ArticlePDF Available

A Novel Diagnostic Target in the Hepatitis C Virus Genome

PLOS
PLOS Medicine
Authors:

Abstract and Figures

Detection and quantification of hepatitis C virus (HCV) RNA is integral to diagnostic and therapeutic regimens. All molecular assays target the viral 5'-noncoding region (5'-NCR), and all show genotype-dependent variation of sensitivities and viral load results. Non-western HCV genotypes have been under-represented in evaluation studies. An alternative diagnostic target region within the HCV genome could facilitate a new generation of assays. In this study we determined by de novo sequencing that the 3'-X-tail element, characterized significantly later than the rest of the genome, is highly conserved across genotypes. To prove its clinical utility as a molecular diagnostic target, a prototype qualitative and quantitative test was developed and evaluated multicentrically on a large and complete panel of 725 clinical plasma samples, covering HCV genotypes 1-6, from four continents (Germany, UK, Brazil, South Africa, Singapore). To our knowledge, this is the most diversified and comprehensive panel of clinical and genotype specimens used in HCV nucleic acid testing (NAT) validation to date. The lower limit of detection (LOD) was 18.4 IU/ml (95% confidence interval, 15.3-24.1 IU/ml), suggesting applicability in donor blood screening. The upper LOD exceeded 10(-9) IU/ml, facilitating viral load monitoring within a wide dynamic range. In 598 genotyped samples, quantified by Bayer VERSANT 3.0 branched DNA (bDNA), X-tail-based viral loads were highly concordant with bDNA for all genotypes. Correlation coefficients between bDNA and X-tail NAT, for genotypes 1-6, were: 0.92, 0.85, 0.95, 0.91, 0.95, and 0.96, respectively; X-tail-based viral loads deviated by more than 0.5 log10 from 5'-NCR-based viral loads in only 12% of samples (maximum deviation, 0.85 log10). The successful introduction of X-tail NAT in a Brazilian laboratory confirmed the practical stability and robustness of the X-tail-based protocol. The assay was implemented at low reaction costs (US$8.70 per sample), short turnover times (2.5 h for up to 96 samples), and without technical difficulties. This study indicates a way to fundamentally improve HCV viral load monitoring and infection screening. Our prototype assay can serve as a template for a new generation of viral load assays. Additionally, to our knowledge this study provides the first open protocol to permit industry-grade HCV detection and quantification in resource-limited settings.
Content may be subject to copyright.
A Novel Diagnostic Target in the Hepatitis C
Virus Genome
Jan Felix Drexler
1,2,3
,BerndKupfer
2
, Nadine Petersen
1
, Rejane Maria Tommasini Grotto
4
, Silvia Maria Corvino Rodrigues
4
,
Klaus Grywna
1
,MarcusPanning
1
, Augustina Annan
1
, Giovanni Faria Silva
4
, Jill Douglas
5
,EvelynS.C.Koay
6,7
,Heidi
Smuts
8
,EduardoM.Netto
3
,PeterSimmonds
5
, Maria Ine
ˆs de Moura Campos Pardini
4
,W.KurtRoth
9
, Christian Drosten
2*
1Clinical Virology Group, Bernhard Nocht Institute for Tropical Medicine, Hamburg, Germany, 2Institute of Virology, University of Bonn, Bonn, Germany, 3Infectious
Diseases Research Laboratory, University Hospital Prof. Edgard Santos, Federal University of Bahia, Salvador, Brazil, 4University of Sa
˜o Paulo State (UNESP), Botucatu Medical
School, Blood Transfusion Centre - Molecular Biology Laboratory and Internal Medicine Department, Botucatu, Sa
˜o Paulo, Brazil, 5Virus Evolution Group, Centre for
Infectious Diseases, University of Edinburgh, Edinburgh, United Kingdom, 6Department of Pathology, Yong Loo Lin School of Medicine, National University of Singapore, 7
Molecular Diagnosis Centre, National University Hospital, Singapore, 8Division Medical Virology/National Health Laboratory Service, Department of Clinical Laboratory
Sciences, Faculty of Health Sciences, University of Cape Town, Cape Town, South Africa, 9GFE Blut mbH, Frankfurt (Main), Germany
Funding: The Brazilian study was
partially supported by a donation of
reagents from Qiagen, Germany. The
German study was supported by the
German Ministry of Health (BMGS) as
a part of funding of the National
Reference Centre for Tropical
Infections at the Bernhard Nocht
Institute and by European Union
contract number SSPE-CT-2005–
022639. The funders had no role in
study design, data collection and
analysis, decision to publish, or
preparation of the manuscript.
Competing Interests: The authors
have declared that no competing
interests exist.
Academic Editor: Paul Klenerman,
Oxford University, United Kingdom
Citation: Drexler JF, Kupfer B,
Petersen N, Grotto RMT, Rodrigues
SMC, et al. (2009) A novel diagnostic
target in the hepatitis C virus
genome. PLoS Med 6(2): e1000031.
doi:10.1371/journal.pmed.1000031
Received: September 8, 2008
Accepted: December 24, 2008
Published: February 10, 2009
Copyright:Ó2009 Drexler et al. This
is an open-access article distributed
under the terms of the Creative
Commons Attribution License, which
permits unrestricted use,
distribution, and reproduction in any
medium, provided the original
author and source are credited.
Abbreviations: bDNA, branched
DNA; HCV, hepatitis C virus; IU,
international units; LANL, Los
Alamos National Laboratory; LOD,
limit of detection; NAT, nucleic acid
testing; NCR, noncoding region; RT-
PCR, reverse transcriptase PCR; SD,
standard deviation; X-tail RT-PCR, X-
tail real-time RT-PCR assay
* To whom correspondence should
be addressed. E-mail: drosten@
virology-bonn.de
ABSTRACT
Background
Detection and quantification of hepatitis C virus (HCV) RNA is integral to diagnostic and
therapeutic regimens. All molecular assays target the viral 59-noncoding region (59-NCR), and all
show genotype-dependent variation of sensitivities and viral load results. Non-western HCV
genotypes have been under-represented in evaluation studies. An alternative diagnostic target
region within the HCV genome could facilitate a new generation of assays.
Methods and Findings
In this study we determined by de novo sequencing that the 39-X-tail element, characterized
significantly later than the rest of the genome, is highly conserved across genotypes. To prove
its clinical utility as a molecular diagnostic target, a prototype qualitative and quantitative test
was developed and evaluated multicentrically on a large and complete panel of 725 clinical
plasma samples, covering HCV genotypes 1–6, from four continents (Germany, UK, Brazil, South
Africa, Singapore). To our knowledge, this is the most diversified and comprehensive panel of
clinical and genotype specimens used in HCV nucleic acid testing (NAT) validation to date. The
lower limit of detection (LOD) was 18.4 IU/ml (95% confidence interval, 15.3–24.1 IU/ml),
suggesting applicability in donor blood screening. The upper LOD exceeded 10
9
IU/ml,
facilitating viral load monitoring within a wide dynamic range. In 598 genotyped samples,
quantified by Bayer VERSANT 3.0 branched DNA (bDNA), X-tail-based viral loads were highly
concordant with bDNA for all genotypes. Correlation coefficients between bDNA and X-tail
NAT, for genotypes 1–6, were: 0.92, 0.85, 0.95, 0.91, 0.95, and 0.96, respectively; X-tail-based
viral loads deviated by more than 0.5 log10 from 59-NCR-based viral loads in only 12% of
samples (maximum deviation, 0.85 log10). The successful introduction of X-tail NAT in a
Brazilian laboratory confirmed the practical stability and robustness of the X-tail-based
protocol. The assay was implemented at low reaction costs (US$8.70 per sample), short
turnover times (2.5 h for up to 96 samples), and without technical difficulties.
Conclusion
This study indicates a way to fundamentally improve HCV viral load monitoring and infection
screening. Our prototype assay can serve as a template for a new generation of viral load
assays. Additionally, to our knowledge this study provides the first open protocol to permit
industry-grade HCV detection and quantification in resource-limited settings.
The Editors’ Summary of this article follows the references.
PLoS Medicine | www.plosmedicine.org February 2009 | Volume 6 | Issue 2 | e10000310210
P
L
o
S
MEDICINE
Introduction
Hepatitis C virus (HCV) is one of the leading causes of
chronic hepatitis, liver cirrhosis, and hepatocellular carcino-
ma [1,2]. Seroprevalence studies suggest that at least 170
million individuals have been infected worldwide [3]. The
incidence of new HCV infections has decreased in affluent
countries owing to screening of blood products, but an
increase of global patient numbers is still expected [1–3]. At
present, six genotypes and more than 30 subtypes are well
characterized, with an overall nucleotide diversity of 31%–
33% between genotypes and 20%–25% between subtypes [4].
Diagnostic detection of HCV RNA identifies an infection
weeks to months before a detectable antibody response. To
prevent transmission by transfusion of viremic blood, nucleic
acid testing (NAT) of blood donors has become a routine
procedure in industrialized countries. As a marker of treat-
ment success, the decrease of virus RNA concentration (viral
load) as measured by quantitative NAT has become a clinical
gold standard [5,6]. For these molecular tests, several
generations of qualitative and/or quantitative HCV NAT
assays have been in use. The most recent improvement was
the introduction of real-time PCR-based assays [7–10].
However, even the latest versions of assays diverge in
performance between different genotypes [11–21]. This
divergence is of strategic clinical relevance as success of
therapy varies between HCV genotypes [22]. Consistently,
genotypes other than those occurring in industrialized
countries have been under-represented in clinical evaluation
studies [23]. Along with high costs for commercial assays,
genotype variation complicates the initiation of successful
treatment programmes in several less-affluent countries. In
analogy to HIV treatment, open-protocol and low-cost HCV
viral load technology would be desirable [24–28].
Unfortunately, the design of both commercial and in-house
tests for HCV suffers from properties of the viral 59-
noncoding region (59-NCR). This region is thought to be
the most conserved portion of the HCV genome [29] and is
targeted by all assays. Test manufacturers have made
substantial efforts to optimize 59-NCR target sites, keeping
them unpublished for all current commercial assays. Still,
there seem to be remaining issues with these tests, as
exemplified by the discontinuation of a new real-time PCR-
based assay in 2004 after initial introduction in blood
screening [13,30]. One of the most important issues may be
the existence of complex secondary structures due to the 59-
NCR’s function as an internal ribosomal entry site for
genome translation [31]. These secondary structures may
interfere in complex ways with primer or probe binding.
Due to notorious problems with the 59-NCR, a different
target region would be beneficial. However, viral genes
downstream of the 59-NCR are not useful as diagnostic
targets because of their nucleotide variability [4,32]. In the
mid 1990s, molecular biology studies on the HCV replication
cycle revealed that the genome carries an additional element
in its 39-UTR downstream of the poly-U tract, the so-called X-
tail [33,34]. The biological function of the X-tail is not
completely understood, but it has been shown that it is
involved in several crucial steps throughout the HCV life
cycle, involving both RNA-RNA and RNA-protein interac-
tions [35–38]. The X-tail has been suggested to be highly
conserved [35,36], but it is unclear whether other issues (e.g.,
secondary structures) might interfere with its molecular
detection. It may be due to lack of available sequence
information across all genotypes that the X-tail has been
neglected as a diagnostic target so far. In this study we
explored its utility for HCV detection and quantification by
sequencing the X-tail from a broad and complete panel of
HCV genotypes. To assess its diagnostic utility, we then
developed an X-tail-based viral load assay that fulfills the
same technical standards as commercial assays currently in
clinical use. Clinical validation involved 725 stored samples
across HCV genotypes 1–6 from patients from four con-
tinents (Germany, UK, South Africa, Singapore, and Brazil).
Technical robustness was demonstrated by implementation
of the assay in a routine viral load laboratory in Brazil.
Materials and Methods
Reference Plasma
The World Health Organization (WHO) international
standard reagent 96/798 was obtained from National Institute
for Biological Standards and Control, (Hertfordshire, UK)
[39]. It contained 50,000 international units (IU) of HCV,
genotype 1a, per milliliter. A genotype reference plasma panel
was obtained from the German Hepatitis C Reference Center
at the University of Essen, including lyophilized human
plasma containing HCV genotypes 1a, 1b, 2a, 2b, 2c, 2i, 3a,
4d, 5a, and 6e. Genotypes as determined by commercial assays
(Inno-Lipa HCV2, GEN-ETI-K DEIA) and in-house sequenc-
ing, as well as viral loads as determined by both the Roche
Amplicor 2.0 (Amplicor) and the Bayer VERSANT bDNA 3.0
(branched DNA [bDNA]) systems were provided with this
panel (http://www.uni-essen.de/virologie/hc_hcv-panel.html).
Only the samples mentioned in this paragraph were used
for the technical development of the assay.
Patient Plasma and Commercially Available Viral Load and
Genotyping Assays
For validation of the assay, a total of 725 human plasma
samples (one sample per patient) were studied. Of these, 598
were obtained from routine clinical testing in Germany, UK,
South Africa, Singapore, and Brazil (n¼319, 62, 102, 52, 46,
and 17 of HCV genotypes 1, 2, 3, 4, 5, and 6, respectively).
These stored samples had been selected because of their
known genotypes, and were collected at the University of
Bonn, Germany (n¼530, HCV genotypes 1 to 4, sampled from
2003 to 2007), the University of Cape Town, South Africa (n¼
46, genotype 5, sampled from 1996 to 2007), the National
University of Singapore (n¼9, genotypes 3k and 6, sampled
from 2006 to 2007), and the University of Edinburgh, UK (n¼
13, genotype 6, sampled from 1999 to 2004). Samples were
generally taken at the time of genotyping, i.e., immediately
before the beginning of therapy. We did not accept recorded
genotype information from earlier or later samples from a
given patient. An additional set of plasma samples, collected
from 2005 to 2007 at the University of Sa
˜o Paulo State, Brazil,
contained 127 nongenotyped samples from patients during
HCV therapy. All samples were anonymized and approval was
obtained by the ethical committees of the University of Bonn
Medical Centre, Bonn, Germany; the University of Sa
˜o Paulo
State (UNESP); Botucatu Medical School, Botucatu, Sa
˜o
Paulo, Brazil; and the Federal University of Bahia Medical
School, Salvador, Bahia, Brazil. Viral loads in all samples were
PLoS Medicine | www.plosmedicine.org February 2009 | Volume 6 | Issue 2 | e10000310211
Novel Diagnostic HCV Target
determined at the University of Bonn with bDNA on a
VERSANT 440 Molecular System. Samples were genotyped by
the VERSANT HCV Genotype assay (Line Probe assay [LiPA])
(Bayer Diagnostics). Genotyping of the South African and
British samples was done by restriction fragment length
polymorphism (RFLP), as previously described [40,41] and
confirmed by LiPA. The genotypes of all samples from
Singapore were confirmed by sequencing of the 59-NCR and
nonstructural protein 5b (NS5b) domains [42]. Samples were
stored at 20 8C until quantification with the X-tail real-time
reverse transcriptase (RT)-PCR assay (X-tail RT-PCR). All
Brazilian samples were tested by bDNA on a VERSANT 340
Molecular System and X-tail RT-PCR in the local laboratory
in Brazil.
Quantification of HCV Viral Load by X-Tail RT-PCR
Reactions of 50 ll contained 20 ll of RNA extract, 53
reaction buffer (Qiagen One-step RT-PCR kit), 200 lMof
each dNTP, 200 nM of primer XTF5 (GTGGCTCCATCT-
TAGCCCTAGT), 300 nM of primer HCMgR2
(TGCGGCTCACGGACCTTT), 100 nM of HCV-specific probe
HCVMGB2 (FAM-CACGGCTAGCTGTG-Black Hole Quench-
er 2/Minor grove binder), 100 nM of Internal Control-probe
YFPY (VIC-ATCGTTCGTTGAGCGATTAGCAG-Black Hole
Quencher 1), and 2 ll of Enzyme Mix. Thermal cycling was
done on Applied Biosystems 7700 or 7500 SDS instruments
under the following conditions: 55 8C, 10 min; 95 8C, 15 min;
45 cycles at 94 8C, 10 s and 58 8C, 15 s.
Computing
Statistical analyses were performed with the SPSS 13 (SPSS).
Alignments were generated using BioEdit (http://www.mbio.
ncsu.edu/BioEdit/bioedit.html). Sequences were analyzed us-
ing the Lasergene software package (DNAStar). All database
workwasperformedonlineonLosAlamosNational
Laboratory (LANL) (http://hcv.lanl.gov/content/hcv-index)
and euHCV (http://euhcvdb.ibcp.fr/euHCVdb) domains.
Supplemental Materials and Methods
A STARD diagram and checklist are included in Figure S1
and Text S1. A more detailed description of the X-tail NAT
assay including protocols for RNA extraction and data on
technical performance are provided in Text S2. A bench
protocol and instructions for requesting controls is provided
in Text S3.
Accession Numbers
Sequences determined in this study have been submitted to
GenBank (http://www.ncbi.nlm.nih.gov/Genbank) under acces-
sion numbers EU835523-EU835532.
Results
Analysis of the X-Tail Region
In order to examine the conservedness of the X-tail region,
a nucleic acid sequence alignment was generated that
contained all X-tail sequences stored in the LANL and
euHCV databases as of January 2007. These databases
comprised sequences classified as genotypes 1, 2, and 3. Sets
of sequencing primers for the ultimate 39-end of the genome
were identified, and the X-tail was sequenced from reference
plasma samples of HCV genotypes 1–6. Together with one
outlier sequence taken from the initial characterization study
on the X-tail [33] that was found later on to correspond to
genotype 4 (PS, unpublished data), these novel sequences
were added to the alignment. As shown in Figure 1, the
degree of conservedness in the X-tail was comparable to that
of the 59-NCR, which is the target of all existing viral load
assays.
Experimental Identification of a Diagnostic Target Region
within the X-Tail
The complete 39-UTR alignment was evaluated as a target
for real-time PCR design. Since primer design software was
not able to identify suitable candidate oligonucleotides, in a
first approach five forward primers, five reverse primers, and
two probes were selected upon inspection of the alignment.
All of the 50 resulting real-time PCR sets were tested
experimentally for reaction efficiency. However, even after
selection of the most effective oligonucleotides and optimi-
zation of reaction conditions, the lower limit of detection
(LOD) did not fall below 150 IU/ml (see Text S2 for additional
Figure 1. Nucleotide Similarities within HCV 59- and 39-Genome Ends
Top panel: schematic representation of the HCV reference genome H77 as given in the 2008 LANL database.
Bottom panel: Percent nucleotide identity along the 59-NCR and the X-tail, as calculated by a sliding window analysis with VectorNTI software. Window
size was 2 nucleotides. The alignment used for the sliding window analysis contained the complete HCV genotype reference panel (n¼60 sequences)
from the LANL HCV database (available at http://hcv.lanl.gov/content/sequence/NEWALIGN/align.html) and all X-tail sequences available in GenBank as
shown in Text S2, including the sequences determined de novo in this study (EU835523-EU835532).
doi:10.1371/journal.pmed.1000031.g001
PLoS Medicine | www.plosmedicine.org February 2009 | Volume 6 | Issue 2 | e10000310212
Novel Diagnostic HCV Target
information on technical evaluation procedures). Since
primers had a perfect match with the target sequence, the
limitation in sensitivity was ascribed to RNA secondary
structures possibly interfering with oligonucleotide hybrid-
ization.
RNA secondary structures of both the X-tail and the 59-
NCR were modeled at PCR primer annealing conditions (58
8C and 50 mM ion concentration) using MFOLD [43]. In the
59-NCR, secondary structures in all genotypes were almost
completely dissolved, but still a few stable stem-loop elements
remained. Of note, some of these fell in the conserved regions
targeted by a prototypic 59-NCR assay, the Roche Amplicor
Monitor, and probably also the later generation COBAS
TaqMan system [13]. For the X-tail, a 10-nucleotide-stem
element in the stem-loop 1 region towards the 39end of the X-
tail was predicted to be stable at the same conditions. Figure 2
shows the structure predictions with exemplified oligonucleo-
tides for X-tail and 59-NCR. It was concluded that stable stem
structures were likely to interfere with binding of all antisense
primers and most of the probes used up to this point.
Prototype Viral Load Assay Based on the HCV X-Tail
Consequently, the antisense primer and probe binding sites
were moved further upstream into the amplicon. Out of
several new candidate primers and probes, an oligonucleotide
combination was determined that provided very high
amplification efficiency (final assay oligonucleotides as shown
in the Materials and Methods section). These primers and
probe were directly adjacent, without unoccupied nucleo-
tides between them. Very few nucleotide mismatches
occurred with any of the sequences taken from GenBank or
determined from reference plasma (see Figure 2). According
to established data, these mismatches were highly unlikely to
interfere with assay performance [44–46].
To assess the diagnostic applicability of the X-tail, this PCR
set was developed into a clinical-grade prototype assay. The
input RNA volume was maximized to 20 ll in a 50-ll assay
volume, and the chemistry was adjusted for use with a semi-
automated nucleic acids extraction. In order to detect PCR
inhibition, a competitive internal control was incorporated. It
was amplified by the same primers as the diagnostic target but
was detected by a probe of different sequence composition
and different fluorescent labeling. To enhance assay stability,
control RNA was formulated to be nuclease resistant and
added at working concentration to all reactions at the lysis
buffer stage. As a reference standard for viral loads, a
noninfectious and stable copy of the HCV 1a X-tail was cloned
Figure 2. Secondary Structure Predictions
Secondary structure of the HCV 59-NCR (A) and the 39-X-tail (B) ofHCV 1a reference genome H77 as predicted by MFOLD [43]. Prediction was done at 50 mM
salt concentration and PCR primer annealing temperature (58 8C). Free energies DGwere1.6 310
3
J/mol for (A) and 5,06 310
4
J/mol for (B). 59-NCR
structure predictions were identical for all genotypes in the LANL genotype reference panelexcept genotype 2, which had one additional stem-loop element
(not covered by primers, not shown in [A]). Structure predictions for all X-tail sequences (database and new sequences) were identical (not shown in [B]).
(A) Binding sites of oligonucleotides used by the Roche Amplicor assay are shown in red (sense primer KY80, hybridization probe KY150, antisense
primer KY78). Nucleotide variability at these binding sites is shown in Text S2).
(B) Outer lines in red identify the hybridization sites of oligonucleotides used in initial X-tail RT-PCR set, yielding limited sensitivity (forward primer F5,
TaqMan probe P1, reverse primer R7). Interior lines in green identify the final X-tail prototype assay (forward primer F5, reverse primer R2, probe MGB2).
Nucleotide variability at these sites is shown in Text S2.
doi:10.1371/journal.pmed.1000031.g002
PLoS Medicine | www.plosmedicine.org February 2009 | Volume 6 | Issue 2 | e10000310213
Novel Diagnostic HCV Target
in an armored RNA phage (refer to Text S2 for synthesis and
calibration of internal control and reference standard).
Before evaluation of its technical and clinical performance,
the specificity of the assay for HCV was confirmed on cell
culture supernatants or sera containing the following
Flaviviridae species: dengue virus 1, 2, 3, and 4; yellow fever
virus; tick-borne encephalitis virus; St. Louis encephalitis
virus; West Nile virus; and hepatitis G virus (HCV)/GB virus-C.
Due to frequent co-infection of patients with HIV-1 and
HCV, an HIV-1 genotype reference panel comprising 11 HIV-
1 genotypes from groups M, N, and O [47] was acquired from
National Institute for Biological Standards and Control
(product code 01/466) and tested with X-tail RT-PCR. High
RNA content in all of these materials was confirmed by a
flavivirus genus-specific RT-PCR and an HIV-1 in-house real
time RT-PCR assay [28,48]. No nonspecific amplification
occurred with any of these materials (unpublished data).
LOD and Quantification Range
A common technical specification for the sensitivity of
NAT is the 95% LOD, i.e., the concentration down to which
.95% of iterative tests will detect virus. 95% LODs for the X-
tail-based prototype assay were determined for each individ-
ual genotype by parallel limiting dilution testing as described
in Text S2. Results indicated high sensitivity across all
genotypes (Table 1). The overall LOD was as low as 18.4 IU/
ml (Figure 3).
Next, the upper quantification limit was evaluated, i.e., the
maximal RNA concentration that can be measured without
technical bias contributed by the system. This is relevant
because HCV patients may show extraordinarily high viral
loads. Synthetic Armored RNA standards exceeding clinically
observed viral loads (1,890,000, 18,900,000, and 189,000,000
IU/ml) were tested. The determined quantities did not deviate
from those expected by dilution factor (unpublished data).
Accuracy in Viral Load Monitoring
Intra-assay coefficient of variation (CV) at 7,245,000 IU/ml
of HCV RNA was 5.76% with a standard deviation (SD) of
0.03 log10. Inter-assay CVs ranged from 6.11 to 18.42% (SDs,
0.03–0.09 log10) at 1,890–189,000 IU/ml (refer to Text S2).
To evaluate the utility of the X-tail in viral load monitor-
ing, the prototype assay was compared against bDNA. The
latter was chosen as a clinical gold standard because it
represented a well-established assay that is more robust
against genotype bias than other clinical assays [49–51]. All of
the 598 plasma samples, covering HCV genotypes 1–6 and all
ranges of viral loads occurring in HCV patients, were tested
with X-tail RT-PCR and results compared to bDNA. Samples
with results below or above the quantification ranges of
bDNA or X-tail RT-PCR (n¼47) were excluded from
correlation analysis, leaving 551 samples. As shown in Figure
3, X-tail RT-PCR correlated well with bDNA for all genotypes
tested, with correlation coefficients of 0.92, 0.85, 0.95, 0.91,
0.95, and 0.96 for HCV genotypes 1 to 6, respectively. All
correlations were highly significant (p,0.01 for all, Pearson’s
goodness of fit test).
Figure 4 shows differences in absolute quantification
results for genotypes 1–6. Differences of more than 0.5
log10 occurred infrequently (12.0% of samples in total;
maximally observed deviation: 0.85 log10), with a balanced
distribution to both higher and lower viral loads (5.1% and
6.9%, respectively). Mean and median log10 differences and
SDs were small for all six genotypes, and all were below
clinical significance (Figure 4). X-tail RT-PCR yielded
minimally higher viral loads than bDNA for all genotypes
except genotype 4 (mean log10 differences were 0.00, 0.01,
0.17, 0.09, 0.07, and 0.12 for genotypes 1 to 6, respectively).
In order to assess whether the slight underquantification of
genotype 4 could have been caused by a systematic error, the
two samples that showed strongest underquantification (up to
0.7 log10) were sequenced. They showed no nucleotide
mismatches at the oligonucleotide binding sites, suggesting
other reasons for the observed deviations (handling, storage,
error in the gold standard, etc.).
Clinical Sensitivity
Because correlation analysis only included samples that
were positive in both tests, it did not reflect overall clinical
sensitivity. A comparison of qualitative detection rates is
shown in Table 2. Sixteen of a total of 598 samples (2.68%)
were below the detection limit of bDNA but detectable by X-
tail RT-PCR at a median viral load of 284 IU/ml. Three of
these samples had viral loads that should have been
detectable with bDNA, according to its LOD (615 IU/ml).
Two samples were negative by X-tail RT-PCR despite RNA
detection by bDNA at viral loads of 989 and 4,681 IU/ml,
respectively.
A total of 15 samples had viral loads above the upper cut-
off of bDNA, requiring predilution and repetition of the
bDNA assay. All of them were quantifiable by X-tail RT-PCR
upon first testing.
Table 1. Lower LOD by HCV Genotype
HCV RNA Input (IU/ml)
a
Hit Rate
c
in Genotype Reference Samples
1a 1b 2a 2b 2c 2i 3a 4 5a 6 All
54 5/5 5/5 5/5 5/5 5/5 5/5 5/5 5/5 5/5 5/5 50/50
18 5/5 5/5 5/5 4/5 5/5 4/5 5/5 5/5 4/5 4/5 46/50
63/5 2/5 5/5 1/5 4/5 3/5 5/5 2/5 3/5 2/5 30/50
21/5 0/5 0/5 0/5 2/5 3/5 3/5 0/5 0/5 0/5 9/50
LOD
b
11.0 7.5 4.5 22.5 8.9 33.6 3.1 7.5 22.6 22.7 18.4
a
Based on Bayer VERSANT bDNA 3.0 viral load assay.
b
LOD (IU/ml) at 95% probability, Probit analysis (see Text S2).
c
nPositive test results per five parallel tests performed.
doi:10.1371/journal.pmed.1000031.t001
PLoS Medicine | www.plosmedicine.org February 2009 | Volume 6 | Issue 2 | e10000310214
Novel Diagnostic HCV Target
Potential for Implementation in Resource-Limited Settings
Affordable viral load monitoring would be desirable in
resource-limited settings with high HCV prevalence. To
evaluate whether the X-tail prototype assay would provide
adequate stability and quality in comparison to 59-NCR-based
assays, it was implemented in a diagnostic center involved in
the Brazilian HCV treatment program. Protocols, controls,
and standards for X-tail NAT were provided to the laboratory
in Brazil. All other reagents were purchased locally. Plasma
samples from 127 patients were tested by bDNA and X-tail
NAT. The correlation coefficient between viral loads obtained
with both assays was 0.97 (p,0.001) at a mean quantitative
difference of 0.05 log10 (SD 0.018). As shown in Figure 5,
almost all quantitative differences were below 0.5 log10. Only
four outlier samples occurred, with deviations to higher and
lower quantification results of up to 1 log10 in X-tail RT-PCR.
Discussion
In this study we have identified a new diagnostic target
region in the HCV genome. A prototype RT-PCR assay based
on the HCV X-tail proved to be highly sensitive (18.4 IU/ml,
95% probit probability), robust against genotype variation,
and highly precise in virus quantification. The assay was
efficiently implemented and projected to be highly cost
efficient in an emerging country setting. These data may assist
in translating state-of-the art diagnostic technology to less
affluent settings.
The X-tail region of the HCV genome has been known for
several years [33,34] but has not been used as a diagnostic
target so far. Its biological functions are now becoming
clearer, and it is likely that its role in the virus life cycle
involves both RNA and protein interactions. The three X-tail
stem-loop structures (SL1–3) seem to be involved in negative-
strand synthesis and significantly enhance HCV replication
by means of binding to ribosomal and other cellular proteins
[35–38]. A kissing-loop interaction between an X-tail stem-
loop structure (SL2) and an RNA loop within the non-
structural protein 5b (NS5b) region was shown to be essential
for HCV replication [52]. These multiple mechanisms of
interaction imply a high degree of conservedness in the X-
Figure 3. Correlation of Viral Loads as Determined by X-Tail RT-PCR (x-Axis) and bDNA (y-Axis)
Genotypes and numbers of samples (n) are given in the bottom right corner of each panel. Pearson’s bivariate correlation coefficients were 0.92, 0.85, 0.95,
0.91, 0.95, and 0.96 for HCV genotypes 1 to 6, respectively. The dashed lines represent ideal correlations. Genotype 5, 0.07/0.31; and genotype 6, 0.12/0.19.
doi:10.1371/journal.pmed.1000031.g003
PLoS Medicine | www.plosmedicine.org February 2009 | Volume 6 | Issue 2 | e10000310215
Novel Diagnostic HCV Target
tail, which could indeed be confirmed for all genotypes in this
study. As anticipated [32,53], the region proved as conserved
as the 59-NCR across all genotypes, and seemed to exhibit less
problematic structural features than the 59-NCR.
To circumvent the limitations that the 59-NCR presents in
diagnostics, we have established the HCV X-tail as a target for
molecular detection and quantification. Initially we could not
be sure about its utility for several reasons implied by its
biological functions. A possible limitation could have been its
position at the end of the genome and beyond the poly-U
tract, making the X-tail prone to nuclease degradation.
Moreover, due to its functions as a 39-element it could have
been associated strongly with viral or cellular proteins.
Although we could not investigate these issues in our study,
we are confident about its diagnostic utility from the clinical
part of our study, showing that X-tail-based viral loads were
highly concordant with results from bDNA testing. bDNA was
chosen as a gold standard because it uses multiple probes
along the 59-NCR and initial core region and has proven to be
the most robust viral load test compared to other assays [49–
51]. We used a large panel of clinical specimens that included
sufficient numbers of samples of genotypes 1–6 (n¼725 in
total, 598 genotyped samples). To our knowledge, this is the
most diversified and comprehensive panel of clinical speci-
mens used in the validation of HCV NAT so far [7,10,14,17–
21,49,54,55]. For genotypes 4–6, earlier studies relied on
sparse genotype reference samples, which have also been used
for the original design of assays and may under-represent the
genetic diversity observed in clinical samples. Our study
included field clinical samples of all genotypes in substantial
numbers, sampled in four different geographic locations. The
genotype-related robustness of the prototype X-tail assay was
at least equivalent with that of proprietary assays
[12,13,15,17–21,23,49,51,56,57]. Quantitative correlation and
lower limits of detection were highly consistent for all
genotypes by X-tail RT-PCR. The lower LOD (18.4 IU/ml)
and the upper quantification range of our assay
(.189,000,000 IU/ml) were equivalent to that of the last
generation Roche COBAS TaqMan HCV test and the Abbott
HCV RealTime assay [7–10,23,58]. Critically, X-tail-based viral
loads and bDNA results were highly congruent, making the
assay compatible with other quantification systems. This
compatibility is critical when patients switch their treating
institution, which may entail switching between viral load
tests. The observed quantitative deviations were generally
,0.5 log10, i.e., below clinical relevance.
In resource-limited settings our prototype assay could be
used instead of more costly proprietary assays in HCV
treatment and testing. It is not patented, has a simple and
accessible formulation (refer to the bench protocol and
instructions for requesting controls in Text S3), and is
appropriate for rare HCV genotypes. Several less-affluent
countries have established successful HIV treatment pro-
grams, but suffer from considerable HCV prevalence as well.
One important example is Brazil, where networks of well-
managed public laboratories conduct viral load testing as a
part of the national HIV-1 treatment program [59]. Intensive
efforts have been made to develop low-cost viral load assays
for HIV-1 in Brazil and elsewhere [24–28]. The most
important issue with such in-house assays is their robustness.
In proprietary commercial assays robustness is contributed
by advanced features like automated RNA preparation,
nuclease resistant calibrators, and synthetic internal controls.
We have incorporated all of these features in the open
protocol HCV X-tail assay. A second issue in using in-house
Table 2. Detection of HCV by X-Tail RT-PCR Versus bDNA
Sample Testing Results X-Tail RT-PCR bDNA nSamples Median Viral Load (IU/ml)
a
Range Viral Load (IU/ml)
Positive Positive 551
Negative Negative 14
Positive Negative 16 284 43–1,816,128
Negative Positive 2 2,835 989–4,681
Positive Overflow
b
15 16,124,130 251,662–196,797,600
Total — 598
a
Median HCV RNA IU/ml in the positive assay when below LOD in compared assay (,615 IU/ml for bDNA 3.0 and ,46 IU/ml for X-tail RT-PCR).
b
Could not be quantified due to viral loads .7,692,308 IU/ml (viral load was determined a posteriori by predilution).
doi:10.1371/journal.pmed.1000031.t002
Figure 4. Quantitative Deviations between X-Tail–Based and bDNA Viral
Loads, per Genotype (y-Axis)
Genotypes are indicated below the x-axis, the number of samples tested
per genotype (n) above the x-axis. Each box shows the median,
interquartile range (box length, containing 50% of data) and whiskers
show extreme values (there were no statistical outliers). Deviations
between X-tail RT-PCR and bDNA [log10 X-tail RT-PCR log10 bDNA]
were genotype 1, 0.00/0.31 (mean/SD); genotype 2, 0.01/0.33; genotype
3, 0.17/0.32; genotype 4, 0.09/0.37; genotype 5, 0.07/0.31; and
genotype 6, 0.12/0.19.
doi:10.1371/journal.pmed.1000031.g004
PLoS Medicine | www.plosmedicine.org February 2009 | Volume 6 | Issue 2 | e10000310216
Novel Diagnostic HCV Target
assays is quality control, which is shifted largely to the
production process in commercial kits, but can be managed
at the laboratory level if appropriate quality control
procedures are in place. In the context of HIV viral load
monitoring, we and others have demonstrated that in-house
testing in emerging countries is feasible in well-managed
laboratories [24,25,27,28]. The potential for translation of the
assay was demonstrated by implementation in a Brazilian
laboratory and on-site comparison against the Bayer bDNA
assay. X-tail RT-PCR had short turnover times (2.5 h for up to
96 samples) and was implemented from provided protocols
without technical difficulties. Overall results suggest that X-
tail RT-PCR is suitable for the Brazilian setting, even though
genotyping was not feasible at the study site. The overall HCV
genotype distribution in Sa
˜o Paulo state is known to be
approximately 65%, 5%, and 30% of genotypes 1, 2, and 3,
respectively [60]. Reaction costs of the assay ranged around
US$8.70 per sample, or 8.1% of the current viral load costs in
Brazil. It should be mentioned that under European
conditions, an additional service license fee would apply,
increasing the cost of one assay to US$18 per sample (refer to
Table 3 for an estimation of costs).
Finally, HCV is an important issue for blood safety in
emerging countries. Again, Brazil provides an example
representative of many other emerging countries. An urgent
federal decree in 2003 demanded the general testing of
donated blood for HCV by NAT [61], but this strategy has not
yet been implemented. According to estimates by the Brazil-
ian Ministry of Health, it would cost about US$40 million to
screen 4 million blood donations annually [62].
We previously demonstrated the feasibility of in-house
testing of blood donors for HCV by NAT methods [63]. Due
to the high sensitivity of X-tail RT-PCR (18.4 IU/ml), the
prototype assay could be used for testing pooled blood donor
samples. With the well-established approach of pooling 24
donors prior to NAT testing [64], the projected LOD would
be 441 IU/ml per donor (18.4 IU/ml 324). This projection
would match the sensitivity achieved with pooled blood
donor screening in Europe [65,66] and North America [64,67].
Calculating the cost for one X-tail NAT assay at US$10
(US$8.70 for RT-PCR according to Table 3, plus US$1.30 for
extra consumables due to pooling), and assuming 20%
additional costs for controls and confirmatory testing, 24-
donor pool testing by X-tail RT-PCR would amount to US$2
Table 3. Approximate Pricing of HCV Viral Load Assays, US Dollars, without Taxes
Item X-Tail RT-PCR Germany
a
Roche Cobas Amplicor, Brazil
b
Bayer bDNA, Brazil
b
RNA preparation 2.70
a
Included
c
Included
c
Reagents 4.20 Included
c
Included
c
Consumables and controls 1.80 0.80 0.80
License fees
d
10.00 Included
c
Included
c
Total 18.70 106.80 106.80
a
Based on current conditions in Germany.
b
Current conditions for the Sa
˜o Paulo State Viral Hepatitis Network (unpublished data based on personal communications [RMTG and MIMCP]; see also [68]). Purchase by laboratories
outside this privileged network may involve higher prices up to about US$190 per test.
c
These items are included in the cost of the commercial assay.
d
Service licence fees for using the generic 59-nuclease probe format.
doi:10.1371/journal.pmed.1000031.t003
Figure 5. Results Obtained with X-Tail Viral Load Monitoring, Implemented in Brazil
127 samples were measured in a Brazilian HIV-1 viral load monitoring centre by Bayer VERSANT 3.0 HCV bDNA assay and by X-tail in-house PCR. (A)
Correlation of viral loads between X-tail RT-PCR (x-axis) and Bayer bDNA 3.0 (y-axis). The dashed line represents an ideal correlation. Pearson’s bivariate
correlation coefficient was 0.97.
(B) Quantitative differences. The box shows the median and interquartile range (box length). The whiskers represent an extension of the 25th or 75th
percentiles by 1.5 times the interquartile range. Datum points beyond the whisker range are considered as outliers and marked as asterisks.
doi:10.1371/journal.pmed.1000031.g005
PLoS Medicine | www.plosmedicine.org February 2009 | Volume 6 | Issue 2 | e10000310217
Novel Diagnostic HCV Target
million per 4 million donations annually (4 million/24 3
US$10 31.2). This means a reduction down to 5% of the
current economic estimate. In view of the technical robust-
ness and transferability of the assay described in this report,
we believe that such a strategy could be a realistic approach
in emerging countries aiming to increase blood safety
through the introduction of HCV NAT.
Supporting Information
Alternative Language Abstract S1. Translation of the Abstract into
French by Felix Drexler
Found at doi:10.1371/journal.pmed.1000031.sd001 (29 KB DOC).
Alternative Language Abstract S2. Translation of the Abstract into
German by Felix Drexler and Christian Drosten
Found at doi:10.1371/journal.pmed.1000031.sd002 (29 KB DOC).
Alternative Language Abstract S3. Translation of the Abstract into
Portuguese by Felix Drexler and Maria Inez Pardini
Found at doi:10.1371/journal.pmed.1000031.sd003 (31 KB DOC).
Alternative Language Abstract S4. Translation of the Abstract into
Spanish by Oscar Martinez Martinez
Found at doi:10.1371/journal.pmed.1000031.sd004 (27 KB DOC).
Figure S1. STARD Diagram
Found at doi:10.1371/journal.pmed.1000031.sg001 (11 KB PDF).
Text S1. STARD Checklist
Found at doi:10.1371/journal.pmed.1000031.sd005 (18 KB PDF).
Text S2. Detailed Description of the X-tail NAT Assay
Found at doi:10.1371/journal.pmed.1000031.sd006 (696 KB DOC).
Text S3. Bench Protocol and Instructions for Requesting Controls
Found at doi:10.1371/journal.pmed.1000031.sd007 (86 KB PDF).
Acknowledgments
We are grateful to technical staff at the involved institutions.
Author contributions. JFD and CD conceived and designed the
experiments. JFD, BK, NP, RMTG, SMCR, KG, MP, AA, GFS, JD,
ESCK, HS, PS, MIdMCP, and WKR performed the experiments. JFD,
EMN, and CD analyzed the data. BK, RMTG, SMCR, GFS, JD, ESCK,
HS, PS, MIdMCP, and WKR contributed materials. JFD and CD wrote
the paper.
References
1. Poynard T, Yuen MF, Ratziu V, Lai CL (2003) Viral hepatitis C. Lancet 362:
2095–2100.
2. Perz JF, Alter MJ (2006) The coming wave of HCV-related liver disease:
dilemmas and challenges. J Hepatol 44: 441–443.
3. Anonymous (1999) Global surveillance and control of hepatitis C. Report
of a WHO Consultation organized in collaboration with the Viral Hepatitis
Prevention Board, Antwerp, Belgium. J Viral Hepat 6: 35–47.
4. Simmonds P, Bukh J, Combet C, Deleage G, Enomoto N, et al. (2005)
Consensus proposals for a unified system of nomenclature of hepatitis C
virus genotypes. Hepatology 42: 962–973.
5. Berg T, Sarrazin C, Herrmann E, Hinrichsen H, Gerlach T, et al. (2003)
Prediction of treatment outcome in patients with chronic hepatitis C:
significance of baseline parameters and viral dynamics during therapy.
Hepatology 37: 600–609.
6. Terrault NA, Pawlotsky JM, McHutchison J, Anderson F, Krajden M, et al.
(2005) Clinical utility of viral load measurements in individuals with
chronic hepatitis C infection on antiviral therapy. J Viral Hepat 12: 465–
472.
7. Beld M, Sentjens R, Rebers S, Weegink C, Weel J, et al. (2002) Performance
of the New Bayer VERSANT HCV RNA 3.0 assay for quantitation of
hepatitis C virus RNA in plasma and serum: conversion to international
units and comparison with the Roche COBAS Amplicor HCV Monitor,
Version 2.0, assay. J Clin Microbiol 40: 788–793.
8. Halfon P, Bourliere M, Penaranda G, Khiri H, Ouzan D (2006) Real-time
PCR assays for hepatitis C virus (HCV) RNA quantitation are adequate for
clinical management of patients with chronic HCV infection. J Clin
Microbiol 44: 2507–2511.
9. Lee SC, Antony A, Lee N, Leibow J, Yang JQ, et al. (2000) Improved version
2.0 qualitative and quantitative AMPLICOR reverse transcription-PCR
tests for hepatitis C virus RNA: calibration to international units, enhanced
genotype reactivity, and performance characteristics. J Clin Microbiol 38:
4171–4179.
10. Sizmann D, Boeck C, Boelter J, Fischer D, Miethke M, et al. (2007) Fully
automated quantification of hepatitis C virus (HCV) RNA in human plasma
and human serum by the COBAS AmpliPrep/COBAS TaqMan system. J
Clin Virol 38: 326–333.
11. Morishima C, Chung M, Ng KW, Brambilla DJ, Gretch DR (2004) Strengths
and limitations of commercial tests for hepatitis C virus RNA quantifica-
tion. J Clin Microbiol 42: 421–425.
12. Colson P, Motte A, Tamalet C (2006) Broad differences between the
COBAS AmpliPrep total nucleic acid isolation-COBAS TaqMan 48
hepatitis C virus (HCV) and COBAS HCV monitor v2.0 assays for
quantification of serum HCV RNA of non-1 genotypes. J Clin Microbiol
44: 1602–1603.
13. Gelderblom HC, Menting S, Beld MG (2006) Clinical performance of the
new Roche COBAS TaqMan HCV Test and High Pure System for
extraction, detection and quantitation of HCV RNA in plasma and serum.
Antivir Ther 11: 95–103.
14. Mellor J, Hawkins A, Simmonds P (1999) Genotype dependence of hepatitis
C virus load measurement in commercially available quantitative assays. J
Clin Microbiol 37: 2525–2532.
15. Chevaliez S, Bouvier-Alias M, Brillet R, Pawlotsky JM (2007) Overestimation
and underestimation of hepatitis C virus RNA levels in a widely used real-
time polymerase chain reaction-based method. Hepatology 46: 22–31.
16. Tang YW, Sefers SE, Li H (2005) Primer sequence modification enhances
hepatitis C virus genotype coverage. J Clin Microbiol 43: 3576–3577.
17. Colucci G, Ferguson J, Harkleroad C, Lee S, Romo D, et al. (2007) Improved
COBAS TaqMan hepatitis C virus test (Version 2.0) for use with the High
Pure system: enhanced genotype inclusivity and performance character-
istics in a multisite study. J Clin Microbiol 45: 3595–3600.
18. Pittaluga F, Allice T, Abate ML, Ciancio A, Cerutti F, et al. (2008) Clinical
evaluation of the COBAS Ampliprep/COBAS TaqMan for HCV RNA
quantitation in comparison with the branched-DNA assay. J Med Virol 80:
254–260.
19. Sandres-Saune K, Abravanel F, Nicot F, Peron JM, Alric L, et al. (2007)
Detection and quantitation of HCV RNA using real-time PCR after
automated sample processing. J Med Virol 79: 1821–1826.
20. Sarrazin C, Dragan A, Gartner BC, Forman MS, Traver S, et al. (2008)
Evaluation of an automated, highly sensitive, real-time PCR-based assay
(COBAS Ampliprep/COBAS TaqMan) for quantification of HCV RNA. J
Clin Virol 43: 162–168.
21. Vermehren J, Kau A, Gartner BC, Gobel R, Zeuzem S, et al. (2008)
Differences between two real-time PCR-based hepatitis C virus (HCV)
assays (RealTime HCV and Cobas AmpliPrep/Cobas TaqMan) and one
signal amplification assay (Versant HCV RNA 3.0) for RNA detection and
quantification. J Clin Microbiol 46: 3880–3891.
22. Pawlotsky JM (2003) Mechanisms of antiviral treatment efficacy and failure
in chronic hepatitis C. Antiviral Res 59: 1–11.
23. Sarrazin C, Gartner BC, Sizmann D, Babiel R, Mihm U, et al. (2006)
Comparison of conventional PCR with real-time PCR and branched DNA-
based assays for hepatitis C virus RNA quantification and clinical
significance for genotypes 1 to 5. J Clin Microbiol 44: 729–737.
24. de Lamballerie X, Colson P (2006) The battle against infectious diseases in
developing countries: the inseparable twins of diagnosis and therapy. Clin
Chem 52: 1217.
25. Fiscus SA, Cheng B, Crowe SM, Demeter L, Jennings C, et al. (2006) HIV-1
viral load assays for resource-limited settings. PLoS Med 3: e417. doi:10.
1371/journal.pmed.0030417
26. Phillips AN, Pillay D, Miners AH, Bennett DE, Gilks CF, et al. (2008)
Outcomes from monitoring of patients on antiretroviral therapy in
resource-limited settings with viral load, CD4 cell count, or clinical
observation alone: a computer simulation model. Lancet 371: 1443–1451.
27. Preiser W, Drexler JF, Drosten C (2006) HIV-1 viral load assays for
resource-limited settings: clades matter. PLoS Med 3: e538; author reply
e550. doi:10.1371/journal.pmed.0030538
28. Drosten C, Panning M, Drexler JF, Hansel F, Pedroso C, et al. (2006)
Ultrasensitive monitoring of HIV-1 viral load by a low-cost real-time
reverse transcription-PCR assay with internal control for the 59long
terminal repeat domain. Clin Chem 52: 1258–1266.
29. Bukh J, Purcell RH, Miller RH (1992) Sequence analysis of the 59noncoding
region of hepatitis C virus. Proc Natl Acad Sci U S A 89: 4942–4946.
30. Anonymous (2004) Verbot der Verwendung des HPS COBAS TaqMan
HCV Test (Roche Diagnostics GmbH, 68298 Mannheim) bei der Her-
stellung von zellula
¨ren Blutkomponenten und gefrorenem Frischplasma.-
Paul-Ehrlich-Institut, editor. Bundesanzeiger Nr 166: Paul-Ehrlich-Institut.
pp. 19770.
31. Honda M, Brown EA, Lemon SM (1996) Stability of a stem-loop involving
the initiator AUG controls the efficiency of internal initiation of
translation on hepatitis C virus RNA. RNA 2: 955–968.
32. Simmonds P (2004) Genetic diversity and evolution of hepatitis C virus–15
years on. J Gen Virol 85: 3173–3188.
33. Kolykhalov AA, Feinstone SM, Rice CM (1996) Identification of a highly
PLoS Medicine | www.plosmedicine.org February 2009 | Volume 6 | Issue 2 | e10000310218
Novel Diagnostic HCV Target
conserved sequence element at the 39terminus of hepatitis C virus genome
RNA. J Virol 70: 3363–3371.
34. Tanaka T, Kato N, Cho MJ, Shimotohno K (1995) A novel sequence found at
the 39terminus of hepatitis C virus genome. Biochim Biophysic Res Comm
215: 744–749.
35. Friebe P, Bartenschlager R (2002) Genetic analysis of sequences in the 39
nontranslated region of hepatitis C virus that are important for RNA
replication. J Virol 76: 5326–5338.
36. Ito T, Tahara SM, Lai MM (1998) The 39-untranslated region of hepatitis C
virus RNA enhances translation from an internal ribosomal entry site. J
Virol 72: 8789–8796.
37. Tsuchihara K, Tanaka T, Hijikata M, Kuge S, Toyoda H, et al. (1997)
Specific interaction of polypyrimidine tract-binding protein with the
extreme 39-terminal structure of the hepatitis C virus genome, the 39X. J
Virol 71: 6720–6726.
38. Wood J, Frederickson RM, Fields S, Patel AH (2001) Hepatitis C virus 39X
region interacts with human ribosomal proteins. J Virol 75: 1348–1358.
39. Saldanha J, Heath A, Aberham C, Albrecht J, Gentili G, et al. (2005) World
Health Organization collaborative study to establish a replacement WHO
international standard for hepatitis C virus RNA nucleic acid amplification
technology assays. Vox Sang 88: 202–204.
40. McOmish F, Yap PL, Dow BC, Follett EA, Seed C, et al. (1994) Geographical
distribution of hepatitis C virus genotypes in blood donors: an interna-
tional collaborative survey. J Clin Microbiol 32: 884–892.
41. Davidson F, Simmonds P, Ferguson JC, Jarvis LM, Dow BC, et al. (1995)
Survey of major genotypes and subtypes of hepatitis C virus using RFLP of
sequences amplified from the 59non-coding region. J Gen Virol 76: 1197–
1204.
42. Laperche S, Lunel F, Izopet J, Alain S, Deny P, et al. (2005) Comparison of
hepatitis C virus NS5b and 59noncoding gene sequencing methods in a
multicenter study. J Clin Microbiol 43: 733–739.
43. Zuker M (2003) Mfold web server for nucleic acid folding and hybridization
prediction. Nucleic Acids Res 31: 3406–3415.
44. Christopherson C, Sninsky J, Kwok S (1997) The effects of internal primer-
template mismatches on RT-PCR: HIV-1 model studies. Nucleic Acids Res
25: 654–658.
45. Kwok S, Kellogg DE, McKinney N, Spasic D, Goda L, et al. (1990) Effects of
primer-template mismatches on the polymerase chain reaction: human
immunodeficiency virus type 1 model studies. Nucleic Acids Res 18: 999–
1005.
46. Peyret N, Seneviratne PA, Allawi HT, SantaLucia J Jr. (1999) Nearest-
neighbor thermodynamics and NMR of DNA sequences with internal A.A,
C.C, G.G, and T.T mismatches. Biochemistry 38: 3468–3477.
47. Holmes H, Davis C, Heath A (2008) Development of the 1st International
Reference Panel for HIV-1 RNA genotypes for use in nucleic acid-based
techniques. J Virol Methods 154: 86–91.
48. Pierre V, Drouet M, Deubel V (1994) Identification of mosquito-borne
flavivirus sequences using universal primers and reverse transcription/
polymerase chain reaction. Res Virol 145: 93–104.
49. Caliendo AM, Valsamakis A, Zhou Y, Yen-Lieberman B, Andersen J, et al.
(2006) Multilaboratory comparison of hepatitis C virus viral load assays. J
Clin Microbiol 44: 1726–1732.
50. Elbeik T, Surtihadi J, Destree M, Gorlin J, Holodniy M, et al. (2004)
Multicenter evaluation of the performance characteristics of the Bayer
VERSANT HCV RNA 3.0 assay (bDNA). J Clin Microbiol 42: 563–569.
51. Tuaillon E, Mondain AM, Ottomani L, Roudiere L, Perney P, et al. (2007)
Impact of hepatitis C virus (HCV) genotypes on quantification of HCV
RNA in serum by COBAS AmpliPrep/COBAS TaqMan HCV test, Abbott
HCV realtime assay and VERSANT HCV RNA assay. J Clin Microbiol 45:
3077–3081.
52. You S, Rice CM (2008) 39RNA elements in hepatitis C virus replication:
kissing partners and long poly(U). J Virol 82: 184–195.
53. Saito T, Owen DM, Jiang F, Marcotrigiano J, Gale M Jr. (2008) Innate
immunity induced by composition-dependent RIG-I recognition of
hepatitis C virus RNA. Nature 454: 523–527.
54. Germer JJ, Harmsen WS, Mandrekar JN, Mitchell PS, Yao JD (2005)
Evaluation of the COBAS TaqMan HCV test with automated sample
processing using the MagNA pure LC instrument. J Clin Microbiol 43: 293–
298.
55. Highbarger HC, Hu Z, Kottilil S, Metcalf JA, Polis MA, et al. (2007)
Comparison of the Abbott 7000 and Bayer 340 systems for measurement of
hepatitis C virus load. J Clin Microbiol 45: 2808–2812.
56. Sabato MF, Shiffman ML, Langley MR, Wilkinson DS, Ferreira-Gonzalez A
(2007) Comparison of performance characteristics of three real-time
reverse transcription-PCR test systems for detection and quantification
of hepatitis C virus. J Clin Microbiol 45: 2529–2536.
57. Gelderblom HC, Beld MG (2007) Hepatitis C virus RNA quantification by
the COBAS AmpliPrep/COBAS TaqMan System: averages do not tell the
whole story. J Clin Virol 39: 326–327.
58. Konnick EQ, Williams SM, Ashwood ER, Hillyard DR (2005) Evaluation of
the COBAS Hepatitis C Virus (HCV) TaqMan analyte-specific reagent assay
and comparison to the COBAS Amplicor HCV Monitor V2.0 and Versant
HCV bDNA 3.0 assays. J Clin Microbiol 43: 2133–2140.
59. Greco DB, Simao M (2007) Brazilian policy of universal access to AIDS
treatment: sustainability challenges and perspectives. Aids 21 Suppl 4: S37–
45.
60. Campiotto S, Pinho JR, Carrilho FJ, Da Silva LC, Souto FJ, et al. (2005)
Geographic distribution of hepatitis C virus genotypes in Brazil. Braz J Med
Biol Res 38: 41–49.
61. Anonymous (2003) Portaria no. 79, de 31 de janeiro de 2003.Saude Md
editor. Portarias (Brazil): Ministerio da Saude.
62. Anonymous (2002): Brazilian Ministry of Health, National Viral Hepatitis
Program. Available: http://sistemas.aids.gov.br/imprensa/Noticias.
asp?NOTCod¼45078. Accessed 8 November 2008.
63. Roth WK, Weber M, Seifried E (1999) Feasibility and efficacy of routine
PCR screening of blood donations for hepatitis C virus, hepatitis B virus,
and HIV-1 in a blood-bank setting. Lancet 353: 359–363.
64. Stramer SL, Glynn SA, Kleinman SH, Strong DM, Caglioti S, et al. (2004)
Detection of HIV-1 and HCV infections among antibody-negative blood
donors by nucleic acid-amplification testing. N Engl J Med 351: 760–768.
65. Roth WK, Weber M, Buhr S, Drosten C, Weichert W, et al. (2002) Yield of
HCV and HIV-1 NAT after screening of 3.6 million blood donations in
central Europe. Transfusion 42: 862–868.
66. Hourfar MK, Jork C, Schottstedt V, Weber-Schehl M, Brixner V, et al.
(2008) Experience of German Red Cross blood donor services with nucleic
acid testing: results of screening more than 30 million blood donations for
human immunodeficiency virus-1, hepatitis C virus, and hepatitis B virus.
Transfusion 48: 1558–1566.
67. Grant PR, Busch MP (2002) Nucleic acid amplification technology methods
used in blood donor screening. Transfus Med 12: 229–242.
68. Dia
´rio Oficial Poder Executivo (2006) Sec¸a
˜oI,Sa
˜o Paulo, 116: 161, 16
February 2006.
PLoS Medicine | www.plosmedicine.org February 2009 | Volume 6 | Issue 2 | e10000310219
Novel Diagnostic HCV Target
Editors’ Summary
Background. About 3% of the world’s population (170 million people)
harbor long-term (chronic) infections with the hepatitis C virus (HCV) and
about 3–4 million people are newly infected with this virus every year.
HCV—a leading cause of chronic hepatitis (inflammation of the liver)—is
spread through contact with the blood of an infected person. Globally,
the main routes of transmission are the use of unscreened blood for
transfusions and the reuse of inadequately sterilized medical instru-
ments, including needles. In affluent countries, where donated blood is
routinely screened for the presence of HCV, most transmission is through
needle sharing among drug users. The risk of sexual and mother-to-child
transmission of HCV is low. Although HCV infection occasionally causes
an acute (short-lived) illness characterized by tiredness and jaundice
(yellow eyes and skin), most newly infected people progress to a
symptom-free, chronic infection that can eventually cause liver cirrhosis
(scarring) and liver cancer. HCV infections can be treated with a
combination of two drugs called interferon and ribavirin, but these drugs
are expensive and are ineffective in many patients.
Why Was This Study Done? An effective way to limit the global spread
of HCV might be to introduce routine screening of the blood that is used
for transfusions in developing countries. In developed countries, HCV
screening of blood donors use expensive, commercial ‘‘RT-PCR’’ assays
to detect small amounts of HCV ribonucleic acid (RNA; HCV stores the
information it needs to replicate itself—its genome—as a sequence of
‘‘ribonucleotides’’). All the current HCV assays, which can also quantify
the amount of viral RNA in the blood (the viral load) during treatment,
detect a target sequence in the viral genome called the 59-noncoding
region (59-NCR). However, there are several different HCV ‘‘genotypes’’
(strains). These genotypes vary in their geographical distribution and,
even though the 59-NCR sequence is very similar (highly conserved) in
the common genotypes (HCV genotypes 1–6), the existing assays do not
detect all the variants equally well. This shortcoming, together with their
high cost, means that 59-NCR RT-PCR assays are not ideal for use in many
developing countries. In this study, the researchers identify an alternative
diagnostic target sequence in the HCV genome—the 39-X-tail element—
and ask whether this sequence can be used to develop a new generation
of tests for HCV infection that might be more appropriate for use in
developing countries.
What Did the Researchers Do and Find? The researchers determined
the RNA sequence of the 39-X-tail element in reference samples of the
major HCV genotypes and showed that this region of the HCV genome is
as highly conserved as the 59-NCR. They then developed a prototype X-
tail RT-PCR assay and tested its ability to detect small amounts of HCV
and to measure viral load in genotype reference samples and in a large
panel of HCV-infected blood samples collected in Germany, the UK,
Brazil, South Africa, and Singapore. The new assay detected low levels of
HCV RNA in all of the genotype reference samples and was also able to
quantify high RNA concentrations. The viral load estimates it provided
for the clinical samples agreed well with those obtained using a
commercial assay irrespective of the sample’s HCV genotype. Finally, the
X-tail RT-PCR assay gave similar results to a standard assay at a fraction of
the cost when used to measure viral loads in a Brazilian laboratory in an
independent group of 127 patient samples collected in Brazil.
What Do These Findings Mean? These findings suggest that the HCV
39-X-tail element could provide an alternative target for screening blood
samples for HCV infection and for monitoring viral loads during
treatment, irrespective of HCV genotype. In addition, they suggest that
X-tail RT-PCR assays may be stable and robust enough for use in
laboratories in emerging countries. Overall, these findings should
stimulate the development of a new generation of clinical HCV assays
that, because the protocol used in the X-tail assay is freely available,
could improve blood safety in developing countries by providing a
cheap and effective alternative to existing proprietary HCV assays.
Additional Information. Please access these Web sites via the online
version of this summary at http://dx.doi.org/10.1371/journal.pmed.
1000031.
The World Health Organization has a fact sheet about hepatitis C (in
English and French)
The US Centers for Disease Control and Prevention provides
information on hepatitis C for the public and for health professionals
(information is also available in Spanish)
The US National Institute of Diabetes and Digestive and Kidney
Diseases provides basic information on hepatitis C (in English and
Spanish)
The MedlinePlus Encyclopedia has a page on hepatitis C; MedlinePlus
also provides links to further information on hepatitis C (in English and
Spanish)
PLoS Medicine | www.plosmedicine.org February 2009 | Volume 6 | Issue 2 | e10000310220
Novel Diagnostic HCV Target

Supplementary resources (18)

Nucleotide Sequence
June 2008
C. Drosten · J.F. Drexler
Nucleotide Sequence
June 2008
C. Drosten · J.F. Drexler
Nucleotide Sequence
June 2008
C. Drosten · J.F. Drexler
Nucleotide Sequence
June 2008
C. Drosten · J.F. Drexler
Nucleotide Sequence
June 2008
C. Drosten · J.F. Drexler
Nucleotide Sequence
June 2008
C. Drosten · J.F. Drexler
Nucleotide Sequence
June 2008
C. Drosten · J.F. Drexler
Nucleotide Sequence
June 2008
C. Drosten · J.F. Drexler
Nucleotide Sequence
June 2008
C. Drosten · J.F. Drexler
Nucleotide Sequence
June 2008
C. Drosten · J.F. Drexler
... RNA was purified from approximately 20 mg of faecal material suspended in 500 μL RNAlater stabilising solution using the MagNA Pure 96 system (Roche) with elution volumes set at 100 μL. We used a real-time reverse transcription-PCR assay designed to detect several alpha-and beta-CoVs and genetically related bat CoVs using the SSIII RT-PCR kit (Life Technologies) and a cycling protocol in a LightCycler 480 (Roche) as described previously (Corman et al., 2015;Drexler et al., 2009;Pfefferle et al., 2009; see Appendix S1; Table S2). Bats ...
Article
Understanding the immunogenetic basis of coronavirus (CoV) susceptibility in major pathogen reservoirs, such as bats, is central to inferring their zoonotic potential. Members of the cryptic Hipposideros bat species complex differ in CoV susceptibility, but the underlying mechanisms remain unclear. The genes of the major histocompatibility complex (MHC) are the best understood genetic basis of pathogen resistance, and differences in MHC diversity are one possible reason for asymmetrical infection patterns among closely related species. Here, we aimed to link asymmetries in observed CoV (CoV-229E, CoV-2B and CoV-2Bbasal) susceptibility to immunogenetic differences amongst four Hipposideros bat species. From the 2072 bats assigned to their respective species using the mtDNA cytochrome b gene, members of the most numerous and ubiquitous species, Hipposideros caffer D, were most infected with CoV-229E and SARS-related CoV-2B. Using a subset of 569 bats, we determined that much of the existent allelic and functional (i.e. supertype) MHC DRB class II diversity originated from common ancestry. One MHC supertype shared amongst all species, ST12, was consistently linked to susceptibility with CoV-229E, which is closely related to the common cold agent HCoV-229E, and infected bats and those carrying ST12 had a lower body condition. The same MHC supertype was connected to resistance to CoV-2B, and bats with ST12 were less likely be co-infected with CoV-229E and CoV-2B. Our work suggests a role of immunogenetics in determining CoV susceptibility in bats. We advocate for the preservation of functional genetic and species diversity in reservoirs as a means of mitigating the risk of disease spillover.
... A highly sensitive reverse transcription-quantitative polymerase chain reaction assay targeting the HCV X-tail domain was used for confirmation and for viral load estimations, relying on an armoured RNA standard as a quantitative control, as described previously (Drexler et al. 2009). ...
Article
Full-text available
The hepatitis C virus genotype 2 (HCV2) is endemic in Western and Central Africa. The HCV2 evolutionary origins remain uncertain due to the paucity of available genomes from African settings. In this study, we investigated the molecular epidemiology of HCV infections in rural Guinea, Western Africa, during 2004 and 2014. Broadly reactive nested RT-PCR-based screening of sera from 1,571 asymptomatic adults resulted in the detection of 25 (1.5%, 95% CI 0.9-2.3) positive samples, with a median viral load of 2.54E+05 IU/mL (IQR 6.72E+05). HCV-infected persons had a median age of 47 years and 62.5% were male and 37.5% female. The full polyprotein-encoding genes were retrieved by a combination of high throughput and Sanger sequencing from 17 samples showing sufficiently high viral loads. Phylogenetic analysis and sequence distances ≥13% averaged over the polyprotein genes compared to other HCV2 subtypes revealed 9 previously unknown HCV2 subtypes. The time to the most recent common ancestor of the Guinean HCV2 strains inferred in a Bayesian framework was 493 years (95% HPD 453-532). Most of the Guinean strains clustered poorly by location both on the level of sampling sites within Guinea and the level of countries in the phylogenetic reconstructions. Ancestral state reconstruction provided decisive support (Bayes factor >100) for an origin of HCV2 in Western Africa. Phylogeographic reconstructions in a Bayesian framework pointed to a radial diffusion of HCV2 from Western African regions encompassing today’s countries like Ghana, Guinea Bissau, or Burkina Faso, to Central and Northern African regions that took place from the 16th century onwards. The spread of HCV2 coincided in time and space with the main historic slave trade and commerce routes, supported by Bayesian tip-association significance testing (p = 0.01). Our study confirms the evolutionary origins of HCV2 in Western Africa and provides a potential link between historic human movements and HCV2 dispersion.
... As sequence information about SARS-CoV-2 has recently become available, quantitative nucleic acid testing has become the gold standard for clinical decisions regarding the use of antiviral therapy [20][21][22]. The most recently developed nucleic acid testing method is quantitative real-time PCR (RT-qPCR) [23,24], although RT-qPCR assay design optimization can be a complicated process. ...
Article
Full-text available
In December 2019, a new coronavirus disease (COVID-19) outbreak occurred in Wuhan, China. Severe acute respiratory syndrome-coronavirus-2 (SARS-CoV-2), which is the seventh coronavirus known to infect humans, is highly contagious and has rapidly expanded worldwide since its discovery. Quantitative nucleic acid testing has become the gold standard for diagnosis and guiding clinical decisions regarding the use of antiviral therapy. However, the RT-qPCR assays targeting SARS-CoV-2 have a number of challenges, especially in terms of primer design. Primers are the pivotal components of a RT-qPCR assay. Once virus mutation and recombination occur, it is difficult to effectively diagnose viral infection by existing RT-qPCR primers. Some primers and probes have also been made available on the WHO website for reference. However, no previous review has systematically compared the previously reported primers and probes and described how to design new primers in the event of a new coronavirus infection. This review focuses on how primers and probes can be designed methodically and rationally, and how the sensitivity and specificity of the detection process can be improved. This brief review will be useful for the accurate diagnosis and timely treatment of the new coronavirus pneumonia.
... Viral RNA from serum and cell supernatant was quantified using a strain-specific real-time RT-PCR with the following primers FSHV-rtF (5′-TGTTGGTGATTGTGCATATATTGG), FSHV-rtR (5′-GGCGGACAACCATGTTTAATACT) and the probe FSHV-rtP (5′-ATCTAGCCAGTAGGTTATCTGCCACGCAGC), labelled with fluorescein amidite (FAM) at the 5′-end and a dark quencher at the 3′-end. The assay was controlled by photometrically quantified in vitro-transcribed RNA controls generated from synthesized DNA fragments (IDT) containing a T7-RNA polymerase promoter region as described previously (Drexler et al., 2009). Thermocycling involved reverse transcription at 55°C for 20 min followed by 94°C for 3 min and then 45 cycles of 94°C for 15 s and 58°C for 30 s. RT-PCR was done using the OneStep SuperScript III RT-PCR kit (Thermo Fisher) with 5 µl RNA input and reaction components according to the manufacturer's instructions. ...
Article
Full-text available
An orthobunyavirus termed Fort Sherman virus (FSV) was isolated in 1985 from a febrile US soldier in Panama, yet potential animal reservoirs remained unknown. We investigated sera from 192 clinically healthy peri‐domestic animals sampled in northeastern Brazil during 2014‐2018 by broadly reactive RT‐PCR for orthobunyavirus RNA, including 50 cattle, 57 sheep, 35 goats, and 50 horses. One horse sampled in 2018 was positive (0.5%; 95% CI, 0.01‐3.2) at 6.2x103 viral RNA copies/mL. Genomic comparisons following virus isolation in Vero cells and deep sequencing revealed high identity of translated amino acid sequences between the new orthobunyavirus and the Panamanian FSV prototype (genes: L, 98.8%; M, 83.5%; S, 100%), suggesting these viruses are conspecific. Database comparisons revealed even higher genomic identity between the Brazilian FSV and Argentinian mosquito‐ and horse‐derived viruses sampled in 1965, 1982 and 2013 with only 1.1% maximum translated amino acid distances. The Panamanian FSV strain was an M gene reassortant relative to all Southern American FSV strains, clustering phylogenetically with Cache Valley virus (CVV). Mean dN/dS ratios among FSV genes ranged from 0.03‐0.07, compatible with strong purifying selection. FSV‐specific neutralizing antibodies occurred at relatively high end‐point titers in the range of 1:300 in 22.0% of horses (11 out of 50 animals), 8.0% of cattle (4/50 animals), 7.0% of sheep (4/57 animals) and 2.9% of goats (1/35 animals). High specificity of serologic testing was suggested by significantly higher overall FSV‐specific compared to CVV‐ and Bunyamwera virus‐specific end‐point titers (p=0.009), corroborating a broad vertebrate host range within peri‐domestic animals. Growth kinetics using mosquito‐, midge‐ and sandfly‐derived cell lines suggested Aedes mosquitos as potential vectors. Our findings highlight the occurrence of FSV across a geographic range exceeding 7000 kilometers, surprising genomic conservation across a timespan exceeding 50 years, M gene‐based reassortment events, and the existence of multiple animal hosts of FSV.
Article
Currently, PCR is the gold standard for the detection of hepatitis C virus (HCV). However, the PCR technique is complicated and time-consuming, which prevents its application and, clinical point-of-care testing (POCT). Herein, we report a POCT method with simplicity, high sensitivity and specificity, which consists of a catalytic hairpin assembly (CHA) signal amplification system coupled with a lateral flow immunochromatographic (LFIA) test strip for the detection of HCV. Two ingeniously designed hairpin probes were hybridized to form the H1–H2 duplex in the presence of the target DNA. The catalytic hairpin assembly which was characterized of isothermal and enzyme-free, was accomplished within 40 min and the reaction was then applied to a LFIA test strip. Only the H1–H2 duplex labeled with both digoxin and biotin could be captured by the test strip, and the fluorescence value was determined. In addition, we evaluated the application potential for the detection of clinical samples. The reported method demonstrated high sensitivity with a detectable minimum concentration at 1 fM and showed a good linear range from 10 nM to 10pM, and high specificity for various mismatched sequences. The results demonstrated that clinically positive samples could be successfully detected. In conclusion, the reported method is simple, rapid, and free of large-scale equipment. POCT is expected to be useful for HCV detection in clinic.
Article
Full-text available
In humans, co-infection of hepatitis B and C virus (HBV, HCV) is common, and aggravates disease outcome. Infection-mediated disease aggravation is poorly understood, partly due to lack of suitable animal models. Carnivores are understudied for hepatitis virus homologues. We investigated Mexican carnivores (ringtails, Bassariscus astutus) for HBV and HCV homologues. Three out of eight animals were infected with a divergent HBV termed ringtail HBV (RtHBV) at high viral loads of 5×10⁹-1.4×10¹⁰ copies/mL serum. Two of the RtHBV-infected animals were co-infected with a divergent hepacivirus termed ringtail hepacivirus (RtHV) at 4×10⁶-7.5×10⁷ copies/mL in strain-specific qRT-PCR assays. Immunofluorescence assays relying on HBV core and RtHV NS3/4a proteins indicated that none of the animals had detectable hepadnavirus core-specific antibodies, whereas one RtHV-infected animal had concomitant RtHV-specific antibodies at 1:800 end-point titer. RtHBV and RtHV complete genomes showed typical HBV and HCV structure and length. All RtHBV genomes were identical, whereas RtHV genomes showed 4 amino acid substitutions located predominantly in the E1/E2-encoding genomic regions. Both RtHBV (>28% genomic nucleotide sequence distance) and RtHV (>30% partial NS3/NS5B amino acid sequence distance) formed new species within their virus families. Evolutionary analyses showed that RtHBV grouped with HBV homologues from different Laurasiatherian hosts (carnivores, bats, and ungulates), whereas RtHV grouped predominantly with rodent-borne viruses. Ancestral state reconstructions showed that RtHV, but not RtHBV, likely emerged via non-recent host switch involving rodent-borne hepacivirus ancestors. Conserved hepatitis virus infection patterns in naturally infected ringtails indicate that carnivores may be promising animal models to understand HBV/HCV co-infection. This article is protected by copyright. All rights reserved
Article
Full-text available
Rapid antigen-detecting tests (Ag-RDTs) can complement molecular diagnostics for COVID-19. The recommended temperature for storage of SARS-CoV-2 Ag-RDTs ranges between 2-30 °C. In the global South, mean temperatures can exceed 30 °C. In the global North, Ag-RDTs are often used in external testing facilities at low ambient temperatures. We assessed analytical sensitivity and specificity of eleven commercially-available SARS-CoV-2 Ag-RDTs using different storage and operational temperatures, including short- or long-term storage and operation at recommended temperatures or at either 2-4 °C or at 37 °C. The limits of detection of SARS-CoV-2 Ag-RDTs under recommended conditions ranged from 1.0 × 10⁶-5.5 × 10⁷ genome copies/ml of infectious SARS-CoV-2 cell culture supernatant. Despite long-term storage at recommended conditions, 10 minutes pre-incubation of Ag-RDTs and testing at 37 °C resulted in about ten-fold reduced sensitivity for five out of 11 SARS-CoV-2 Ag-RDTs, including both Ag-RDTs currently listed for emergency use by the World Health Organization. After 3 weeks of storage at 37 °C, eight of the 11 SARS-CoV-2 Ag-RDTs exhibited about ten-fold reduced sensitivity. Specificity of SARS-CoV-2 Ag-RDTs using cell culture supernatant from common respiratory viruses was not affected by storage and testing at 37 °C, whereas false-positive results occurred at outside temperatures of 2-4 °C for two out of six tested Ag-RDTs. In summary, elevated temperatures impair sensitivity, whereas low temperatures impair specificity of SARS-CoV-2 Ag-RDTs. Consequences may include false-negative test results at clinically relevant virus concentrations compatible with inter-individual transmission and false-positive results entailing unwarranted quarantine assignments. Storage and operation of SARS-CoV-2 Ag-RDTs at recommended conditions is essential for successful usage during the pandemic.
Article
Full-text available
Information on severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) spread in Africa is limited by insufficient diagnostic capacity. Here, we assessed the coronavirus disease (COVID-19)-related diagnostic workload during the onset of the pandemic in the central laboratory of Benin, Western Africa; characterized 12 SARS-CoV-2 genomes from returning travelers; and validated the Da An RT-PCR-based diagnostic kit that is widely used across Africa. We found a 15-fold increase in the monthly laboratory workload due to COVID-19, dealt with at the cost of routine activities. Genomic surveillance showed near-simultaneous introduction of distinct SARS-CoV-2 lineages termed A.4 and B.1, including the D614G spike protein variant potentially associated with higher transmissibility from travelers from six different European and African countries during March-April 2020. We decoded the target regions within the ORF1ab and N genes of the Da An dual-target kit by MinION-based amplicon sequencing. Despite relatively high similarity between SARS-CoV-2 and endemic human coronaviruses (HCoVs) within the ORF1ab target domain, no cross-detection of high-titered cell culture supernatants of HCoVs was observed, suggesting high analytical specificity. The Da An kit was highly sensitive, detecting 3.2 to 9.0 copies of target-specific in vitro transcripts/reaction. Although discrepant test results were observed in low-titered clinical samples, clinical sensitivity of the Da An kit was at least comparable to that of commercial kits from affluent settings. In sum, virologic diagnostics are achievable in a resource-limited setting, but unprecedented pressure resulting from COVID-19-related diagnostics requires rapid and sustainable support of national and supranational stakeholders addressing limited laboratory capacity. IMPORTANCE Months after the start of the COVID-19 pandemic, case numbers from Africa are surprisingly low, potentially because the number of SARS-CoV-2 tests performed in Africa is lower than in other regions. Here, we show an overload of COVID-19-related diagnostics in the central laboratory of Benin, Western Africa, with a stagnating average number of positive samples irrespective of daily sample counts. SARS-CoV-2 genomic surveillance confirmed a high genomic diversity in Benin introduced by travelers returning from Europe and other African countries, including early circulation of the D614G spike mutation associated with potentially higher transmissibility. We validated a widely used RT-PCR kit donated by the Chinese Jack Ma Foundation and confirmed high analytical specificity and clinical sensitivity equivalent to tests used in affluent settings. Our assessment shows that although achievable in an African setting, the burden from COVID-19-related diagnostics on national reference laboratories is very high.
Preprint
Full-text available
Rapid antigen-detecting tests (Ag-RDTs) can complement molecular diagnostics for COVID-19. The recommended temperature for storage of SARS-CoV-2 Ag-RDTs ranges between 5-30°C. In many countries that would benefit from SARS-CoV-2 Ag-RDTs, mean temperatures exceed 30°C. We assessed analytical sensitivity and specificity of eleven commercially available SARS-CoV-2 Ag-RDTs using different storage and operational temperatures, including (i) long-term storage and testing at recommended conditions, (ii) recommended storage conditions followed by 10 minutes exposure to 37°C and testing at 37°C and (iii) 3 weeks storage followed by testing at 37°C. The limits of detection of SARS-CoV-2 Ag-RDTs under recommended conditions ranged from 8.2x105-7.9x107 genome copies/ml of infectious SARS-CoV-2 cell culture supernatant. Despite long-term storage at recommended conditions, 10 minutes pre-incubation of Ag-RDTs and testing at 37°C resulted in about ten-fold reduced sensitivity for 46% of SARS-CoV-2 Ag-RDTs, including both Ag-RDTs currently listed for emergency use by the World Health Organization. After 3 weeks of storage at 37°C, 73% of SARS-CoV-2 Ag-RDTs exhibited about ten-fold reduced sensitivity. Specificity of SARS-CoV-2 Ag-RDTs using cell culture-derived human coronaviruses HCoV-229E and HCoV-OC43 was not affected by storage and testing at 37°C. In summary, short- and long-term exposure to elevated temperatures likely impairs sensitivity of several SARS-CoV-2 Ag-RDTs that may translate to false-negative test results at clinically relevant virus concentrations compatible with inter-individual transmission. Ensuring appropriate transport and storage conditions, and development of tests that are more robust across temperature fluctuations will be important for accurate use of SARS-CoV-2 Ag-RDTs in tropical settings.
Article
Full-text available
The genealogy of the hepatitis C virus (HCV) and the genus Hepacivirus remains elusive despite numerous recently discovered animal hepaciviruses (HVs). Viruses from evolutionarily ancient mammals might elucidate the HV macro-evolutionary patterns. Here, we investigated sixty-seven two-toed and nine three-toed sloths from Costa Rica for HVs using molecular and serological tools. A novel sloth HV was detected by reverse transcription polymerase chain reaction (RT-PCR) in three-toed sloths (2/9, 22.2%; 95% confidence interval (CI), 5.3-55.7). Genomic characterization revealed typical HV features including overall polyprotein gene structure, a type 4 internal ribosomal entry site in the viral 5 0-genome terminus, an A-U-rich region and X-tail structure in the viral 3 0-genome terminus. Different from other animal HVs, HV seropositivity in two-toed sloths was low at 4.5 per cent (3/67; CI, 1.0-12.9), whereas the RT-PCR-positive three-toed sloths were seronegative. Limited cross-reactivity of the serological assay implied exposure of seropositive two-toed sloths to HVs of unknown origin and recent infections in RT-PCR-positive animals preceding seroconversion. Recent infections were consistent with only 9 nucleotide exchanges between the two sloth HVs, located predominantly within the E1/E2 encoding regions. Translated sequence distances of NS3 and NS5 proteins and host comparisons suggested that the sloth HV represents a novel HV species. Event-and sequence distance-based reconciliations of phylogenies of HVs and of their hosts revealed complex macro-evolutionary patterns, including both long-term evolutionary associations and host switches, most strikingly from rodents into sloths. Ancestral state reconstructions corroborated rodents as predominant sources of HV host switches during the ge-nealogy of extant HVs. Sequence distance comparisons, partial conservation of critical amino acid residues associated with HV entry and selection pressure signatures of host genes encoding entry and antiviral protein orthologs were consistent V C The Author(s) with HV host switches between genetically divergent mammals, including the projected host switch from rodents into sloths. Structural comparison of HCV and sloth HV E2 proteins suggested conserved modes of hepaciviral entry. Our data corroborate complex macro-evolutionary patterns shaping the genus Hepacivirus, highlight that host switches are possible across highly diverse host taxa, and elucidate a prominent role of rodent hosts during the Hepacivirus genealogy.
Article
Full-text available
Hepatitis C virus (HCV) RNA detection and quantification are the key diagnostic tools for the management of hepatitis C. Commercially available HCV RNA assays are calibrated to the HCV genotype 1 (gt1)-based WHO standard. Significant differences between assays have been reported. However, it is unknown which assay matches the WHO standard best, and little is known about the sensitivity and linear quantification of the assays for non-gt1 specimens. Two real-time reverse transcriptase PCR-based assays (RealTime HCV and Cobas Ampliprep/Cobas TaqMan HCV [CAP/CTM]) and one signal amplification-based assay (the Versant HCV RNA, version 3.0, branched DNA [bDNA] assay) were compared for their abilities to quantify HCV RNA in clinical specimens (n = 65) harboring HCV isolates of gt1 to g5. The mean differences in the amounts detected by RealTime HCV in comparison to those detected by the bDNA assay and CAP/CTM were -0.02 and 0.72 log(10) IU/ml HCV RNA, respectively, for gt1; -0.22 and 0.03 log(10) IU/ml HCV RNA, respectively, for gt2; -0.27 and -0.22 log(10) IU/ml HCV RNA, respectively, for gt3; -0.19 and -1.27 log(10) IU/ml HCV RNA, respectively, for gt4; and -0.03 and 0.09 log(10) IU/ml HCV RNA, respectively, for gt5. The lower limits of detection for RealTime HCV and CAP/CTM were 16.8 and 10.3 IU/ml, respectively, for the WHO standard and in the range of 4.7 to 9.0 and 3.4 to 44.4 IU/ml, respectively, for clinical specimens harboring gt1 to gt6. Direct comparison of the two assays with samples of the WHO standard (code 96/798) with high titers yielded slightly smaller amounts by RealTime HCV (-0.2 log(10) at 1,500 IU/ml and -0.3 log(10) at 25,000 IU/ml) and larger amounts by CAP/CTM (0.3 log(10) at 1,500 IU/ml and 0.2 log(10) at 25,000 IU/ml). Finally, all three tests were linear between 4.0 x 10(3) and 1.0 x 10(6) IU/ml (correlation coefficient, >or=0.99). In conclusion, the real-time PCR based assays sensitively detected all genotypes and showed comparable linearities for the quantification of HCV RNA, with the exception of gt1 and gt4. The previously reported differences in the absolute quantification of samples harboring gt1 were confirmed and may be explained by different calibrations to the WHO standard.
Article
Full-text available
We have determined the nucleotide sequence of the 5' noncoding (NC) region of the hepatitis C virus (HCV) genome in 44 isolates from around the world. We have identified several HCV isolates with significantly greater sequence heterogeneity than reported previously within the 5' NC region. The most distantly related isolates were only 90.1% identical. Nucleotide insertions were seen in three isolates. Analysis of the nucleotide sequence from 44 HCV isolates in this study combined with that of 37 isolates reported in the literature reveals that the 5' NC region of HCV consists of highly conserved domains interspersed with variable domains. The consensus sequence was identical to the prototype HCV sequence. Nucleotide variations were found in 45 (16%) of the 282 nucleotide positions analyzed and were primarily located in three domains of significant heterogeneity (positions -239 to -222, -167 to -118, and -100 to -72). Conversely, there were three highly conserved domains consisting of 18, 22, and 63 completely invariant nucleotides (positions -263 to -246, -199 to -178, and -65 to -3, respectively). Two nucleotide domains within the 5' NC region, conserved among all HCV isolates studied to date, shared statistically significant similarity with pestivirus 5' NC sequences, providing further evidence for a close evolutionary relationship between these two groups of viruses. Additional analysis revealed the presence of short open reading frames in all HCV isolates. Our sequence analysis of the 5' NC region of the HCV genome provides additional information about conserved elements within this region and suggests a possible functional role for the region in viral replication or gene expression. These data also have implications for selection of optimal primer sequences for the detection of HCV RNA by the PCR assay.
Article
Full-text available
We investigated the effects of various primer-template mismatches on DNA amplification of an HIV-1 gag region by the polymerase chain reaction (PCR). Single internal mismatches had no significant effect on PCR product yield while those at the 3′-terminal base had varied effects. A:G, G:A, and C:C mismatches reduced overall PCR product yield about 100-fold, A:A mismatches about 20-fold. All other 3′-terminal mismatches were efflcientiy amplified, although the G:G mismatches appeared to be more sensitive to sequence context and dNTP concentrations than other mismatches. It shouid be noted that mismatches of T with either G, C, or T had a minimal effect on PCR product yield. Double mismatches within the last four bases of a primer-template duplex where one of the mismatches is at the 3′ terminal nucleotide, in general, reduced PCR product yield dramatically. The presence of a mismatched T at the 3′-terminus, however, ailowed significant amplification even when coupled with an adjacent mismatch. Furthermore, even two mismatched Ta at the 3′-terminus allowed efficient ampiification.
Article
Full-text available
A method is described for identifying different genotypes of hepatitis C virus (HCV) by restriction endonuclease cleavage of sequences amplified by PCR from the 5' non-coding region. Using the enzymes HaeIII-RsaI and HinfI-MvaI, followed by cleavage with BstU1 or ScrFI, it was possible to identify and distinguish HCV genotypes 1a, 1b, 2a, 2b, 3a, 3b, 4, 5 and 6. The method was used to investigate the prevalence of these genotypes in 723 blood donors in 15 countries, the largest survey to date, and one which covered a wide range of geographical regions (Europe, America, Africa and Asia). These results, combined with a review of the existing literature, indicate the existence of several distinct regional patterns of HCV genotype distribution, and provide the framework for future detailed epidemiological investigations of HCV transmission.
Article
We previously identified a highly conserved 98-nucleotide (nt) sequence, the 3*X, as the extreme 3*-terminal structure of the hepatitis C virus (HCV) genome (T. Tanaka, N. Kato, M.-J. Cho, and K. Shimotohno, Biochem. Biophys. Res. Commun. 215:744-749, 1995). Since the 3* end of positive-strand viral RNA is the initiation site of RNA replication, the 3*X should contribute to HCV negative-strand RNA synthesis. Cellular factors may also be involved in this replication mechanism, since several cellular proteins have been shown to interact with the 3*-end regions of other viral genomes. In this study, we found that both 38- and 57-kDa proteins in the human hepatocyte line PH5CH bound specifically to the 3*-end structure of HCV positive-strand RNA by a UV-induced cross-linking assay. The 57-kDa protein (p57), which had higher affinities to RNA probes, recognized a 26-nt sequence including the 5*-terminal 19 nt of the 3*X and 7 flanking nt, designated the transitional region. This sequence contains pyrimidine-rich motifs and shows similarity to the consensus binding sequence of the polypyrimidine tract-binding protein (PTB), which has been implicated in alternative pre-mRNA splicing and cap-independent translation. We found that this 3*X-binding p57 is identical to PTB. The 3*X-binding p57 was immunoprecipitated by anti-PTB antibody, and recombinant PTB bound to the 3*X RNA. In addition, p57 bound solely to the 3*-end region of positive-strand RNA, not to this region of negative-strand RNA. We suggest that 3*X-PTB interaction is involved in the specific initiation of HCV genome replication.
Article
Hepatitis C is a global health problem caused by infection with the hepatitis C virus. Although representative prevalence data are not a available from many countries, available data indicate that approximately 3% of the world's population is infected with HCV, It is estimated that as many as 170 million persons worldwide may be infected with HCV, In many countries, the exact magnitude of the problem and the relative contribution of the various routes of transmission have not been defined with population-based studies. Wherever possible such studies should be performed to enable countries to estimate the burden of hepatitis C disease, to prioritize their preventative measures and to make the most appropriate use of available resources. To assess hepatitis C on a global scale the World Health Organization (WHO) organized a consultation of international experts, in order to review the public health aspects related to hepatitis C infection and to make recommendations for its prevention and control.
Article
Twenty-eight laboratories from 16 countries participated in a collaborative study to evaluate an HIV-1 RNA Genotype Reference Panel for use with nucleic acid-based tests (NAT). The Reference Panel consisted of 11 coded samples representing different HIV-1 genotypes (subtypes A-D, AE, F, G, AA-GH, groups N and O) as well as a negative diluent control. Each laboratory assayed the eleven panel members concurrently with the 1st International Standard for HIV-1 RNA (NIBSC Code 97/656) on at least three separate occasions and the data collated and analysed at NIBSC. Twenty-nine sets of data from NAT were received, 19 from quantitative and 10 from qualitative assays, with six different commercial assays and five "in-house" assays represented. The results showed that viruses from subtypes A-D and recombinant virus AE [CRF01_AE] were detected consistently, but that some assays had difficulty with the detection and quantification of viruses from subtypes F and G, a mixed recombinant virus AA-GH and a representative of group N. Furthermore, most assays failed to detect the group O representative. The study illustrated the limitations of some molecular assays particularly in detection of certain non-B genotypes which are important viruses in the global AIDS pandemic and illustrated the value of a well-characterised genotype panel. The panel has been established by the World Health Organisation's Expert Committee on Biological Standardisation as the 1st International Reference Panel HIV-1 RNA Genotypes (code 01/466).
Article
Diagnosis of hepatitis C virus (HCV) infection and its therapy is based on qualitative and quantitative measurement of HCV RNA. A new assay that employs automated specimen extraction and real-time RT-PCR (COBAS Ampliprep/COBAS TaqMan, "CAP/CTM", Roche Diagnostics, Pleasanton, USA) was designed for linear quantification and highly sensitive detection of HCV RNA. The performance characteristics of CAP/CTM were compared to standard RT-PCR-based COBAS Amplicor Monitor 2.0 (CAM) assay in a multicenter study. The limit of detection of CAP/CTM was 7.4 IU/ml (95% CI 6.2-10.6) and clinical specificity was 99%. The linear range of HCV RNA quantification by CAP/CTM was between 28 and 1.4 x 10(7) IU/ml, with a correlation coefficient between expected and observed results of >0.99. A fivefold dilution of serum- or plasma-samples showed a linear correlation of HCV RNA levels in undiluted and diluted samples. Analyses of the mean intra- and inter-assay imprecision within the linear range of quantification showed a coefficient of variation of 3% and 3%, respectively. HCV genotypes 1a/b, 2b, 3a, 4, 5 and 6 were equally quantified by the CAP/CTM and CAM assay with mean deviations ranging from -0.29log(10) to 0.32log(10) IU/ml. HCV RNA quantification by CAP/CTM and CAM was highly concordant (correlation coefficient of 0.96). The CAP/CTM assay is a reliable and robust assay for highly sensitive detection and quantification of HCV RNA within a broad linear range.
Article
The 3' end region of positive-strand RNA-virus genomes is implicated in the initiation of genomic replication. We analyzed the extreme 3' end of the hepatitis C virus (HCV) genome by primer extension of the 5' end region of the antigenomic strand RNA found in infected liver. We discovered a novel sequence present downstream of the poly(U) stretch that was previously considered to be the 3' end structure of the HCV genome. The novel sequence was 98 nucleotides long and had no significant homology with any known sequences, viral or nonviral. The discovery of this novel tail on the HCV genome should contribute to the study of HCV replication.
Article
A reverse transcription/polymerase chain reaction (RT/PCR) protocol for the rapid detection and identification of flaviviruses was developed using a set of universal oligonucleotide primers. These primers correspond to sequences in the 3' non-coding region and in the NS5 gene which are highly conserved among the mosquito-borne flaviviruses. The sequences of the resulting amplified products were analysed for dengue 1, dengue 2, dengue 3, dengue 4, Japanese encephalitis, West Nile, yellow fever and Zika viruses, and compared with the published sequences of other flaviviruses. The 291-297 nucleotides corresponding to the C-terminus of NS5 gene showed 56 to 76% similarity, whereas the 3' non-coding region (190 to 421 nucleotides) showed only 20 to 36% similarity. Genetic classification of the Zika virus supported its traditional serological grouping. Recombinant plasmids containing the flavivirus sequences were used in a nucleic acid hybridization test to identify the RT/PCR products derived from viral RNA extracted from experimentally infected mosquitoes. The plasmids were dotted on a strip of nitrocellulose membrane and incubated with the RT/PCR product labelled with digoxigenin during the PCR step. This is a valuable method for the rapid and specific identification of mosquito-borne flaviviruses in biological specimens and for subsequent sequence analysis.