You are on page 1of 20

PRIMER

Hepatitis B virus infection


Man-Fung Yuen1,2*, Ding-Shinn Chen3,4, Geoffrey M. Dusheiko5, Harry L. A. Janssen6,
Daryl T. Y. Lau7, Stephen A. Locarnini8, Marion G. Peters9 and Ching-Lung Lai1,2
Abstract | Hepatitis B virus (HBV) is a hepatotropic virus that can establish a persistent and chronic
infection in humans through immune anergy. Currently, 3.5% of the global population is chronically
infected with HBV, although the incidence of HBV infections is decreasing owing to vaccination and,
to a lesser extent, the use of antiviral therapy to reduce the viral load of chronically infected
individuals. The course of chronic HBV infection typically comprises different clinical phases, each of
which potentially lasts for decades. Well-defined and verified serum and liver biopsy diagnostic
markers enable the assessment of disease severity, viral replication status, patient risk stratification
and treatment decisions. Current therapy includes antiviral agents that directly act on viral
replication and immunomodulators, such as interferon therapy. Antiviral agents for HBV include
reverse transcriptase inhibitors, which are nucleoside or nucleotide analogues that can profoundly
suppress HBV replication but require long-term maintenance therapy. Novel compounds are being
actively investigated to achieve the goal of HBV surface antigen seroclearance (functional cure),
a serological state that is associated with a higher remission rate (thus, no viral rebound) after
treatment cessation and a lower rate of cirrhosis and hepatocellular carcinoma. This Primer addresses
several aspects of HBV infection, including epidemiology, immune pathophysiology, diagnosis,
prevention and management.

Hepatitis B virus (HBV), a partially double-stranded HBV infection is managed with reverse transcriptase
hepatotropic DNA virus, is the aetiological agent (BOX 1; inhibitors, which are nucleoside or nucleotide analogues
FIG.  1) of acute and chronic hepatitis B in humans. (NUCs), and, less commonly interferon (IFN) therapy.
In countries where HBV prevalence is high, infection The total elimination of HBV from patients (HBV cure)
usually occurs early in life, leading to lifelong chronic remains a remote treatment goal, but a more practi-
infection and carriage of the virus, whereas infection later cal goal is the achievement of HBsAg seroclearance
in life usually leads to an acute, self-resolving illness fol- (functional cure) as early as possible after infection.
lowed by viral clearance or, in rare fulminant cases, liver Although HBsAg seroclearance occurs in only a small
failure. Chronic hepatitis B infection is defined as the proportion of patients who have undergone treatment
detection of serum hepatitis B surface antigen (HBsAg; with NUCs, it is associated with a low rate of clinical
the viral glycoprotein) after 6 months of infection. The complications and a high chance of disease remission
transmission route of HBV is primarily through blood after treatment cessation.
and bodily fluids and includes perinatal and early infant Vaccination remains the most effective tool to pre-
transmission (mother to child t­ ransmission (MTCT)) vent HBV infection. HBsAg produced in yeast has
as well as sexual and parenteral modes. been engineered as an effective recombinant vaccine
Chronic HBV infection is a major public health prob- for HBV and was licensed for use in the United States
lem. In chronic HBV infection, there may be ongoing in 1986. An updated generation of HBV vaccines has
low-grade liver inflammation, with episodes of transient been developed that consists of genetically engineered
high-grade liver inflammation and activation of fibro- viral proteins, which can achieve a primary neutral-
genic processes, leading to liver fibrosis and cirrhosis, izing antibody response after vaccination in >95% of
which may culminate in decompensated (symptomatic) individuals2. In this Primer, we cover the mechanisms
liver disease and/or the development of hepatocellular and pathophysio­logy of HBV infection. Furthermore,
*e-mail:
mfyuen@hkucc.hku.hk carcinoma (HCC) in 25–40% of HBV carriers1. In addi- we consider the epidemiology, diagnosis, prevention
tion, HBV infection has further oncogenic potential and management of infection from a global perspective.
Article number: 18035
doi:10.1038/nrdp.2018.35 after HBV integration into the host genome, which is an Finally, we discuss promising future research directions
Published online 7 Jun 2018 ­additional pathway contributing to HCC. and new drug development.

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 18035 | 1


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Author addresses to low prevalence countries9,10. Moreover, mass immigra-


tion has been a constant feature in recent history, owing
1
Department of Medicine, The University of Hong Kong, Queen Mary Hospital, Hong Kong. to the globalization of the economy, increasing air and
2
State Key Laboratory of Liver Research, The University of Hong Kong, Queen Mary sea travel, and persistent military or political conflicts
Hospital, Hong Kong. in different parts of the world. After immigration to a
3
Hepatitis Research Center, National Taiwan University Hospital, Department of Internal
new country, many patients with chronic HBV infection
Medicine, National Taiwan University College of Medicine, Taipei, Taiwan.
4
Genomics Research Center, Academia Sinica, Taipei, Taiwan. remain in minority groups with suboptimal health care,
5
University College London Medical School and Kings College Hospital, London, UK. which can hinder the control of HBV infection. Thus,
6
Toronto Center for Liver Diseases, Toronto General Hospital, University Health Network, in countries with high levels of immigration, a hepati-
University of Toronto, Toronto, Ontario, Canada. tis B control programme needs to address the related
7
Division of Gastroenterology, Beth Israel Deaconess Medical Center, Harvard Medical cultural, social and logistical hurdles to case f­ inding
School, Boston, MA, USA. and referral9–11.
8
Victorian Infectious Diseases Reference Laboratory, Melbourne, Victoria, Australia.
9
Department of Gastroenterology, University of California, San Francisco, CA, USA. HBV genotype distribution
There are ten different genotypes of HBV (A–J), and
Epidemiology the geographical distribution of each HBV genotype is
Global prevalence distinct 12,13 (FIG. 2). Infection with different genotypes of
The distribution of chronic HBV infection varies substan- HBV is associated with different outcomes in chroni­
tially in the world (FIG. 2). The prevalence of chronic HBV city after infection, disease progression and responses
infection is highest in the Western Pacific (6.2% or 115 to IFNα treatment; however, the approved HBV vaccines
million individuals have a chronic HBV infection) and are effective against all genotypes12,13. The characteriza-
African regions (6.1% or 60 million individuals). The tion of different HBV genotypes in given populations
Eastern Mediterranean region has fairly high prevalence may have epidemiological importance, as HBV genotype
(3.3% or 21 million individuals); prevalence is lower in generally reflects its country of origin and, therefore,
the South–East Asian (2.0% or 39 million individ­uals) and may be used to track patterns of transmission; as such,
European regions (1.6% or 15 million i­ ndividuals) and is an immigrant’s HBV genotype is generally consistent
lowest in the North and South American regions (0.7% with their country of origin14.
or 7 million individuals)3. These prevalence data are
instrumental in estimating the disease burden of HBV Long-term complications and mortality
and in guiding regional health policies. The prevalence On average, ~25% of untreated individuals with chronic
of chronic HBV infection has been recently meticulously HBV infection will die of cirrhosis complications and/or
reviewed and estimated4. According to the most recent HCC; this number increases to 50% if only men are
European report 5, prevalence of chronic HBV infection taken into account 15. However, there is a lack of recent
varies between different parts of Europe with the follow- studies on the natural history of untreated HBV infec-
ing figures: <1.5% in Northern Europe, <2% in Southern tion, as it is unethical to deny treatment for the virus.
Europe, <1% in Western Europe and <5% in Eastern The incidence of long-term sequelae of hepatitis B,
Europe (except for Uzbekistan: 8%). There has been a which are cirrhosis and HCC, generally parallels the
steady decline of prevalence of chronic HBV infection in occurrence of chronic HBV infection16. The incidence
European countries overall in the past 30 years. of the cirrhosis and HCC can be difficult to assess owing
The epidemiology of HBV infection has evolved sub- to the slow evolution of disease from chronic hepatitis to
stantially since the 1980s owing to the widely used hep- cirrhosis and HCC. Furthermore, cirrhosis and HCC
atitis B vaccine and demographic changes as a result of are asymptomatic in the early stage17,18 and currently
population migration6,7. However, despite the advent of lack good non­invasive biomarkers, making diagnosis
effective vaccines, as well as other disease control meas- difficult. Moreover, the aetiology of cirrhosis and HCC
ures, including antiviral therapy, HBV infection is still is heterogeneous worldwide19,20, adding further diffi-
a serious global health problem. The WHO estimated culties to the estimation of the incidence of liver dis-
in 2015 that 257 million individuals (3.5% of the global ease caused by a given specific aetiology. Nevertheless,
population) are chronically infected with HBV3, most as cirrhosis and HCC will ultimately mani­fest with
of whom were born before the widespread use of the decompensated liver disease or death, the frequency
vaccine3,6,7. Numbers of infected children have declined of HBV complications can be extrapolated from mor-
since the introduction of the expanded vaccination pro- bidity and mortality data. According to the WHO3,
gramme promoted by the WHO; in 2015, only ~1.3% mortality due to viral hepatitis is increasing. In 2000,
of children <5 years of age developed chronic infection 1.1 million individuals died of viral hepatitis globally;
worldwide3 (FIG. 2b). The WHO African region still has a this number increased by 22% to 1.3 million individuals
high prevalence of chronic hepatitis B infection in chil- in 2015. In 2015, approximately 0.8 million deaths were
dren, owing to the low coverage of HBV vaccination and ­attributed to HBV infection.
the high viral load (and, therefore, increased ­infectivity) In a large systematic analysis for the Global Burden
in pregnant mothers (discussed below)8. of Disease Study 2010 (REF.21), deaths at all ages associ-
The prevalence of infection in different regions has ated with acute HBV were 68,600 in 1990 and 132,200
also been affected by immigration, owing to the trend of in 2010. For cirrhosis, deaths at all ages attributable to
individuals from HBV-endemic countries immigrating HBV were 241,700 in 1990 and 312,400 in 2010, and

2 | ARTICLE NUMBER 18035 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

for HCC, deaths attributed to HBV were 210,200 in triggering viral internalization. The regions of inter­
1990 and 341,400 in 2010. In 2010, HBV, therefore, action between the NTCP and pre‑S1 proteins have been
caused a total of 786,000 deaths as a result of acute mapped26,27, permitting future rational drug design for
HBV infection, cirrhosis and HCC. In the systematic blocking viral entry.
analysis for the Global Burden of Disease Study 2016 Following entry, the rcDNA genome is transported
(REF.22), deaths at all ages caused by acute HBV infection into the nucleus and converted into covalently closed
were 100,300 in 2016. By contrast, for HBV-associated circular DNA (cccDNA) using the host cell DNA repair
HCC, deaths at all ages were 349,500 in 2016, and cir- response28. HBV cccDNA molecules associate with his-
rhosis due to HBV was associated with 365,600 deaths tones H3 and H4 and non-histone proteins, forming a
at all ages. These data emphasize the key burden of typical viral minichromosome29; the acetylation status
disease due to hepatitis B worldwide and the need for of histones H3 and H4 directly drives transcriptional
appropriate interventions. activity from HBV cccDNA30, whereas methylation
of H4 represses cccDNA transcription31. Pregenomic
Mechanisms/pathophysiology RNA (pgRNA) is exported from the nucleus and forms
HBV is an enveloped, DNA virus that belongs to the the template for reverse transcription from which the
Hepadnaviridae family 23. The envelope surrounds an two strands of HBV DNA are formed. The pgRNA also
icosahedral nucleocapsid, which encloses a partially serves as the translational template for the core protein
double-­stranded, relaxed circular DNA (rcDNA) genome and polymerase. Replication complexes of core parti-
of ~3.2 kilobases. Four partially overlapping open read- cles containing pgRNA and HBV polymerase assemble
ing frames (ORFs), termed P (poly­merase), S (surface), in the cytosol following the binding of polymerase to
C (core) and X (HBx protein), define the coding capac- the 5ʹ stem–loop structure (ε) in pgRNA. The binding
ity of the HBV genome (FIG. 3). Phylogenetic analyses event induces a packaging reaction, which results in the
of HBV strains isolated from various regions of the formation of immature nucleocapsids in which reverse
world have identified ten major genotypes (A–J) that transcription of the pgRNA occurs; the circular double-­
differ at the nucleotide level across full-length genomes stranded rcDNA genome is generated after a complex
by >8%24. multistep process of priming and template shifts within
the nucleocapsid32 (FIG. 4). To egress from cells, HBV
HBV infection and replication genome-containing capsid binds to HBV surface pro-
HBV infects and replicates largely in hepatocytes (FIG. 4). teins in the endoplasmic reticulum (ER) and is trans­
Initiation of infection is a multistep process and involves located in the ER before exiting the hepatocytes through
the binding of a highly conformational determinant the secretory pathway.
region (a) of the HBsAg glycoprotein to heparan sulfate A minority of nucleocapsids (~10%) enclose a
proteoglycans on the cell surface25; this is a low-affinity double-­stranded linear genome. Double-stranded linear
binding reaction that is reversible. Subsequently, HBV HBV genomes are efficient precursors for HBV DNA
interacts with its specific major receptor, the hepatic bile integration into the host cell genome33. The produc-
acid transporter sodium taurocholate co‑transporting tion of double-stranded linear DNA increases with the
polypeptide (NTCP)26, with high affinity via the large progression of liver disease, including in the develop-
HBsAg (L) protein (encoded by pre‑S1) of the virus, ment of HCC34, which corresponds to viral integration
being a tumorigenic event. As HBV DNA integration
typically leaves only the whole HBsAg ORF intact, this
Box 1 | Discovery of HBV has implications for achieving therapeutic HBsAg loss
Hepatitis was initially described inappropriately as catarrhal jaundice in 1865, a
(see below).
description that was later corrected by redefinition as hepatic inflammation in 1930s184. A characteristic feature of HBV replication is the
However, it was not until 1965 that geneticist Baruch Blumberg discovered evidence of secretion into the blood of a vast excess of subviral par-
hepatitis B virus (HBV); he observed that an immunoprecipitant formed after mixing the ticles of HBsAg, predominantly in the form of 17–25 nm
serum of a patient with haemophilia with that of an individual of aboriginal origin185. enveloped spherical vesicles. In addition, the translation
He was later awarded a Nobel Prize for this discovery in 1976. The initially named of the precore ORF can result in the presence of a secre-
Australia antigen marked the beginning of the journey of discovery of HBV. Another tory protein required for the establishment of chronicity,
researcher, Alfred M. Prince, linked Australia antigen with serum hepatitis and published hepatitis B e antigen (HBeAg; produced by processing of
his findings 3 years later186. Later, Australia antigen was commonly found in individuals precore protein), in the serum during certain infection
living in institutions187, suggesting that crowded conditions facilitated virus acquisition.
phases (FIG. 4).
Results from subsequent studies on Australia antigen in patients with abnormal serum
aminotransferases confirmed the link between the antigen and hepatitis.
In 1967, Krugman and colleagues188 identified two types of hepatitis, which were Immunopathogenesis
termed MS‑1 and MS‑2. MS‑1 was typically acquired through the oral route after a Chronic HBV infection can be considered a disease
short incubation period, whereas MS‑2 was ostensibly transmitted parenterally with that is characterized by the failure to mount an effective
a long incubation period. MS‑1 and MS‑2 hepatitis were subsequently classified as and coordinated adaptive immune response, resulting
hepatitis A and B, respectively. Serial chimpanzee and human experiments confirmed in viral persistence, which may be a consequence of
that Australia antigen was responsible for hepatitis B; hence, it was renamed as HBV. the ability of HBV to manipulate key innate immune
In 1970, Dane and colleagues189 used electron microscopy to visualize diverse subviral responses. Coordinated innate and adaptive immune
particles (hepatitis B surface antigen existing in circular and filamentous forms (FIG. 1)) responses are important for clearance of HBV from the
in addition to the complete enveloped virions, which were termed Dane particles.
liver. Importantly, under most circumstances, HBV is

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 18035 | 3


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

noncytopathic (does not directly kill hepatocytes); thus, intra­hepatic spread probably require a combination of
the liver inflammation and fibrosis that complicate long- cytolytic and noncytolytic mechanisms43 (FIG. 5). Thus,
term chronic infection are predominantly ­mediated by during acute resolving hepatitis B, there is a robust,
the host immune response. coordin­ated adaptive immune response with virus-­
specific CD8+ T cells mediating clearance of infected
Initial events after HBV infection. Following HBV infec- cells, B cells secreting neutralizing antibodies against
tion, there is a lag period of 4–7 weeks before HBV DNA HBsAg (anti-HBs) that block further infection and CD4+
and HBsAg become detectable in serum or the liver T cells supporting effective viral clearance44. Importantly,
(incubation period). Gene array studies during the acute hepatitis B with full clinical recovery does not usu-
incubation period have not revealed a strong interferon-­ ally occur when the infection occurs perinatally and in
stimulated gene response, leading to the proposition very early childhood because of the immaturity of the
that HBV might be a stealth virus capable of evading immune system, which can lead to chronic infection.
initial innate immune responses35,36. In addition, virus-
driven suppressive responses have been demonstrated, Mechanisms leading to HBV chronicity. Persistent
including: blocking recruitment of natural killer (NK) chronic HBV infection occurs mainly in infections
cells, NK T cells and macrophages37; decreasing cytokine acquired at birth or during the first 2 years of life.
secretion (interferon-γ (IFNγ), tumour necrosis fac- Virological factors that can influence persistence include
tor (TNF), IL‑6, IL‑8 and IL‑1β) from Kupffer cells38; the continuous production of HBsAg in concentrations
and blocking retinoic-­acid inducible gene I protein at least 1,000‑fold higher than that of infectious virions45.
(RIG‑I; also known as DDX58) and Toll-like receptor Excessive quantities of subviral particles containing
(TLR)-signalling, within the hepatocyte39,40. HBsAg may act as a decoy for HBV-specific humoral
In acute infection in adults, once viraemia becomes immunity and have been shown to promote a state of
detectable, the HBV DNA level increases exponentially virus-specific T cell anergy (exhaustion triggered by
in serum to peak levels of 108–109 viral copies per ml constant triggering of antigen-specific T cell receptor
but then typically declines, preceding the onset of clin­ signalling pathways) and deletion. Subviral particles
ical hepatitis41. Peripheral HBV-specific CD4+ and CD8+ can also negatively modulate innate immune signalling
T cell-mediated responses become detectable with the pathways, including nuclear factor-κB (NF‑κB) and
exponential increase in HBV replication. In the liver, mitogen-activated protein kinase (MAPK), suppress-
there is a lack of any detectable cellular infiltration of ing inflammatory cytokines and interferon-stimulated
immune cells, suggesting a process of noncytolytic gene transcription normally upregulated by TLR signal-
(no direct killing of infected hepatocytes) clearance of ling 45–48. Furthermore, the multifunctional HBx protein
HBV, which mainly involves cytokine-mediated inhib­ can inhibit proteasome-based degradation of viral pro-
ition of HBV replication via IFNγ and TNF, which are teins, thereby reducing HBV antigen presentation49 and
secreted from virus-specific CD8+ T cells42. The clear- decreasing the host antiviral response. HBx also blocks
ance of infected hepatocytes and prevention of further key host viral restriction factors, such as SMC 5/6 (two
of the six proteins in sister chromatid cohesion), which
silence HBV cccDNA. Both mechanisms possibly con-
tribute to establishment and/or maintenance of viral per-
sistence49, whereas expression of the HBV polymerase
protein suppresses the production of the myeloid dif-
ferentiation primary response protein MYD88, which
is crucial for TLR signalling 50. The precore or HBeAg
proteins can generate viral tolerance51 by downregulat-
Subviral particle
ing TLR2 expression on hepatocytes, Kupffer cells and
peripheral monocytes of patients during HBeAg-positive
disease52 and by downregulating the co‑stimulatory mol-
Dane particle
ecules CD28 on HBV-specific cytotoxic T lymphocytes
(CTLs)53 and CD86 on peripheral blood monocytes and
Kupffer cells in the liver 52 (FIG. 5). Interestingly, a recent
study reported that the expression of co-stimulatory
molecule OX40 ligand (OX40L) in innate immune cells
is important for the clearance of HBV from a mouse
model of infection54. The age-dependent expression of
OX40L may partly explain the high chronicity rate
of infection in children.
Figure 1 | Hepatitis B virions and subviral particles under electron
Nature microscopy.
Reviews | Disease Primers
Most immunological studies of chronic HBV infec-
Electromicrograph of hepatitis B virus (HBV) showing complete double-shelled virions
(Dane particles) and excess subviral particles (containing hepatitis B surface antigen tion have been performed on the peripheral blood com-
(HBsAg)) in tubular and small spherical form. Dane particles measure approximately 45 nm partment; however, the cellular composition of the liver
in diameter. The subviral particles do not contain the HBV genome. Taken from the US is different from that of blood. Intrahepatic immune
Centers for Disease Control and Prevention’s (CDC) Public Health Image Library (PHIL), responses are described in BOX 2. On the adaptive arm
identification number #5631. of the immune response, HBV-specific CD4+ and CD8+

4 | ARTICLE NUMBER 18035 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

a c 120
D

A 100

D
B and C
H E

Number of persons (in millions)


80
A, B, C and D A
F B, C and D
A and D 60

b
40

20

ric io c

on

ed gi n

ro eg an

on

gi th
Af reg cifi

M re Asia

Eu r rane
gi

gi
an n

ite on

an n

s
re ou
on
pe io
Pa

re

re

an S
st
rn

ic nd
Ea
te

er h a
h–
es

Am rt
ut
W

No
So

n
er
Rate of chronic hepatitis B virus infection

st
Ea
Low (<2%) Moderate (2–8%) High (≥8%) WHO region
Figure 2 | Global distribution of chronic hepatitis B infection and viral genotypes. Regions with different population
prevalence of chronic hepatitis B infection, categorized as high (>8%), moderate (2–8%) and low Reviews
Nature (<2%), are| Disease
depicted in
Primers
part a. High prevalence of chronic hepatitis B infection is found in the Western Pacific and Africa. Prevailing genotypes
(letters) in different countries are also depicted in part a. Part b illustrates the global prevalence of chronic hepatitis B
infection in children <5 years old. There has been a substantial reduction of prevalence of chronic hepatitis B infection
in children as a result of a successful vaccination programme, for example, in Asian countries. Part c shows the exact
population size of chronic hepatitis B carriers in different WHO regions.

T cell responses are substantially decreased and barely the suppression of signalling molecules associated with
functional relative to individuals with resolved acute T cell proliferation64 and result in increased production
infection. Notably, in chronic HBV infection, the fre- of ­immunosuppressive IL‑10 (REF.62).
quency and function of both intrahepatic and periph- Finally, humoral immunity may also play an impor-
eral HBV-specific T cells are inversely proportional tant but incompletely understood part in chronic HBV
to the level of circulating HBV DNA55. Furthermore, infection. This is evidenced by the treatment of individ-
these blunted T cell responses are coupled to the sus- uals with resolved chronic HBV with specific anti‑B cell
tained presence of high levels of virus and viral anti- therapy (for example, rituximab or ofatumumab) to treat
gens, therefore, leading to progressive exhaustion and unrelated diseases; these patients have a substantial risk
dysfunction of HBV-specific T cells56,57. The upregu- (30–60%) of viral reactivation. This risk persists even in
lation of mol­ecules responsible for suppressing CTL patients who have achieved HBsAg sero­clearance and
function, such as the programmed cell death protein 1 seroconversion to anti-HBs. The rate of reactivation after
(PD1) receptor, has been demonstrated on HBV-specific rituximab is higher than for similarly matched patients
CTLs in patients with chronic HBV infection58,59, with treated with T cell immunosuppressive agents65,66, hint-
a corresponding increased expression of PD1 ligand 1 ing at the importance of the B cell response for the
(PDL1) on the surface of dendritic cells60. Moreover, the ­control of HBV.
inhibitory co‑stimulatory marker cytotoxic T lympho-
cyte antigen 4 (CTLA4) is also upregulated on CD8+ Natural history of HBV infection
T cells61, and the positive co‑stimulatory ligand CD40 New terminology has been suggested for the phases
is downregulated on CD4+ T cells62. In addition, in of chronic HBV infection. The previously used terms of
chronic HBV infection, hepatic macrophages expressing immune tolerance, immune active and/or clearance
the C‑type lectin pattern recognition receptor CD206 and immune control and/or residual phases have been
have been found to produce amphiregulin, which pro- replaced by HBeAg-positive chronic infection, HBeAg-
motes the immunosuppressive activity of intrahepatic positive chronic hepatitis and HBeAg-negative chronic
regulatory T cells63. Collectively, these changes result in infection, respectively 67 (FIG. 6).

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 18035 | 5


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

may contribute to the development of HCC by trans-


PreS2
S1 activating human oncogenes. Detectable HBV-specific
Pre T cell immunity could reflect immune tolerance (which
determines virus persistence and chronicity) if immune
Po tolerance to HBV is defined operationally as the dele-
– lym
er tion or silencing of a sufficient number of HBV-specific
+ a
T cell clones69.

S
se
Immune clearance phase — HBeAg-positive chronic
hepatitis. The second phase, historically termed
immune clearance or immune active phase, most com-
monly occurs in the third or fourth decade of life in
patients who were infected with HBV at an early age.
In this phase, the triggering of immune T cell-mediated
responses against infected hepatocytes leads to the cyto­
lytic release of alanine transaminases (ALT) and reduc-
tion of HBV DNA levels. However, the intensity of the
immune response fluctuates over time, resulting in fluc-
tuating levels of ALT (hepatitis flares, which are defined
C

3' differently in different studies; for example, twofold ele-


5'
3' vation from the baseline ALT levels70,71) and HBV DNA.
5' The liver damage caused by hepatitis flares may lead to
P different degrees of fibrosis and cirrhosis. This phase
Pre can be of varying duration and ends with a reduction of
C HBV DNA concentrations and HBeAg seroconversion
to anti‑HBe positivity (antibodies specific for HBeAg).

X Immune control phase — HBeAg-negative chronic


infection. A proportion of patients (annual rate of ~15%)
Figure 3 | Genomic structure of HBV. The hepatitis B virus (HBV)
Nature genome
Reviews is organized
| Disease in
Primers will spontaneously seroconvert from HBeAg to anti-
a circular form with the positive-strand DNA (with variable base length) forming the inner HBe; these patients (previously referred to as inactive
circle and the negative-strand DNA (completely circular) forming the outer circle. The carriers) typically show near normal serum aminotrans-
four overlapping genes are P for translating polymerase, PreS1/PreS/S for hepatitis B ferases and HBV DNA concentrations of <2,000 IU per
surface antigen (HBsAg), PreC/C for core protein and X for HBx protein. The P protein has ml. A more rapid seroconversion to anti-HBe (occur-
reverse transcriptase activity. The C and S open reading frames (ORFs) have extensions at
ring before the accumulation of marked hepatitis fibro-
their 5ʹ ends termed pre‑C and pre‑S. The pre‑S region is divided into pre‑S1 and pre‑S2
domains, and translation of the S ORF leads to the production of the large (L), medium
sis) with continued low levels of replication and hepatic
(M) and small (S) HBsAg. Translation of the pre‑C ORF results in a secretory protein, inflammation confers good prognosis. However, in gen-
hepatitis B e antigen (HBeAg), which is an accessory protein required establishing eral, HBeAg-negative patients are older, and if patients
chronicity, whereas the C ORF results in the capsid protein. HBx is required for still have moderate or high levels of viral replication,
establishment of infection and maintenance of active replication by inhibiting the host they would have more-progressive histological disease,
nuclear restriction factor, sister chromatid cohesion 5/6 (SMC 5/6)202. Overexpression of and progression to cirrhosis can be more-rapid.
HBx has been shown to result in cellular transformation via promiscuous transactivation. Notably, in a substantial proportion of patients in
Asian and Mediterranean countries, HBeAg seroconver-
Immune tolerance phase — HBeAg-positive chronic sion is associated with the occurrence of precore and/or
infection. The natural history and immunopatho­genesis core promoter mutations72,73. In these patients, HBeAg
of HBV infection are still being defined. Chronic HBV is absent as a result of the selection of HBV virions that
infection can be divided into several phases (FIG. 6). do not express HBeAg (precore mutant HBV).
The first immune tolerance phase (also termed the Finally, patients may have spontaneous HBsAg sero-
high replicative, low inflammatory phase or HBeAg- clearance during the natural course of disease. Various
positive chronic infection) typically occurs during studies have shown that patients with spontaneous
childhood chronic infection. This phase is character- HBsAg seroclearance at an earlier age have a better
ized by high viral load (with HBV DNA levels often outcome than those without HBsAg sero­clearance. In a
>107 international units (IU) per ml), near normal liver cohort of 298 Chinese patients, the cumulative risk of
histology owing to a minimal host immune-mediated HCC was higher in patients with HBsAg sero­clearance
reaction and serum positivity for HBeAg. The concept at ages ≥50 years compared with those with HBsAg sero­
of immunological tolerance in young patients who are clearance at ages <50 years. Moreover, fibrosis measured
HBeAg positive has been criticized as recent studies by liver stiffness assessment was also substantially higher
have indicated that adolescents may have detectable in the former group (29.5% and 7.9%, respectively).
HBV-specific functionally active T cells. Studies of this Intrahepatic total HBV DNA and cccDNA (occult
phase of the disease have also revealed clonal expansion infection; BOX 3) were detected in 100% and 79.3% of
of hepatocytes and discernible HBV integrants68, which patients, respectively 74.

6 | ARTICLE NUMBER 18035 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Diagnosis, screening and prevention to go through with liver biopsy because of its invasive
Diagnosis nature. Algorithms for the use of serum biomarkers and
The key serological marker of acute and chronic hepati- liver stiffness measurements (if transient elastography is
tis B is the detection of HBsAg in serum. However, sev- available) can be used in clin­ical practice to adjudge the
eral other HBV serological markers are clinically useful degree of fibrosis in chronic hepatitis B75.
in HBV infection. For example, in patients with clinical
recovery of HBV infection, antibodies to HBsAg (anti- Detection of HBsAg and HBsAg-specific antibodies.
HBs) and hepatitis B core antigen (HBcAg; anti-HBc) Detection of HBsAg in serum is the standard serolog-
may be detectable. Anti-HBc antibodies may remain ical test to confirm HBV infection and is also used for
detectable in patients with chronic HBV infection, and screening in population-based epidemiological studies
other serum markers can assess the viro­logical status of to establish the prevalence of chronic HBV infection76.
the disease (FIG. 4). In addition to serum tests, an assess- Conventionally, a chronic HBV infection is diagnosed
ment of the fibrosis and cirrhosis status in patients with by a repeat positive test for HBsAg, 6 months after the
chronic HBV infection is also important for disease initial positive test. By contrast, after recovery from
prognostication, treatment indication and manage- acute HBV infection, the levels of HBsAg become
ment. An accurate but invasive test of liver disease is undetectable. HBsAg concentrations differ during the
liver biopsy, whereas noninvasive tests include tran- varying longi­tudinal phases of disease and are generally
sient elastography (to measure liver stiffness) or vari­ous higher in individuals with detectable HBeAg (discussed
serum biomarkers. With the availability of the latter two below) 77. Importantly, HBsAg remains measurable
methodo­logies, most patients and clinicians are reluctant when serum HBV DNA falls to undetectable levels by

Hepatocyte Bloodstream
HBV
Cytoplasm HBV DNA HBsAg

rcDNA
pgRNA
Nucleus Complete virons Subviral
with rcDNA particles
A(n)
cccDNA A(n)
A(n) HBcAg HBsAg
HBeAg HBV RNA

ATG ATG
PreC HBV core gene
C-terminal
Core protein domain
Translation HBcAg
HBcAg
HBcrAg HBeAg
Precore protein
p22cr
Incomplete Protein processing
protein
processing HBeAg
p22cr

Figure 4 | The HBV replication cycle and key viral markers. After viral entry into hepatocytes
Naturevia the high
Reviews affinity Primers
| Disease
receptor sodium taurocholate co‑transporting polypeptide (left upper diagram), hepatitis B virus (HBV) relaxed circular
DNA (rcDNA) enters the nucleus and is converted into covalently closed circular DNA (cccDNA) in the form of a
minichromosome — the major transcriptional template of the virus. The transcription products are exported from the
nucleus, with the larger pregenomic RNA (pgRNA) incorporated into replication complexes in the cytoplasm comprising
the viral polymerase and core protein. Within these replication complexes, pgRNA is reverse-transcribed into HBV DNA,
which can replenish cccDNA or undergo further packaging. The HBV DNA-containing capsid binds to the HBV surface
proteins on the endoplasmic reticulum, is translocated into the lumen before exiting the hepatocytes through the
secretory pathway and is then released as mature virus particles. mRNAs transcribed from cccDNA also produce various
viral antigens. Except for cccDNA, all the other viral products (HBV rcDNA, HBV RNA, hepatitis B e antigen (HBeAg),
hepatitis B surface antigen (HBsAg), hepatitis B core antigen (HBcAg) and 22 kDa precore protein (p22cr)) are easily
measurable in the blood (right diagram). The lower left part of the diagram illustrates the three antigens, HBcAg, HBeAg
and p22cr (collectively known as hepatitis B core-related antigen (HBcrAg)), produced from translation of different starting
codons of the preC core gene and differential protein processing afterwards. A(n), polyadenosine at the end of mRNAs.

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 18035 | 7


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

sensitive PCR. Moreover, quantitative measurement associated with a response to HBV vaccination, recov-
of HBsAg has prognostic importance and has been ery from acute hepatitis or HBsAg seroconversion in
incorporated into risk scores to predict the risk of chronic HBV infection. HBsAg seroclearance occurs
HCC, and may also indicate the risk of viral rebound in 0.1–2% of patients with chronic HBV infection per
after stopping NUCs78. A positive test result for anti- year. Thus, the c­ umulative rate can be substantial in
HBs in patients who tested negative for HBsAg can be long‑term follow‑up.

Innate immune response Adaptive immune response

Kupffer B cell
Monocyte
cell
Antibodies
TNF TLR2 TH2 CD4+
IL-6 T cell
IL-8
IL-1β HBeAg+ CD4+
T cell
TCR
APC
Cell HBV
death MHC II
TH1 CD4+ Treg cell
MHC I T cell (FOXP3 CD4+
Hepatocyte
CD25+)
TRAIL
TNF
IL-10 CD8+
T cell Cytolytic
PD1 hepatocyte
NO PDL1
Perforins
Granzyme
FAS

NK cell NK T cell Macrophage FASL


Noncytolytic
hepatocyte

HBV-specific IFNγ
HBV CD8+ T cell
Subviral particles
Nature Reviews | Disease Primers
Figure 5 | HBV-specific immune responses. Successful clearance of hepatitis B virus (HBV) from infected hepatocytes
requires the coordinated effort of both the innate and adaptive arms of the immune response. During infection, HBV
­pathogen-associated molecular patterns are sensed by cells of the innate response, including Kupffer cells and monocytes
(top left-hand side of figure), resulting in the release of antiviral cytokines, including tumour necrosis factor (TNF) and
interferon-γ (IFNγ). The virus counters this initial response with the expression of its own proteome, particularly
hepatitis B e antigen (HBeAg) and hepatitis B surface antigen (HBsAg). The HBeAg downregulates Toll-like receptor 2
(TLR2) signalling on Kupffer cells, and monocytes and hepatocytes promote viral persistence. Excess production of HBsAg
subviral particles (FIG. 4) also substantially blunts the responses of natural killer (NK) cells, NK T cells and macrophages with
the overproduction of the immunosuppressive cytokine IL‑10 (bottom left-hand side of figure). Viral clearance and disease
pathogenesis are linked to the appearance of a vigorous T cell response to the entire HBV proteome. By contrast, viral
persistence is associated with a markedly diminished HBV-specific T cell response. The HBV-specific CD8+ T cell is
regarded as the main immune effector cell clearing HBV infection (lower right-hand side of figure), but other mechanisms
of control and clearance, including antibody-dependent cell cytotoxicity and intracellular neutralization, could also
contribute (top right-hand side of figure). The onset of viral clearance in acute hepatitis B is strongly linked to the
appearance of virus-specific T cells as well as CD3, CD8 and IFNγ‑mRNA in the liver. Whether this temporal relationship
also holds true during the steps towards functional cure needs to be established because exhaustion markers, such as
programmed cell death protein 1 (PD1), are found at much higher frequencies in chronic hepatitis B (central part of figure).
Thus, the host immune response directed against HBV requires the coordinated action of both the innate immunity and
cellular and humoral adaptive immunity to affect both noncytolytic and cytolytic clearance and control. The coordinating
cell that bridges the innate and adaptive response is the antigen-presenting cell (APC) (central part of figure), and
virus-specific blocking of antigen presentation is an effective mechanism of persistence. FAS, tumour necrosis factor
receptor superfamily member 6; FASL, tumour necrosis factor ligand superfamily member 6 (also known as FASLG);
FOXP3, forkhead box protein P3; MHC, major histocompatibility complex; NO, nitric oxide; PDL1, programmed cell death
protein 1 ligand 1; TCR, T cell receptor; TH1, T helper 1; TH2, T helper 2; TRAIL, TNF-related apoptosis-inducing ligand
(also known as TNFSF10); Treg, regulatory T cell. Adapted with permission from REF.203, Baishideng Publishing Group.

8 | ARTICLE NUMBER 18035 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Box 2 | Intrahepatic immune responses in chronic HBV association between HBV DNA concentrations and the
risk of HCC82. A decline in HBV DNA concentrations
Studies of the immune response that focus on the peripheral blood compartment can predict the efficacy of treatment, and an increase in
rather than intrahepatic events have resulted in concepts and paradigms that are not HBV DNA in serum is observed with the development
universally accepted. Furthermore, studies using liver-derived material from untreated of resistance to NUCs18 (discussed below).
patients have established a pivotal role for innate immunity in the immunopathology of
hepatitis B virus (HBV) infection and challenge earlier concepts, including the concept
that HBV is a stealth virus that evades innate immunity. The discovery that HBV is Prevention
recognized in hepatocytes by nucleic acid sensors, such as the Toll-like receptors and Vaccination. The age of HBV infection is of vital impor-
the retinoic acid-inducible gene protein I (RIG‑I) systems, and that other cell tance in determining the risk of chronicity. The risk is
populations have important roles within the liver microenvironment have all >90% for infants and children <1 years of age, 30% for
contributed to these new insights. For example, evidence suggests that HBV is a weak children 1–5 years of age and 0–2% for adults83–85. Thus,
inducer and active repressor of innate proinflammatory cytokines, such as interferon-γ most chronic HBV infections are acquired through
(IFNγ), in hepatocytes190 rather than having evasive properties. Furthermore, it is MTCT at the time of birth or during infancy and child-
recognized that natural killer (NK) cell numbers are increased several-fold in the liver hood. In addition, transmission to children can occur
compartment compared with the periphery191 and that other innate lymphoid cell from other close relatives who are HBV positive86.
subsets, such as myeloid-derived suppressor cells and mucosal-associated invariant
MTCT is particularly important in the South-East
T cells, play a key part in HBV immunopathogenesis. A landmark study performed
biopsies on >100 patients with untreated chronic HBV infection and demonstrated a Asian and Western Pacific regions, where HBV preva-
strong downregulation of the major innate immune response pathways, including lence is still high and the young female population has
interferon-­stimulated gene, Toll-like and pathogen recognition receptor pathways, a high HBV load, and Africa, where coverage of uni-
and antiviral sensors40. Thus, this work identified a strong impairment of innate immune versal vaccination for HBV is still suboptimal. In par-
responses in the livers of patients with chronic HBV. The investigators also ticular, HBeAg-positive infection increases the risk of
demonstrated an association of low quantifiable levels of hepatitis B surface antigen MTCT owing to high viral load. Additionally, horizontal
(HBsAg) with decreased innate immune gene repression compared with patients with household and community transmission are problem-
higher HBsAg levels, in whom some innate genes were substantially more suppressed. atic in endemic regions87. Effective and safe vaccines are
available that prevent HBV infection and are the key
preventive measure88,89. The introduction of childhood
Detection of HBeAg. HBeAg serological testing, usu- vaccination has pivotally led to a decline in the preva-
ally performed after establishing HBsAg positivity lence of hepatitis B in children3,6,90 (FIG. 2). The world-
for >6 months (chronic HBV-infected patients), dis- wide health threat from chronic hepatitis B has led to the
tinguishes two forms of hepatitis B: HBeAg-positive drafting of aspirational targets to eliminate HBV infec-
and HBeAg-negative chronic HBV infection. HBeAg- tion. The WHO elimination goals include 90% complete
negative, anti-HBe-positive chronic HBV infection is coverage of HBV vaccination, a reduction in the preva-
characterized by the absence of HBeAg in serum. HBeAg lence of HBsAg in children <5 years of age to 0.1% and,
seroconversion marks the transition from the immune finally, an i­ mprovement in the rates of treatment to 80%
clearance phase or HBeAg-positive chronic hepatitis to by 2030 (REF.91).
the immune control phase or HBeAg-negative chronic The WHO recommends universal vaccination for
infection or hepatitis (FIG. 6). HBV and that all infants should receive the first dose
of vaccine soon after birth, preferably within 24 hours.
Detection of anti-HBc. Total anti-HBc measured by The birth dose, or vaccination within the expanded
immunoassay is a marker of acute, chronic and resolved programme on immunization (EPI) dose, is followed
HBV infection, or occult hepatitis B (BOX 3). In addition, by an additional two to three doses, depending upon
the presence of anti-HBc can predict reactivation of national policies and prevalence. In 2016, 186 of the 194
HBV associated with rituximab or immunosuppressive WHO member countries included the HBV vaccine
therapy. Immunoglobulin M (IgM) anti-HBc is detected in their EPI. However, many births in Global Alliance
during acute HBV infection and is also detected during for Vaccines and Immunization (GAVI)-supported
exacerbations of chronic HBV, and importantly, it may countries still occur outside of hospitals and, hence,
be the only serological marker detectable in fulminant these infants do not receive the standard vaccination.
acute hepatitis B. Thus, measurement of anti-HBc is usu- Moreover, only monovalent HBV vaccine can be used
ally recommended for diagnosis in patients suspected at birth, and the monovalent vaccine is not funded by
to have an acute exacerbation of chronic HBV infection GAVI. As such, in 2014, <38% of babies born world-
or to decide if patients who are planning to undergo wide received HBV birth-dose vaccine within 24 hours
immunosuppressive therapy will require prophylactic of birth92.
antiviral therapy. In addition to the high infection risk faced by neo-
nates and infants, HBV transmission is also more
Detection of HBV DNA. HBV DNA tests should be per- common in some groups of people and, therefore, vacci-
formed on a regular basis (for example, every 6 months) nation is more important in these specific groups. These
in all patients with chronic HBV. Tests for HBV DNA in include health-care providers, laboratory technicians
serum directly assess viral replication. The concentra- who handle bodily fluid samples, immunocompromised
tions of HBV DNA provide an indication for therapy patients, institutionalized individuals (for example, pris-
within current guidelines67,79–81 and for monitoring treat- oners) and patients on regular fluid exchange procedures
ment efficacy. Indications for treatment are based on the (for example, dialysis).

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 18035 | 9


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

HBV DNA HBsAg


ALT HBeAg
Anti-HBe

HBV DNA levels

ALT levels
HBeAg-positive HBeAg-positive HBeAg-negative HBeAg-negative
New terminology
chronic infection chronic hepatitis chronic infection chronic hepatitis
Immune tolerance Immune active Immune control Residual phase
Old terminology
phase and/or clearance phase and/or residual phase and/or disease flares

Treatment indicated Treatment indicated


Figure 6 | Hepatitis B disease phases and treatment indications. Diagram showing the relationship between hepatitis B
Nature Reviews | Disease Primers
virus (HBV) DNA and alanine transaminase (ALT) levels and the relation of these levels to different phases of chronic HBV
infection using new and old terminology. Some patients (solid lines) experience intermittent flares in HBV DNA and ALT
levels before achieving HBeAg seroconversion, whereas other patients (dashed line) may have a less frequent flares.
Treatment is indicated when the HBV DNA levels are >2,000 or >20,000 international units (IU) per litre and ALT levels are
higher than one or two times the upper limit of normal according to different regional guidelines. anti-HBe, antibodies
against HBeAG; HBeAg, hepatitis B e antigen; HBsAg, hepatitis B surface antigen.

Antiviral prophylaxis. Vaccine protection may fail in Indications for treatment in chronic HBV infection are
infants born to highly viraemic mothers, typically deline­ based on the levels of serum HBV DNA, ALT and severity
ated as maternal HBV DNA concentrations of 2 × 105 IU of liver disease. Treatment indications are generally sim-
per ml. Thus, despite prophylaxis, MTCT is more likely ilar for HBeAg-positive and HBeAg-negative diseases.
to be observed in HBeAg-positive women93, occurring Treatment is generally indicated for those with HBV
in 8–32% of infants94. The rationale and effectiveness DNA levels >2,000 IU per ml, or depending on regional
of NUC prophylaxis in highly viraemic ­mothers have guidelines, at a higher level, for example, 20,000 IU per
been recognized, and clinical trials have shown reduced ml for HBeAg-positive disease67,80,81 (TABLE 1). Treatment
MTCT with telbivudine, lamivudine or tenofovir dis- is also generally indicated for elevated ALT levels, which
oproxil fumarate (TDF) prophylaxis93,95–97. In a recent are defined, depending on the regional guidelines, as ALT
study, despite vaccination at birth, 7% of infants became levels greater than the upper limit of normal (ULN) or
HBsAg positive; however, this was reduced to 0% of greater than twice the ULN. The presence of moderate
infants whose mothers received TDF from 30–32 weeks fibrosis by liver biopsy or noninvasive testing may also be
of gestation95. At present, the American Association taken into consideration in some conditions, for example,
for the Study of Liver Diseases (AASLD) recommends in elderly patients with relatively normal ALT levels. For
the administration of antiviral prophylaxis to HBsAg- patients with HBV DNA >20,000 IU per ml and ALT
positive mothers whose serum HBV DNA concentra- levels greater than twice the ULN, no assessment of fibro-
tions exceed 2 × 105 IU per ml in the third trimester of sis is needed for treatment initiation67. In the presence
pregnancy. TDF is the recommended NUC for prophy- of cirrhosis, treatment is often recommended in cases of
laxis in pregnancy, as the drug is safe in pregnancy, with detectable HBV DNA levels regardless of the ALT levels.
birth defect rates similar to those of the general popula- Other factors to take into account for the indication
tion. TDF can be stopped 3 months after delivery if the of treatment include the patient’s age (older age should
only indication is to prevent MTCT. be a factor to consider when starting treatment), a fam-
Antiviral prevention of HBV MTCT requires screen- ily history of liver cancer, comorbidities (for example,
ing to identify HBV-infected mothers with high levels obesity or the presence of alcoholic liver disease), risk of
of viraemia using HBV DNA testing. However, in many HBV transmission and extrahepatic manifestations of
countries where universal vaccination is administered, hepatitis B. Another situation in which treatment would
HBV DNA concentrations are not tested in pregnancy. be indicated is for the rapid control of viral replication
Implementation of this preventive strategy would require in patients with severe HBV flares from chronic HBV
additional testing, which remains largely unattainable in infection, the parameters of which include ALT levels
resource-limited countries. >10-times the ULN with or without clinical symptoms,
including malaise, loss of appetite and tea-coloured
Management urine. By contrast, treatment for acute HBV infection
Indications for treatment is primarily supportive, and antiviral treatment is not
The relationship between different phases of chronic indicated under normal circumstances unless patients
HBV and indications for therapy is depicted in FIG. 6. have fulminant liver failure.

10 | ARTICLE NUMBER 18035 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Disease progression patient monitoring and for informing treatment deci-


In patients who acquire HBV infection during early child- sions; for example, patients calculated to have a rela-
hood, the majority sequelae of cirrhosis and HCC occur tively high risk of developing HCC should be strongly
after HBeAg seroconversion. In a study of 3,233 Chinese ­recommended to receive long-term antiviral treatment.
patients, the median age of HBeAg seroconversion was
35 years, whereas the median age for the development End points of treatment
of complications was 57.2 years, and of those with com- HBV treatment is typically aimed to achieve profound
plications, 73.5% were positive for anti-HBe98. However, virological suppression, which in turn will lead to bio-
serum HBV DNA level is an independent and important chemical remission (return of ALT values to the ­normal
factor for the development of disease ­complications and range), histological improvement and prevention of
is a surrogate marker for HBV replication. the complications of liver disease, such as cirrhosis,
Multiple studies have shown a direct proportionality liver failure and HCC. A more advanced functional
between HBV DNA levels and the risk of development cure is HBsAg seroclearance, as spontaneous or treat-
of cirrhosis and HCC; the most rigorous are the two ment-associated HBsAg seroclearance substantially
REVEAL studies from Taiwan82,99. In these two studies, decreases the risk of HCC development, provided the
81.6% and 80.1% of patients were HBeAg negative with patient is <50 years of age and has not developed cir-
normal ALT levels. For patients with HBV DNA levels rhosis74,107,108. However, HBsAg seroclearance is rarely
>106 copies per ml (~200,000 IU per ml), the incidence achieved with the currently available anti-HBV agents
of cirrhosis and HCC were 2,498 and 1,152 per 100,000 (TABLE 2). A more realistic end point is, therefore, the
person-years, respectively. However, for patients with induction of sustained or maintained virological remis-
HBV DNA levels <300 copies per ml (~60 IU per ml), sion, as indicated by undetectable HBV DNA levels and
the incidence of complications per 100,000 person-years HBeAg seroconversion.
was still moderately high at 338.8 for cirrhosis and 108 Current potent NUCs have a profound effect on viro-
for HCC. This finding implies that antiviral treatment logical suppression with continued therapy. For NUCs:
should aim to maintain sustained suppression of HBV a virological response is defined as undetectable serum
DNA to very low, preferably undetectable, levels100. HBV DNA (10–60 IU per ml); a serological response
Other risk factors for HCC include male sex, increasing as HBeAg and/or HBsAg seroconversion; a biochemi-
age and the presence of core promoter mutations and cal response as normalization of serum ALT level; and a
pre‑S deletions in HBV101,102. The carcinogenic effects histological response as a decrease in necroinflam­mation
of HBV are potentiated in patients with several risk of the liver 67,80,81. Worldwide, pegylated ­interferon-α
factors concomitantly 103. (PEG-IFNα) therapy is given for a finite duration to a
Several risk calculators that use common clinical and small subset of patients (discussed below) with the aim
laboratory parameters have been derived to predict the of achieving sustained off-treatment immune control.
risk of HCC in patients with chronic HBV. These include The response is assessed 6 months post-treatment,
GAG (guide with age, gender, HBV DNA, core promoter with criteria similar to those of NUCs, except for the
mutations and cirrhosis)–HCC score101, CU (Chinese viro­logical response, which is defined as HBV DNA
University)–HCC score104, REACH–B (risk estimation <2,000 IU per ml.
for HCC in chronic hepatitis B) score105 and the PAGE–B Long-term follow‑up studies of sustained treatment
(platelet, age, gender) score106. These scores are useful with NUCs demonstrate that a profound reduction in
for the strategic planning for disease complications and development of cirrhosis can be achieved with the early
initiation of treatment. Consequently, this has resulted
in an almost complete elimination of liver failure in the
Box 3 | Occult HBV infection setting of treated patients with chronic HBV infection.
Occult hepatitis B is defined as the detection of hepatitis B virus (HBV) DNA in the livers In addition, there is a clear reduction, but not complete
(with or without detectable HBV DNA in the serum) of individuals negative for HBV prevention, of the development of HCC109,110.
surface antigen (HBsAg)192. The incidence tends to depend on the endemicity of the
hepatitis B infection. These individuals are positive for antibodies specific for hepatitis Treatment agents
B core antigen (anti-HBc) and/or anti-HBs, indicating previous exposure to HBV. Treatment agents approved by the US FDA, European
HIV co-infection may be associated with occult HBV infection193,194. Medicines Agency and many countries in Asia are
Studies performed in France demonstrated that a large proportion of patients with broadly classified into immunomodulatory agents
alcohol-related hepatocellular carcinoma (HCC) had HBV DNA integrated into the
and antiviral agents. The former includes conven-
genome of the neoplastic liver cells195. A second study of cryptogenic HCC in Chinese
individuals also found that 73% had HBV in non-tumour and/or tumour liver tissues196.
tional IFNα‑2b and PEG-IFNα‑2a. The latter includes
Both studies imply that occult HBV infection may be an important cause of HCC in NUCs, namely, lamivudine, adefovir dipivoxil, ente-
HBsAg-negative patients. cavir, telbivudine, TDF and tenofovir alafenamide
There is a resurgence of interest in occult hepatitis B because of the widespread use (TAF). The mainstay of treatment in most countries are
of potent B cell-depleting monoclonal antibodies against CD20, such as rituximab NUCs (FIG. 7).
(and anti-CD52). These agents are associated with severe, sometimes fatal HBV
reactivation in patients with occult HBV infection up to 12 months after cessation of NUCs. Of the six NUCs licensed by the FDA (FIG. 7),
rituximab therapy. It is, therefore, important to monitor the HBV DNA levels in these entecavir, TDF and TAF are the advocated first-line
patients and initiate antiviral therapy once HBV DNA levels become detectable. agents because of their potency and low resistance
Prophylactic antiviral therapy can also be considered in these patients197.
rates (TABLE 2). Entecavir has negligible adverse effects

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 18035 | 11


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

and is an excellent agent for treatment-naive patients. For the majority of patients, NUCs should be taken on
However, patients previously treated with lamivudine a long-term, probably lifelong, basis. NUCs can be
who have developed lamivudine resistance will already stopped after HBsAg seroclearance, but monitoring for
have two of the three mutations required for the develop­ the development of HCC should be continued, especially
ment of entecavir resistance; 51% of these individ­uals for patients with HBsAg seroclearance who are >50 years
will develop entecavir resistance within 5  years111. of age or who have cirrhosis.
In lamivudine-­resistant patients, TDF and TAF are The cost-effectiveness of using NUCs for chronic
the best treatment options. Compared with TDF, TAF HBV infection is also the subject of research. A recent
can achieve a higher intrahepatic level of tenofovir; study carried out community-based screening using
thus, a lower dose of TAF can be used to exert the same an HBsAg rapid test followed by antiviral treatment in
degree of viral suppression. As such, the off-target effects individuals who were HBsAg positive in a highly HBV
of TAF are decreased112. According to one study, TAF prevalent area, The Gambia118. The study showed an
has comparable viral suppression to TDF with improved incremental cost-effective ratio of US$540 per disabil-
rates of ALT normalization, substantially fewer declines ity adjusted life-year and $640 quality-adjusted life-year
in hip and spine bone mineral density (an adverse effect saved compared with no intervention.
of TDF) and smaller decreases in estimated glomerular
filtration rate, indicating TAF has less of an off-target Interferon-based therapy. PEG-IFNα has both immune
effect on renal function113. However, in general, clin­ modulatory and antiviral properties and is administered
ical adverse effects are considered to be minimal during subcutaneously once per week in subsets of patients who
long-term NUC therapy. are indicated for treatment (see below). One year of
Short and medium term (1–5 years) treatment with PEG-IFNα treatment leads to a sustained off-treatment
NUCs can achieve sustained suppression of HBV DNA response in ~30% of treated patients119,120; treatment
to very low, often undetectable levels in all patients100. response is defined as serum HBV DNA <2,000 IU per ml
Normal ALT levels (30 U per litre for men; 19 U per together with ALT levels within the normal range and in
litre for women) can be achieved in the majority of HBeAg-positive patients, HBeAg loss or seroconversion.
patients. Importantly, long-term treatment with NUCs HBsAg loss (functional cure) is the desired outcome,
has been shown to reverse cirrhosis of the liver 114,115 and but this occurs in only 10% of treated patients. A longer
decrease the risk of HCC109. These improvements are not duration of therapy could lead to higher response rates
­dependent on the genotype of HBV. but is difficult to tolerate for many patients121. Adverse
HBsAg seroclearance occurs in 3–11% of patients events can be prominent, and patients undergoing
receiving NUCs (TABLE  2). According to one study, PEG-IFNα therapy require careful monitoring for, for
HBsAg seroclearance is unlikely to occur with NUCs example, clinical symptoms of malaise, fever, thyrotoxic
treatment 116. This is corroborated by another study in features, depression and lab­oratory tests of blood cell
42 patients on long-term NUCs (median duration of count, liver biochemistry and thyroid function. In addi-
treatment of 126 months) with three liver biopsies per- tion, PEG-IFNα is ­contraindicated in patients with
formed over the course of the study 117. In this study, decompensated cirrhosis.
long-term treatment with NUCs for >5 years caused As the response rate is modest and there are con-
a reduction of intrahepatic HBV DNA and cccDNA siderable adverse effects, PEG-IFNα therapy should be
by >99%, whereas HBsAg titres were reduced by only individualized and given to only a subset of patients who
71.5%; notably, there was no correlation between HBsAg are predicted to have the highest response rates to PEG-
levels and intrahepatic HBV DNA and cccDNA levels at IFNα therapy. In HBeAg-positive patients, PEG-IFNα
the time of the third biopsy. Both studies postulate that is a reasonable choice as first-line therapy for patients
after potent suppression of HBV replication (resulting infected with HBV genotype A or B (which are more
in undetectable HBV DNA), HBsAg continues to be responsive to IFN therapy) who are young, lack major
produced by other sources, most likely from integrated comorbidities and have lower serum HBV DNA levels
viral sequences117. Currently, no large systematic analy- and ALT levels greater than twice the ULN122. In patients
ses exist for HBsAg seroclearance rates on NUC therapy. who are HBeAg negative, pretreatment factors that

Table 1 | Indications for initiating treatment in chronic HBV infection


Guideline Individuals who are HBeAg positivea Individuals who are HBeAg negativea
HBV DNA IU/ml ALT U/l HBV DNA IU/ml ALT U/l
EASL67 2017 ≥2,000 >ULN and/or at least moderate liver ≥2,000 >ULN and/or at least moderate liver
necroinflammation or fibrosis necroinflammation or fibrosis
≥20,000 >2× ULN irrespective of fibrosis ≥20,000 >2× ULN irrespective of fibrosis
AASLD80 2016 >20,000 >2× ULN or significant histological disease >2,000 >2× ULN or significant histological disease
APASL 2015
81
≥20,000 >2× ULN or significant histological disease ≥2,000 >2× ULN or significant histological disease
AASLD, American Association for the Study of Liver Diseases; ALT, alanine aminotransferase; APASL, Asian Pacific Association for the Study of Liver; EASL,
European Association for the Study of the Liver; HBeAg, hepatitis B e antigen; HBV, hepatitis B virus; IU, international units; ULN, upper limit of normal. aPatients
with cirrhosis with detectable HBV DNA should be treated irrespective of the HBV DNA and ALT levels.

12 | ARTICLE NUMBER 18035 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

Table 2 | Resistance rates and HBsAg seroclearance rates of anti-HBV drugs therapy has substantially improved the management of
HBV-associated decompensated liver disease. Prolonged
Drug Resistance HBsAg seroclearance Refs suppression of HBV DNA with NUCs can lead to sub-
Entecavir 1.2% (7 years follow-up) 2.5% (7 years follow-up) 204
stantial improvements in hepatic fibrosis and hepatic
TDF 0% (9 years follow-up) 11.8% for HBeAg+ and 1.3% for 205 synthetic function and reduced HCC risks; if patients
HBeAg– (7 years follow-up) with decompensated disease are stabilized, it can obviate
TAF 0% (96 weeks follow-up)a NA 206 the need for liver transplantation130–133. Treating patients

, negative; +, positive; HBeAg, hepatitis B e antigen; HBsAg, hepatitis B surface antigen; HBV,
with decompensated cirrhosis in a timely manner is
hepatitis B virus; NA, not yet available; TAF, tenofovir alafenamide; TDF, tenofovir disoproxil critical. There is evidence that higher pretreatment bil-
fumarate. aBecause TAF is newly licensed, only 96‑week data are available. irubin, creatinine and HBV DNA levels are associated
with the highest risk of mortality, despite the initiation
predict response at 24 weeks after treatment included of NUC therapy 134,135.
younger age, female sex, high baseline ALT levels and Ethical considerations mean it is not possible to per-
low HBV DNA; i­ nterferon treatment would be indicated form large randomized clinical trials of NUC treatment
in these patients123. in decompensated disease, as NUCs are the standard of
On‑treatment termination rules have been established care. However, one study compared the clinical outcomes
for patients who do not respond to PEG-IFNα treatment, of 70 decompensated patients who received enteca-
which will prevent unneeded costs for continuation of vir with 144 compensated patients who received ente-
treatment in the context of a very low chance of response cavir daily as first-line therapy for 1 year 135. At the end
and adverse effects. In HBeAg-positive patients, no of therapy, both groups had similar degrees of viral sup-
decline in levels of HBsAg at week 12 of PEG-IFNα ther- pression and rates of undetectable HBV DNA levels and
apy is a strong indication of non-­response in genotypes A HBeAg seroconversion. However, among the decom-
and D, whereas HBsAg >20,000 IU per ml is an indica- pensated patients, six patients died of liver failure and
tion for halting therapy in patients infected with geno- three had liver transplantation within the first 6 months
types B and C. At week 24 of treatment, all patients with of therapy. At baseline, all nine of these patients had high
a serum HBsAg >20,000 IU per ml are classified as having Child–Pugh–Turcotte (CPT) scores (>10, grade C: the
a very high chance of treatment failure, irrespective of most advanced grade). Transplantation, therefore, should
HBV genotype124. In these patients, PEG-IFNα treatment not be postponed for patients with grade C CPT score
should be discontinued or treatment with NUCs initiated. or a model of end-stage liver disease (MELD) score >15,
HBeAg-negative patients who have no decline in HBsAg as these patients may be at a high risk of liver failure.
and <2 log decline in HBV DNA at week 12 of treatment A prospective randomized study compared the effi-
have a very low chance of achieving a treatment response, cacy of TDF versus TAF plus emtricitabine (another
and therapy should be similarly discontinued125. NUC similar to lamivudine with anti-HBV effect) ver-
sus entecavir in 112 treatment-experienced and treat-
Combined therapy. Studies using combination ther- ment-naive patients with decompensated liver disease
apy aim to determine whether drug combinations can for 48 weeks136. Although the primary end point of this
increase the rate of HBsAg seroclearance. According study was safety, secondary outcomes were virological
to one study with 740 treatment-naive patients, a com- markers and CTP scores. The rates of HBV DNA sup-
bination of PEG-IFNα with TDF resulted in a higher pression were similar in all groups, and improvements
rate of HBsAg seroclearance: 9.1% (TDF and 48‑week in CPT (≥2 points) were observed in 26–48% of patients.
PEG-IFNα), 2.8% (TDF and 16‑week PEG-IFNα), 0% Overall survival was >90% in all treatment arms. The
(120‑week TDF) and 2.8% (48‑week PEG-IFNα). Thus, rates of adverse events, including nephrotoxicity, were
the combination therapy is superior to both TDF mono- similar at 1 year of therapy in all groups. The study sug-
therapy and PEG-IFNα monotherapy126. Notably, most of gests that, in general, NUCs are effective at suppress-
the patients who had HBsAg seroclearance were infected ing HBV, with an acceptable 1‑year survival in patients
with HBV genotype A or B. In another study in NUC- with decompensated liver disease. Although TAF has
treated patients who were HBeAg negative with undetect- lower risks of bone and renal toxic effects, its efficacy
able HBV DNA, the addition of 48 weeks of PEG-IFNα in patients with advanced liver disease still needs to
was not associated with a significant increase in HBsAg be evaluated.
seroclearance rate (7.8% with PEG-IFNα versus 3.2% Regular HCC surveillance is essential for patients
without PEG-IFNα, P = 0.15)127. It is important to note with advanced baseline hepatic fibrosis, even if they
that these two studies had discrepant results. It might show a favourable virological response after treatment.
be that combination therapy using NUCs and PEG-IFNα A study examined the efficacy of long-term entecavir
works better in treatment-naive patients who are infected therapy in reducing the risk of HCC in a large cohort
with genotype A or B. In general, combination therapy is of HBV-related cirrhotic patients137. The suppression of
not routinely advocated at the present time. HBV DNA to <20 IU per ml with entecavir in the first
year of treatment was associated with a 38% reduction in
Antiviral therapy for decompensated liver disease. HCC. However, the annual HCC rate remained signifi­
It is estimated that ~15–30% of patients with chronic cant at 2.4%. Older age, male sex, HBeAg positivity and
HBV infection progress to cirrhosis and decompen- baseline α-fetoprotein ≥7 ng per ml were predictors
sated liver disease128,129. The availability of potent NUC of HCC.

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 18035 | 13


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

IFNα-2b Lamivudine Adefovir Telbivudine TDF the introduction of TDF and ETV, optimal suppression
of HBV DNA before transplant can be achieved. Thus,
post-transplant prophylaxis with TDF or ETV alone may
be adequate to prevent infection reoccurrence, but not
1990 1998 2002 2005 2006 2008 2016 enough data currently exist to recommend this strat-
egy. TAF likely will replace TDF for all post-transplant
prophylaxis owing to its improved safety profile.
Interferon therapy Entecavir

NUCs therapy Special hepatitis B populations


PEG-IFNα TAF
Immunosuppressed individuals. There is a high rate
Figure 7 | Approved treatment agents for chronic HBVNature Reviews
infection. | Disease
A timeline thatPrimers
shows of HBV reactivation in patients who have previously
the year of US FDA approval for individual hepatitis B virus (HBV) treatment agents. All the achieved HBV suppression after treatment but then
treatment agents have also been approved by the European Medicines Agency and in become immunosuppressed146,147. This may occur dur-
various Asian countries. The boxes in green refer to nucleoside or nucleotide analogues ing chemotherapy, in patients with HIV after immune
(NUCs), and the pale yellow boxes refer to interferon-based therapy. In 2018, the reconstitution (after highly active antiretroviral ther-
recommended first-line agents are pegylated interferon α-2a (PEG-IFNα), entecavir,
apy), after organ transplant and haematopoietic stem cell
tenofovir disoproxil fumarate (TDF) and tenofovir alafenamide (TAF).
transplant, or with B cell-depleting agents such as rituxi-
mab (anti‑CD20) or with biological response modifiers
Liver transplantation such as TNF inhibitors. HBV reactivation can also occur
Since 2000, HCC has become the primary indication in immunocompetent patients treated with steroids and
for liver transplantation in patients with chronic HBV biological response modifiers.
infection. However, the rate for liver transplantation for Rituximab is the most potent reactivator of HBV.
decompensated cirrhosis has decreased and is likely the Chemotherapy-induced HBV reactivation can result in
result of the excellent efficacy of antiviral therapy 138. the abrupt reappearance or increase in HBV replication
The use of prophylactic therapy has a major impact with liver damage occurring during and/or following
in preventing the risk of recurrent HBV infection post-­ immune reconstitution. Clinically, this may range from
transplant 139. In many transplant centres, the standard subclinical to severe or even fatal hepatitis (with con-
prophylactic combination therapy is hepatitis B immuno­ current rises in virological and biochemical markers)
globulin (HBIg) given intramuscularly, with dosing and may progress to liver failure and death. The rate of
aimed at maintaining a protective serum anti-HBs level, HBV reactivation after chemotherapy for solid tumours
and NUC therapy. With such an approach, the efficacy is varies148. Of concern, HBV DNA may be undetect­able
>90% in preventing recurrent HBV infection after liver at the time of the ALT peak, and HBV reactivation
transplantation139. With the availability of NUCs with a should always be considered in immunosuppressed
low risk of HBV resistance developing, such as entecavir patients who are HBV positive. A lack of pre-emptive
and TDF, protection can be achieved with lower dose and antiviral therapy can lead to liver damage, resulting in
shorter duration of HBIg 140. In fact, some post-transplant chemotherapy treatment interruptions and premature
centre protocols recommend HBIg use only during the termination of chemotherapy 148. All patients who are
perioperative period (time period immediately after sur- to receive immuno­suppressive therapy should be tested
gery). Researchers in Hong Kong pioneered HBIg-free for HBsAg, anti-HBs and anti-HBc, and the initiation of
prophylaxis, which uses continuous entecavir mono- antiviral therapy before immunosuppression is recom-
therapy from the day of transplant 141,142. This approach mended if HBsAg is positive. If anti-HBc is positive but
provided encouraging results with progressive rates of HBsAg is negative, anti-HBV therapy should be admin-
HBsAg seroclearance from 85% in year 1 to 92% by year istered pre-emptively in the case of rituximab treatment
8 shown in 265 patients. or haematopoietic stem cell transplant, or HBV DNA
The selection of HBV prophylaxis needs to be indi- should be monitored monthly, as the risk of seroreversion
vidualized. Before the era of prophylaxis, patients with (to HBsAg positive status) is at least 10%146. For other
HBV and hepatitis D virus (HDV) co-infection (BOX 4) forms of chemotherapy, close monitoring of HBV DNA
had improved outcomes after liver transplantation as should be considered146,149. In patients who are HBsAg
they have lower HBV DNA levels before transplant 143. negative, anti-HBs positivity seems to be associated with
However, recurrent HDV infection can occur if HBsAg a decreased risk of HBV reactivation150. Reactivation
levels rebound, as the HDV replicative cycle requires may also occur in patients with rheumatoid arthritis
HBsAg. For patients with HDV co-infection, it would who are treated with methotrexate or biological response
be advisable to use a combination of HBIg and NUCs to ­modifiers, including anti-TNF agents151.
ensure long-term suppression of HBsAg post-­transplant,
therefore, preventing graft loss by recurrent infec- HIV co-infection. Co-infection with HBV occurs in ~10%
tion139,144. Liver transplantation has been successfully of HIV-infected patients owing to similar modes of trans-
performed for patients with HIV and HBV co-­infection mission. Globally, HBV is generally acquired in child-
in the United States and Europe since 2001. The combin­ hood, whereas HIV infection occurs later in life owing
ation of HBIg and NUCs was previously used in these to sexual transmission; however, both viruses can be
co-infected patients, especially those with detectable acquired sexually, via MTCT or via intravenous drug use.
HBV DNA levels before transplant 145. However, with Adults with HIV infection who are exposed to HBV have

14 | ARTICLE NUMBER 18035 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

a higher risk (~20%) of developing chronic HBV infec- extensive hepatic fibrosis and an increased risk of
tion152. HBV serology may be atypical in HIV patients: end-stage liver disease, HCC and death than in HBV
HIV and HBV co-infected patients may have a negative monoinfected patients. HCV treatment is as safe and
HBsAg with positive anti-HBc. Isolated anti-HBc (in the effective in co-infected patients as it is in monoinfected
absence of any other serum viral marker) occurs in 7–19% patients and should be considered following the same
of HIV-infected patients and is more likely in patients rules as HCV monoinfected patients. However, HBV
with underlying hepatitis C virus (HCV) co-infection153. reactivation can occur in patients with HBV and HCV
With the immunosuppression that can accompany HIV co‑infection who are treated with direct-acting anti­
infection, seroreversion may occur from HBsAg negative virals (DAAs)158,159. HCV patients co‑infected with HBV
to positive. The US Department of Health and Human should be treated for HBV before initiating DAA therapy.
Service guidelines state that all HIV patients who are Furthermore, HBsAg-negative patients with serological
HBsAg positive should be treated with antiretroviral evidence of prior HBV should be monitored closely for
therapy that include two drugs effective against HBV, HBV DNA and ALT levels during and after receiving
regardless of HBV DNA level and ALT level. TDF-based DAA therapy for HBV reactivation.
or TAF-based therapy plus lamivudine or TDF or TAF
plus emtricitabine is preferred as initial therapy; HIV– Extrahepatic manifestations
HBV treatment should be indefinite, as interruption may Extrahepatic manifestations of HBV are uncommon and
lead to severe HBV flares and death154. Anti-HBV drugs include rheumatological, neurological, renal and vascu-
should not be used alone in HIV patients, as most HBV litic complications, and all are associated with immune
drugs also have anti-HIV activity, and using them alone complexes160. Both membranous glomerulonephritis
may result in HIV drug resistance155. and membranoproliferative glomerulonephritis are seen
in chronic HBV patients. Notably, 20–30% of patients
HCV co-infection. HCV co-infection in HBV-infected with polyarteritis nodosa have chronic HBV. Antiviral
patients is uncommon156 as is triple infection with therapy for HBV can improve both renal and vasculitic
HBV, HCV and HIV157. Co-infected patients have more manifestations of disease.

Quality of life
Box 4 | HDV infection Convincing evidence exists that advanced chronic
liver disease, especially with the presence of ascites and
Virology
encephalopathy, is associated with significantly impaired
Hepatitis D virus (HDV) is a defective RNA virus that contains a circular single-stranded
negative-sense RNA genome. The viroid-like 1.7 kb genome encodes two isoforms health-related quality of life (HRQOL) scores161. By con-
(large or small) of the hepatitis D antigen (HDAg). The viral life cycle requires trast, patients with mild disease can present with subtle,
concomitant hepatitis B virus (HBV) infection for HDV to assemble and propagate within nonspecific symptoms. Fatigue is the mostly commonly
the host, as the HDV ribonucleoprotein complex requires an envelope of the three HBV recorded symptom among patients with viral hepatitis.
(small, middle and large hepatitis B surface antigen (HBsAg)) envelope proteins. However, malaise and lethargy are generally common,
Epidemiology with prevalence of these symptoms being as high as 25%
Globally, 6–8% of chronic HBV-infected carriers are also infected with HDV. There are in the general population162. A validated fatigue score is,
several different HDV genotypes: genotype 1 is prevalent worldwide; genotypes 2 and therefore, necessary to systematically evaluate patients
4 are prevalent in Japan and Taiwan; genotype 3 is prevalent in the Amazon basin; with chronic HBV infection.
and genotypes 5–8 are prevalent in individuals of African origin. The NIH Patient-Reported Outcomes Management
Disease course Information System (PROMIS) is highly reliable in
Infection occurs either as co-infection (simultaneous HBV–HDV infection) or as a measuring physical, mental and social functioning. The
superinfection in a patient with chronic HBV infection. Fulminant hepatitis can occur PROMIS Fatigue Short Form was applied in addition to
in 3–4% of co-infected individuals. Chronic HDV and HBV superinfection lead to a high the Medical Outcomes Study 36‑item Short Form Mental
rate of chronic disease and to severe and progressive chronic viral hepatitis with a Health Functioning subscale to evaluate a large cohort of
high rate of cirrhosis. patients with chronic HBV in North America under the
Management NIH-sponsored Hepatitis B Research Network163. This
HDV infection is managed with pegylated interferon-α (PEG-IFNα) given weekly for cohort included 948 participants who were predominantly
12–18 months. Serum HDV RNA remains undetectable after 6 months of therapy in of Asian descent (71%), and only 2% had advanced hepatic
~25% of patients198. fibrosis with aspartate-to‑platelet ratio index (APRI;
Novel treatment on trial an index of liver fibrosis) scores >1.5. The frequency of
Lonafarnib is an orally ingested tricyclic derivative of carboxamide that inhibits fatigue in this cohort was lower than that reported in the
prenylation of the large HDAg by inhibiting farnesyltransferase199. Myrcludex B general population. The mean fatigue scores were simi-
(a 47‑mer myristoylated pre‑S1 lipopeptide) is a subcutaneously delivered allosteric lar among Asian, black and white individ­uals. The study
inhibitor of sodium taurocholate co‑transporting polypeptide200. Nucleic acid polymers confirmed that patients with more advanced fibrosis and
(NAPS) (phosphorothioated oligonucleotides) interfere with viral entry via heparin those with comorbid conditions, such as diabetes melli-
sulfate proteoglycans201. All these therapies for HDV are currently in clinical trials.
tus and low mental health functioning, tend to experience
Outlook more fatigue. The findings underscore the importance of
The clearance of HDV RNA depends on the clearance of HBsAg. Other potential being alerted to the possibility of advanced liver disease
antiviral strategies, including silencing RNAs or immunomodulatory therapies, may be or concurrent mental and medical conditions if patients
required for the effective treatment of HDV.
expressed fatigue during clinic visits.

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 18035 | 15


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

In another population of patients with chronic HBV With HBV RNA testing, it has become apparent that
infection in the United States, differences in disease and RNA-containing virus like particles are secreted from
treatment perceptions were reported between different the hepatocytes during HBV replication (FIG. 4). Recent
racial groups, with non-Asian patients experiencing reports have demonstrated the detection of HBV RNA in
more symptoms164. Interestingly, patients with a history serum samples of HBV-infected patients. The nature of
of interferon-based therapy reported greater symptom- the HBV RNA in the serum is uncertain but could repre-
atology compared with patients treated with only oral sent pgRNA. However, there are technical difficulties in
antiviral therapy. discriminating HBV RNA from a large excess of rcDNA.
Future studies of HRQOL among patients with Serum HBV RNA is being assessed as a surrogate to
chronic HBV infection should take into account the monitor cccDNA, to assess suppression of DNA synthe-
health attitudes of the different racial groups, the influ- sis, to determine treatment efficacy using new agents
ence of the various comorbid ethnic and the effects of such as capsid inhibitors and to define end points173–177.
therapy in both mild and severe disease states.
Novel treatment
Outlook The ideal treatment goal is the total eradication of HBV
Disease control from carriers, that is, the elimination of cccDNA and
As of 2016, a universal hepatitis B vaccination pro- integrated HBV DNA from the liver. A definition exists
gramme has been implemented in 186 countries for such a goal: the sterilizing cure of HBV. To achieve
worldwide3. Global coverage with three doses of HBV this, very preliminary research in cell culture and murine
vaccine is estimated to be 84% (>90% in the Western experiments is using targeted endonucleases (CRISPR)
Pacific, regions of the Americas and the South–East Asia that specifically cleave highly selected HBV DNA
region). In addition, strategic planning is emerging for sequences. However, it is anticipated important safety
antiviral prophylaxis in pregnant HBV-infected women, considerations need to be e­ valuated before these agents
which effectively decreases MTCT. As a result of these can be tested in humans.
efforts, the global prevalence of HBV infection and also In the near future, we may have several novel
the incidence of HBV-associated liver decompensation thera­peutic approaches to enhance the rate of HBsAg
and HCC are expected to decline. However, it may take seroclearance, hence achieving the functional cure.
at least 30 years for this favourable outcome to emerge. These include agents acting through different target­
Until then, we should improve awareness of the dis- able steps in the viral life cycle (FIG. 4). One promising
ease, case finding, surveillance strategies and treatment approach entering clinical trials is through viral tran-
optimization for the existing 257 million HBV chron- scription silencing using short interfering RNAs. In
ically infected people. These challenging steps should selected patient popu­lations, a profound reduction of
be implemented without delay in all countries to reach HBsAg, HBcAg and HBeAg has been observed178,179.
the goal set in May 2016 by the WHO: the elimination Amelioration of these immune-exhausting viral pro-
of viral hepatitis as a public threat, with 90% reduction of teins will hopefully restore innate and specific HBV
new hepatitis infections and 65% reduction in mortality immune responses, leading to higher rates of HBsAg
by 2030. Although health-care workers, health-related sero­clearance. HBsAg seroclearance is a disease state
statutory personnel, paramedics and patient groups are in which there is minimal chance of viral rebound on
actively involved in the first three aspects, HBV disease treatment cessation and is associated with the lowest
clinicians and researchers are the key persons to improve risk of developing disease c­ omplications and death if
the treatment for HBV disease. achieved early in life.
Another group of promising novel antiviral com-
New HBV markers. Improved and supplementary HBV pounds is the capsid protein assembly modifiers or
markers, in addition to pre-existing markers, are con- inhibitors. Disrupting core proteins may affect mul-
tinuously being sought for better disease prognostica- tiple vital functions for viral replication, including
tion and treatment management. The two most recently viral RNA packaging, immune modulating func-
developed markers are hepatitis B core-related antigen tions, replenishment of cccDNA and expression of
(HBcrAg; FIG. 4) and HBV RNA. HBcrAg consists of cccDNA180. In addition, viral protein release inhibitors
three related proteins sharing an identical 149 amino and viral entry inhibitors are also being actively inves-
acid sequence: HBcAg, HBeAg and a truncated 22 kDa tigated with preliminary promising effects. All these
precore protein (p22Cr) (FIG. 4). Monoclonal antibodies DAAs are developed with the aim of the restoration of
can be used to detect all three proteins, and there is a immune control181,182.
good correlation between HBcrAg, serum HBV DNA Direct immune-modulating agents, for example,
and cccDNA165,166. Detectable HBcrAg may demonstrate therapeutic vaccines and TLR agonists, have promising
the persistence of intrahepatic cccDNA. HBcrAg detec- effects in animal models but have yet to show major
tion can differentiate HBeAg-negative from HBeAg- positive effects when tested in humans. However, there
positive disease and may be used to predict HBeAg is an interesting upcoming group of compounds acting
seroconversion, sustained response to NUCs treatment, through activation of viral sensor proteins, for example,
the presence of necroinflammatory disease, the risk of RIG‑I, to stimulate type I and III interferon production
reactivation of hepatitis B with immunosuppression and and boost antiviral immunity. An initial phase I study
the risk of HCC development 166–172. has shown promising results183.

16 | ARTICLE NUMBER 18035 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

The agents discussed above are now under active seroclearance. Another challenge will be the time required
early clinical phase I and phase II studies. Once their to achieve this goal. The ideal goal is to have therapy
efficacies and safety profiles are established, the remain- leading to HBsAg seroclearance at a younger age, which
ing challenge is to determine the best combination and may result in a considerable reduction of ­development of
sequence of agents to achieve the practical goal of HBsAg ­cirrhosis and HCC in HBV-infected persons.

1. Marynard, J., Kare, M. A. & Alter, M. in Viral Hepatitis variability in envelope capsid interactions. EMBO J. 46. Wu, J. et al. Hepatitis B virus suppresses toll-like
and Liver Disease (ed. Zukerman, A.) 967–969 26, 4160–4167 (2007). receptor-mediated innate immune responses in
(Alan R. Liss, New York, 1988). 24. Okamoto, H. et al. Typing hepatitis B virus by murine parenchymal and nonparenchymal liver cells.
2. Keating, G. M. & Noble, S. Recombinant hepatitis B homology in nucleotide sequence: comparison Hepatology 49, 1132–1140 (2008).
vaccine (Engerix‑B): a review of its immunogenicity of surface antigen subtypes. J. Gen. Virol. 69, 47. Jiang, M. et al. Toll-like receptor-mediated immune
and protective efficacy against hepatitis B. Drugs 63, 2575–2583 (1988). responses are attenuated in the presence of high
1021–1051 (2003). 25. Sureau, C. & Salisse, J. A conformational heparan levels of hepatitis B virus surface antigen. J. Viral
3. World Health Organization. Global Hepatitis Report, sulfate binding site essential to infectivity overlaps Hepat. 21, 860–872 (2014).
2017 (WHO, Geneva, 2017). with the conserved hepatitis B virus A‑determinant. 48. Wang, S. et al. Hepatitis B virus surface antigen
4. Schweitzer, A., Horn, J., Mikolajczyk, R. T., Krause, G. Hepatology 57, 985–994 (2013). selectively inhibits TLR2 ligand-induced IL‑12
& Ott, J. J. Estimations of worldwide prevalence of 26. Yan, H. et al. Sodium taurocholate cotransporting production in monocytes/macrophages by interfering
chronic hepatitis B virus infection: a systematic review polypeptide is a functional receptor for human with JNK activation. J. Immunol. 190, 5142–5151
of data published between 1965 and 2013. Lancet hepatitis B and D virus. eLife 1, e00049 (2012). (2013).
386, 1546–1555 (2015). 27. Ni, Y. et al. Hepatitis B and D viruses exploit sodium 49. Seeger, C. & Mason, W. S. Hepatitis B virus biology.
5. European Association for the Study of the Liver. EASL- taurocholate co‑transporting polypeptide for species- Microbiol. Mol. Biol. Rev. 64, 51–68 (2000).
HEPAHEALTH Project Report: Risk Factors and the specific entry into hepatocytes. Gastroenterology 50. Wu, M. et al. Hepatitis B virus polymerase inhibits the
Burden of Liver Disease in Europe and Selected 146, 1070–1083 (2014). interferon-inducible MyD88 promoter by blocking
Central Asian Countries (EASL, Geneva, 2018). 28. Schreiner, S. & Nassal, M. A. Role for the host DNA nuclear translocation of Stat1. J. Gen. Virol. 88,
6. Ott, J. J., Stevens, G. A., Groeger, J. & Wiersma, S. T. damage response in hepatitis B virus cccDNA 3260–3269 (2007).
Global epidemiology of hepatitis B virus infection: new formation — and beyond? Viruses 9, 125 (2017). 51. Milich, D. Exploring the biological basis of hepatitis B
estimates of age-specific HBsAg seroprevalence and 29. Newbold, J. E. et al. The covalently closed duplex form e antigen in hepatitis B virus infection. Hepatology
endemicity. Vaccine 30, 2212–2219 (2012). of the hepadnavirus genome exists in situ as a 38, 1075–1086 (2003).
7. Ni, Y.‑H. et al. Continuing decrease in hepatitis B virus heterogeneous population of viral minichromosomes. 52. Visvanathan, K. et al. Regulation of Toll-like receptor‑2
infection 30 years after initiation of infant vaccination J. Virol. 69, 3350–3357 (1995). expression in chronic hepatitis B by the precore
program in Taiwan. Clin. Gastroenterol. Hepatol. 14, 30. Levrero, M. et al. Control of cccDNA function in protein. Hepatology 45, 102–110 (2006).
1324–1330 (2016). hepatitis B virus infection. J. Hepatol. 51, 581–592 53. Li, X. et al. Changes of costimulatory molecule CD28
This study describes the positive long-term effects (2009). on circulating CD8+T cells correlate with disease
of universal vaccination on the reduction of 31. Zhang, W. et al. PRMT5 restricts hepatitis B virus pathogenesis of chronic hepatitis B. Biomed. Res. Int.
hepatitis B prevalence. replication through epigenetic repression of covalently 2014, 423181 (2014).
8. Tamandjou, C. R., Maponga, T. G., Chotun, N., closed circular DNA transcription and interference 54. Publicover, J. et al. An OX40/OX40L interaction
Preiser, W. & Andersson, M. I. Is hepatitis B birth dose with pregenomic RNA encapsidation. Hepatology 66, directs successful immunity to hepatitis B virus. Sci.
vaccine needed in Africa? Pan Afr. Med. J. 27, 18 398–415 (2017). Transl Med. 10, eaah5766 (2018).
(2017). 32. Bremer, C. & Glebe, D. The molecular virology of 55. Webster, G. J. M. et al. Longitudinal analysis of CD8+
9. Sharma, S., Carballo, M., Feld, J. J. hepatitis B virus. Semin. Liver Dis. 33, 103–112 T cells specific for structural and nonstructural
& Janssen, H. L. A. Immigration and viral hepatitis. (2013). hepatitis B virus proteins in patients with chronic
J. Hepatol. 63, 515–522 (2015). 33. Tu, T., Budzinska, M., Shackel, N. & Urban, S. HBV hepatitis B: implications for immunotherapy. J. Virol.
10. Ward, J. W. & Byrd, K. K. Hepatitis B in the United DNA integration: molecular mechanisms and clinical 78, 5707–5719 (2004).
States: a major health disparity affecting many implications. Viruses 9, 75 (2017). 56. Wherry, E. J., Blattman, J. N., Murali-Krishna, K.,
foreign-born populations. Hepatology 56, 419–421 34. Zhao, X.‑L. et al. Serum viral duplex-linear DNA van der Most, R. & Ahmed, R. Viral persistence alters
(2012). proportion increases with the progression of liver CD8 T‑cell immunodominance and tissue distribution
11. Kao, J.‑H. & Chen, D.‑S. Universal hepatitis B disease in patients infected with HBV. Gut 65, and results in distinct stages of functional impairment.
vaccination: killing 2 birds with 1 stone. Am. J. Med. 502–511 (2015). J. Virol. 77, 4911–4927 (2003).
121, 1029–1031 (2008). 35. Wieland, S. F., Spangenberg, H. C., Thimme, R., 57. Zhou, S., Ou, R., Huang, L., Price, G. E.
12. Kao, J.‑H. & Chen, D.‑S. HBV genotypes: epidemiology Purcell, R. H. & Chisari, F. V. Expansion and & Moskophidis, D. Differential tissue-specific
and implications regarding natural history. contraction of the hepatitis B virus transcriptional regulation of antiviral CD8+ T‑cell immune
Curr. Hepat. Rep. 5, 5–13 (2006). template in infected chimpanzees. Proc. Natl Acad. responses during chronic viral infection. J. Virol. 78,
13. Sunbul, M. Hepatitis B virus genotypes: global Sci. USA 101, 2129–2134 (2004). 3578–3600 (2004).
distribution and clinical importance. 36. Wieland, S., Thimme, R., Purcell, R. H. & Chisari, F. V. 58. Boni, C. et al. Characterization of hepatitis B virus
World J. Gastroenterol. 20, 5427 (2014). Genomic analysis of the host response to hepatitis B (HBV)-specific T‑cell dysfunction in chronic HBV
14. Mixson-Hayden, T. et al. Hepatitis B virus and virus infection. Proc. Natl Acad. Sci. USA 101, infection. J. Virol. 81, 4215–4225 (2007).
hepatitis C virus infections in United States-bound 6669–6674 (2004). 59. Fisicaro, P. et al. Combined blockade of programmed
refugees from Asia and Africa. Am. J. Trop. Med. Hyg. 37. Fisicaro, P. et al. Early kinetics of innate and adaptive death‑1 and activation of CD137 increase responses
90, 1014–1020 (2014). immune responses during hepatitis B virus infection. of human liver T cells against HBV, but not HCV.
15. Beasley, R. P. Hepatitis B virus as the etiologic agent Gut 58, 974–982 (2009). Gastroenterology 143, 1576–1585.e4 (2012).
in hepatocellular carcinoma — epidemiologic 38. Hösel, M. et al. Not interferon, but interleukin‑6 60. Chen, L. et al. B7‑H1 up‑regulation on myeloid
considerations. Hepatology 2, 21S–26S (1982). controls early gene expression in hepatitis B virus dendritic cells significantly suppresses T cell immune
16. Shafritz, D., Sherman, M. & Tur-Kaspa, R. in infection. Hepatology 50, 1773–1782 (2009). function in patients with chronic hepatitis B.
Hepatology: A Textbook of Liver Disease 945–958 39. Revill, P. & Yuan, Z. New insights into how HBV J. Immunol. 178, 6634–6641 (2007).
(WB Saunders, Philadelphia, 1990). manipulates the innate immune response to establish 61. Schurich, A. et al. Role of the coinhibitory receptor
17. Chen, D. From hepatitis to hepatoma: lessons from acute and persistent infection. Antivir. Ther. 18, 1–15 cytotoxic T lymphocyte antigen‑4 on apoptosis-Prone
type B viral hepatitis. Science 262, 369–370 (1993). (2013). CD8 T cells in persistent hepatitis B virus infection.
18. Yuen, M.‑F. et al. Chronic hepatitis B virus infection: 40. Lebossé, F. et al. Intrahepatic innate immune response Hepatology 53, 1494–1503 (2011).
disease revisit and management recommendations. pathways are downregulated in untreated chronic 62. Barboza, L. et al. Altered T cell costimulation during
J. Clin. Gastroenterol. 50, 286–294 (2016). hepatitis B. J. Hepatol. 66, 897–909 (2017). chronic hepatitis B infection. Cell. Immunol. 257,
19. Mokdad, A. A. et al. Liver cirrhosis mortality in 187 41. Webster, G. Incubation phase of acute hepatitis B 61–68 (2009).
countries between 1980 and 2010: a systematic in man: dynamic of cellular immune mechanisms. 63. Dai, K., Huang, L., Sun, X., Yang, L. & Gong, Z. Hepatic
analysis. BMC Med. 12, 145 (2014). Hepatology 32, 1117–1124 (2000). CD206‑positive macrophages express amphiregulin
20. El‑Serag, H. B. Hepatocellular carcinoma. N. Engl. 42. Xia, Y. et al. Interferon‑γ and tumor necrosis factor‑α to promote the immunosuppressive activity of
J. Med. 365, 1118–1127 (2011). produced by T cells reduce the HBV persistence form, regulatory T cells in HBV infection. J. Leukoc. Biol. 98,
21. Lozano, R. et al. Global and regional mortality from cccDNA, without cytolysis. Gastroenterology 150, 1071–1080 (2015).
235 causes of death for 20 age groups in 1990 and 194–205 (2016). 64. Chen, C.‑F. et al. Regulation of T cell proliferation
2010: a systematic analysis for the Global Burden of 43. Xia, Y. & Protzer, U. Control of hepatitis B virus by by JMJD6 and PDGF‑BB during chronic hepatitis B
Disease Study 2010. Lancet 380, 2095–2128 cytokines. Viruses 9, 18 (2017). infection. Sci. Rep. 4, 6359 (2014).
(2012). 44. Shin, E.‑C., Sung, P. S. & Park, S.‑H. Immune 65. Shibolet, O. & Shouval, D. Immunosuppression and
22. GBD 2016 Causes of Death Collaborators. Global, responses and immunopathology in acute and chronic HBV Reactivation. Semin. Liver Dis. 33, 167–177
regional, and national age-sex specific mortality for viral hepatitis. Nat. Rev. Immunol. 16, 509–523 (2013).
264 causes of death, 1980‑2016: a systematic (2016). 66. Seto, W.‑K. et al. Hepatitis B reactivation in patients
analysis for the Global Burden of Disease Study 2016. 45. Shi, B. et al. HBsAg inhibits IFN‑α production in with previous hepatitis B virus exposure undergoing
Lancet 390, 1151–1210 (2017). plasmacytoid dendritic cells through TNF‑α and IL‑10 rituximab-containing chemotherapy for lymphoma:
23. Seitz, S., Urban, S., Antoni, C. & Böttcher, B. induction in monocytes. PLoS ONE 7, e44900 a prospective study. J. Clin. Oncol. 32, 3736–3743
Cryo‑electron microscopy of hepatitis B virions reveals (2012). (2014).

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 18035 | 17


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

67. European Association for the Study of the Liver. EASL 89. World Health Organization. Immunization, vaccines hepatitis B virus therapy. Antimicrob. Agents
2017 Clinical Practice Guidelines on the management and biologicals: Hepatitis B. WHO http://www.who.int/ Chemother. 59, 3563–3569 (2015).
of hepatitis B virus infection. J. Hepatol. 67, 370–398 immunization/monitoring_surveillance/burden/vpd/ 113. Buti, M. et al. Tenofovir alafenamide versus tenofovir
(2017). surveillance_type/passive/hepatitis/en/ (2017). disoproxil fumarate for the treatment of patients with
68. Mason, W. S. et al. HBV DNA integration and clonal 90. Liang, X. et al. Epidemiological serosurvey of hepatitis HBeAg-negative chronic hepatitis B virus infection:
hepatocyte expansion in chronic hepatitis B patients B in China — declining HBV prevalence due to hepatitis a randomised, double-blind, phase 3, non-inferiority
considered immune tolerant. Gastroenterology 151, B vaccination. Vaccine 27, 6550–6557 (2009). trial. Lancet Gastroenterol. Hepatol. 1, 196–206
986–998.e4 (2016). 91. World Health Organization. Global Health Sector (2016).
This study reports that HBV DNA integration Strategy On Viral Hepatitis 2016–2021 114. Marcellin, P. et al. Regression of cirrhosis during
occurs even in the early phase of chronic (WHO, Geneva, 2017). treatment with tenofovir disoproxil fumarate for
hepatitis B infection. 92. World Health Organization. Data, Statistics and chronic hepatitis B: a 5‑year open-label follow‑up
69. Milich, D. R. The concept of immune tolerance in Graphics (WHO, Geneva, 2018). study. Lancet 381, 468–475 (2013).
chronic hepatitis B virus infection is alive and well. 93. Patton, H. & Tran, T. T. Management of hepatitis B 115. Chang, T.‑T. et al. Long-term entecavir therapy results
Gastroenterology 151, 801–804 (2016). during pregnancy. Nat. Rev. Gastroenterol. Hepatol. in the reversal of fibrosis/cirrhosis and continued
70. Yuen, M. et al. Prognostic factors in severe 11, 402–409 (2014). histological improvement in patients with chronic
exacerbation of chronic hepatitis B. Clin. Infect. Dis. 94. Lin, X. et al. Immunoprophylaxis failure against hepatitis B. Hepatology 52, 886–893 (2010).
36, 979–984 (2003). vertical transmission of hepatitis B virus in the Chinese 116. Chevaliez, S., Hézode, C., Bahrami, S., Grare, M.
71. Yuen, M.‑F. et al. Role of hepatitis B virus genotypes in population. Pediatr. Infect. Dis. J. 33, 897–903 & Pawlotsky, J.‑M. Long-term hepatitis B surface
chronic hepatitis B exacerbation. Clin. Infect. Dis. 37, (2014). antigen (HBsAg) kinetics during nucleoside/nucleotide
593–597 (2003). 95. Pan, C. Q. et al. Tenofovir to prevent hepatitis B analogue therapy: Finite treatment duration unlikely.
72. Yuan, H.‑J. et al. Precore and core promoter mutations transmission in mothers with high viral load. N. Engl. J. Hepatol. 58, 676–683 (2013).
at the time of HBeAg seroclearance in Chinese J. Med. 374, 2324–2334 (2016). 117. Lai, C.‑L. et al. Reduction of covalently closed circular
patients with chronic hepatitis B. J. Infect. 54, This study demonstrates the efficacy of nucleotide DNA with long-term nucleos(t)ide analogue treatment
497–503 (2007). analogue therapy on reducing the rate of MTCT of in chronic hepatitis B. J. Hepatol. 66, 275–281
73. Laras, A., Koskinas, J., Avgidis, K. & Hadziyannis, S. J. HBV. (2017).
Incidence and clinical significance of hepatitis B virus 96. Zhang, H. et al. Telbivudine or lamivudine use in late This study reports the virological status in the liver
precore gene translation initiation mutations in e pregnancy safely reduces perinatal transmission of after long-term nucleos(t)ide analogue treatment.
antigen-negative patients. J. Viral Hepat. 5, 241–248 hepatitis B virus in real-life practice. Hepatology 60, 118. Nayagam, S. et al. Cost-effectiveness of community-
(1998). 468–476 (2014). based screening and treatment for chronic hepatitis B
74. Yuen, M. et al. HBsAg seroclearance in chronic 97. Brown, R. S. et al. Antiviral therapy in chronic hepatitis in The Gambia: an economic modelling analysis.
hepatitis B in Asian patients: replicative level and risk B viral infection during pregnancy: a systematic review Lancet Glob. Health 4, e568–e578 (2016).
of hepatocellular carcinoma. Gastroenterology 135, and meta-analysis. Hepatology 63, 319–333 (2015). 119. Janssen, H. L. A. et al. Pegylated interferon alfa‑2b
1192–1199 (2008). 98. Yuen, M.‑F. Prognostic determinants for chronic alone or in combination with lamivudine for HBeAg-
This is the first study to report the persistent risk hepatitis B in Asians: therapeutic implications. Gut 54, positive chronic hepatitis B: a randomised trial. Lancet
of development of HCC in patients with HBsAg 1610–1614 (2005). 365, 123–129 (2005).
seroclearance >50 years of age. 99. Iloeje, U. H. et al. Predicting cirrhosis risk based 120. Marcellin, P. et al. Peginterferon alfa‑2a alone,
75. European Association for Study of Liver & Asociacion on the level of circulating hepatitis B viral load. lamivudine alone, and the two in combination in
Latinoamericana para el Estudio del Higado. EASL- Gastroenterology 130, 678–686 (2006). patients with HBeAg-negative chronic hepatitis B.
ALEH Clinical Practice Guidelines: Non-invasive tests 100. Lai, C.‑L. & Yuen, M.‑F. Chronic hepatitis B — new N. Engl. J. Med. 351, 1206–1217 (2004).
for evaluation of liver disease severity and prognosis. goals, new treatment. N. Engl. J. Med. 359, 121. Lampertico, P. et al. Randomised study comparing 48
J. Hepatol. 63, 237–264 (2015). 2488–2491 (2008). and 96 weeks peginterferon α-2a therapy in genotype
76. Botha, J. F., Dusheiko, G. M., Ritchie, M. J. J., 101. Yuen, M.‑F. et al. Independent risk factors and D HBeAg-negative chronic hepatitis B. Gut 62,
Mouton, H. W. K. & Kew, M. C. Hepatitis B virus predictive score for the development of hepatocellular 290–298 (2012).
carrier state in black children in ovamboland: role carcinoma in chronic hepatitis B. J. Hepatol. 50, 122. Buster, E. H. C. J. et al. Factors that predict response
of perinatal and horizontal infection. Lancet 323, 80–88 (2009). of patients with hepatitis B e antigen–positive chronic
1210–1212 (1984). 102. Zhang, A.‑Y. et al. Deep sequencing analysis of hepatitis B to peginterferon-alfa. Gastroenterology
77. Seto, W.‑K. et al. Linearized hepatitis B surface quasispecies in the HBV pre‑S region and its 137, 2002–2009 (2009).
antigen and hepatitis B core-related antigen in the association with hepatocellular carcinoma. 123. Bonino, F. et al. Predicting response to
natural history of chronic hepatitis B. Clin. Microbiol. J. Gastroenterol. 52, 1064–1074 (2017). peginterferon‑2a, lamivudine and the two combined
Infect. 20, 1173–1180 (2014). 103. Yuen, M.‑F. et al. Risk for hepatocellular carcinoma for HBeAg-negative chronic hepatitis B. Gut 56,
This study reports the profile of HBsAg levels in with respect to hepatitis B virus genotypes B/C, 699–705 (2007).
different phases of chronic hepatitis B disease. specific mutations of enhancer II/core promoter/ 124. Sonneveld, M. J. et al. Response-guided peginterferon
78. Cornberg, M. et al. The role of quantitative hepatitis B precore regions and HBV DNA levels. Gut 57, 98–102 therapy in hepatitis B e antigen-positive chronic
surface antigen revisited. J. Hepatol. 66, 398–411 (2007). hepatitis B using serum hepatitis B surface antigen
(2017). This study reports the synergistic effects of levels. Hepatology 58, 872–880 (2013).
79. World Health Organization. Guidelines for the different risk factors on the development of HCC. 125. Rijckborst, V. et al. Early on‑treatment prediction of
Prevention, Care and Treatment of Persons With 104. Wong, V. W.‑S. et al. Clinical scoring system to predict response to peginterferon alfa‑2a for HBeAg-negative
Chronic Hepatitis B Infection (WHO, Geneva, 2015). hepatocellular carcinoma in chronic hepatitis B chronic hepatitis B using HBsAg and HBV DNA levels.
80. Terrault, N. A. et al. AASLD guidelines for treatment carriers. J. Clin. Oncol. 28, 1660–1665 (2010). Hepatology 52, 454–461 (2010).
of chronic hepatitis B. Hepatology 63, 261–283 105. Yang, H.‑I. et al. Risk estimation for hepatocellular 126. Marcellin, P. et al. Combination of tenofovir disoproxil
(2015). carcinoma in chronic hepatitis B (REACH‑B): fumarate and peginterferon α-2a increases loss of
81. Sarin, S. K. et al. Asian-Pacific clinical practice development and validation of a predictive score. hepatitis B surface antigen in patients with chronic
guidelines on the management of hepatitis B: a 2015 Lancet Oncol. 12, 568–574 (2011). hepatitis B. Gastroenterology 150, 134–144.e10
update. Hepatol. Int. 10, 1–98 (2016). 106. Papatheodoridis, G. et al. PAGE‑B predicts the risk of (2016).
82. Chen, C.‑J. Risk of hepatocellular carcinoma across developing hepatocellular carcinoma in Caucasians 127. Bourlière, M. et al. Effect on HBs antigen clearance of
a biological gradient of serum hepatitis B virus DNA with chronic hepatitis B on 5‑year antiviral therapy. addition of pegylated interferon alfa‑2a to nucleos(t)
level. JAMA 295, 65 (2006). J. Hepatol. 64, 800–806 (2016). ide analogue therapy versus nucleos(t)ide analogue
This study reports the direct proportional 107. Kim, G.‑A. et al. Incidence of hepatocellular carcinoma therapy alone in patients with HBe antigen-negative
relationship between HBV DNA level and the risk of after HBsAg seroclearance in chronic hepatitis B chronic hepatitis B and sustained undetectable plasma
development of HCC and cirrhosis. patients: a need for surveillance. J. Hepatol. 62, hepatitis. Lancet Gastroenterol. Hepatol. 2, 177–188
83. Seeff, L. B. et al. A serologic follow‑up of the 1942 1092–1099 (2015). (2017).
epidemic of post-vaccination hepatitis in the United 108. Kim, G.‑A. et al. HBsAg seroclearance after nucleoside 128. Fattovich, G., Bortolotti, F. & Donato, F. Natural
States Army. N. Engl. J. Med. 316, 965–970 (1987). analogue therapy in patients with chronic hepatitis B: history of chronic hepatitis B: Special emphasis on
84. Ozasa, A. et al. Influence of genotypes and precore clinical outcomes and durability. Gut 63, 1325–1332 disease progression and prognostic factors. J. Hepatol.
mutations on fulminant or chronic outcome of acute (2013). 48, 335–352 (2008).
hepatitis B virus infection. Hepatology 44, 326–334 This study reports the sustained clinical benefits in 129. Chen, C.‑J. & Yang, H.‑I. Natural history of chronic
(2006). patients with HBsAg seroclearance after cessation hepatitis B REVEALed. J. Gastroenterol. Hepatol. 26,
85. Stevens, C. E., Beasley, R. P., Tsui, J. & Lee, W.‑C. of therapy. 628–638 (2011).
Vertical transmission of hepatitis B antigen in Taiwan. 109. Lai, C.‑L. & Yuen, M.‑F. Prevention of hepatitis B 130. Buti, M. et al. Long-term clinical outcomes in cirrhotic
N. Engl. J. Med. 292, 771–774 (1975). virus-related hepatocellular carcinoma with antiviral chronic hepatitis B patients treated with tenofovir
86. Lin, H. J. et al. Evidence for intrafamilial transmission therapy. Hepatology 57, 399–408 (2013). disoproxil fumarate for up to 5 years. Hepatol. Int. 9,
of hepatitis B virus from sequence analysis of mutant 110. Hosaka, T. et al. Long-term entecavir treatment 243–250 (2015).
HBV DNAs in two Chinese families. Lancet 336, reduces hepatocellular carcinoma incidence in patients 131. Papatheodoridis, G. V., Chan, H. L.‑Y., Hansen, B. E.,
208–212 (1990). with hepatitis B virus infection. Hepatology 58, Janssen, H. L. A. & Lampertico, P. Risk of
87. Li, Z., Hou, X. & Cao, G. Is mother-to‑infant 98–107 (2013). hepatocellular carcinoma in chronic hepatitis B:
transmission the most important factor for persistent 111. Tenney, D. J. et al. Long-term monitoring shows assessment and modification with current antiviral
HBV infection? Emerg. Microbes Infect. 4, e30 (2015). hepatitis B virus resistance to entecavir in nucleoside- therapy. J. Hepatol. 62, 956–967 (2015).
88. Thio, C. L., Guo, N., Xie, C., Nelson, K. E. naïve patients is rare through 5 years of therapy. 132. Ahn, J. et al. Lower observed hepatocellular carcinoma
& Ehrhardt, S. Global elimination of mother-to‑child Hepatology 49, 1503–1514 (2009). incidence in chronic hepatitis B patients treated with
transmission of hepatitis B: revisiting the current 112. Murakami, E. et al. Implications of efficient hepatic entecavir: results of the ENUMERATE study.
strategy. Lancet Infect. Dis. 15, 981–985 (2015). delivery by tenofovir alafenamide (GS‑7340) for Am. J. Gastroenterol. 111, 1297–1304 (2016).

18 | ARTICLE NUMBER 18035 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

133. Yao, F. Lamivudine treatment is beneficial in patients 155. McMahon, M. A. et al. The HBV drug entecavir — 178. Wooddell, C. I. et al. RNAi-based treatment of
with severely decompensated cirrhosis and actively effects on HIV‑1 replication and resistance. N. Engl. chronically infected patients and chimpanzees reveals
replicating hepatitis B infection awaiting liver J. Med. 356, 2614–2621 (2007). that integrated hepatitis B virus DNA is a source of
transplantation: a comparative study using a matched, 156. Pol, S. et al. The negative impact of HBV/HCV HBsAg. Sci. Transl Med. 9, eaan0241 (2017).
untreated cohort. Hepatology 34, 411–416 (2001). coinfection on cirrhosis and its consequences. This is the first study showing the prevailing HBV
134. Fontana, R. J. et al. Determinants of early mortality in Aliment. Pharmacol. Ther. 46, 1054–1060 (2017). integration in HBeAg-negative patients resulting in
patients with decompensated chronic hepatitis B 157. Bonacini, M., Louie, S., Bzowej, N. & Wohl, A. R. relatively suboptimal efficacy of short interfering
treated with antiviral therapy. Gastroenterology 123, Survival in patients with HIV infection and viral RNA in knocking down virus transcriptions.
719–727 (2002). hepatitis B or C. AIDS 18, 2039–2045 (2004). 179. Yuen, M.‑F. et al. RNA interference therapy with
135. Shim, J. H. et al. Efficacy of entecavir in treatment- 158. Bersoff-Matcha, S. J. et al. Hepatitis B Virus ARC‑520 Injection results in long term off-therapy
naïve patients with hepatitis B virus-related reactivation associated with direct-acting antiviral antigen reductions in treatment naïve, HBeAg positive
decompensated cirrhosis. J. Hepatol. 52, 176–182 therapy for chronic hepatitis C virus: a review of cases and negative patients with chronic HBV. J. Hepatol.
(2010). reported to the U. S. Food and Drug Administration 68, S526 (2018).
136. Liaw, Y.‑F. et al. Tenofovir disoproxil fumarate (TDF), Adverse Event Reporting System. Ann. Intern. Med. 180. Mak, L.‑Y., Wong, D. K.‑H., Seto, W.‑K., Lai, C.‑L.
emtricitabine/TDF, and entecavir in patients with 166, 792 (2017). & Yuen, M. F. Hepatitis B core protein as a therapeutic
decompensated chronic hepatitis B liver disease. 159. Chen, G. et al. Hepatitis B reactivation in hepatitis B target. Expert Opin. Ther. Targets 21, 1153–1159
Hepatology 53, 62–72 (2010). and C coinfected patients treated with antiviral (2017).
137. Su, T.‑H. et al. Four-year entecavir therapy reduces agents: a systematic review and meta-analysis. 181. Bazinet, M. et al. Safety and efficacy of REP 2139 and
hepatocellular carcinoma, cirrhotic events and Hepatology 66, 13–26 (2017). pegylated interferon alfa‑2a for treatment-naive
mortality in chronic hepatitis B patients. Liver Int. 36, 160. Han, S. H. Extrahepatic manifestations of chronic patients with chronic hepatitis B virus and hepatitis D
1755–1764 (2016). hepatitis B. Clin. Liver Dis. 8, 403–418 (2004). virus co‑infection (REP 301 and REP 301‑LTF):
138. Flemming, J. A., Kim, W. R., Brosgart, C. L. 161. Orr, J. G. et al. Health related quality of life in people a non-randomised, open-label, phase 2 trial.
& Terrault, N. A. Reduction in liver transplant wait- with advanced chronic liver disease. J. Hepatol. 61, Lancet Gastroenterol. Hepatol. 2, 877–889 (2017).
listing in the era of direct-acting antiviral therapy. 1158–1165 (2014). 182. Seto, W.‑K. & Yuen, M.‑F. New pharmacological
Hepatology 65, 804–812 (2016). 162. Cullen, W., Kearney, Y. & Bury, G. Prevalence of fatigue approaches to a functional cure of hepatitis B.
139. Zhou, K. & Terrault, N. Management of hepatitis B in general practice. Ir. J. Med. Sci. 171, 10–12 Clin.  Liver Dis. 8, 83–88 (2016).
in special populations. Best Pract. Res. Clin. (2002). 183. Yuen, M.‑F. et al. Dose response and safety of the
Gastroenterol. 31, 311–320 (2017). 163. Evon, D. M. et al. Fatigue in patients with chronic daily, oral RIG‑I agonist Inarigivir (SB 9200) in
140. Cholongitas, E. & Papatheodoridis, G. V. High genetic hepatitis B living in North America: results from the treatment naïve patients with chronic hepatitis B:
barrier nucleos(t)ide analogue(s) for prophylaxis from Hepatitis B Research Network (HBRN). Dig. Dis. Sci. results from the 25 mg and 50 mg cohorts in the
hepatitis B virus recurrence after liver transplantation: 61, 1186–1196 (2016). ACHIEVE trial. J. Hepatol. 68, S509–S510 (2018).
a systematic review. Am. J. Transplant. 13, 353–362 164. Hann, H.‑W. et al. Symptomatology and health 184. Roholm, K. & Iversen, P. Changes in the liver in acute
(2012). attitudes of chronic hepatitis B patients in the USA. epidemic hepatitis (catarrhal jaundice) based on
141. Fung, J. et al. Entecavir monotherapy is effective J. Viral Hepat. 15, 42–51 (2008). 38 aspiration biopsies. Acta Pathol. Microbiol. Scand.
in suppressing hepatitis B virus after liver 165. Chen, E.‑Q. et al. Serum hepatitis B core-related 16, 427–442 (2010).
transplantation. Gastroenterology 141, 1212–1219 antigen is a satisfactory surrogate marker of 185. Blumberg, B. S. A ‘new’ antigen in leukemia sera.
(2011). intrahepatic covalently closed circular DNA in chronic JAMA 191, 541 (1965).
142. Fung, J. et al. Long-term outcomes of entecavir hepatitis B. Sci. Rep. 7, 173 (2017). This is the first publication reporting the discovery
monotherapy for chronic hepatitis B after liver 166. Wong, D. K.‑H. et al. Hepatitis B virus core-related of Australia antigen.
transplantation: results up to 8 years. Hepatology 66, antigen as a surrogate marker for covalently closed 186. Prince, A. M. An antigen detected in the blood during
1036–1044 (2017). circular DNA. Liver Int. 37, 995–1001 (2017). the incubation period of serum hepatitis. Proc. Natl
This study reports a high rate of HBsAg negativity 167. Toyoda, H., Kumada, T. & Tada, T. Hepatitis B core- Acad. Sci. USA 60, 814–821 (1968).
using long-term entecavir monotherapy after liver related antigen: a possible indicator for the 187. Sutnick, A. I. Anicteric hepatitis associated with
transplantation. termination of prophylactic nucleos(t)ide analogue Australia antigen. JAMA 205, 670 (1968).
143. Samuel, D. et al. Liver transplantation in European therapy in patients after immunosuppressive therapy. 188. Krugman, S. Infectious hepatitis. Evidence for two
patients with the hepatitis B surface antigen. N. Engl. Am. J. Gastroenterol. 112, 969–970 (2017). distinctive clinical, epidemiological, and immunological
J. Med. 329, 1842–1847 (1993). 168. Riveiro-Barciela, M. et al. Serum hepatitis B core- types of infection. JAMA 200, 365–373 (1967).
144. Adil, B. et al. Hepatitis B virus and hepatitis D virus related antigen is more accurate than hepatitis B 189. Dane, D. S., Cameron, C. H. & Briggs, M. VIRUS-like
recurrence in patients undergoing liver transplantation surface antigen to identify inactive carriers, regardless particles in serum of patients with Australia-antigen-
for hepatitis B virus and hepatitis B virus plus of hepatitis B virus genotype. Clin. Microbiol. Infect. associated hepatitis. Lancet 295, 695–698 (1970).
hepatitis D virus. Transplant. Proc. 48, 2119–2123 23, 860–867 (2017). This is the first study to describe the morphology
(2016). 169. Zhang, Z.‑Q. et al. Measurement of the hepatitis B of HBV particles under electron microscopy.
145. Coffin, C. S. et al. Virologic and clinical outcomes core-related antigen is valuable for predicting the 190. Maini, M. K. & Gehring, A. J. The role of innate
of hepatitis B virus infection in HIV-HBV coinfected pathological status of liver tissues in chronic hepatitis immunity in the immunopathology and treatment of
transplant recipients. Am. J. Transplant. 10, B patients. J. Virol. Methods 235, 92–98 (2016). HBV infection. J. Hepatol. 64, S60–S70 (2016).
1268–1275 (2010). 170. Jung, K. S. et al. Clinical outcomes and predictors for 191. Doherty, D. G. et al. The human liver contains multiple
146. Loomba, R. & Liang, T. J. Hepatitis B reactivation relapse after cessation of oral antiviral treatment in populations of NK cells, T cells, and CD3+CD56+
associated with immune suppressive and biological chronic hepatitis B patients. J. Gastroenterol. 51, natural T cells with distinct cytotoxic activities and
modifier therapies: current concepts, management 830–839 (2015). Th1, Th2, and Th0 cytokine secretion patterns.
strategies, and future directions. Gastroenterology 171. Honda, M. et al. Hepatitis B virus (HBV) core-related J. Immunol. 163, 2314–2321 (1999).
152, 1297–1309 (2017). antigen during nucleos(t)ide analog therapy is related 192. Raimondo, G. et al. Statements from the Taormina
147. Dong, H.‑J. et al. Risk of hepatitis B virus (HBV) to intra-hepatic HBV replication and development expert meeting on occult hepatitis B virus infection.
reactivation in non-Hodgkin lymphoma patients of hepatocellular carcinoma. J. Infect. Dis. 213, J. Hepatol. 49, 652–657 (2008).
receiving rituximab-chemotherapy: a meta-analysis. 1096–1106 (2015). 193. Wagner, A. A. et al. Serological pattern ‘anti-hepatitis
J. Clin. Virol. 57, 209–214 (2013).
 172. Mak, L.‑Y. et al. Review article: hepatitis B core-related B core alone’ in HIV or hepatitis C virus-infected
148. Voican, C. S. et al. Hepatitis B virus reactivation in antigen (HBcrAg): an emerging marker for chronic patients is not fully explained by hepatitis B surface
patients with solid tumors receiving systemic hepatitis B virus infection. Aliment. Pharmacol. Ther. antigen mutants. AIDS 18, 569–571 (2004).
anticancer treatment. Ann. Oncol. 27, 2172–2184 47, 43–54 (2018). 194. Carimo, A. A. et al. First report of occult hepatitis B
(2016). This review summarizes the application of HBcrAg infection among ART naïve HIV seropositive
149. Coluccio, C. et al. Hepatitis B in patients with level in different scenarios of hepatitis B disease. individuals in Maputo, Mozambique. PLoS ONE 13,
hematological diseases: an update. World J. Hepatol. 173. Lam, A. M. Hepatitis B virus capsid assembly e0190775 (2018).
9, 1043 (2017). modulators, but not nucleoside analogs, inhibit the 195. Bréchot, C. et al. Evidence that hepatitis B virus has a
150. Paul, S. et al. Role of surface antibody in hepatitis B production of extracellular pregenomic RNA and role in liver-cell carcinoma in alcoholic liver disease.
reactivation in patients with resolved infection and spliced RNA variants. Antimicrob. Agents Chemother. N. Engl. J. Med. 306, 1384–1387 (1982).
hematologic malignancy: a meta-analysis. Hepatology 61, e00680‑17 (2017). 196. Wong, D. K. H. et al. Occult hepatitis B infection and
66, 379–388 (2017). 174. Giersch, K., Allweiss, L., Volz, T., Dandri, M. HBV replicative activity in patients with cryptogenic
151. Cheung, K.‑S., Seto, W.‑K., Lai, C.‑L. & Yuen, M.‑F. & Lütgehetmann, M. Serum HBV pgRNA as a clinical cause of hepatocellular carcinoma. Hepatology 54,
Prevention and management of hepatitis B virus marker for cccDNA activity. J. Hepatol. 66, 460–462 829–836 (2011).
reactivation in cancer patients. Hepatol. Int. 10, (2017). This study reports a high percentage of the
407–414 (2016). 175. Block, T. M., Locarnini, S., McMahon, B. J., presence of HBV in patients with HCC without
152. Koziel, M. J. & Peters, M. G. Viral hepatitis in HIV Rehermann, B. & Peters, M. G. Use of current and new obvious identifiable causes.
infection. N. Engl. J. Med. 356, 1445–1454 (2007). endpoints in the evaluation of experimental hepatitis 197. Yuen, M.‑F. Need to improve awareness and
153. Bhattacharya, D. et al. Isolated hepatitis B core B therapeutics. Clin. Infect. Dis. 64, 1283–1288 management of hepatitis B reactivation in patients
antibody is associated with advanced hepatic fibrosis (2017). receiving immunosuppressive therapy. Hepatol. Int.
in HIV/HCV infection but not in HIV infection alone. 176. Wang, J. et al. Serum hepatitis B virus RNA is 10, 102–105 (2016).
J. Acquir. Immune Def. Syndr. 72, e14–e17 (2016). encapsidated pregenome RNA that may be associated 198. Hughes, S. A., Wedemeyer, H. & Harrison, P. M.
154. Panel on Antiretroviral Guidelines for Adults and with persistence of viral infection and rebound. Hepatitis delta virus. Lancet 378, 73–85 (2011).
Adolescents. Guidelines for the Use of Antiretroviral J. Hepatol. 65, 700–710 (2016). 199. Koh, C. et al. Oral prenylation inhibition with lonafarnib
Agents in Adults and Adolescents Living with HIV. 177. Lok, A. S., Zoulim, F., Dusheiko, G. & Ghany, M. G. in chronic hepatitis D infection: a proof‑of‑concept
AIDSinfo https://aidsinfo.nih.gov/contentfiles/ Hepatitis B cure: From discovery to regulatory randomised, double-blind, placebo-controlled phase 2A
lvguidelines/adultandadolescentgl.pdf (2018) approval. Hepatology 66, 1296–1313 (2017). trial. Lancet Infect. Dis. 15, 1167–1174 (2015).

NATURE REVIEWS | DISEASE PRIMERS VOLUME 4 | ARTICLE NUMBER 18035 | 19


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.
PRIMER

200. Yan, H., Liu, Y., Sui, J. & Li, W. NTCP opens the door 206. Ogawa, E., Furusyo, N. & Nguyen, M. H. Tenofovir H.L.A.J. has received grants and/or served as consultant for
for hepatitis B virus infection. Antiviral Res. 121, alafenamide in the treatment of chronic hepatitis B: AbbVie, Arbutus, Bristol-Myers Squibb, Gilead Sciences,
24–30 (2015). design, development, and place in therapy. Drug Des. Janssen, Medimmune, Merck and Roche Molecular Systems.
201. Vaillant, A. Nucleic acid polymers: Broad spectrum Devel. Ther. 11, 3197–3204 (2017). D.T.Y.L. has received research grants from Bristol-Myers
antiviral activity, antiviral mechanisms and Squibb, Gilead Sciences and Roche Molecular Systems and has
optimization for the treatment of hepatitis B and Author contributions served as consultant for AbbVie and Gilead Sciences. S.A.L.
hepatitis D infection. Antiviral Res. 133, 32–40 Introduction (M.-F.Y.); Epidemiology (D.-S.C.); Mechanisms/ has served as an adviser and received consulting fees from
(2016). pathophysiology (S.A.L. and C.-L.L.); Diagnosis, screening and Arrowhead Pharmaceuticals, AusBio Ltd., Gilead Sciences,
202. Livingston, C., Ramakrishnan, D., Strubin, M., prevention (G.M.D.); Management (C.-L.L., D.T.Y.L., H.L.A.J. Roche Molecular Systems and Janssen and has contract
Fletcher, S. & Beran, R. Identifying and characterizing and M.G.P.); Quality of life (D.T.Y.L.); Outlook (M.-F.Y.); research grants with Arrowhead Pharmaceuticals, Gilead
interplay between hepatitis B virus X protein and Overview of Primer (M.-F.Y.). Sciences and Spring Bank Pharmaceuticals, Inc. M.G.P.’s
Smc5/6. Viruses 9, 69 (2017). spouse is employed at Hoffman‑La Roche. C.-L.L. has received
203. You, C. R., Lee, S. W., Jang, J. W. & Yoon, S. K. Update Competing interests speaker fees from Bristol-Myers Squibb, Gilead Sciences and
on hepatitis B virus infection. World J. Gastroenterol. M.-F.Y. has received research grants and/or served as an Novartis and is an advisory board member of Gilead Sciences.
20, 13293–13305 (2014). adviser for AbbVie, Arrowhead Pharmaceuticals, Biocartis,
204. Lam, Y.‑F. et al. Seven-year treatment outcome of B r i s to l M
­ ye rs S q u i b b, F u j i re b i o , G i l e a d S c i e n c e s, Publisher’s note
entecavir in a real-world cohort: effects on clinical GlaxoSmithKline, LF Asia Limited, Merck Sharp & Dohme, Springer Nature remains neutral with regard to jurisdictional
parameters, HBsAg and HBcrAg levels. Clin. Transl Novartis Pharmaceuticals, Roche Molecular Systems and claims in published maps and institutional affiliations.
Gastroenterol. 8, e125 (2017). Sysmex Corporation. D.S.C. has served as an adviser for
205. Buti, M. et al. Seven-year efficacy and safety of Bristol-Myers Squibb and Merck Sharp & Dohme. G.M.D. has Reviewer information
treatment with tenofovir disoproxil fumarate for received research grants and/or served as an adviser for Nature Reviews Disease Primers thanks M. Buti and the
chronic hepatitis B virus infection. Dig. Dis. Sci. 60, Abbott Laboratories, AbbVie, Bristol-Myers Squibb, Gilead other, anonymous referee(s) for their contribution to the peer
1457–1464 (2015). Sciences, Janssen, Merck Sharp & Dohme and Transgene. review of this work.

20 | ARTICLE NUMBER 18035 | VOLUME 4 www.nature.com/nrdp


©
2
0
1
8
M
a
c
m
i
l
l
a
n
P
u
b
l
i
s
h
e
r
s
L
i
m
i
t
e
d
,
p
a
r
t
o
f
S
p
r
i
n
g
e
r
N
a
t
u
r
e
.
A
l
l
r
i
g
h
t
s
r
e
s
e
r
v
e
d
.

You might also like