Sie sind auf Seite 1von 19

Ultrasound Obstet Gynecol 2013; 42: 1533

Published online in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/uog.12513


Non-invasive prenatal testing for aneuploidy: current status
and future prospects
P. BENN*, H. CUCKLE and E. PERGAMENT
*Department of Genetics and Developmental Biology, University of Connecticut Health Center, Farmington, CT, USA; Department of
Obstetrics and Gynecology, Columbia University Medical Center, New York, NY, USA; Northwestern Reproductive Genetics, Chicago,
IL, USA
KEYWORDS: amniocentesis; aneuploidy; chorionic villus sampling; Down syndrome; fetal DNA; maternal plasma; screening;
sequencing; trisomy
ABSTRACT
Non-invasive prenatal testing (NIPT) for aneuploidy
using cell-free DNA in maternal plasma is revolutionizing
prenatal screening and diagnosis. We review NIPT in the
context of established screening and invasive technologies,
the range of cytogenetic abnormalities detectable, cost,
counseling and ethical issues. Current NIPT approaches
involve whole-genome sequencing, targeted sequencing
and assessment of single nucleotide polymorphism (SNP)
differences between mother and fetus. Clinical trials
have demonstrated the efcacy of NIPT for Down and
Edwards syndromes, and possibly Patau syndrome, in
high-risk women. Universal NIPT is not cost-effective,
but using NIPT contingently in women found at moderate
or high risk by conventional screening is cost-effective.
Positive NIPT results must be conrmed using invasive
techniques. Established screening, fetal ultrasound and
invasive procedures with microarray testing allow the
detection of a broad range of additional abnormalities
not yet detectable by NIPT. NIPT approaches that
take advantage of SNP information potentially allow the
identication of parent of origin for imbalances, triploidy,
uniparental disomy and consanguinity, and separate
evaluation of dizygotic twins. Fetal fraction enrichment,
improved sequencing and selected analysis of the most
informative sequences should result in tests for additional
chromosomal abnormalities. Providing adequate prenatal
counseling poses a substantial challenge given the broad
range of prenatal testing options nowavailable. Copyright
2013 ISUOG. Published by John Wiley & Sons Ltd.
INTRODUCTION
The past quarter century has been witness to a series of
remarkable advances in the screening of pregnancies for
aneuploidy, particularly in the identication of Down
syndrome (trisomy 21). During the 1970s and early
Correspondence to: Dr P. Benn, Department of Genetics and Developmental Biology, University of Connecticut Health Center, Farmington,
CT, USA (e-mail: benn@nso1.uchc.edu)
Accepted: 14 May 2013
1980s, advanced maternal age, dened in most localities
as over 35years, was the only means by which the
general population was assessed as to risk of a fetal
chromosomal abnormality. Fewer than one third of
Down syndrome pregnancies were diagnosed prenatally
and of those undergoing invasive prenatal diagnosis only
about 2% had fetal karyotype abnormalities
1
, a gure
comparable to the 0.51% chance of procedure-related
fetal loss associated with amniocentesis or chorionic villus
sampling (CVS)
2
. In the late 1980s and early 1990s, the
introduction of second-trimester maternal serummarkers,
in the form of double, triple and quad marker testing,
improved signicantly the screening performance for ane-
uploidy. The proportion of Down syndrome pregnancies
diagnosed more than doubled and a chromosomal abnor-
mality was found in as many as 4% of those designated
as screen-positive
3
. In the late 1990s and early 2000s,
aneuploidy screening shifted to the rst trimester with
the combined test, which uses ultrasound measurement
of nuchal translucency thickness (NT) together with
maternal serum concentration of placental proteins
human chorionic gonadotropin (hCG) (free , intact
or total) and pregnancy-associated plasma protein-A
(PAPP-A). Currently available screening protocols also
incorporate additional ultrasound markers and sequential
screening using two blood samples, one in the rst
and one in the second trimester, with or without NT.
Consequently, screening performance has improved such
that more than nine-tenths of Down syndrome cases can
be diagnosed prenatally
4
and the yield from invasive
testing has risen to about 6%
5
.
Recently, analysis of cell-free (cf)DNA in maternal
blood for non-invasive prenatal testing (NIPT) has been
shown to be highly accurate in the detection of common
fetal autosomal trisomies. In 1997, Lo et al.
6
rst reported
their seminal discovery that plasma frompregnant women
contained cfDNA, including a fraction designated fetal,
although this is thought to be placental in origin, resulting
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. REVI EW
16 Benn et al.
fromapoptotic trophoblasts. This and other early studies
7
suggested that the fetal fraction was only 36%, but
more recent studies have found that it may be closer
to 1020%
8
. Fetal cfDNA can be detected as early as
4weeks gestation
9
and exceeds 4% of all the cfDNA in
nearly all women from 10weeks onward. The typical
size of the cfDNA fragments is approximately 150
basepairs
10,11
and, importantly, the entire fetal genome
is represented. It has been shown that the half-life of
fetal cfDNA is very short
12
; fetal fragments are no longer
detectable very soon after birth
13,14
. There is therefore
no serious concern that a prenatal cfDNA test could be
confounded by a prior pregnancy in multigravid women.
Maternal plasma cfRNA, unlike RNA extracted directly
from cells, appears to be relatively stable and can also
potentially be used in screening
1517
.
The commercial introduction of NIPT raises several
concerns. One major issue is the optimal integration of
NIPT into current screening practice
18
. Should NIPT and
conventional screening tests be treated as related but
independent measures when assessing aneuploidy risk or
should they be complementary to one another? How can
healthcare providers be educated and prospective patients
counseled about all the prenatal testing options now
available?
In this Review we provide the information needed by
clinicians and public health providers when considering
to whom NIPT will be offered and regarding the
consequences of doing so, practical suggestions on how to
implement this new and powerful technology into routine
clinical practice, and some indications of how we expect
this testing to expand. We focus primarily on the detection
of fetal chromosomal abnormalities, recognizing that the
technology will eventually be used for testing for a broad
range of other genetic disorders.
ESTABLISHED SCREENING MODALITIES
In the current context it is important to distinguish pre-
natal diagnosis of aneuploidy from antenatal screening. A
diagnostic test performed on chorionic villi, amniotic uid
or fetal blood needs to have very fewfalse negatives (aneu-
ploid pregnancies misdiagnosed as euploid) and false pos-
itives (euploid pregnancies misdiagnosed as aneuploid),
since the result will inform the decision as to whether to
terminate the pregnancy. In contrast, antenatal screening
does not aim to be denitive; rather, it is designed to iden-
tify women who are at sufciently high risk of common
aneuploidies as to warrant invasive prenatal diagnosis.
Since the invasive diagnostic procedures, mainly CVS and
amniocentesis, are hazardous and expensive, only a rela-
tively small group of women are identied as being at high
risk. Established screening programs determine the risk of
Down syndrome, many also including Edwards syndrome
(trisomy 18) and some including Patau syndrome (trisomy
13) and Turner syndrome (45,X). The model-predicted
aneuploidy detection rates (DR) for xed false-positive
rates (FPR) of established prenatal screening protocols
has been reviewed elsewhere
4
.
Among those undergoing invasive prenatal diagnosis
because of a high risk of Down and Edwards syndromes,
a large proportion have a different chromosomal abnor-
mality. Alamillo et al.
19
reported on 10years experience
of a combined test, giving risks of Down syndrome
and Edwards or Patau syndromes. The 97 cases with
chromosomal abnormalities included Down syndrome
(48%), Edwards syndrome (16%), Patau syndrome (6%),
Turner syndrome (9%), other sex chromosomal aneu-
ploidies (4%), other aneuploidies including mosaics (9%)
and chromosomal rearrangements (6%). The California
state-wide screening program
20
used a second-trimester
quad test for Down and Edwards syndromes and detected
a total of 1316 chromosomal abnormalities. Excluding
balanced translocations, these were Down syndrome
(48%), Edwards syndrome (14%), Patau syndrome
(2%), Turner syndrome (8%), triploidy (2%), Klinefelter
syndrome (47,XXY) (1%) and other abnormalities
(25%). Some of these were chance ndings but many
were related to maternal age or abnormal marker proles.
Conventional aneuploidy screening modalities also
need to be considered in the context of the substantial
number of non-chromosomal fetal abnormalities and
pregnancy complications that may be identied as a result
of this testing. Particularly important is the role of rst-
trimester ultrasound, which has the potential to identify
a non-chromosomal abnormality in approximately 1%
of cases
5
.
ESTABLISHED INVASIVE PRENATAL
DIAGNOSIS
Conventional cytogenetics
Chromosomal analysis of cultured chorionic villi and
amniotic uid cells can identify fetal aneuploidy,
structural rearrangements, such as translocations and
inversions, and relatively large duplications and deletions
(generally exceeding 5Mb in size)
21
. Chromosomal
analysis of amniotic uid cells is considered to be the
gold standard in prenatal testing because error rates are
exceedingly low, probably less than 0.010.02%, and
mostly appear to be due to maternal cell contamination,
laboratory error or typographic mistakes
22
. Error rates
with CVS are higher than those seen with amniocentesis,
and these can be due to conned placental mosaicism
(CPM), maternal cell contamination or lower resolution
banding of chromosomes
23
. In a small proportion of
cases in which CVS karyotyping is carried out there
are ambiguous results which may require resampling or
amniocentesis for clarication.
Molecular cytogenetics
Because of the time taken to obtain a complete chromo-
somal analysis, the occasional nding of a karyotype of
minimal or uncertain signicance and cost considerations,
some centers perform a rapid aneuploidy test (RAT)
using molecular genetic techniques as an adjunct to or
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
NIPT for aneuploidy 17
replacement for karyotyping. Methods include the use
of uorescence in-situ hybridization (FISH)
24,25
, quanti-
tative uorescence polymerase chain reaction (QF-PCR)
testing
26
and multiplex ligation-dependent amplication
(MLPA)
27
. These tests are generally designed to test only a
restricted range of cytogenetic abnormalities
28
. Although
proven to be very accurate, conrmatory testing using
conventional cytogenetics has been recommended
29
.
In contrast, relative to conventional karyotyping, other
molecular approaches can now improve substantially on
the detection of clinically signicant genetic imbalances.
In high-risk pregnancies with a normal chromosomal
constitution, the increased resolution of array compara-
tive genome hybridization (aCGH) coupled with single
nucleotide polymorphism (SNP) array has the potential to
identify submicroscopic copy number variations (CNVs),
specically microdeletions and microduplications, as well
as regions of homozygosity.
When compared to conventional chromosomal analy-
sis, aCGHand SNP array are purported to have signicant
advantages in terms of: (1) faster turnaround time;
(2) higher throughput while being less labor-intensive;
(3) higher resolution independent of the ability of the
cells to grow and/or generate good metaphase spreads;
(4) amenability to automation and quality control; and,
(5) direct mapping of aberrations to the genome sequence.
Faster turnaround time is possible because arrays can
be performed on DNA obtained from uncultured villi
and amniotic uid cells, thus avoiding the need for tis-
sue culture and the associated time delay required for
conventional chromosomal analysis.
A National Institute of Health (NICHD) multicenter,
prospective, blinded study
30
evaluated the application
of microarrays in over 4000 high-risk pregnancies
undergoing prenatal invasive testing largely because of
advanced maternal age, a positive rst-trimester screen
or the presence of an ultrasound anomaly. The use of
microarrays resulted in the identication of clinically
relevant CNVs that were unrecognizable by conventional
karyotyping in 2.5% of all cases. In those cases referred
for testing because of an ultrasound abnormality and
showing a normal karyotype, 6% were found to have
a clinically relevant CNV. These studies were performed
on low-resolution aCGH platforms; newer arrays have
much higher resolution. The use of SNP arrays has
expanded the ability to detect triploidy and signicant
regions of homozygosity, thereby detecting uniparental
disomy and potentially identifying disorders associated
with consanguinity. The NIH multicenter study also
revealed the limitations of all aCGH and SNP array
platforms, the principal limitation being the identication
of variants of uncertain signicance (VOUS).
Hazards of invasive testing
The principal hazard of amniocentesis is miscarriage, but
the excess risk associated with the procedure is difcult
to quantify precisely. Some 34% of mid-trimester
pregnancies will miscarry without amniocentesis and in
a particular case of fetal loss following the procedure it
is possible only rarely to attribute directly the adverse
outcome to the procedure; cases of amnionitis or chronic
amniotic uid leakage would be attributable but these
are relatively rare consequences. In a randomized trial of
amniocentesis at a single center
31
, the fetal loss rate was
0.8% higher in the amniocentesis group compared with
controls. There have been no similar randomized trials
for CVS, but a number of studies have compared late
rst-trimester CVS with second-trimester amniocentesis.
A Cochrane Review
32
showed that, when performed by a
skilled operator, the fetal loss rates of the two procedures
were comparable. An updated review
2
concluded that the
excess miscarriage rate for either procedure is between
0.5% and 1%. An initial concern that CVS may cause
fetal limb reduction defects has not been sustained, as
the reported excesses occurred when CVS was performed
early in gestation, before 10weeks
32
.
NIPT METHODS
The fact that fetal cfDNA and cfRNA are present only
as minority components of the total cell-free nucleic acids
in maternal plasma specimens has posed a signicant
technical obstacle in the development of NIPT. A
number of approaches have been suggested that allow
enrichment of the fetal component. These include taking
advantage of size differences between fetal and maternal
DNA fragments
11
, using formaldehyde in the samples
33
,
taking advantage of the nucleosome binding property
of the DNA
34
and immunoprecipitation of hypo- or
hypermethylated DNA
35
, but none is currently being
used in clinical practice. Even without enrichment a
number of technical approaches have now been proven
to be effective. Each approach relies on the extraordinary
advances in molecular biology and sequencing achieved
in recent years. We briey review the major methods
below, without discussing in detail all of the possible
renements and other developments that are currently
under consideration.
Shotgun massively parallel sequencing (s-MPS)
This approach relies on the identication and counting
of large numbers of the DNA fragments in the plasma
specimen. Using massively parallel sequencing (MPS),
millions of both fetal and maternal DNA fragments can
be sequenced simultaneously and, since the entire human
genome sequence is known, each piece that maps to a
discrete locus can be assigned to the chromosome from
which it came
36,37
. If fetal aneuploidy is present, there
should be a relative excess or decit for the chromosome
in question. The difference in counts will be small; for
example, if fetal trisomy 21 is present and the fetal
fraction is 20%, the relative excess in chromosome
21 DNA fragments will be (0.82) +(0.2 3) =2.2
compared with the situation for a euploid fetus
(0.8 2) +(0.2 2) =2, i.e. a relative increase in number
of chromosome 21 counts of only 10%. To be able to
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
18 Benn et al.
detect reliably these differences, large numbers of counts
are necessary, the fetal fraction needs to be appreciable
and the counts need to be compared with the expected
counts for euploid cases. The latter can be achieved either
by normalizing counts for the chromosome of interest
against other chromosomes that are expected to be
disomic within the same test run
36
or by comparing the
fraction of counts assigned to a particular chromosome
against the fractions seen for a set of known euploid
cases
38
. Results can be expressed as a z-score which
can be converted into a probability that a specic
chromosome result departs from the normal diploid
situation. Although it is possible to convert the z-score
to a patient-specic risk, major commercial providers
of this testing currently choose to present their results
only in terms of positive or negative based on z-scores
exceeding predened thresholds. Another variation of the
statistical analysis calculates two Student t-test statistics,
one based on the hypothesis that the result is from a
euploid fetus and the other that it is from an aneuploid
fetus and then calculates a further test statistic, L,
which is the logarithmic likelihood ratio
39
. Renements
in the statistical analysis also include adjustment for
guanine-cytosine (GC) base content of the sequences
40,41
.
The approach is referred to as shotgun because it relies
on sequencing and counting all informative chromosome
regions. Sequences which do not map to a unique locus
are uninformative. In some studies, only those fragments
that have sequences identical to the reference genome
sequence are accepted in the analysis, while in others
one or two mismatches in the sequence will be accepted.
Because of costs, it is common to run multiple samples
simultaneously (multiplexing) by adding a bar-code
sequence to each fragment that will allow identication
of the sample origin. The number of cases run together
can be variable and is an important consideration because
high-level multiplexing places a limitation on the total
number of sequence reads that are carried out on each
case (i.e. the depth of sequencing). The number of
fetal sequence reads will also vary according to the
fetal fraction and therefore a minimum fetal fraction,
typically 4%, is currently required. Most studies have
used Illumina sequence analyzers but there are also studies
using Solid
42,43
and Helicos
44
platforms.
Targeted massively parallel sequencing (t-MPS)
It is also possible to amplify selectively only those chro-
mosomal regions that are of interest (e.g. chromosomes
21, 18 and 13) and then to assess whether there is a
departure from euploidy based on the relative number of
DNA fragment counts for this subset of chromosomes
45
.
This targeted approach has the advantage of requiring
considerably less sequencing and thereby reduces costs.
The major commercial provider of this approach includes
an adjustment to allow for the variability in the propor-
tion of DNA that is fetal and then combines the results of
the laboratory test with maternal age to provide a patient-
specic risk for Down, Edwards and Patau syndromes
45
.
In theory, results could also be combined with other fac-
tors, for example results of other screening tests or prior
history of an aneuploid pregnancy.
Single nucleotide polymorphism (SNP)-based
approaches
The feasibility of a non-invasive aneuploidy test that takes
advantage of DNApolymorphisms was rst demonstrated
by Dhallan et al.
46
. In their approach, the paternal,
maternal and fetal SNPs present in blood (paternal),
buffy coat (maternal) and maternal plasma (maternal
and fetal) were evaluated as bands on sequencing gels.
Quantication of band intensities of the uniquely inherited
paternal SNPs in the plasma allowed an estimation of
the fetal fraction. Comparison of the maternal plus
fetal bands with the unique fetal band intensity also
allowed estimation of the fetal chromosome 21 dosage.
A similar approach was used by Ghanta et al.
47
, who
identied combinations of paired SNPs to focus on highly
informative regions and then used cycling temperature
capillary electrophoresis for evaluating fetal fraction and
dosage. The method was thought to be useful in situations
in which the fetal fraction was as low as 2%.
A more complex approach that relies on polymor-
phisms has been proposed by Zimmermann et al.
48
. The
approach involves a multiplex amplication of 11000
SNP sequences in a single PCR reaction carried out on
the plasma DNA followed by sequencing. Each product is
evaluated based on the hypothesis that the fetus is mono-
somic, disomic or trisomic. After considering the positions
of the SNPs on the chromosomes and the possibility
that there may have been recombination, a maximum
likelihood is calculated that the fetus is either normal,
aneuploid (chromosome 21, 18, 13 or sex chromosome)
or triploid, or that uniparental disomy is present.
In theory, with SNPs there should be advantages over
methods based on total sequence counting. SNPs can
provide information about parent of origin of aneuploidy,
recombination and inheritance of mutations. On the
other hand, SNPs account for only 1.6% of the human
genome and therefore enrichment for fetal DNA, deeper
sequencing or higher levels of high-delity amplication
may be required to identify unambiguously affected
pregnancies with small imbalances
49
. When few loci are
used in the analysis, greater attention needs to be paid to
the exclusion of regions in which there are rare benign
CNVs. The SNP approach could result in the inadvertent
detection of non-paternity or perhaps consanguinity. In
the case of assisted reproduction with an egg from an
unrelated donor, SNP-based approaches would need to
be modied to take into consideration the presence of
additional maternally derived fetal alleles not present in
the surrogate mother.
The laboratory methods and data analyses used for
SNP-based NIPT are substantially different from those
used in shotgun and targeted counting approaches. In this
Review we therefore consider the emerging clinical trial
data for SNP-based testing separately.
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
NIPT for aneuploidy 19
Methylated DNA-based approaches
This approach takes advantage of the fact that there
will be epigenetic differences throughout the genome
which will result in the differential expression of genes,
depending on the tissue type
50
. These differences can be
associated with either hyper- or hypomethylation of the
genes. Some genes present in placental cells (specically,
trophoblasts) will therefore differ from maternal cells
in the methylation pattern. Methylated DNA can be
chemically modied, thus allowing its characterization
separate from the unmethylated DNA
51
. Alternatively,
the hypermethylated DNA can be analyzed separately
following restriction enzyme digest of the hypomethylated
DNA
52
. If, additionally, there are polymorphic differences
in the fetal alleles, disturbance from a 1:1 ratio can be
used to identify trisomy. This has been demonstrated for
detection of Edwards syndrome but the approach has not
been tested in clinical trials
51
. Differentially methylated
regions may be relatively common
53
so the method may
also be applied to detect other aneuploidies.
It is also possible to use immunoprecipitation with
antibodies specic to 5-methylcytidine to enrich for the
methylated DNA. For example, the identication of a set
of specic DNA sequences on chromosome 21 that are
hypermethylated in placental cells allows the selective
enrichment of fetal cfDNA from maternal plasma
35
.
Following quantitative real-time PCR, there may be
measurable differences in the amount of chromosome 21-
derived DNA when fetal trisomy 21 is present, compared
with in euploid fetuses. One proof-of-principle study has
demonstrated the efcacy of such an approach to NIPT
54
.
A follow-up study using a modied set of differentially
methylated chromosome 21 DNA sequences also showed
encouraging results
55
. This approach is currently the
subject of further clinical trials. If an approach such
as this can be fully developed, it could be substantially
cheaper than sequencing-based methods.
Digital PCR
Another way in which the costs of sequencing can be
avoided is through the use of digital PCR. This technology
is based on the dilution of test materials such that
individual DNA fragments of interest (e.g. chromosome
21) can be amplied separately and counted
5658
.
Theoretical estimates of the number of amplication
events have been computed
59
and this approach could
become particularly attractive if reliable fetal cfDNA
enrichment methods were developed.
RNA-based testing
Identication of genes that are expressed by trophoblasts
but not maternal cells should result in the presence of
the corresponding cfRNA species in maternal plasma,
free of contaminating maternal RNA with a similar
sequence. If the cfRNA also carries polymorphisms that
distinguish the fetal alleles, the presence of aneuploidy
could be deduced from departures from a normal 1:1
ratio in the relative amount of the inherited alleles
60
. This
approach to NIPT showed some success in the initial
proof-of-principle studies, but unpublished clinical trial
data were disappointing. Nevertheless, with the discovery
of additional RNAs that can be analyzed, this approach
may yet play an important role in NIPT
61
.
NIPT IN CLINICAL TRIALS:
NON-MOSAIC DOWN, EDWARDS AND
PATAU SYNDROMES
Proof-of-principle studies demonstrated the feasibility of
NIPT using small series of maternal plasma samples
fromaneuploid and euploid pregnancies
3638,4046,6264
.
However, generally, the samples were not tested blind
to the outcome of pregnancy, ascertainment criteria were
not fully described and some samples were obtained after
an invasive diagnostic procedure which could have led
to increased transfer of fetal DNA into the maternal
circulation. There have nowalso been several large clinical
trials of MPS which have overcome these limitations and
potential biases
39,6575
. These have established NIPT as a
potentially highly effective screening test but not as a test
that could replace current invasive prenatal diagnosis. We
therefore prefer the term NIPT or cfDNA screening rather
than non-invasive prenatal diagnosis, which is misleading.
High-risk studies
There have been seven cfDNA studies involving
shotgun or targeted sequencing in women who had
invasive prenatal diagnosis because of high risk for:
Down syndrome
65,67
, Down syndrome or another
aneuploidy
66,68,69
or any disorder
70,71
. In one study, a
screen-positive screening test was the only indication
for high risk
69
, whilst in the four other studies in which
the reason for invasive testing was high risk for Down
syndrome or another aneuploidy, the indications were
mixed. The reason for invasive testing was a screen-
positive screening test in 77%
65
, 43%
67
, 30%
66
and
22%
68
of cases; advanced maternal age in 68%
66
, 37%
(only indication in these cases)
68
and 33%
67
of cases;
ultrasound in 13%
66
, 11%
67
and 7%
68
of cases; family
history in 5%
66
, 3%
67
and 3%
68
of cases; and multiple
reasons in 22%
68
and 9%
67
of cases. Five studies included
results on both Down and Edwards syndromes
6872
and
three also included Patau syndrome
68,72,75
.
The study design details for these trials on high-risk
women are summarized in Table 1. One study designated
as unclassied results with z-scores of 2.54.0 for
autosomal chromosomes and excluded them from the
DR and FPR calculations
68
. Doing so is not valid
76
,
particularly when no protocol was suggested for the
unclassied group, and in this Review we have reclassied
them as negative. However, the unclassied group had a
ve-fold increased likelihood of aneuploidy (3.5%(5/144)
of trisomies unclassied compared with 0.7% (9/1358) of
controls) and it could be argued that they should in fact
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
20 Benn et al.
Table 1 Eleven cell-free DNA studies of shotgun or targeted sequencing: methodological differences
Trial Design Average GA (weeks) Method Algorithm Cut-off*
High risk
Chiu et al.
65
Prospective and casecontrol 13 Shotgun z-score 3.0
Ehrich et al.
66
Casecontrol 16 Shotgun z-score 2.5
Palomaki et al.
67,72
Casecontrol 15 Shotgun z-score 3.0
Bianchi et al.
68
Casecontrol 15 Shotgun z-score 4.0
Sparks et al.
70
Casecontrol 18 Targeted Risk 1%
Ashoor et al.
69,75
Casecontrol 13 Targeted Risk 1%
Norton et al.
71
Prospective 16 Targeted Risk 1%
Other than high risk
Lau et al.
73
Prospective 14 Shotgun z-score? ?
Nicolaides et al.
74
Retrospective 12 Targeted Risk 1%
Dan et al.
39
Prospective 20 Shotgun z-score & L-score 2.5 or 1.0
Gil et al.
78
Prospective 10 Targeted Risk 1%
*For autosomal chromosomes. Includes maternal and gestational age. Results with scores 2.54.0 were unclassied. Positive if z <2.5
but L>1. GA, gestational age at testing.
be reclassied as positive, so those rates are also given in
this Review.
Table 2 summarizes the Down and Edwards syndromes
results for these seven cfDNA studies carried out in high-
risk pregnancies. In total these studies compared 594
samples from pregnancies with Down syndrome with
5745 that did not, and 193 samples from pregnancies
with Edwards syndrome with 5459 that did not, although
in both cases some had other aneuploidies. The results
fromthe various studies, notwithstanding methodological
differences, are consistent with each other and their
combined data represent the best estimates of DR
and FPR. For Down syndrome, the overall DR of
99.3% has a 95% CI of 98.299.8% and FPR of
0.16% (95% CI, 0.080.31%). Even with the most
extreme CI values this performance exceeds, by far,
anything achievable by current Down syndrome screening
protocols. However, this performance falls short of that
for current diagnostic tests. Furthermore, reclassifying as
positive the unclassied results in the study of Bianchi
et al.
68
would have increased the overall DR slightly to
99.5%, but almost doubled the FPR to 0.26%. Hence,
we can conclude that based on present evidence cfDNA
testing is not a diagnostic test of Down syndrome but
rather is a very good secondary screening test. A woman
with a positive cfDNA test will be at high risk of
Down syndrome but depending on her pretest risk it
may not be extremely high. This can be illustrated from
the combined data in Table 3, by estimating a positive
likelihood ratio (LR) using the DR divided by the FPR,
i.e. 634. For example, a woman with a pre-test risk of
1 in 270 following a combined test would have a post-
test odds of 634:269 or a risk of 2 in 3. This is not
sufciently high as to warrant termination of pregnancy
without a conrmatory invasive diagnostic test. Similarly,
the negative LR, estimated from the false-negative rate
divided by the true negative rate, is 1/148. A woman
with a pre-NIPT risk of 1 in 10 would have a 1 in 1333
risk after a negative test, which might be insufciently
reassuring for some women. These LRs are illustrative
and cannot be used to counsel individual women because
some of the studies had already incorporated prior risk
based on maternal age into their algorithms. However,
it is reasonable to conclude that using NIPT following a
screen-positive test for any of the established screening
protocols followed by conrmatory invasive diagnostic
testing if the NIPT result is positive will probably hardly
affect the overall screening DR but will reduce the FPR
about 300-fold.
For Edwards syndrome, the overall DR is 97.4%
(95% CI, 93.799.0%) and the FPR 0.15% (95% CI,
0.070.31%), with positive LR of 665 and negative
Table 2 Down syndrome and Edwards syndrome results in seven cell-free DNA studies carried out in high-risk pregnancies
Down syndrome Edwards syndrome
Trial DR (n (%)) FPR (n (%)) DR (n (%)) FPR (n (%))
Chiu et al.
65
* 86/86 (100) 3/146 (2.1)
Ehrich et al.
66
39/39 (100) 1/410 (0.24)
Palomaki et al.
67,72
209/212 (98.6) 3/1471 (0.20) 59/59 (100) 5/1688 (0.30)
Bianchi et al.
68
89/90 (98.9) 0/410 (0.00) 35/38 (92.1) 0/463 (0.00)
Sparks et al.
70
36/36 (100) 1/123 (0.81) 8/8 (100) 1/123 (0.81)
Ashoor et al.
69
50/50 (100) 0/297 (0.00) 49/50 (98.0) 0/297 (0.00)
Norton et al.
71
81/81 (100) 1/2888 (0.03) 37/38 (97.4) 2/2888 (0.06)
Total 590/594 (99.3) 9/5745 (0.16) 188/193 (97.4) 8/5459 (0.15)
*2-plex and 8-plex data were reported but only the more favorable 2-plex results are included here. DR, detection rate; FPR, false-positive
rate.
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
NIPT for aneuploidy 21
Table 3 Failure rates in 11 studies of cell-free DNA testing*
Trial Failure rate (n (%)) Reasons for failure
Chiu et al.
65
11/764 (1.4) Low total DNA (n=2), low DNA library concentration (n=8), low matched DNA sequence
reads (n=1)
Ehrich et al.
66
18/467 (3.8) Low % fetal DNA (< 4%) (n=7), low total DNA (< 556 copies) (n=7), low DNA library
concentration (< 32.3 nM) (n=15), low number unique DNA sequence counts (< 3
million) (n =11); some failed more than one criteria
Palomaki et al.
67
13/1696 (0.8) Low % fetal DNA (< 4%) (n=6), other QC parameters (n=7): low DNA library
concentration (< 25 nM) and low matched DNA sequence reads (< 12.5 million)
Bianchi et al.
68
16/532 (3.0) No fetal DNA
Sparks et al.
70
8/338 (2.4) Low % fetal DNA (< 3%), low DNA sequence counts, evidence from SNPs of non-singleton
pregnancy
Ashoor et al.
69
3/400 (0.8) Amplication and sequencing
Norton et al.
71
148/3228 (4.6) Low % fetal DNA (< 4%) (n=57), assay failure (n=91): inability to measure % fetal, high
variation in DNA counts and failed sequencing
Lau et al.
73
0/567 (0.0)
Nicolaides et al.
74
100/2049 (4.9) Low % fetal DNA (< 4%) (n=46), assay failure (n=54)
Dan et al.
39
79/11184 (0.7) Quality of separation, extraction, sample preparation and sequencing: low peak DNA library
size, low library concentration (< 30 nM) and low matched DNA sequence reads (< 2
million)
Gil et al.
78
40/997 (4.0) Low % fetal DNA (< 4%) (n=23), assay failure (n=17)
*Excluding tests rejected because of inadequate sample quality.
LR 1/38. Including the unclassied results in the study
of Bianchi et al.
68
as positive would have increased
the DR to 98.4% and the FPR to 0.20%. Since a
cfDNA test is unlikely to be aimed at detecting Edwards
syndrome alone, these false-positives should be regarded
as additional to those generated by cfDNA testing for
Down syndrome. There are no data with which to assess
whether there is any tendency for those testing falsely
positive for one type of aneuploidy to be falsely positive
for another. Assuming these are uncorrelated, the FPR for
both aneuploidies together is 0.30%, or 0.46% if those
unclassied by Bianchi et al.
68
are included.
Combining the results from the three studies which
included Patau syndrome
68,72,75
gives a DR of 30/38
(78.9%; 95% CI, 65.991.9%) and FPR of 17/4112
(0.41%; 95% CI, 0.220.61%), with a positive LR of
191 and a negative LR of 1/4.7. Including the unclassied
results as positive, the DR increased to 80.0% with no
increase in the FPR. On the basis of these results, an MPS
test for all three trisomies would have a FPRof 0.70.9%,
depending on the inclusion of the unclassied results.
Initial prospective clinical experience has been reported
for a commercial laboratory, Verinata Inc, offering
tests to high-risk women
77
. While follow-up information
was incomplete, among 5974 samples there were false
positives for Down, Edwards and Patau syndromes and
false negatives for Down and Edwards syndromes, with
rates compatible with those seen in the trial data.
Non-high risk
There have been four NIPT studies involving shotgun
or targeted sequencing in which the subjects were not
specically selected because they had invasive prenatal
diagnosis. Their methods are summarized in Table 1.
Lau et al.
73
reported prospective cfDNA screening
results of 567 patients, 49% of whom were tested
at 1213weeks gestation. These women comprised
179 (32%) who were screen-positive by conventional
testing, 194 (34%) who were screen-negative or awaiting
results and 194 (34%) who were unscreened. There was
insufcient follow-up to report DRs reliably, but eight
cases of Down syndrome and one of Edwards syndrome
were detected and there were no false positives. Seven of
the detected cases were among the screen-positive women
(1 in 25) and two, both Down syndrome, were among
the remainder (1 in 194).
Nicolaides et al.
74
carried out a retrospective study
using stored plasma samples from women who had a
combined test at 1113weeks gestation. After excluding
non-viable pregnancies, those terminated for reasons
other than Down or Edwards syndromes and cases lost
to follow-up, 1949 were successfully tested for cfDNA,
including eight with Down syndrome and two with
Edwards syndrome. All aneuploidies were detected and,
while there were no false-positive Down syndrome cases
among 1939 tested, there were two (0.05%) false-positive
Edwards syndrome cases among 1937 tested. Invasive pre-
natal diagnosis had been carried out in 4.3% (86/2049) of
those selected, including all of the cases with aneuploidy.
A large prospective study of cfDNA screening was
carried out in China
39
. A total of 11105 pregnancies
were tested successfully, including 4522 (41%) that
were screen-positive following conventional screening and
2770 (25%) with risk factors such as advanced maternal
age, ultrasound abnormalities and family history. All
of the known Down syndrome (n=140) and Edwards
syndrome (n=42) cases were detected by the cfDNA
test. However, approximately one third of the 10915
with negative test results were lost to follow-up; 2818
(26%) had invasive testing, a further 70 (< 1%) were
terminated, and follow-up was available for 4524 (41%).
The protocol included an insurance scheme whereby the
family would be compensated in the event of unpredicted
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
22 Benn et al.
delivery of either a Down or an Edwards syndrome
baby. This is likely to be a considerable incentive to
report false-negative results so the lack of follow-up may
not be a major problem. There were two proven false-
positive cfDNA results, one each for Down and Edwards
syndromes. However, a further eight positive cases did
not have a karyotype following fetal loss or termination
and therefore the possibility of additional false positives
cannot be excluded.
Gil et al.
78
carried out a prospective study of cfDNA
screening at 10weeks gestation in 997 women with
median age 37years; combined tests were also carried
out. A total of 15 cases of aneuploidy were detected by
cfDNA and conrmed by invasive testing (10 cases of
Down syndrome, four of Edwards syndrome and one
of Patau syndrome) and there was one false positive
(for trisomy 18). One case was positive for trisomy 21
but miscarried before invasive conrmation and no false
negatives had emerged at the time of writing.
These studies provide some, albeit limited, evidence
to suggest that cfDNA screening may be as effective
in the general population as it is in those already
scheduled for invasive testing: the FPR is not dissimilar
to that in the high-risk studies; all aneuploidies appear
to have been detected, although most of them were in
pregnancies which were already candidates for invasive
testing. Moreover, among high-risk women, performance
does not appear to be related to the indication for invasive
testing. Palomaki et al.
67
found that the DRs and FPRs
were very similar according to indication: screen-positive
99% (96/97) and 0.2% (1/637), respectively; advanced
maternal age or family history 100% (24/24) and 0.3%
(2/587); ultrasound 100% (51/51) and 0% (0/130);
and multiple reasons 95% (37/39) and 0% (0/112).
Nevertheless, until much larger general population studies
are published, the limited direct evidence of efcacy needs
to be supplemented by indirect evidence. The separation
in the distributions of z-scores or risks between aneuploid
and euploid pregnancies is necessarily dependent on the
fetal fraction and it is therefore important to consider
whether this might be lower in low-risk women.
Several covariables for fetal fraction have been
investigated, including indication for invasive prenatal
diagnosis, maternal age, screening marker level and
estimated aneuploidy risk. In one published series of
pregnancies at high risk of Down syndrome or other
aneuploidy
67
, there was no signicant difference in mean
z-score, in either Down syndrome or euploid pregnancies,
according to indication or correlated with maternal age,
but there was a signicant negative correlation with
maternal weight. The latter effect can be explained by
there being a xed mass of fetal DNA, from the placenta,
mixing with an increasing mass of maternal DNA; such
an effect is also seen in conventional maternal serum
markers. In another published series of pregnancies at
high risk of Down syndrome
69
, a multivariate regression
analysis was carried out which conrmed that maternal
weight was a covariable but maternal age was not
79
. In
addition it showed that there was a signicant positive
correlation between fetal fraction and both PAPP-A and
free -hCG, which might be related to placental volume,
although there was no association with gestational age in
either study. Since in Down syndrome pregnancies PAPP-
A is reduced on average and free -hCG is increased, the
correlations with fetal fraction may cancel each other out.
In Edwards syndrome pregnancies, both serum analyte
markers are reduced, consistent with reduced placental
volume
80
. There could therefore be a somewhat increased
chance for cfDNA test failure in affected pregnancies.
On the other hand, those cases with higher PAPP-A and
-hCG which might not be detected by the combined test
may showhigher fetal fractions and these may therefore be
more amenable to detection by a cfDNA-based NIPT. In a
further published series of those having invasive prenatal
diagnosis due to high risk generally
71
, the mean fetal
fraction was computed for deciles of maternal age, NT
and aneuploidy risk
81
. There was no signicant difference
in means between the highest and lowest deciles. The
highest decile for risk comprised 100 women with risk
greater than 1 in 10 and a mean fetal fraction of 11.4%
compared with 135 women in the lowest decile with risk
less than 1 in 6500 and a mean fetal fraction of 10.8%.
Failed results
Several of the MPS studies discussed above specied the
number of pregnancies not considered for cfDNA testing
because of sample quality. Overall, among 13260 eligible
pregnancies there were 814 (6.1%) untested. This rate
ranged from 0.8% to 9.9% between the studies and may
reect the stringent standards required when carrying
out a research project. Of more concern is the 2.0% of
pregnancies (436/22 222) in which a cfDNA test was
performed but failed because of insufcient fetal DNA
or other technical reasons (Table 3). The failure rate
may be even higher in those resampled after an initial
failure: it was 32% (13/40) in one study
78
. Quality
control criteria varied between the studies but in those
that included a minimum acceptable percentage of fetal
DNA, about half the failures were due to this not being
met. In the prospective laboratory experience reported
by Futch et al.
77
, a result was not possible in 2.4% of
cases. This is relevant to comparisons with conventional
screening, which rarely reports failure to obtain a result.
Three prospective studies of women not specically
selected because they had invasive prenatal diagnosis
provide information on the extent to which a successful
test result was achieved following an initial failure. In one
study
73
, among 567 successful cfDNA tests, four (0.7%)
had required a repeat blood sample and no other tests had
failed during that period (T.K. Lau, pers. comm.). The
time from the initial blood draw to the cfDNA report was
greater than 14days for 16 (2.8%) tests, including those
requiring a repeat sample. In another study
39
, among
11105 tests, 97 (0.9%) needed resampling. In this study
as a whole, the turnaround time was within 10 working
days for 95% of pregnancies and within 15 for 99%;
those requiring resampling took a further 1015days to
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
NIPT for aneuploidy 23
produce a report. In the third study
82
, a request for a
patient redraw was made in 1.9% of referrals and among
those who did undergo a redraw there was sufcient fetal
cfDNA in the second sample in 56% of cases. The success
rate was strongly dependent on maternal weight.
SNP studies
Following the initial proof-of-concept study using poly-
morphisms and maximum LRs for test interpretation
48
,
the results of one small clinical trial have been published
83
and additional information has been presented at scien-
tic meetings in the form of an abstract
84
and a poster
85
.
The study of Nicolaides et al.
83
involved 242 singleton
pregnancies with results available for 229 (95%). All 32
aneuploid and 210 euploid cases were identied correctly.
The abstract
84
provided information on 407 pregnan-
cies including 44 with aneuploidies and the results were
100% correct among samples that passed quality control.
In addition to the maternal plasma, paternal blood or
buccal mucosa samples were genotyped. Whilst this was
not an absolute requirement (paternally inherited alleles
in the fetus can be deduced), the failure rate was lower
when a paternal sample was available. In the poster
85
,
there were 763 samples, of which 45 (5.9%) failed to pass
quality control, and among the remaining 717 (including
47 Down, 15 Edwards, 7 Patau and 12 Turner syndromes)
samples all except one result, for Turner syndrome, was
correct.
Multiple pregnancies
There is a potential problem with using cfDNA to
screen for aneuploidy in multiple pregnancies. As with
biochemical screening, in twins discordant for trisomy the
excess in fragments from a specic chromosome in the
affected twin may be masked by the normal proportion
of those fragments in the euploid cotwin. This would also
be the case for higher order multiple pregnancies. This
problem might be overcome by using SNPs to measure
the variations in the fetal fraction between genomic
regions
86,87
. The method allows determination of zygosity
and, in dizygous twins, examination of the distribution of
fetal fragments for each fetus.
At present, only two small multiple-pregnancy cfDNA
series have been published, using the same method as
for singletons
88,89
. The rst series was derived from one
of the high-risk studies
67
and included three discordant
twins (two with Down and one Patau syndrome), ve
concordant Down syndrome twins, 17 euploid twins and
two euploid triplets
88
; there were no false-positive or false-
negative cfDNA test results. The second series was from
one of the studies of women not selected because they
were at high risk
73
and included only one discordant
twin with Down syndrome and 11 euploid twins
89
;
all pregnancies were correctly classied by the cfDNA
test.
In a singleton pregnancy that had had a vanishing twin,
it is theoretically possible that apoptosis of cells from
the fetoplacental remains of the non-viable fetus could
interfere with the cfDNA result; there is a relatively high
chance that the fetal loss would have been associated
with aneuploidy and this could lead to a false-positive
result
77
.
COST OF NIPT
Currently, the cost of a cfDNA test is high, ranging from
about $800 to $2000 in the USA, and from $500 to
$1500 elsewhere. Whilst this may not necessarily deter
individuals from being tested, it is likely to determine
whether public health planners provide the testing and
will be the basis for whether health insurance schemes
reimburse those tested. We have published a sensitivity
analysis
90
in which the factors contributing to these
overall costs are elucidated for four cfDNA screening
policies: (1) cfDNA screening for those screen-positive on
combined test; (2) universal cfDNA screening, replacing
all current screening modalities; (3) contingent cfDNA
screening, for the 1020% of women with the highest
risk based on a combined test; (4) cfDNA screening for
women of advanced maternal age and younger women
with a positive combined test
90
.
Restricting cfDNA testing to screen-positive cases leads
to a small reduction in detection, a massive reduction in
false positives and is the least expensive way in which
to utilize cfDNA testing. With this policy the average
cost per Down syndrome birth avoided is dependent only
on the difference in unit cost of a cfDNA test and of
invasive prenatal diagnosis. For a combined test with
FPR of 3%, the average cost would be reduced by using
secondary cfDNA testing, if the unit cost was less than or
equal to three-quarters that of invasive testing. Even at a
higher unit cost of cfDNA, the average cost per fetal loss
prevented would be relatively low.
For other policies in which the aim of cfDNA testing
is to increase detection, the most important public health
consideration will be not the total or average cost but
the marginal cost compared with the existing screening
provision. The marginal cost is the average cost per Down
syndrome birth avoided or loss prevented over and beyond
that which would have been avoided by the existing
service. This is shown in Table 4 for xed unit costs for
the combined test and invasive testing and with variable
unit costs of cfDNA. The table also shows screening
performance.
Universal cfDNA screening yields both a very high DR
and a small FPR but the marginal cost is very high,
exceeding the lifetime cost to the system of a Down
syndrome birth
9194
. With universal screening the cost
per fetal loss prevented when replacing conventional
screening is also very high. Contingent protocols, in which
all women receive the combined test and the 1020%with
the highest risk receive cfDNA testing, are efcient and
can reduce massively the marginal cost compared with
universal screening. The average cost of preventing a fetal
loss is also several times lower than that for universal
cfDNA testing. A policy of cfDNA testing only for older
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
24 Benn et al.
Table 4 Four cell-free (cf)DNA policies*: performance, marginal cost per Down syndrome (DS) birth avoided and average cost per fetal loss
prevented, according to unit cost of cfDNA testing
Policy Performance
Marginal cost per DS birth
avoided ($ million) for
unit cost cfDNA of:
Average cost per fetal loss
prevented ($ million) for
unit cost cfDNA of:
Combined cfDNA DR (%) FPR (%) OAPR $1500 $1000 $500 $1500 $1000 $500
None All 99.3 0.16 1:1 5.84 3.63 1.42 9.47 5.89 2.31
All Contingent:
10% 90.4 0.016 7:1 1.11 0.65 0.19 0.82 0.48 0.14
15% 92.8 0.024 5:1 1.39 0.86 0.32 1.33 0.82 0.30
20% 94.4 0.032 4:1 1.67 1.05 0.44 1.85 1.16 0.48
< 35 years 35 years 85.0 2.1 1:19 2.90 1.73 0.57 2.35 1.41 0.46
< 35 years 35 years or positive
on combined test
84.7 0.017 6:1 3.69 2.02 0.37 0.81 0.44 0.08
*Three reported in a modeling analysis
90
and the fourth, cfDNA restricted to older women, using the same model. Combined test,
false-positive rate (FPR) = 3%, unit cost =$150; invasive prenatal diagnosis, unit cost =$1000, iatrogenic fetal loss rate =0.5%. DR,
detection rate; OAPR, odds of being affected given a positive result.
women was not substantially more effective than was
the combined test and is more costly than contingent
screening. A hybrid policy of offering cfDNA testing to
all those of advanced maternal age plus all combined test
screen-positive women was better as it reduced the FPR
massively but it was less effective and more expensive
than was a contingent testing policy.
The detection rate for a contingent policy can be
increased by including additional serum and ultrasound
markers. Nicolaides et al.
95
suggest markers which are
also of value in screening for pre-eclampsia and fetal
cardiac abnormalities. Cost-benet analyses for such
variants have not yet been carried out.
Two other published studies have estimated costs when
cfDNA is applied to women with positive conventional
screening results, using the estimated NIPT DR and FPR
from their own single study. Palomaki et al.
67
considered
total costs following screening in 100000 screen-positive
women, with an average risk of 1 in 33 for a Down
syndrome fetus. They concluded that if the unit cost of
cfDNA was less than 96% of the unit cost of invasive
testing, there could be a net saving (4% of costs need
to be assigned to test failures and conrmatory invasive
testing following cfDNA positive results). The second
study also considered 100000 screen-positive women,
modeled as a mixture of combined test or second-trimester
serum screening
96
. They estimated total costs including
both the conventional screen and subsequent testing and
concluded that with a unit cost of cfDNA $1200, CVS
$1700 and amniocentesis $1400, there would be a 1%
cost reduction. One other study has estimated costs
for cfDNA screening in older women and in younger
women with screen-positive conventional screening tests,
using DR and FPR estimates from the literature
97
. The
authors claimed that this policy, in the USA, would have
lower total cost compared with conventional screening.
However, this conclusion appears to have been reached
following incorrect assumptions about the DR and FPR in
younger women and a very low uptake of invasive testing
for women screen-positive by conventional screening
compared with cfDNA screening
90
.
FUTURE PROSPECTS: NIPT FOR
ADDITIONAL CYTOGENETIC
ABNORMALITIES
Other autosomal abnormalities
With the widely anticipated rapid improvements in
sequencing, it may soon be possible to screen for
other autosomal trisomies in situations in which the
fetal fraction is very low, thereby allowing detection of
aneuploidy even earlier in pregnancy. Early pregnancy
is associated with very high levels of chromosomal
abnormality and although nearly all abnormal cases
will spontaneously abort, early identication could be
benecial to women. First-trimester autosomal trisomy
can involve any chromosome, but some occur much more
frequently than others
22
. Detection of these through NIPT
will be facilitated by the larger size of the additional
chromosome in the non-viable trisomies. Early studies
suggested that there was greater variability in the fragment
counts for some of the larger chromosomes
36
, but this
could be overcome by GC correction and selective use of
the most consistent regions of chromosomes
37,62
.
Screening for a large number of potential abnormalities
together does require very high specicity for each
component so that the aggregate FPRfor all chromosomes
combined is low.
Sex-chromosome abnormality
Establishing NIPT for sex-chromosome imbalances is
desirable because some specic sex-chromosomal abnor-
malities are associated with a sufciently severe phenotype
as to fully justify prenatal diagnosis: for example, many
cases of Turner syndrome, most of which will not sur-
vive to full term, gain of multiple X chromosomes and
some unbalanced structural rearrangements of the X-
chromosome. Collectively, unbalanced sex-chromosomal
abnormalities constitute over a third of the unbalanced
karyotypes seen in amniocenteses performed in women
under 35years of age and nearly a quarter of those in
older women
98
. There is mosaicism in approximately one
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
NIPT for aneuploidy 25
third of sex chromosomal aneuploidies detected through
amniocentesis
98
.
One commercial laboratory that provides NIPT,
Sequenom Inc, was routinely providing information on
the presence or absence of Y-chromosome DNA in their
reporting. Based on a series of 1639 samples, the DR
for a Y-chromosome was 828/833 (99.4%), and the
accuracy for the prediction of a female fetus was 803/806
(99.6%)
99
. More recently, they have been providing
NIPT for Turner syndrome, Klinefelter syndrome,
47,XXX and 47,XYY based on a study by Mazloom
et al.
100
, with DRs of 83% (25/30), 85% (11/13), 83%
(5/6) and 75% (3/4), respectively, for each of the these
abnormalities. The FPR for Turner syndrome was 0.2%
(4/1945) and there was one other false positive (47,XXX)
in the series
100
. Of the normal cases in which gender was
interpreted, it was inconsistent with the karyotype result
in 0.7% (13/1814)
100
. Bianchi et al.
68
developed a com-
plex algorithm for evaluating sex-chromosome-derived
cfDNA in plasma. Their study derived z-scores for the
X-chromosome (z
X
) and Y-chromosome (z
Y
). Aresult was
classied as positive for Turner syndrome if z
X
<4.0 and
z
Y
<2.5; it was classied as 46,XX or 46,XY or 47,XXX
or 47,XXY or 47,XYY using different cut-offs, and all
others were unclassied. Among the non-mosaic Turner
syndrome cases the DRwas 75%(15/20) and the FPRwas
0.2% (1/462), with 49 remaining unclassied (including
four affected and 45 controls). Among the mosaic cases
with a 45,X cell line, the DR for the monosomy was
29% (2/7). For the other non-mosaic sex chromosomal
abnormalities the DRs were: XXX, 75% (3/4); XYY,
100%(3/3); and Klinefelter syndrome, 67%(2/3). Robust
estimates for the efcacy of sex chromosomal aneuploidy
detection are therefore not provided, although Verinata
Inc. does offer testing on the basis of this study.
There are several factors that could confound the
development of highly effective NIPT for X- and Y-
chromosomal imbalances. First, many fetal sex chromo-
somal aneuploidies are mosaic, with either a normal cell
line or an additional abnormal cell line. This includes
situations in which one abnormal cell line may mask
another when cfDNA is analyzed. For example, fetal
45,X/46,XXX mosaicism may show approximately nor-
mal proportions of X-chromosome cfDNA if the propor-
tions of the two cell types are similar. Second, although
the detection of Y-chromosome-specic DNA is relatively
simple and reliable in establishing gender
101
, nding suf-
cient Y-chromosome loci that are informative for copy
number quantication may be difcult. The potentially
informative euchromatic part of the Y-chromosome is
relatively small, it contains many sequences that show
polymorphic CNVs, and some segments show partial
homology with X-chromosomal sequences
102
. A third
difculty is that normal adult females show an age-related
loss of X-chromosomes
103
. In women at normal reproduc-
tive age, the proportion of maternal cells with a missing
X-chromosome can be comparable to the proportion of
cfDNA in a maternal plasma specimen and there may be
some women with unusually high levels of 45,X cells
104
.
The age-related loss of the X-chromosome is thought
to involve preferentially the inactive X-chromosome
105
.
Some women may also show cells that have gained an
X-chromosome
106
. Clonal populations of maternal cells
with X-chromosome gain or loss in hematological cells
(which are believed to be the source maternal cfDNA in
plasma), maternal constitutional mosaicismor even a non-
mosaic X-chromosomal abnormality such as 47,XXX are
also possible
107
. The fetal X-chromosomal dosage there-
fore has to be identied against a background of maternal
X-chromosomal DNA that is likely to be quantitatively
more variable than is seen for the autosomes. These con-
siderations suggest that NIPT based on simply counting
the total X- and Y-chromosomal sequences in maternal
plasma and comparing them with the expected normal
karyotype values may be less effective than that which is
achievable for autosomal gain or loss. Furthermore, it may
be misleading to quote a DR that is based on all sex chro-
mosomal abnormalities combined for genetic counseling.
Approaches based on quantication of SNPs present
in the plasma DNA may be informative for at least some
fetal sex chromosomal aneuploidies. Y-chromosome
gain, most cases of Turner syndrome
108
, approximately
half of cases with Klinefelter syndrome
109
and some other
imbalances are attributable to erroneous segregation of
a paternally derived sex chromosome. These cases could
therefore be identied in a maternal plasma specimen
through the over- or under-representation of paternal X-
and Y-chromosomal SNPs that differ from the mothers
SNPs. On the other hand, most cases of 47,XXX
110
,
and the remaining cases of Turner syndrome, Klinefelter
syndrome and others are attributed to erroneous segre-
gation of a maternally derived chromosome. The reliable
quantication of cfDNA with these fetal X-chromosomal
SNPs, against a background of the identical maternal
SNPs, is more difcult.
Mosaicism
Two of the NIPT clinical trials have specically addressed
the ability to detect mosaic cytogenetic abnormalities in
the fetus
68,111
. In one study
68
, of 11 cases in which a non-
complex mosaicism was known to be present based on
karyotyping, ve were called affected and six unaffected.
In additional studies involving deeper sequencing,
four mosaic abnormalities recognized by conventional
chromosomal analysis were not detected by NIPT
112
. In
the second study involving ve cases
111
, the abnormal line
was detected in one case and not detected in four cases.
Approximately 14% of all cytogenetic abnormalities
identied through karyotyping of amniotic uid cells are
mosaic
98
. For CVS, the proportion showing mosaicism
is much higher; approximately 45%
23
. This excludes
cases in which there is pseudomosaicism (possible cell
culture artifact) and includes cases in which there is a
discrepancy between CVS direct or short-term culture
preparation results (cytotrophoblasts) and longer termcell
culture results (mesenchymal tissue). Discrepancy usually
involves abnormal cells in cytotrophoblasts with normal
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
26 Benn et al.
mesenchyme, but the reverse situation can also be present.
An abnormal cell line can also be present in both direct and
long-termcultures but be undetectable in the fetus. Longer
termcell culture is usually considered to be a more reliable
indicator of the true karyotype of the fetus compared
with direct preparations. Overall, approximately 1% of
all CVS specimens will show a second cell line that is
not detectable in the fetus, a situation termed conned
placental mosaicism (CPM)
113
. This may sometimes
result in spontaneous loss, low birth weight or preterm
delivery, particularly when it involves both trophoblastic
and mesenchymal cell lineages
114
and/or when it involves
specic karyotype abnormalities such as trisomy 16
115
.
However, in most pregnancies in which CPM is detected
the pregnancy outcome is normal
116
.
The frequency and signicance of CPM is relevant to
NIPT because cfDNA is believed to originate primarily
from non-viable trophoblasts. CPM could result in either
false-positive or false-negative NIPT results. It has been
suggested that viable pregnancies with Edwards or Patau
syndrome may have a substantial euploid cell line in
the trophoblasts
117,118
and therefore NIPT could fail to
detect some viable trisomic cases in which there is a
substantial population of abnormal cells in the fetus but
not in the placenta. This would lead to an apparent false-
negative result by NIPT. Conversely, the trophoblasts
may be mostly abnormal and the fetal cells normal,
leading to an apparent false-positive result. Moreover, if
chromosomally abnormal trophoblasts were more likely
to undergo apoptosis, NIPT could preferentially identify
cases in which the proportion of abnormal cells in the
placenta is low.
A number of cases have already been described which
illustrate how differences in the distribution of normal
and abnormal cells may lead to test result discrepancies
between NIPTand karyotyping. Faas et al.
43
described the
identication of Turner syndrome based on NIPT when
the abnormal cell line was conrmed by cytogenetic anal-
ysis of short-term CVS cultures but the abnormality was
not seen in the long-term CVS cultures or in the newborn
blood. Choi et al.
119
described a case involving possible
trisomy 22 detected by cfDNA testing. The trisomy 22
cell line was conrmed by analyzing the term placenta
but the abnormality was not detected in a blood specimen
from the dysmorphic baby. Hall et al.
120
describe a
case with a positive Patau syndrome cfDNA test, mosaic
47,XY,+13/46,XYresult for CVS, normal result for amni-
otic uid cells, normal result for cord blood, but trisomy
13 mosaicism conrmed in the placenta. These examples
can each be interpreted as CPM although the presence of
low-level true fetal mosaicism cannot be excluded.
Early trials on chromosomal analysis following CVS
provide an indication as to how often CPM might lead to
an apparent false-positive or false-negative result. Because
milder abnormalities such as sex chromosomal aneuploidy
or mosaicism involving abnormalities not typically seen
in livebirths may not have been fully pursued in follow-
up studies, this analysis needs to be conned to Down,
Edwards and Patau syndrome. Ledbetter et al.
23
noted six
cases (four involving chromosome 13, one chromosome
18 and one chromosome 21) interpreted as a false-positive
CPM result and one false-negative result (chromosome
18) among 6560 cases that had analysis of a direct CVS
preparation. Similarly, Smith et al.
121
reported nine cases
(two involving chromosome 13, four chromosome 18 and
three chromosome 21) interpreted as a false-positive CPM
result and ve false-negative results (two chromosome 18,
three chromosome 21) in approximately 18195 cases
that had direct preparation analysis. Combining these
data, for chromosome 21, 18 or 13, the incidence of
false positives was 1 in 1650 and the incidence of false
negatives was 1 in 4126. These estimates are crude because
the number of cells analyzed on direct CVS preparation
is often small, some women receiving an abnormal result
terminated their pregnancy without conrmation, and
low-level true fetal mosaicism may have gone undetected.
It should be noted that most cases of CPM do not
involve chromosomes 21, 18 and 13. Therefore, as NIPT
tests are expanded to include additional chromosomes or
chromosomal regions, the issue of erroneous results due
to CPM is likely to become increasingly important.
For those cases that do show true fetal mosaicism, the
phenotype can be highly variable but will generally be less
severe than that present with the non-mosaic abnormal
karyotype
22
. Although the proportion of abnormal cells
in amniotic uid or chorionic villi can be a poor
guide to phenotype in individual cases, in general, a
high proportion of abnormal cells is associated with
adverse pregnancy outcome
122
. NIPT tests that only
present a result categorically as either positive or
negative obviously cannot provide any information
about the proportion of abnormal cells or clinical severity.
Furthermore, since some of the clinical trials have
specically excluded mosaic cases and optimized their
z-score cut-off on the basis of non-mosaic cases, testing is
presumably biased towards not detecting mosaic cases. It
is not yet clear whether patients with z-scores close to the
cut-off should be offered invasive testing to help rule out
mosaic trisomy. To present this problem appropriately
to clinicians, and also to allow better integration of the
results with other screening and diagnostic techniques,
we recommend that NIPT reports should be formatted
so that they provide patient-specic risks for each non-
mosaic trisomy, the z-score and the fetal fraction, rather
than just a categorical positive or negative nding.
The possibility of maternal mosaicism also needs to be
considered. This includes the rare situation in which the
abnormal maternal cfDNA is derived from a malignant
cell population
123
. Maternal mosaicism could potentially
be recognized using SNP-based NIPT methodology.
Translocations
Unbalanced translocations, in theory, should be identi-
able in NIPT as a result of the partial trisomy and
monosomy that is present. In a retrospective analysis of
six plasma specimens from pregnant women with known
fetal imbalances, Srinivasan et al.
112
were able to identify
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
NIPT for aneuploidy 27
the imbalances in maternal plasma in all six cases. In two
of their cases the karyotype showed additional material of
unknown origin and the NIPTanalysis was able to provide
a more precise characterization of the extra chromosomal
material. The ability to detect routinely smaller imbal-
ances would require deeper sequencing and improvements
in bioinformatic analysis of sequencing data
41,124126
.
For SNP-based approaches to NIPT, detection of small
imbalances requires identication of sufcient informative
polymorphisms in the region of interest.
Conventional chromosomal analysis does allow detec-
tion of most apparently balanced translocations and large
inversions. Prenatally identied de-novo rearrangements
are associated with an increased risk for abnormality
127
.
It would therefore be desirable to detect these non-
invasively. Paired-end sequencing of DNA fragments that
span a chromosomal breakpoint can identify at least some
translocations through the analysis of cellular DNA. For
example, Schluth-Bolard et al.
128
used a whole-genome
paired-end protocol to identify the specic sequence dis-
ruptions in four cases with apparently balanced transloca-
tions and abnormal phenotypes. Breakpoints were mostly
in repeat sequences and most breakpoints were associated
with gain or loss of some basepairs. Talkowski et al.
129
used sequencing to rene breakpoints and establish a
diagnosis of CHARGE syndrome based on disruption of
the CHD7 gene by analyzing amniotic uid cells. These
studies involved the sequencing of larger pieces of DNA
and were carried out in cases in which the karyotype
information was used to help identify candidate DNA
fragment sequences that spanned the breakpoints.
Although the routine detection of de-novo rearrange-
ments using cfDNA, without knowledge of where to look,
would not seem to be technically achievable at this time,
it is potentially possible. The challenge is similar to the
detection of other mutations
130132
, but is made sim-
pler with the presence of two related reciprocal events.
By sequencing cfDNA it should be possible to construct
the haplotype sequence in which one end matches per-
fectly to a chromosome, the other end matches perfectly
a second chromosomal location and there are similar
hybrid sequences present that correspond to the recipro-
cal translocation product(s). Improved sequencing delity,
deeper sequencing and the development of advanced bio-
metric algorithms that specically look for the hybrid
sequences would be needed.
Inherited rearrangements are generally considered
benign, but their detection through karyotyping follow-
ing CVS or amniocentesis is considered advantageous
because it identies a risk factor for future pregnancies.
When a balanced translocation is already known to be
segregating in a family and the phase of SNPs on the
chromosomes in the region of interest can be established,
the presence or absence of the breakpoint can be inferred
from the SNPs in the cfDNA in a plasma specimen. In the
case of a paternally inherited translocation, the presence
of the translocation chromosomal SNPs and absence of
the normal homolog SNPs identies the translocation in
the conceptus. In maternal inheritance, detection of the
translocation relies on the identication of a relative excess
of the translocation chromosomal SNPs and relative de-
ciency of the normal homolog SNPs in the plasma sample.
Triploidy
Triploid pregnancies can arise as a result of an
additional set of maternal (digynic) or paternal (diandric)
chromosomes. Most recognized cases show a 69,XXX
or 69,XXY karyotype; 69,XYY forms are thought to
undergo early rst-trimester miscarriage
133
. Most cases
of triploidy will be identied by ultrasound abnormality
and characteristic rst- or second trimester-maternal
serum marker levels
134136
. It is exceptional for these
pregnancies to survive into the third trimester.
Since all chromosomes are proportionately represented
in a 69,XXX pregnancy, NIPT based on the counting
of all chromosomal sequences would yield a normal
result. Pregnancies with either a 69,XXY or a 69,XYY
karyotype might be identiable through unexpected sex
chromosomal sequence ratios. Digynic triploidy might
be particularly difcult to detect by NIPT because the
placenta is very small and the fetal fraction may therefore
be reduced
137
. Consistent with this, Bianchi et al.
68
reported that of nine cases of fetal 69,XXXtriploidy, there
was insufcient fetal cfDNA for analysis in three (33%).
Using the approach that takes into consideration the
SNPs present in the cfDNA, diandric triploidy could be
identied through the presence of two different paternally
inherited SNPs at some loci. Digynic cases with sufcient
fetal cfDNA for analysis might be recognized through
the quantitative excess of maternally inherited fetal DNA
fragments relative to paternally inherited fragments.
CNVs
Small duplications and deletions have the potential to be
detected through NIPT. Examples that have been detected
include a 4-Mb deletion on chromosome 12
138
, and two
cases in which 3-Mb 22q11.2 deletions were present
139
.
In a series of 11 prenatal plasma specimens, Srinivasan
et al.
112
identied small gains or losses not reported in
the clinical karyotypes in eight cases, including four with
two such variations. However, it is not clear whether they
were true or false positives.
A variety of strategies have been proposed to detect the
small deletions, insertions, inversions, and duplications
that are common in the human genome
140
. Much of
the analysis is contingent on the ability to identify
departures in the location and orientation of sequenced
DNAfragments fromwhere they are expected to be found
in reference mapped genomes. The ability to do this type
of analysis in a non-invasive test is a long-term goal in
molecular cytogenetics.
Overview of future testing
Currently, only methods based on MPS have been the
subject of clinical trials and can be viewed as being
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
28 Benn et al.
suitable for clinical use. An optimal NIPT for fetal
aneuploidy will be accurate, simple, cheap, applicable
early in pregnancy and fully compatible with existing
prenatal screening methods so that risks developed using
the different approaches can be combined. Many of the
methods under consideration can be rened to allow for
the detection of other chromosomal imbalances. MPS
renements that will aid this process include selection
of optimal sequences for analysis, adjustments for base
composition and the use of sequencing platforms that
provide higher read accuracy.
Sequencing also provides opportunities to detect the
presence or absence of single gene disorders
130
. The
methods can even be applied to the construction of an
entire fetal genotype
131,132
. On the other hand, methods
that do not require sequence information fromthe parents
or which do not generate unwanted additional genotypic
status of the parents (for example resulting in information
about paternity or other disease risks) should also be
considered advantageous. The optimal technologies may
ultimately be those which are most amenable to answering
only those clinical questions that are being asked, rather
than those which provide the most information.
COUNSELING
The internationally accepted approach to dealing with
patients diverse ethical, religious and cultural values is to
provide individual non-directive counseling, testing and
other clinical information at each step in the screening and
diagnosis process. This model, which maximizes patient
autonomy in reproductive decision making, has been
adopted by diverse populations throughout the world.
Counseling of women who are considering NIPT is
challenging. The range of cytogenetic abnormalities cur-
rently detectable through cfDNA testing is smaller than
that detectable by conventional karyotyping and substan-
tially less than that achievable through microarray testing.
In the absence of public health policies that dene testing
strategies, each high-risk woman who previously would
have made a decision about amniocentesis or CVS and
karyotyping, will now need to choose between no testing,
NIPT, invasive testing with conventional cytogenetics or
invasive testing with microarray testing. These women will
need to be counseled carefully regarding the complex ben-
ets, hazards and trade-offs associated with each option.
In part, these decisions will be determined by the
indication for testing and perceived risk, but they will also
be determined by the available alternative testing services.
For example, in a situation in which routine invasive
testing is based on RAT for a limited set of disorders, the
choice of NIPT that detects the same range of aneuploidies
might seem preferable. On the other hand, offering array
technology may provide much more valuable information
concerning fetal wellbeing than can current NIPT. This
benet has to be weighed against the risk of losing a
normal pregnancy following an invasive procedure or the
heightened anxiety caused by the presence of a CNV of
unknown signicance, poorly dened penetrance and/or
variable expressivity. In addition, the testing could expose
consanguinity or non-paternity. Given these realities, the
challenge now facing those providing pretest counseling
cannot be understated. The National Society of Genetic
Counselors recognizes that due to limited resources,
pretest counseling cannot always be provided by a
genetic counselor and this information will therefore also
need to be communicated by other qualied healthcare
providers
141
.
The availability of professional guidelines, such
those from the American College of Obstetricians and
Gynecologists, together with the Society for Maternal-
Fetal Medicine
142
, the Society of Obstetricians and
Gynaecologists of Canada
143
, the National Society of
Genetic Counselors
144
, the American College of Medical
Genetics and Genomics
145
and the International Society
for Prenatal Diagnosis (ISPD)
146
, can greatly facilitate
the process by clarifying which processes are scientically
valid and clinically acceptable standards of care.
ETHICAL ISSUES
The promise of NIPT is that it will reduce the number of
women with an indication for invasive testing and only
those women who are at very high risk for dened disor-
ders such as Down syndrome will then need to deal with
the ethical challenges associated with procedure-related
loss and pregnancy termination. However, the ease of
providing NIPT will potentially result in substantially
more prenatal diagnoses and terminations of affected
pregnancies
18
. The issue is confounded by an inability to
provide adequate counseling for all women who might be
considering NIPT
147
. There may also be a shift in patients
and/or healthcare providers views on prenatal diagnosis,
with a greater expectation for pregnancy termination
as the usual response to an abnormal diagnosis. This
has been referred to as normalization of testing and
termination
148
. Potential consequences could include a
sharply reduced incidence of Down syndrome neonates,
greater emphasis on prevention and less on support and
accommodation of their disabilities, and altered attitudes
to individuals with Down syndrome and their parents
who chose not to test or terminate an affected pregnancy.
The currently available NIPT for aneuploidy is the
consequence of a massive investment by private companies
and venture capital, and so far is only provided by private,
for-prot companies. Aspects of the testing are the subject
of contested patents and other proprietary ownership.
Claims of exclusive intellectual property ownership can
inate costs, restrict access to test protocols and limit
further test development. Although there is currently
no direct-to-consumer testing, there are press releases,
patient information materials and extensive marketing to
healthcare providers, many of whom may have a poor
understanding of the true performance of the tests.
In the future, NIPT may also be encouraged by govern-
ments or insurance carriers who view genetic disorders as
an economic burden
149
. The concern is that these nancial
pressures will be the dominating factor in determining
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
NIPT for aneuploidy 29
who has access to testing, without adequately considering
medical need. Changes in nancial support for testing,
counseling and clinical services can also indirectly
compromise individual patient autonomy when deciding
whether to continue or terminate an affected pregnancy.
It is also worth noting that the early history of NIPT
shows how easily test quality can be compromised when
development and introduction is left entirely to unregu-
lated free market forces
150,151
. ISPD, in its latest Position
Statement on NIPT has strongly emphasized the impor-
tance of adoption of laboratory standards for NIPT
146
.
DISCLOSURES
H. Cuckle is a consultant to PerkinElmer Inc, Ariosa
Diagnostics Inc, Natera Inc, and director of Genome Ltd;
E. Pergament is a consultant to PerkinElmer Inc and
Natera Inc, and director of Northwestern Reproductive
Genetics Inc.
REFERENCES
1. Ferguson-Smith MA, Yates JRW. Maternal age specic
rates for chromosome aberrations and factors inuencing
them: report of a collaborative European study on 52965
amniocenteses. Prenat Diagn 1984; 4 Spec No: 544.
2. Tabor A, Alrevic Z. Update on procedure-related risks for
prenatal diagnosis techniques. Fetal Diagn Ther 2010; 27:
17.
3. Benn PA, Egan JF, Fang M, Smith-Bindman R. Changes in the
utilization of prenatal diagnosis. Obstet Gynecol 2004; 103:
12551260.
4. Cuckle H, Benn P. Multianalyte maternal serum screening
for chromosomal defects. In Genetic Disorders and the Fetus:
Diagnosis, Prevention and Treatment (6
th
edn), Milunsky A,
Milunsky JM (eds). Wiley-Blackwell, Chichester, UK, 2010;
771818.
5. Syngelaki A, Chelemen T, Dagklis T, Allan L, Nicolaides
KH. Challenges in the diagnosis of fetal non-chromosomal
abnormalities at 1113weeks. Prenat Diagn 2011; 31:
90102.
6. Lo YMD, Corbetta N, Chamberlain PF, Rai V, Sargent IL,
Redman CWG, Wainscoat JS. Presence of fetal DNA in
maternal plasma and serum. Lancet 1997; 350: 485487.
7. Lo YMD, Tein MS, Lau TK, Haines CJ, Leung TN, Poon
PM, Wainscoat JS, Johnson PJ, Chang AM, Hjelm NM.
Quantitative analysis of fetal DNA in maternal plasma and
serum: implications for noninvasive prenatal diagnosis. Am J
Hum Genet 1998; 62: 768775.
8. Lun FM, Chiu RW, Allen Chan KC, Yeung Leung T, Kin Lau
T, Dennis Lo YM. Microuidics digital PCR reveals a higher
than expected fraction of fetal DNA in maternal plasma. Clin
Chem 2008; 54: 16641672.
9. Illanes S, Denbow M, Kailasam C, Finning K, Soothill PW.
Early detection of cell-free fetal DNA in maternal plasma.
Early Hum Dev 2007; 83: 563566.
10. Chan KC, Zhang J, Hui AB, Wong N, Lau TK, Leung TN, Lo
KW, Huang DW, Lo YM. Size distributions of maternal and
fetal DNA in maternal plasma. Clin Chem 2004; 50: 8892.
11. Li Y, Zimmermann B, Rusterholz C, Kang A, Holzgreve
W, Hahn S. Size separation of circulatory DNA in maternal
plasma permits ready detection of fetal DNA polymorphisms.
Clin Chem 2004; 50: 10021011.
12. Lo YMD, Zhang J, Leung TN, Lau TK, Chang AM, Hjelm
NM. Rapid clearance of fetal DNA from maternal plasma.
Am J Hum Genet 1999; 64: 218224.
13. Smid M, Galbiati S, Vassallo A, Gambini D, Ferrari A, Viora
E, Pagliano M, Restagno G, Ferrari M, Cremonesi L. No
evidence of fetal DNA persistence in maternal plasma after
pregnancy. Hum Genet 2003; 112: 617618.
14. Hui L, Vaughan JI, Nelson M. Effect of labor on postpartum
clearance of cell-free fetal DNA from the maternal circulation.
Prenat Diagn 2008; 28: 304308.
15. Poon LL, Leung TN, Lau TK, Lo YM. Presence of fetal RNA
in maternal plasma. Clin Chem 2000; 46: 18321834.
16. Ng EK, Tsui NB, Lam NY, Chiu RW, Yu SC, Wong SC, Lo
ES, Rainer TH, Johnson PJ, Lo YM. Presence of lterable
and nonlterable mRNA in the plasma of cancer patients and
healthy individuals. Clin Chem 2002; 48: 12121217.
17. Ng EK, Tsui NB, Lau TK, Leung TN, Chiu RW, Panesar
NS, Lit LC, Chan KW, Lo YM. mRNA of placental origin is
readily detectable in maternal plasma. Proc Natl Acad Sci U S
A 2003; 100: 47484753.
18. Norton ME, Rose NC, Benn P. Noninvasive prenatal testing
for fetal aneuploidy: clinical assessment and a plea for
restraint. Obstet Gynecol 2013; 121: 847850.
19. Alamillo CML, Krantz D, Evans M, Fiddler M, Pergament
E. Nearly a third of abnormalities found after rst-trimester
screening are different than expected: 10-year experience from
a single center. Prenat Diagn 2013; 33: 251256.
20. Kazerouni NN, Currier RJ, Flessel M, Goldman S, Hennigan
C, Hodgkinson C, Lorey F, MalmL, Tempelis C, Roberson M.
Detection rate of quadruple-marker screening determined by
clinical follow-up and registry data in the statewide California
program, July 2007 to February 2009. Prenat Diagn 2011;
31: 901906.
21. Hult en MA, Dhanjal S, Pertl B. Rapid and simple prenatal
diagnosis of common chromosome disorders: advantages and
disadvantages of the molecular methods FISH and QF-PCR.
Reproduction 2003; 126: 279297.
22. Benn PA. Prenatal diagnosis of chromosome abnormalities
through amniocentesis. In Genetic Disorders and the Fetus:
Diagnosis, Prevention and Treatment (6
th
edn), Milunsky A,
Milunsky JM (eds). Wiley-Blackwell: Chichester, UK, 2010:
771818.
23. Ledbetter DH, Zachary JM, Simpson JL, Golbus MS,
Pergament E, Jackson L, Mahoney MJ, Desnick RJ, Schulman
J, Copeland KL, Verlinsky Y, Yang-Feng T, Schonberg SA,
Babu A, Tharapel A, Dorfmann A, Lubs HA, Rhoads GG,
Fowler SE, De La Cruz F. Cytogenetic results from the
U.S. Collaborative Study on CVS. Prenat Diagn 1992; 12:
317345.
24. Evans MI, Henry GP, Miller WA, Bui TH, Snidjers RJ,
Wapner RJ, Miny P, Johnson MP, Peakman D, Johnson A,
Nicolaides K, Holzgreve W, Ebrahim SA, Babu R, Jackson
L. International, collaborative assessment of 146,000 prenatal
karyotypes: expected limitations if only chromosome-specic
probes and uorescent in-situ hybridization are used. Hum
Reprod 1999; 14: 12131216.
25. Vialard F, Simoni G, Aboura A, De Toffol S, Molina Gomes
D, Marcato L, Serero S, Clement P, Bouhanna P, Rouleau E,
Grimi B, Selva J, Gaetani E, Maggi F, Joseph A, Benzacken
B, Grati FR. Prenatal BACs-on-Beads: a new technology for
rapid detection of aneuploidies and microdeletions in prenatal
diagnosis. Prenat Diagn 2011; 31: 500508.
26. Mann K, Hills A, Donaghue C, Thomas H, Ogilvie CM.
Quantitative uorescence PCR analysis of >40,000 prenatal
samples for the rapid diagnosis of trisomies 13, 18 and 21 and
monosomy X. Prenat Diagn 2012; 32: 11971204.
27. Gerdes T, Kirchhoff M, Lind AM, Vestergaard Larsen G,
Kjaergaard S. Multiplex ligation-dependent probe amplica-
tion (MLPA) in prenatal diagnosis-experience of a large series
of rapid testing for aneuploidy of chromosomes 13, 18, 21, X,
and Y. Prenat Diagn 2008; 28: 11191125.
28. Caine A, Maltby AE, Parkin CA, Waters JJ, Crolla JA;
UK Association of Clinical Cytogeneticists (ACC). Prenatal
detection of Downs syndrome by rapid aneuploidy testing for
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
30 Benn et al.
chromosomes 13, 18, and 21 by FISH or PCR without a full
karyotype: a cytogenetic risk assessment. Lancet 2005; 366:
123128.
29. Test and Technology Transfer Committee. Technical and
clinical assessment of uorescence in situ hybridization: An
ACMG/ASHGposition statement. I. Technical considerations.
Genet Med 2000; 2: 356361.
30. Wapner RJ, Martin CL, Levy B, Ballif BC, Eng CM, Zachary
JM, Savage M, Platt LD, Saltzman D, Grobman WA, Klugman
S, Scholl T, Simpson JL, McCall K, Aggarwal VS, Bunke
B, Nahum O, Patel A, Lamb AN, Thom EA, Beaudet
AL, Ledbetter DH, Shaffer LG, Jackson L. Chromosomal
microarray versus karyotyping for prenatal diagnosis. N Engl
J Med 2012; 367: 21752184.
31. Tabor A, Madsen M, Obel EB, Philip J, Bang J, Norgaard-
Pedersen B. Randomised controlled trial of genetic amniocen-
tesis in 4606 low-risk women. Lancet 1986; i: 12871292.
32. Alrevic Z, Mujezinovic F, Sundberg K, Brigham S.
Amniocentesis and chorionic villus sampling for prenatal
diagnosis. Cochrane Database System Rev 2003; Issue 3:
CD003252.
33. Dhallan R, Au WC, Mattagajasingh S, Emche S, Bayliss P,
Damewood M, Cronin M, Chou V, Mohr M. Methods to
increase the percentage of free fetal DNA recovered from the
maternal circulation. JAMA 2004; 291: 11141119.
34. Go AT, van Vugt JM, Oudejans CB. Non-invasive aneuploidy
detection using free fetal DNA and RNA in maternal plasma:
recent progress and future possibilities. Hum Reprod Update
2011; 17: 372382.
35. Papageorgiou EA, Fiegler H, Rakyan V, Beck S, Hulten M,
Lamnissou K, Carter NP, Patsalis PC. Sites of differential DNA
methylation between placenta and peripheral blood: molecular
markers for noninvasive prenatal diagnosis of aneuploidies.
Am J Pathol 2009; 174: 16091618.
36. Fan HC, Blumenfeld YJ, Chitkara U, Hudgins L, Quake
SR. Noninvasive diagnosis of fetal aneuploidy by shotgun
sequencing DNA from maternal blood. Proc Natl Acad Sci U
S A 2008; 105: 1626616271.
37. Chiu RW, Chan KC, Gao Y, Lau VY, Zheng W, Leung TY, Foo
CH, Xie B, Tsui NB, Lun FM, Zee BC, Lau TK, Cantor CR,
Lo YM. Noninvasive prenatal diagnosis of fetal chromosomal
aneuploidy by massively parallel genomic sequencing of DNA
in maternal plasma. Proc Natl Acad Sci U S A 2008; 105:
2045820463.
38. Sehnert AJ, Rhees B, Comstock D, Comstock D, de Feo
E, Heilek G, Burke J, Rava RP. Optimal detection of
fetal chromosomal abnormalities by massively parallel DNA
sequencing of cell-free fetal DNA from maternal blood. Clin
Chem 2011; 57: 10421049.
39. Dan S, Wang W, Ren J, Li Y, Hu H, Xu Z, Lau TK, Xie J,
Zhao W, Huang H, Xie J, Sun L, Zhang X, Wang W, Liao S,
Qiang R, Cao J, Zhang Q, Zhou Y, Zhu H, Zhong M, Guo
Y, Lin L, Gao Z, Yao H, Zhang H, Zhao L, Jiang F, Chen F,
Jiang H, Li S, Li Y, Wang J, Wang J, Duan T, Su Y, Zhang
X. Clinical application of massively parallel sequencing-based
prenatal noninvasive fetal trisomy test for trisomies 21 and 18
in 11,105 pregnancies with mixed risk factors. Prenat Diagn
2012; 32: 12251232.
40. Fan HC, Quake SR. Sensitivity of noninvasive prenatal
detection of fetal aneuploidy from maternal plasma using
shotgun sequencing is limited only by counting statistics. PLoS
One 2010; 5: e10439.
41. Liang D, Lv W, Wang H, Xu L, Liu J, Li H, Hu L, Peng Y, Wu
L. Non-invasive prenatal testing of fetal whole chromosome
aneuploidy by massively parallel sequencing. Prenat Diagn
2013; 33: 409415.
42. Chiu RW, Sun H, Akolekar R, Clouser C, Lee C, McKernan
K, Zhou D, Nicolaides KH, Lo YM. Maternal plasma DNA
analysis with massively parallel sequencing by ligation for
noninvasive prenatal diagnosis of trisomy 21. Clin Chem
2010; 56: 459463.
43. Faas BH, de Ligt J, Janssen I, Eggink AJ, Wijnberger LD,
van Vugt JM, Vissers L, Geurts van Kessel A. Non-invasive
prenatal diagnosis of fetal aneuploidies using massively parallel
sequencing-by-ligation and evidence that cell-free fetal DNA
in the maternal plasma originates fromcytotrophoblastic cells.
Expert Opin Biol Ther 2012; 12 (Suppl 1): S1926.
44. van den Oever JM, Balkassmi S, Verweij EJ, van Iterson M,
Adama van Scheltema PN, Oepkes D, van Lith JM, Hoffer
MJ, den Dunnen JT, Bakker E, Boon EM. Single molecule
sequencing of free DNAfrommaternal plasma for noninvasive
trisomy 21 detection. Clin Chem 2012; 58: 699706.
45. Sparks AB, Wang ET, Struble CA, Barrett W, Stokowski R,
McBride C, Zahn J, Lee K, Shen N, Doshi J, Sun M, Garrison J,
Sandler J, Hollemon D, Pattee P, Tomita-Mitchell A, Mitchell
M, Stuelpnagel J, Song K, Oliphant A. Selective analysis of cell-
free DNA in maternal blood for evaluation of fetal trisomy.
Prenat Diagn 2012; 32: 39.
46. Dhallan R, Guo X, Emche S, Damewood M, Bayliss P, Cronin
M, Barry J, Betz J, Franz K, Gold K, Vallecillo B, Varney J.
A non-invasive test for prenatal diagnosis based on fetal DNA
present in maternal blood: a preliminary study. Lancet 2007;
369: 474481.
47. Ghanta S, Mitchell ME, Ames M, Hidestrand M, Simpson
P, Goetsch M, Thilly WG, Struble CA, Tomita-Mitchell A.
Non-invasive prenatal detection of trisomy 21 using tandem
single nucleotide polymorphisms. PLoS One 2010; 5: e13184.
48. Zimmermann B, Hill M, Gemelos G, Demko Z, Banjevic M,
Baner J, Ryan A, Sigurjonsson S, Chopra N, Dodd M, Levy
B, Rabinowitz M. Noninvasive prenatal aneuploidy testing of
chromosomes 13, 18, 21, X, and Y, using targeted sequencing
of polymorphic loci. Prenat Diagn 2012; 32: 12331241.
49. Liao GJ, Chan KC, Jiang P, Sun H, Leung TY, Chiu RW,
Lo YM. Noninvasive prenatal diagnosis of fetal trisomy
21 by allelic ratio analysis using targeted massively parallel
sequencing of maternal plasma DNA. PLoS One 2012; 7:
e38154.
50. Poon LLM, Leung TN, Lau TK, Chow KC, Lo YMD.
Differential DNA methylation between fetus and mother as
a strategy for detecting fetal DNA in maternal plasma. Clin
Chem 2002; 48: 3541.
51. Tong YK, Ding C, Chiu RW, Gerovassili A, Chim SS, Leung
TY, Leung TN, Lau TK, Nicolaides KH, Lo YM. Noninvasive
prenatal detection of fetal trisomy 18 by epigenetic allelic
ratio analysis in maternal plasma: theoretical and empirical
considerations. Clin Chem 2006; 52: 21942202.
52. Tong YK, Jin S, Chiu RW, Ding C, Chan KC, Leung TY, Yu L,
Lau TK, Lo YM. Noninvasive prenatal detection of trisomy 21
by an epigenetic-genetic chromosome-dosage approach. Clin
Chem 2010; 56: 9098.
53. Chim SS, Jin S, Lee TY, Lun FM, Lee WS, Chan LY, Jin Y,
Yang N, Tong YK, Leung TY, Lau TK, Ding C, Chiu RW,
Lo YM. Systematic search for placental DNA-methylation
markers on chromosome 21: toward a maternal plasma-based
epigenetic test for fetal trisomy 21. Clin Chem 2008; 54:
500511.
54. Papageorgiou EA, Karagrigoriou A, Tsaliki E, Velissariou V,
Carter NP, Patsalis PC. Fetal-specic DNA methylation ratio
permits noninvasive prenatal diagnosis of trisomy 21. Nat
Med 2011; 17: 510513.
55. Tsaliki E, Papageorgiou EA, Spyrou C, Koumbaris G,
Kypri E, Kyriakou S, Sotiriou C, Touvana E, Keravnou A,
Karagrigoriou A, Lamnissou K, Velissariou V, Patsalis PC.
MeDIP real-time qPCR of maternal peripheral blood reliably
identies trisomy 21. Prenat Diagn 2012; 32: 9961001.
56. Fan HC, Quake SR. Detection of aneuploidy with digital
polymerase chain reaction. Anal Chem 2007; 79: 75767579.
57. Lo YM, Lun FM, Chan KC, Tsui NB, Chong KC, Lau TK,
Leung TY, Zee BC, Cantor CR, Chiu RW. Digital PCR for the
molecular detection of fetal chromosomal aneuploidy. Proc
Natl Acad Sci U S A 2007; 104: 1311613121.
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
NIPT for aneuploidy 31
58. Fan HC, Blumenfeld YJ, El-Sayed YY, Chueh J, Quake SR.
Microuidic digital PCR enables rapid prenatal diagnosis of
fetal aneuploidy. Am J Obstet Gynecol 2009; 200: 543.e17.
59. Evans MI, Wright DA, Pergament E, Cuckle HS, Nicolaides
KH. Digital PCR for noninvasive detection of aneuploidy:
power analysis equations for feasibility. Fetal Diagn Ther
2012; 31: 244247.
60. Lo YM, Tsui NB, Chiu RW, Lau TK, Leung TN, Heung
MM, Gerovassili A, Jin Y, Nicolaides KH, Cantor CR, Ding
C. Plasma placental RNA allelic ratio permits noninvasive
prenatal chromosomal aneuploidy detection. Nat Med 2007;
13: 218223.
61. Go AT, Visser A, van Dijk M, Mulders MA, Eijk P, Ylstra
B, Blankenstein MA, van Vugt JM, Oudejans CB. A novel
method to identify syncytiotrophoblast-derived RNAproducts
representative of trisomy 21 placental RNA in maternal
plasma. Methods Mol Biol 2008; 444: 291302.
62. Chen EZ, Chiu RW, Sun H, Akolekar R, Chan KC, Leung
TY, Jiang P, Zheng YW, Lun FM, Chan LY, Jin Y, Go AT,
Lau ET, To WW, Leung WC, Tang RY, Au-Yeung SK, Lam
H, Kung YY, Zhang X, van Vugt JM, Minekawa R, Tang
MH, Wang J, Oudejans CB, Lau TK, Nicolaides KH, Lo YM.
Noninvasive prenatal diagnosis of fetal trisomy 18 and trisomy
13 by maternal plasma DNA sequencing. PLoS One 2011; 6:
e21791.
63. Lau TK, Chen F, Pan X, Pooh RK, Jiang F, Li Y, Jiang H,
Li X, Chen S, Zhang X. Noninvasive prenatal diagnosis of
common fetal chromosomal aneuploidies by maternal plasma
DNA sequencing. J Matern Fetal Neonatal Med 2012; 25:
13701374.
64. Guex N, Iseli C, Syngelaki A, Deluen C, Pescia G, Nicolaides
KH, Xenarios I, Conrad B. A robust second-generation
genome-wide test for fetal aneuploidy based on shotgun
sequencing cell-free DNA in maternal blood. Prenat Diagn
2013. DOI: 10.1002/pd.4130. [Epub ahead of print].
65. Chiu RW, Akolekar R, Zheng YW, Leung TY, Sun H, Chan
KC, Lun FM, Go AT, Lau ET, To WW, Leung WC, Tang
RY, Au-Yeung SK, Lam H, Kung YY, Zhang X, van Vugt
JM, Minekawa R, Tang MH, Wang J, Oudejans CB, Lau TK,
Nicolaides KH, Lo YM. Non-invasive prenatal assessment of
trisomy 21 by multiplexed maternal plasma DNA sequencing:
large scale validation study. BMJ 2011; 342: c7401.
66. Ehrich M, Deciu C, Zweifellhofer T, Tynan JA, Cagasan L,
Tim R, Lu V, McCullough R, McCarthy E, Nygren AO, Dean
J, Tang L, Hutchison D, Lu T, Wang H, Angkachatchai
V, Oeth P, Cantor CR, Bombard A, van den Boom D.
Noninvasive detection of fetal trisomy 21 by sequencing of
DNA in maternal blood: a study in a clinical setting. Am J
Obstet Gynecol 2011; 204; 205.e201211.
67. Palomaki GE, Kloza EM, Lambert-Messerlian GM, Haddow
JE, Neveux LM, Ehrich M, van den Boom D, Bombard AT,
Deciu C, Grody WW, Nelson SF, Canick JA. DNA sequencing
of maternal plasma to detect Down syndrome: an international
clinical validation study. Genet Med 2011; 13: 913920.
68. Bianchi DW, Platt LD, Goldberg JD, Abuhamad AZ,
Sehnert AJ, Rava RP, on behalf of the MatErnal BLood IS
Source to Accurately diagnose fetal aneuploidy (MELISSA)
Study Group. Genome-wide fetal aneuploidy detection by
maternal plasma DNA sequencing. Obstet Gynecol 2012;
119; 890901.
69. Ashoor G, Syngelaki A, Wagner M, Birdir C, Nicolaides KH.
Chromosome-selective sequencing of maternal plasma cell-free
DNA for rst-trimester detection of trisomy 21 and trisomy
18. Am J Obstet Gynecol 2012; 206: 322.e15.
70. Sparks AB, Struble CA, Wang ET, Song K, Oliphant A.
Noninvasive prenatal detection and selective analysis of cell-
free DNA obtained from maternal blood: evaluation for
trisomy 21 and trisomy 18. Am J Obstet Gynecol 2012;
206: 319.e19.
71. Norton ME, Brar H, Weiss J, Karimi A, Laurent LC, Caughey
AB, Rodriguez MH, Williams J 3rd, Mitchell ME, Adair CD,
Lee H, Jacobsson B, Tomlinson MW, Oepkes D, Hollemon D,
Sparks AB, Oliphant A, Song K. Non-Invasive Chromosomal
Evaluation (NICE) Study: results of a multicenter prospective
cohort study for detection of fetal trisomy 21 and trisomy 18.
Am J Obstet Gynecol 2012; 207: 137.e18.
72. Palomaki GE, Deciu C, Kloza EM, Lambert-Messerlian GM,
Haddow JE, Neveux LM, Ehrich M, van den Boom D,
Bombard AT, Grody WW, Nelson SF, Canick JA. DNA
sequencing of maternal plasma reliably identies trisomy 18
and trisomy 13 as well as Down syndrome: an international
collaborative study. Genet Med 2012; 14: 296305.
73. Lau TK, Chan MK, Salome Lo PS, Chan HY, Chan WS,
Koo TY, NgHY, Pooh RK. Clinical utility of noninvasive
fetal trisomy (NIFTY) testearly experience. J Matern Fetal
Neonatal Med 2012; 25: 18561859.
74. Nicolaides KH, Syngelaki A, Ashoor G, Birdir C, Touzet G.
Noninvasive prenatal testing for fetal trisomies in a routinely
screened rst-trimester population. Am J Obstet Gynecol
2012; 207: 374.e16.
75. Ashoor G, Syngelaki A, Wang E, Struble C, Oliphant A,
Song K, Nicolaides KH. Trisomy 13 detection in the rst
trimester of pregnancy using a chromosome-selective cell-free
DNA analysis method. Ultrasound Obstet Gynecol 2013; 41:
2125.
76. Benn P, Cuckle H, Pergament E. Genome-wide fetal
aneuploidy detection by maternal plasma DNA sequencing.
Obstet Gynecol 2012; 119: 1270.
77. Futch T, Spinosa J, Bhatt S, de Feo E,Rava RP, Sehnert AJ.
Initial clinical laboratory experience in noninvasive prenatal
testing for fetal aneuploidy from maternal plasma DNA
samples. Prenat Diagn 2013; 33: 569574.
78. Gil MM, Quezada MS, Bregant B, Ferraro M, Nicolaides KH.
Implementation of maternal blood cell-free DNA testing in
early screening for aneuploidies. Ultrasound Obstet Gynecol
2013; 42: 3440.
79. Ashoor G, Poon L, Syngelaki A, Mosimann B, Nicolaides
KH. Fetal fraction in maternal plasma cell-free DNA at
1113weeks gestation: effect of maternal and fetal factors.
Fetal Diagn Ther 2012; 31: 237243.
80. Wegrzyn P, Faro C, Falcon O, Peralta CF, Nicolaides KH.
Placental volume measured by three-dimensional ultrasound
at 11 to 13+6weeks of gestation: relation to chromosomal
defects. Ultrasound Obstet Gynecol 2005; 26: 2832.
81. Brar H, Wang E, Struble C, Musci T, Norton M. The fetal
fraction of cell-free DNA in maternal plasma is not affected by
a priori risk of fetal trisomy. J Met Fetal Neonat Med 2013;
26: 143145.
82. Wang E, Batey A, Struble C, Musci T, Song K, Oliphant
A. Gestational age and maternal weight effects on fetal cell-
free DNA in maternal plasma. Prenat Diagn 2013. DOI:
10.1002/pd.4119. [Epub ahead of print].
83. Nicolaides KH, Syngelaki A, Gil M, Atanasova V, Markova
D. Validation of targeted sequencing of single-nucleotide poly-
morphisms for non-invasive prenatal detection of aneuploidy
of chromosomes 13, 18, 21, X, and Y. Prenat Diagn 2013;
33: 575579.
84. Rabinowitz M, Gemelos G, Hill M, Levy B, McAdoo S, Savage
M, Demko Z. Abstract 3018T. Highly accurate non-invasive
detection of fetal aneuploidy of chromosomes 13, 18, 21, X
and Y by targeted sequencing. American Society of Human
Genetics Annual Meeting. November 2012. San Francisco.
85. Rabinowitz M, Hill M, Demko Z, McAdoo S, Zimmermann
B, Levy B. Abstract 591. Using targeted sequencing of SNPs
to achieve a highly accurate non-invasive detection of fetal
aneuploidy of 13, 18, 21 and sex chromosomes. SMFM
Meeting, 2013.
86. Qu JZ, Leung TY, Jiang P, Liao GJ, Cheng YK, Sun H, Chiu
RW, Chan KC, Lo YM. Noninvasive prenatal determination
of twin zygosity by maternal plasma DNAanalysis. Clin Chem
2013; 59: 427435.
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
32 Benn et al.
87. Leung TY, Qu JZZ, Liao GJW, Jiang P, Cheng YKY, Chan
KCA, Chiu RWK, Lo YMD. Noninvasive twin zygosity
assessment and aneuploidy detection by maternal plasma
DNAsequencing. Prenat Diagn 2013. DOI: 10.1002/pd.4132.
[Epub ahead of print].
88. Canick JA, Kloza, EM, Lambert-Messerlian GM, Haddow JE,
Ehrich M, van den Boom D, Bombard AT, Deciu C, Palomaki
GE. DNA sequencing of maternal plasma to identify Down
syndrome and other trisomies in multiple gestations. Prenat
Diagn 2012; 32: 15.
89. Lau TK, Jiang F, Chan MK, Zhang H, Salome Lo PS, Wang
W. Non-invasive prenatal screening of fetal Down syndrome
by maternal plasma DNA sequencing in twin pregnancies. J
Matern Fetal Neonatal Med 2013; 26: 434437.
90. Cuckle H, Benn P, Pergament E. Maternal cfDNAscreening for
Downs syndromea cost sensitivity analysis. Prenat Diagn
2013. DOI: 10.1002/pd.4157. [Epub ahead of print].
91. Waitzman NJ, Romano PS, Schefer RM. Estimates of the
economic costs of birth defects. Inquiry 1994; 31: 188205.
92. United States Bureau of Labor Statistics. CPI Detailed Report,
Data for January 2011. http: //www.bls.gov/cpi/cpi_dr.htm
[Accessed 8 April 2013].
93. Ball RH, Caughey AB, Malone FD, Nyberg DA, Comstock
CH, Saade GR, Berkowitz RL, Gross SJ, Dugoff L, Craigo
SD, Timor-Tritsch IE, Carr SR, Wolfe HM, Emig D, DAlton
ME; First and Second Trimester Evaluation of Risk (FASTER)
Research Consortium. First- and second-trimester evaluation
of risk for Down syndrome. Obstet Gynecol 2007; 110:
1017.
94. Caughey AB, Kaimal AJ, Odibo AO. Cost-effectiveness of
Down syndrome screening paradigms. Clin Lab Med 2010;
30: 629642.
95. Nicolaides KH, Wright D, Poon LC, Syngelaki A, Gil
MM. First-trimester contingent screening for trisomy 21
by biomarkers and maternal blood cell-free DNA testing.
Ultrasound Obstet Gynecol 2013; 42: 4150.
96. Gareld SS, Armstrong SO. Clinical and cost consequences
of incorporating a novel non-invasive prenatal test into the
diagnostic pathway for fetal trisomies. J Managed Care Med
2012; 15: 3441.
97. Song K, Musci T, Caughey AB. Clinical utility and cost of
non-invasive prenatal testing with cfDNA analysis in high risk
women based on a U.S. population. J Matern Fetal Neonatal
Med 2013 [Epub ahead of print].
98. Forabosco A, Percesepe A, Santucci S. Incidence of non-age-
dependent chromosomal abnormalities: a population-based
study on 88965 amniocenteses. Eur J Human Genet 2009; 17:
897903.
99. Oeth P, Mazloom A, Wang T, Palomaki GE, Canick
JA, Bombard A, van den Boom D, Ehrich M, Deciu C.
Abstract 3016T. Non-invasive prenatal sex determination
using massively parallel sequencing in samples from a large
clinical validation study. American Society of Human Genetics
Annual Meeting. November 2012.
100. Mazloom AR, D zakula Z, Wang H, Oeth P, Jensen T, Tynan
J, McCullough R,Saldivar J-S, Ehrich M, van den Boom
D, Bombard AT, Maeder M, McLennan G, Meschino W,
Palomaki GE, A. Canick JA, Deciu C. Noninvasive prenatal
detection of sex chromosomal aneuploidies by sequencing
circulating cell-free DNA frommaternal plasma. Prenat Diagn
2013; 33: 591597.
101. Devaney SA, Palomaki GE, Scott JA, Bianchi DW. Noninva-
sive fetal sex determination using cell-free fetal DNA. JAMA
2011; 306: 627636.
102. Tilford CA, Kuroda-Kawaguchi T, Skaletsky H, Rozen S,
Brown LG, Rosenberg M, McPherson JD, Wylie K, Sekhon
M, Kucaba TA, Waterston RH, Page DC. A physical map of
the human Y-chromosome. Nature 2001; 409: 943945.
103. Russell LM, Strike P, Browne CE, Jacobs PA. X chromosome
loss and aging. Cytogenet Genome Res 2007; 116: 181155.
104. Devi AS, Metzger DA, Luciano AA, Benn PA. 45,X/46,XX
mosaicism in patients with idiopathic premature ovarian
failure. Fertil Steril 1998; 70: 8993.
105. Abruzzo MA, Mayer M, Jacobs PA. Aging and aneuploidy:
evidence for the preferential involvement of the inactive X
chromosome. Cytogenet Cell Genet 1985; 39: 275278.
106. Holzgreve W, Schonberg RG, Douglas RG, Golbus MS.
X-chromosome hyperdiploidy in couples with multiple
spontaneous abortions. Obstet Gynecol 1984; 63: 237240.
107. Yoo H, Zhang L, Zhang H, Jiang F, Hu H, Chen F, Jiang
H, Mu F, Zhao L, Liang Z, Wang W. Noninvasive prenatal
genetic testing for fetal aneupliody detects maternal trisomy
X. Prenat Diagn 2012; 32: 11141116.
108. Uematsu A, Yorifuji T, Muroi J, Kawai M, Mamada M, Kaji
M, Yamanaka C, Momoi T, Nakahata T. Parental origin of
normal X chromosomes in Turner syndrome patients with
various karyotypes: implications for the mechanism leading to
generation of a 45,X karyotype. Am J Med Genet 2002; 111:
134139.
109. Thomas NS, Hassold TJ. Aberrant recombination and the
origin of Klinefelter syndrome. Hum Reprod Update 2003; 9:
309317.
110. Hassold T, Hunt P. To err (meiotically) is human; the genesis
of human aneuploidy. Nat Rev Genet 2001; 2: 280291.
111. Canick JA, Palomaki GE, Kloza EM, Lambert-Messerlian
GM, Haddow JE. The impact of maternal plasma DNA fetal
fraction on next generation sequencing tests for common fetal
aneuploidies. Prenat Diagn 2013. DOI: 10.1002/pd.4126.
[Epub ahead of print].
112. Srinivasan A, Bianchi DW, Huang H, Sehnert AJ, Rava RP.
Noninvasive detection of fetal subchromosome abnormalities
via deep sequencing of maternal plasma. Am J Hum Genet
2013; 92: 167176.
113. Hahnemann JM, Vejerslev LO. Accuracy of cytogenetic
ndings on chorionic villus sampling (CVS)-diagnostic
consequences of CVS mosaicism and non-mosaic discrepancy
in centeres contributing to EUCROMIC 19861992. Prenat
Diagn 1997; 17: 801820.
114. Toutain J, Cabeau-Ga uzere C, Barnetche T, Horovitz J, Saura
R. Conned placental mosaicism and pregnancy outcome: a
distinction needs to be made between types 2 and 3. Prenat
Diagn 2010; 30: 11551164.
115. Benn PA. Trisomy 16 and trisomy 16 mosaicism: a review.
Am J Med Genet 1998; 79: 121133.
116. Baffero GM, Somigliana E, Crovetto F, Paffoni A, Persico
N, Guerneri S, Lalatta F, Fogliani R, Fedele L. Conned
placental mosaicism at chorionic villous sampling: risk factors
and pregnancy outcome. Prenat Diagn 2012; 32: 11021108.
117. Kalousek DK, Barrett IJ, McGillivray BC. Placental mosaicism
and interuterine survival of trisomies 13 and 18. Am J Hum
Genet 1989; 44: 338343.
118. Schuring-Blom GH, Boer K, Leschot NJ. A placental diploid
cell line is not essential for ongoing trisomy 13 or 18
pregnancies. Europ J Hum Genet 2001; 9: 286290.
119. Choi H, Lau TK, Jiang FM, Chan MK, Zhang HY, Lo PS, Chen
F, Zhang L, Wang W. Fetal aneuploidy screening by maternal
plasma DNA sequencing: False positive due to conned
placental mosaicism. Prenat Diagn 2013; 33: 198200.
120. Hall AL, Drendel HM, Verbrugge JL, Reese AM, Schumacher
KL, Grifth GB, Weaver DD, Abernathy MP, Litton CG,
Vance GH. Positive cell-free fetal DNA testing for trisomy 13
reveals conned placental mosaicism. Genet Med 2013 DOI:
10.1038/gim.2013.26. [Epub ahead of print].
121. Smith K, Lowther G, Maher E, Hourihan T, Wilkinson
T, Wolstenholme J. The predictive value of ndings of the
common aneuploidies, trisomies 13, 18 and 21, and numerical
sex chromosome abnormalities at CVS: experience from
the ACC U.K. Collaborative Study. Association of Clinical
Cytogeneticists Prenatal Diagnosis Working Party. Prenat
Diagn 1999; 19: 817826.
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.
NIPT for aneuploidy 33
122. Wallerstein R, Yu M-T, Neu RL, Benn P, Bowen CL, Crandell
B, Disteche CM, Donaghue R, Harrison B, Hershey D, Higgins
RR, Jenkins LS, Jackson-Cook C, Keitges E, Khodr GS, Lin
CC, Luthardt FW, Meisner L, Mengden G, Patil S, Rodriguez
M, Sciorra LJ, Shaffer LG, Stetten G, Van Dyke DL, Wang
H, Williams F, Zaslav A-L, Hsu LYF. Common trisomy
mosaicism diagnosed in amniocytes involving chromosomes
13, 18, 20, and 21: karyotype-phenotype correlations. Prenat
Diagn 2000; 20: 103122.
123. Osborne CM, Hardisty E, Devers P, Kaiser-Rogers K, Hayden
MA, Goodnight W, Vora NL. Discordant noninvasive prenatal
testing results in a patient subsequently diagnosed with
metastatic disease. Prenat Diagn 2013; 33: 609611.
124. Lau TK, Jiang FM, Stevenson RJ, Lo TK, Chan LW, Chan MK,
Lo PSS, Wang W, Zhang H-Y, Chen F, Choy KW. Secondary
ndings from non-invasive prenatal testing for common fetal
aneuploidies by whole genome sequencing as a clinical service.
Prenat Diagn 2013; 33: 602608.
125. Yu SCY, Jiang P, Choy KW, Chan KCA, Won H-S, Leung
WC, Lau ET, Tang MHY, Leung TY, Lo YMD. Noninvasive
prenatal molecular karyotyping from maternal plasma. PloS
One 2013; 8: e60968.
126. Chen S, Lau TK, Zhang C, Xu C, Xu Z, Hu P, Xu J, Huang H,
Pan L, Jiang F, Chen F, Pan X, Xie W, Liu P, Li X, Zhang L, Li
S, Li Y, Xu X, Wang W, Wang W, Jiang H, Zhang X. Amethod
for non-invasive detection of fetal large deletions/duplications
by low coverage massively parallel sequencing. Prenat Diagn
2013; 33: 584590.
127. Warburton D. De novo balanced chromosome rearrangements
and extra marker chromosomes identied at prenatal
diagnosis: clinical signicance and distribution of breakpoints.
Am J Hum Genet 1991; 49: 9951013.
128. Schluth-Bolard C, Labalme A, Cordier MP, Till M, Nadeau
G, Tevissen H, Lesca G, Boutry-Kryza N, Rossignol S,
Rocas D, Dubruc E, Edery P, Sanlaville D. Breakpoint
mapping by next generation sequencing reveals causative
gene disruption in patients carrying apparently balanced
chromosome rearrangements with intellectual deciency
and/or congenital malformations. J Med Genet 2013; 50:
144150.
129. Talkowski ME, Ordulu Z, Pillalamarri V, Benson CB,
Blumenthal I, Connolly S, Hanscom C, Hussain N, Pereira
S, Picker J, Rosenfeld JA, Shaffer LG, Wilkins-Haug LE,
Gusella JF, Morton CC. Clinical diagnosis by whole-genome
sequencing of a prenatal sample. N Engl J Med 2012; 367:
22262232.
130. Lo, YM, Chan KC, Sun H, Chen EZ, Jiang P, Lun FM,
Zheng W, Leung TY, Lau TK, Cantor CR, Chiu RW.
Maternal plasma sequencing reveals the genome-wide genetic
and mutational prole of the fetus. Sci Trans Med 2010; 2:
61ra91.
131. Kitzman JO, Snyder MW, Ventura M, Lewis AP, Qiu
R, Simmons LE, Gammill HS, Rubens CE, Santillan DA,
Murray JC, Tabor HK, Bamshad MJ, Eichler EE, Shendure J.
Noninvasive whole-genome sequencing of a human fetus. Sci
Transl Med 2012; 4: 137ra76.
132. Fan HC, Gu W, Wang J, Blumenfeld YJ, El-Sayed YY, Quake
SR. Non-invasive prenatal measurement of the fetal genome.
Nature 2012; 487: 320324.
133. Jacobs PA, Szulman AE, Funkhouser J, Matsuura JS, Wilson
CC. Human triploidy: relationship between parental origin
of the additional haploid complement and development
of partial hydatidiform mole. Ann Hum Genet 1982; 46:
223231.
134. Jauniaux E, Campbell S. Ultrasound assessment of placental
abnormalities. Am J Obstet Gynecol 1990; 163: 16501658.
135. Spencer K, Liao AW, Skentou H, Cicero S, Nicolaides KH.
Screening for triploidy by fetal nuchal translucency and
maternal serum free beta-hCG and PAPP-A at 1014weeks
of gestation. Prenat Diagn 2000; 20: 495499.
136. Benn PA, Gainey A, Ingardia CJ, Rodis JF, Egan JF. Second
trimester maternal serumanalytes in triploid pregnancies: Cor-
relation with phenotype and sex chromosome complement.
Prenat Diagn 2001; 21: 680686.
137. McFadden DE, Kalousek DK. Two different phenotypes of
fetuses with chromosomal triploidy: correlation with parental
origin of the extra haploid set. Am J Med Genet 1991; 38:
535538.
138. Peters D, Chu T, Yatsenko SA, Hendrix N, Hogge WA, Surti
U, Bunce K, Dunkel M, Shaw P, Rajkovic A. Noninvasive
prenatal diagnosis of a fetal microdeletion syndrome. N Engl
J Med 2011; 365: 18471848.
139. Jensen TJ, Dzakula Z, Deciu C, van den Boom D, Ehrich
M. Detection of microdeletion 22q11.2 in a fetus by next-
generation sequencing of maternal plasma. Clin Chem 2012;
58: 11481151.
140. Le Scouarnec S, Gribble SM. Charecterising chromosome
rearragements: recent technical advances in molecular cyto-
genetics. Heredity 2012; 108: 7585.
141. Devers PL, Cronister A, Ormond KE, Facio F, Brasington CK,
Flodman P. Noninvasive prenatal testing/noninvasive prenatal
diagnosis: the position of the National Society of Genetic
Counselors. J Genet Couns 2013; 22: 291295.
142. Noninvasive prenatal testing for fetal aneuploidy. Committee
Opinion No. 545. American College of Obstetricians and
Gynecologists. Obstet Gynecol 2012; 120: 15321534.
143. Langlois S, Brock J-A. Current status in non-invasive prenatal
detection of Down syndrome, trisomy 18, and trisomy 13
using cell-free DNA in maternal plasma. J Obstet Gynaecol
Can 2013; 35: 177181.
144. Wilson KL, Czerwinski JL, Hoskovec JM, Noblin SJ, Sullivan
CM, Harbison A, Campion MW, Devary K, Devers P,
Singletary CN. NSGC practice guideline: prenatal screening
and diagnostic testing options for chromosome aneuploidy. J
Genet Couns 2013; 22: 415.
145. Noninvasive Prenatal Screening Work Group of the American
College of Medical Genetics and Genomics, Gregg AR,
Gross SJ, Best RG, Monaghan KG, Bajaj K, Skotko BG,
Thompson BH, Watson MS. ACMGstatement on noninvasive
prenatal screening for fetal aneuploidy. Genet Med 2013; 15:
395398.
146. Benn P, Borell A, Chiu R, Cuckle H, Dugoff L, Faas B, Gross
S, Johnson J, Maymon R, Norton M, Odibo A, Schielen P,
Spencer K, Huang T, Wright D, Yaron Y. Position Statement
from the Aneuploidy Screening Committee on behalf of the
Board of the International Society for Prenatal Diagnosis.
Prenat Diagn 2013. DOI: 10.1002/pd4139. [Epub ahead of
print].
147. Benn PA, Chapman A. Ethical challenges in providing
nonoinvasive prenatal diagnosis. Curr Opin Obstet Gynecol
2010; 22: 128134.
148. de Jong A, Dondorp WJ, de Die-Smulders CEM, Fritz SGM,
de Wert GMWR. Noninvasive prenatal testing: ethical issues
explored. Europ J Hum Genet 2010; 18: 272277.
149. Chachkin CJ. What potent blood: Non-invasive prenatal
genetic diagnosis and the transformation of modern prenatal
care. Am J Law Med 2007; 33: 953.
150. Bianchi D. At-home fetal DNA gender testing. Obstet Gynecol
2006; 107: 216219.
151. Ray T. Sequenom: Bloodied and Unbowed. Barrons. Sepember
29, 2009. http: //online.barrons.com/article/SB12542362487
5349477.html [Accessed 8 April 2013].
Copyright 2013 ISUOG. Published by John Wiley & Sons Ltd. Ultrasound Obstet Gynecol 2013; 42: 1533.

Das könnte Ihnen auch gefallen