B: Overview of hepatitis B virus mutations and their implications in the management of infection. World J Gastroenterol. 22:145–154. 2016. View Article : Google Scholar : PubMed/NCBI

31 

Locarnini SA: Hepatitis B virus surface antigen and polymerase gene variants: Potential virological and clinical significance. Hepatology. 27:294–297. 1998. View Article : Google Scholar : PubMed/NCBI

32 

Locarnini SA and Yuen L: Molecular genesis of drug-resistant and vaccine-escape HBV mutants. Antivir Ther. 15:451–461. 2010. View Article : Google Scholar : PubMed/NCBI

33 

Warner N and Locarnini S: The antiviral drug selected hepatitis B virus rtA181T/sW172* mutant has a dominant negative secretion defect and alters the typical profile of viral rebound. Hepatology. 48:88–98. 2008. View Article : Google Scholar : PubMed/NCBI

34 

World Health Organization (WHO), . Hepatitis B. WHO; Geneva: 2002, http://apps.who.int/iris/bitstream/10665/67746/1/WHO_CDS_CSR_LYO_2002.2_HEPATITIS_B.pdf

35 

Shi Y, Wei F, Hu D, Li Q, Smith D, Li N and Chen D: Mutations in the major hydrophilic region (MHR) of hepatitis B virus genotype C in North China. J Med Virol. 84:1901–1906. 2012. View Article : Google Scholar : PubMed/NCBI

36 

MAP: Hepatitis B virus S gene escape mutants. Asian J Transfus Sci. 1:62–70. 2007. View Article : Google Scholar : PubMed/NCBI

37 

Salpini R, Surdo M, Warner N, Cortese MF, Colledge D, Soppe S, Bellocchi MC, Armenia D, Carioti L, Continenza F, et al: Novel HBsAg mutations correlate with hepatocellular carcinoma, hamper HBsAg secretion and promote cell proliferation in vitro. Oncotarget. 8:15704–15715. 2017. View Article : Google Scholar : PubMed/NCBI

38 

Silva Souza ACD, Souza Marasca G, Kretzmann-Filho NA, Dall-Bello A, Alexandre Kliemann D, Valle Tovo C and Gorini da Veiga AB: Identification of hepatitis B virus A1762T/G1764A double mutant strain in patients in Southern Brazil. Braz J Infect Dis. 21:525–529. 2017. View Article : Google Scholar : PubMed/NCBI

39 

Zhang ZH, Wu CC, Chen XW, Li X, Li J and Lu MJ: Genetic variation of hepatitis B virus and its significance for pathogenesis. World J Gastroenterol. 22:126–144. 2016. View Article : Google Scholar : PubMed/NCBI

40 

Kim H, Lee SA, Do SY and Kim BJ: Precore/core region mutations of hepatitis B virus related to clinical severity. World J Gastroenterol. 22:4287–4296. 2016. View Article : Google Scholar : PubMed/NCBI

41 

Ceres P, Stray SJ and Zlotnick A: Hepatitis B virus capsid assembly is enhanced by naturally occurring mutation F97L. J Virol. 78:9538–9543. 2004. View Article : Google Scholar : PubMed/NCBI

42 

Ning B and Shih C: Nucleolar localization of human hepatitis B virus capsid protein. J Virol. 78:13653–13668. 2004. View Article : Google Scholar : PubMed/NCBI

43 

Hosono S, Tai PC, Wang W, Ambrose M, Hwang DG, Yuan TT, Peng BH, Yang CS, Lee CS and Shih C: Core antigen mutations of human hepatitis B virus in hepatomas accumulate in MHC class II-restricted T cell epitopes. Virology. 212:151–162. 1995. View Article : Google Scholar : PubMed/NCBI

44 

Cheng CP, Lee PF, Liu WC, Wu IC, Chin CY, Chang TT and Tseng VS: Analysis of precore/core covariances associated with viral kinetics and genotypes in hepatitis B e antigen-positive chronic hepatitis B patients. PLoS One. 7:e325532012. View Article : Google Scholar : PubMed/NCBI

45 

Kim H, Lee SA and Kim BJ: X region mutations of hepatitis B virus related to clinical severity. World J Gastroenterol. 22:5467–5478. 2016. View Article : Google Scholar : PubMed/NCBI

46 

Wu Z, Han T, Gao Y, Li Y, Liu T, Jing L, Liu L, Du Z, Wu Z and Medical T: Evolution of hepatitis B virus quasispecies during sequential therapy in patients with lamivudine-resistant hepatitis B virus infection. Shijie Huaren Xiaohua Zazhi. 18:779–785. 2010.

47 

Fu L, Teng X, Ma Y, Chen S, Xu W, Fu B and Gu H: HBV gene mutations in six multidrug-resistant chronic hepatitis B patients in China. Int J Clin Exp Pathol. 9:2134–2140. 2016.

48 

Thedja MD, Muljono DH, Ie SI, Sidarta E, Turyadi Verhoef J and Marzuki S: Genogeography and immune epitope characteristics of hepatitis B virus genotype c reveals two distinct types: Asian and papua-pacific. PLoS One. 10:e01325332015. View Article : Google Scholar : PubMed/NCBI

49 

Lee SA, Kim KJ, Kim DW and Kim BJ: Male-specific W4P/R mutation in the pre-S1 region of hepatitis B virus, increasing the risk of progression of liver diseases in chronic patients. J Clin Microbiol. 51:3928–3936. 2013. View Article : Google Scholar : PubMed/NCBI

50 

Mayaphi SH, Martin DJ, Mphahlele MJ, Blackard JT and Bowyer SM: Variability of the preC/C region of hepatitis B virus genotype A from a South African cohort predominantly infected with HIV. J Med Virol. 85:1883–1892. 2013. View Article : Google Scholar : PubMed/NCBI

51 

Saha D, Pal A, Biswas A, Panigrahi R, Sarkar N, Das D, Sarkar J, Guha SK, Saha B, Chakrabarti S, et al: Molecular characterization of HBV strains circulating among the treatment-naive HIV/HBV co-infected patients of eastern India. PLoS One. 9:e904322014. View Article : Google Scholar : PubMed/NCBI

52 

Liu Z, Luo K, He H and Hou J: Hot-spot mutations in hepatitis B virus core gene: Eliciting or evading immune clearance? J Viral Hepat. 12:146–153. 2005. View Article : Google Scholar : PubMed/NCBI

53 

Akarca US and Lok AS: Naturally occurring hepatitis B virus core gene mutations. Hepatology. 22:50–60. 1995. View Article : Google Scholar : PubMed/NCBI

54 

Kim DW, Lee SA, Hwang ES, Kook YH and Kim BJ: Naturally occurring precore/core region mutations of hepatitis B virus genotype C related to hepatocellular carcinoma. PLoS One. 7:e473722012. View Article : Google Scholar : PubMed/NCBI

55 

Yeh ML, Peng CY, Dai CY, Lai HC, Huang CF, Hsieh MY, Huang JF, Chen SC, Lin ZY, Yu ML, et al: Pegylated-interferon alpha therapy for treatment-experienced chronic hepatitis B patients. PLoS One. 10:e01222592015. View Article : Google Scholar : PubMed/NCBI

56 

Inoue J, Ueno Y, Wakui Y, Fukushima K, Kondo Y, Kakazu E, Ninomiya M, Niitsuma H and Shimosegawa T: Enhanced replication of hepatitis B virus with frameshift in the precore region found in fulminant hepatitis patients. J Infect Dis. 204:1017–1025. 2011. View Article : Google Scholar : PubMed/NCBI

57 

Carman WF, Jacyna MR, Hadziyannis S, Karayiannis P, McGarvey MJ, Makris A and Thomas HC: Mutation preventing formation of hepatitis B e antigen in patients with chronic hepatitis B infection. Lancet. 2:588–591. 1989. View Article : Google Scholar : PubMed/NCBI

58 

Kim H, Jee YM, Song BC, Shin JW, Yang SH, Mun HS, Kim HJ, Oh EJ, Yoon JH, Kim YJ, et al: Molecular epidemiology of hepatitis B virus (HBV) genotypes and serotypes in patients with chronic HBV infection in Korea. Intervirology. 50:52–57. 2007. View Article : Google Scholar : PubMed/NCBI

59 

Kim HJ, Park JH, Jee Y, Lee SA, Kim H, Song BC, Yang S, Lee M, Yoon JH, Kim YJ, et al: Hepatitis B virus X mutations occurring naturally associated with clinical severity of liver disease among Korean patients with chronic genotype C infection. J Med Virol. 80:1337–1343. 2008. View Article : Google Scholar : PubMed/NCBI

60 

Xie Y, Liu S, Zhao Y, Guo Z and Xu J: X protein mutations in hepatitis B virus DNA predict postoperative survival in hepatocellular carcinoma. Tumor Biol. 35:10325–10331. 2014. View Article : Google Scholar

61 

Li W, Goto K, Matsubara Y, Ito S, Muroyama R, Li Q and Kato N: The characteristic changes in hepatitis B virus × region for hepatocellular carcinoma: A comprehensive analysis based on global data. PLoS One. 10:e01255552015. View Article : Google Scholar : PubMed/NCBI

Related Articles

Journal Cover

April-2018
Volume 17 Issue 4

Print ISSN: 1791-2997
Online ISSN:1791-3004

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Liang Y, Yano Y, Putri WA, Mardian Y, Okada R, Tanahashi T, Murakami Y and Hayashi Y: Early changes in quasispecies variant after antiviral therapy for chronic hepatitis B. Mol Med Rep 17: 5528-5537, 2018
APA
Liang, Y., Yano, Y., Putri, W.A., Mardian, Y., Okada, R., Tanahashi, T. ... Hayashi, Y. (2018). Early changes in quasispecies variant after antiviral therapy for chronic hepatitis B. Molecular Medicine Reports, 17, 5528-5537. https://doi.org/10.3892/mmr.2018.8551
MLA
Liang, Y., Yano, Y., Putri, W. A., Mardian, Y., Okada, R., Tanahashi, T., Murakami, Y., Hayashi, Y."Early changes in quasispecies variant after antiviral therapy for chronic hepatitis B". Molecular Medicine Reports 17.4 (2018): 5528-5537.
Chicago
Liang, Y., Yano, Y., Putri, W. A., Mardian, Y., Okada, R., Tanahashi, T., Murakami, Y., Hayashi, Y."Early changes in quasispecies variant after antiviral therapy for chronic hepatitis B". Molecular Medicine Reports 17, no. 4 (2018): 5528-5537. https://doi.org/10.3892/mmr.2018.8551