ArticlePDF AvailableLiterature Review

Abstract

Noncoding RNAs (ncRNAs) are transcripts that have no apparent protein-coding capacity; however, many ncRNAs have been found to play a major biological role in human physiology. Their deregulation is implicated in many human diseases, but their exact roles are only beginning to be elucidated. Nevertheless, ncRNAs are extensively studied as a novel source of biomarkers, and the fact that they can be detected in body fluids makes them extremely suitable for this purpose. The authors mainly focus on ncRNAs as biomarkers in cancer, but also touch on other human diseases such as cardiovascular diseases, autoimmune diseases, neurological disorders and infectious diseases. The authors discuss the established methods and provide a selection of emerging new techniques that can be used to detect and quantify ncRNAs. Finally, the authors discuss ncRNAs as a new strategy for therapeutic interventions.
10.1586 /ERM.12.134 183
ISSN 1473-7159
© 2013 Expert Reviews Ltd
www.expert-reviews.com
Review
Loci harboring protein-coding genes are the
best-studied in the human genome, but com-
prise only less than 2% of it [1] . The fact that at
least 90% of the genome is actively transcribed
adds another layer of complexity to the human
transcriptome, suggesting that the majority of
the transcripts are noncoding RNAs (ncRNAs)
[2]. ncRNAs are transcripts that lack an open-
reading frame and, therefore, do not encode pro-
teins. They can be divided into small ncRNAs,
of which microRNAs (miRNAs) are the most
prominent, and long ncRNA, a very heterogene-
ous group of ncRNAs that are more than 200
nucleotides in length and that include, among
others, long intergenic ncRNAs (lincRNAs),
transcribed ultraconserved regions (T-UCRs),
pseudogenes and antisense RNAs (reviewed
by [3]). ncRNAs were originally thought to be
only ‘transcriptional noise’; however, increas-
ing evidence assigns a major biological role in
many physiological processes, and deregulation
of ncRNAs is implicated in many human dis-
eases. Although a lot of effort is being put into
discovering new ncRNAs and elucidating their
roles, both in normal physiology as well as in
disease, most of the research so far has focused
on miRNAs. Unraveling the roles and mecha-
nisms of action of most other ncRNAs is still in
its infancy and remains challenging. Especially,
long ncRNAs form a novel, poorly characterized
class of ncRNAs, and it is still a key challenge to
understand their precise roles in relation to phys-
iology and disease. Nevertheless, the function
of many ncRNAs has been elucidated in recent
years, pointing to a general role in a plethora of
human diseases.
In this article, the authors summarize some
emerging new roles for miRNAs and long ncR-
NAs, and describe mechanisms that cause abnor-
mal expression of ncRNAs. The authors review
established and new methods to detect RNAs,
and discuss the use of ncRNAs as biomarkers
in human diseases, as well as their targeting as
new strategies for therapeutic intervention. This
review mainly concentrates on, but is not only
restricted to, miRNAs and long ncRNAs in can-
cer; the authors also consider other human dis-
orders including cardiovascular, autoimmune,
neurological and infectious diseases. For reviews
concerning the use of miRNAs as biomarkers, the
reader is redirected to some excellent reviews [4–15 ] .
Emerging roles for ncRNAs
miRNAs
Since the discovery of miRNAs in 1993 [16 ,17]
and the demonstration of miRNA involvement
Katrien
Van Roosbroeck1,
Jeroen Pollet2 and
George A Calin*1
1Department of Experimental
Therapeutics, Unit 1950, The University
of Texas MD Anderson Cancer Center,
1881 East Road, Houston, TX 77054,
USA
2Department of Chemical and
Biomolecular Engineering, University of
Houston, Houston, TX 77204, USA
*Author for correspondence:
Tel.: +1 713 792 5461
Fax: +1 713 745 4526
gcalin@mdanderson.org
Noncoding RNAs (ncRNAs) are transcripts that have no apparent protein-coding capacity;
however, many ncRNAs have been found to play a major biological role in human physiology.
Their deregulation is implicated in many human diseases, but their exact roles are only beginning
to be elucidated. Nevertheless, ncRNAs are extensively studied as a novel source of biomarkers,
and the fact that they can be detected in body fluids makes them extremely suitable for this
purpose. The authors mainly focus on ncRNAs as biomarkers in cancer, but also touch on other
human diseases such as cardiovascular diseases, autoimmune diseases, neurological disorders
and infectious diseases. The authors discuss the established methods and provide a selection of
emerging new techniques that can be used to detect and quantify ncRNAs. Finally, the authors
discuss ncRNAs as a new strategy for therapeutic interventions.
Keywor ds: autoimmune diseases • biomarkers • cancer • cardiovascular diseases • infectious diseases • miRNAs
• molecular diagnostics • neurological diseases • noncoding RNAs
miRNAs and long noncoding
RNAs as biomarkers in human
diseases
Expert Rev. Mol. Diagn. 13(2), 183–204 (2013)
Expert Review of Molecular Diagnostics
© 2013 Expert Reviews Ltd
10.1586/ERM.12.134
1473-7159
1744-8352
Review
THEMED ARTICLE Cancer/Oncology Diagnostics
For reprint orders, please contact reprints@expert-reviews.com
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
Expert Rev. Mol. Diagn. 13(2 ), (2013)
184
Review
in human diseases, a lot of research has focused on elucidating
their roles in physiology as well as in human diseases. By now, it is
common knowledge that miRNAs negatively regulate expression
of protein-coding genes at the post-transcriptional and transla-
tional level through mRNA cleavage (Fig ure 1A) or translational
repression (Figu re 1B), and that miRNAs themselves are regulated
by transcription factors. However, recent studies suggest that
miRNAs work in a much more sophisticated way than initially
assumed and that complex interaction networks exist between
miRNAs, other ncRNAs and mRNA from protein-coding
genes [18]. For example, it has been shown that miRNAs and
their associated protein complexes, microribonucleoproteins or
microRNPs, can post-transcriptionally upregulate the expression
of target genes (Figure 1A) (reviewed in [1 9]).
Recent studies indicate that RNAs can influence each other’s lev-
els by competing for a limited pool of miRNAs. These competing
endogenous RNAs (ceRNAs) act as miRNA ‘sponges’ that divert
the miRNAs away from their mRNA targets, which results in
reduced downregulation of the target mRNA and translation into
protein (Figure 1C) [20] . ceRNA functions have been demonstrated for
pseudogenes including the pseudogenes of the PTEN tumor sup-
pressor gene (PTENP1) and the KRAS oncogene (KR AS1P) [21] for
mRNAs of protein-coding genes such as the ZEB2 mRNA, which
regulates PTEN protein levels through competition for at least four
miRNAs [2 2,2 3], and for long ncRNAs [24 ,25] . The muscle-specific
long ncRNA linc-MD1 has been shown to display decoy activity
for miR-133 and miR-135, and in this way is able to regulate the
expression of the MAML1 and MEF2C transcription factors, which
activate muscle-specific gene expression. In addition, linc-MD1
levels are strongly reduced in Duchenne muscular dystrophy
myeloblasts, pointing to a possible involvement of long non coding
ceRNAs in this hereditary disease [24] . Also, pseudogenes and
mRNA ceRNAs are suggested to be involved in the pathogenesis
of cancer. The PTENP1 locus is selectively lost in human cancer
and this loss is correlated with a decrease in PTEN [2 1] .
miRNAs can also act as an RNA decoy to modulate the func-
tion of regulatory proteins (Figur e 1D), as has been demonstrated
for miR-328 [26] . In addition to its gene-silencing activity [27] ,
miR-328, found to be downregulated in chronic myeloid leu-
kemia in blast crisis, has been shown to strongly compete with
CEBPA mRNA for binding to hnRNP E2. Consequently, CEBPA
mRNA is released from hnR NP E2-mediated translational repres-
sion and CEBPA mRNA translation is rescued [2 6]. Conversely, it
is well established that RNA-binding proteins can modulate the
function of miRNAs by competing for miRNA binding sites on
mRNAs [28 ,29].
Finally, miRNAs can also interact with and regulate other
ncR NAs such as T-UCRs [3 0] and noncoding antisense RNAs [31] .
Long ncRNAs
Our understanding of long ncRNAs and their functions is rap-
idly advancing nowadays [32,33] . However, a general mechanism
of action, as there is for miRNAs, has not been described due to
the heterogeneity of long ncRNAs. Long ncRNAs were originally
defined as RNA molecules longer than 200 nucleotides that do
not encode a protein. This cutoff was arbitrarily chosen and was
not based on functionality, but on RNA purification protocols
that exclude small RNAs [34] . Nevertheless, most long ncRNAs
described to date have been found to be related to transcriptional
regulation or mRNA processing, characteristics that they share
with miRNAs. However, unlike miRNAs, long ncRNAs show a
greater complexity of their functions and have a wider spectrum
of biological contexts. The emerging roles for long ncRNA are
nicely reviewed in [35] and will not be extensively discussed in this
review. However, a few functions will be highlighted. Many long
ncRNAs are involved in epigenetic regulation (Figur e 1e), typically
repression of target genes, which is achieved through coupling
with histone-modifying or chromatin-remodeling protein com-
plexes, most commonly PRC1 and PRC2 polycomb repressive
complexes [36] . The long ncRNAs ANRIL [37] , HOTAIR [38], H19
[39] , KCNQ1OT1 [4 0] and XIST [41] have all been linked to the
PRC2 complex. More rarely, long ncRNAs have been found to
be implicated in activation of epigenetic complexes [42] . In addi-
tion, long ncRNAs may regulate gene expression by influenc-
ing the activity of gene enhancers [43], and long ncRNAs with
enhancer-like functions have been revealed [4 4– 46] . Very recently,
enhancer-templated noncoding RNAs (eRNAs) have been discov-
ered, which are part of a highly specialized enhancer network that
regulates gene transcription at enhancer loci (Figu re 1F) [43] . Besides
transcriptional regulation, long ncRNAs are also involved in the
post-transcriptional regulation of genes and translational control.
For instance, the ncRNAs MALAT1 and NEAT1 are thought to
contribute to the regulation of alternative mRNA splicing, edit-
ing and export (Figur e 1g) [47, 48]. Also, antisense long ncRNAs may
regulate translation of their sense-coding transcript by stabiliz-
ing the coding mRNA through duplex formation (Fig ure 1H), as
was demonstrated for the Alzheimer’s disease-associated antisense
long ncRNA β-site amyloid precursor protein cleaving enzyme
(BACE1-AS) [49]. This increased stability of BACE1 mRNA leads
to elevated BACE1 protein levels.
A recent review [5 0] distilled the myriad functions of long ncR-
NAs into four archetypes of molecular mechanisms, which are
not mutually exclusive: signals; decoys, long ncRNAs that can
act as molecular sponges that pull away RNA-binding proteins,
which are themselves transcription factors, chromatin modifiers
or other regulatory proteins; guides, long ncRNAs that can recruit
chromatin-modifying enzymes to target genes, either in cis (near
the site of long ncRNA production) or in trans to distant target
genes and scaffolds. long ncRNAs may operate as ‘ropes’ that keep
together multiple proteins to form ribonucleoprotein complexes,
which may act on chromatin to affect histone modifications, or
that stabilize nuclear structures or signaling complexes.
Mechanisms causing abnormal expression of ncRNAs
Deregulation of ncRNAs is a general event in human diseases
including cancer, heart diseases, autoimmune diseases and neuro-
degenerative disorders [ 7,14,51– 53], and has been studied most inten-
sively for miRNAs in cancer [54] . miRNAs, and recently also
long ncRNAs, have been found to act as tumor-suppressor genes
and/or oncogenes, similar to the protein-coding genes. Several
Van Roosbroeck, Pollet & Calin
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
185
www.expert-reviews.com
Review
Post-transcriptional regulation Translational regulation
Competing endogenous RNA RNA decoy
Epigenetic regulation Gene enhancer regulation
Post-transcriptional regulation (e.g., RNA splicing) Translational regulation
miRNAsLong ncRNAs
miRNA
Cleavage
complex
Cleavage
complex
Cleavage
complex
mRNA target
mRNA cleavage
(repression)
mRNA stabilization
(upregulation)
mRNA target
mRNA target
miRNA
RNA binding
protein
RNA
binding
protein
miRNA
miRNA
AAAA
AAAA
AAAA
Repression
Protein
Activation Protein
mRNA target
AAAA
mRNA
‘sponge’
Protein
Repression
Activation
Intron
retention
Exon
splicing
Activation
Stabilization
mRNA
AS RNA
Degradation
Enhancer
eRNA
Gene
Gene
Gene
H3K27me3
Repressor
complex
Activation
complex
Spliceosome
long ncRNA
long ncRNA
H3K4me3
Protein
Protein
Protein
mRNA target
AAAA
mRNA target
AAAA
long ncRNA
A AB B
CC
AC
RNA binding protein
A
B
CProtein
AAAA 3´
mRNA AAAA 3´
Expert Rev. Mol. Diagn. © Future Science Group (2013)
Figure 1. General mechanisms of miRNA and long noncoding RNA function. (A) miRNAs can post-transcriptionally regulate gene
expression when binding to target mRNA sequences with perfect complementarity. This binding induces cleavage and degradation of the
targeted mRNAs (upper panel). Less frequently, miRNAs have been found to stabilize mRNAs, protecting them from degradation and
resulting in upregulation of gene expression (lower panel). (B) Imperfect binding of a miRNA to its target typically occurs in the 3
untranslated region of the mRNA and leads to translational repression. In rare cases, this imperfect binding leads to the activation of
translation into proteins. (C) RNAs can influence each other’s levels by competing for a limited pool of miRNAs. These miRNA ‘sponges’ can
be pseudogenes, mRNAs of protein-coding genes or long ncRNAs. They divert the miRNA away from its target, which results in reduced
downregulation of the target mRNA and upregulation of protein translation. (D) miRNAs can also act as an RNA decoy to modulate the
function of regulatory proteins. They are able to bind to RNA-binding proteins, inhibiting the interaction of these proteins with their mRNA
targets (upper panel). Conversely, RNA-binding proteins can modulate the function of miRNAs by competing for miRNA binding sites on
mRNAs (lower panel). (E) Many long ncRNAs have been found to be involved in epigenetic regulation, typically repression, but infrequently
also activation, of target genes. This repression or activation is achieved predominantly through coupling with histone modification protein
complexes, resulting in methylation of lysine residues in histones. Depending on which residue is modified, transcriptional repression (upper
panel) or activation (lower panel) may occur. (F) eRNAs are transcribed from enhancer loci and have been found to facilitate gene expression
of protein-coding genes. (G) Long ncRNAs also regulate gene expression at the post-transcriptional level, for example via RNA splicing. When
the long ncRNA masks the splice site from the spliceosome, this results in intron retention (upper panel). Long ncRNAs can also interact with
RNA-binding proteins that are involved in alternative splicing, facilitating mRNA exon splicing (lower panel). (H) Antisense long nRNAs are
able to bind to complementary mRNA sequences, which either protect the mRNA from degradation, resulting in protein expression
(upper panel), or induce cleavage and degradation of the target mRNA (lower panel).
AS RNA: Antisense RNA; ceRNA: Competing endogenous RNA; eRNA: Enhancer-templated noncoding RNA; ncRNA: Noncoding RNA.
miRNAs & long noncoding RNAs as biomarkers in human diseases
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
Expert Rev. Mol. Diagn. 13(2 ), (2013)
186
Review
mechanisms have been described to be responsible for aberrant
expression of ncRNAs in human diseases. Extensive reviews on
this matter have already been published [32, 35,50,51, 54–57] , and the
authors would like to briefly touch on the different causes of
ncRNA deregulation, which will make it easier for the reader to
understand the use of ncRNAs as biomarkers.
First, miRNAs and ncRNAs, such as T-UCRs, are frequently
located at fragile sites and disease-causing genomic loci includ-
ing regions of loss of heterozygosity (copy number loss), minimal
regions of amplification (copy number gain) and common break-
points (rearrangements) [30 ,58– 60]. Second, alterations in the epi-
genetic regulation of ncRNAs are important in the pathogenesis
of human diseases [14 , 61] . This is not surprising as DNA is globally
hypomethylated in cancer [62] , and miRNAs and long ncRNAs
are inactivated by hypermethylation [63–65] or activated by hypo-
methylation [6 6]. Furthermore, in several human diseases, many
long ncRNAs interact with histone-modifying and chromatin-
remodeling complexes (discussed in the previous section), and
miRNAs can target genes important in the epigenetic machinery
leading to states of hyper- or hypo-methylation as well [67, 68].
ncRNAs can also be deregulated at the transcriptional level. Many
miRNAs and other ncRNAs are regulated by transcription factors
such as p53 [69,70] , MYC [71,7 2] or MeCP2 [73] . Although the occur-
rence of sequence variations in ncRNAs is still an emerging field
of research, numerous recent reports point to a link with disease
susceptibility, an association that will be dealt with in the next
section [74–76]. Finally, miRNAs may be deregulated due to aber-
rations of enzymes involved in the miRNA biogenesis pathway,
for example TARBP2, DICER or XPO5 [32 ,77] .
ncRNAs as biomarkers
Cancer is hallmarked by a remarkable heterogeneity within groups
of patients diagnosed with the same tumor subtype, and even within
different cells of the same tumor mass. This heterogeneity, which
is a consequence of the fact that cancer is a genetic disease that is
caused by the accumulation of different genetic and epigenetic aber-
rations, is a huge obstacle in choosing the most effective treatment
for each individual patient. Biomarkers can aid in overcoming these
obstacles as they can be used to establish a diagnosis, classify tumors,
determine the stage of the disease and predict disease outcome and
evolution. Based on the levels of expression, specific biomarkers help
physicians to forecast the response to a specific therapy and to decide
on the optimal dose of a drug for an individual patient [78] . Besides
protein-coding genes and proteins, miRNAs and other ncRNAs
are extensively investigated for their potential use as biomarkers and
targets for therapeutic interventions in human diseases.
Characteristics of a good biomarker
Given that ‘a poor biomarker can be just as bad for the patient
as a bad drug’ [78], it is clinically very important to identify and
develop biomarkers that are as close to the ‘ideal’ biomarker as
possible. Biomarkers should have a unique expression profile in
the disease compared with healthy individuals and should be
able to differentiate pathologies. They should, depending on how
the biomarker is to be used, show highly increased or decreased
expression levels in the diseased organ or tissue compared with
nondiseased organ or tissue, should give a reliable indication of the
disease before clinical symptoms appear, enabling early detection,
or allow detection of disease progression or therapeutic response.
Furthermore, ideal biomarkers are accessible through noninvasive
methods, have a long half-life in the sample and can be rapidly
detected by simple, accurate and inexpensive methods [8].
Although many current biomarkers are based on protein levels,
increasing interest has been shown in recent years for the use
of miRNAs and other ncRNAs as biological indicators for dis-
ease diagnosis, prognosis and prediction to therapeutic responses.
miRNAs are very stable, even in body fluids such as plasma, serum,
urine and saliva [4]; their expression is specific to tissues, organs or
biological stages; and the expression level can be easily measured
by methods such as quantitative PCR and miRNA microarrays
[8]. Long ncRNAs are less stable than miRNAs, but still have half-
lives comparable to that of mRNAs of protein-coding genes [79 ,80] .
They show a greater tissue specificity compared with protein-cod-
ing mRNAs and miRNAs, which are frequently expressed from
multiple tissues, and show highly increased or decreased expression
levels in disease [81–8 3]. Similar to miRNAs, long ncRNAs are also
detectible in body fluids [84– 86] . These features make them very
suitable as biomarkers, and many studies have been published on
this matter in recent years. Here, the authors highlight some of
the miRNAs and long ncRNAs that are used as biomarkers, both
in cancer and in other human diseases such as cardiovascular dis-
eases (CVDs), autoimmune diseases and neuro logical disorders.
A comprehensive list of long ncRNAs that have been or might be
linked to cancer can be found in [87] and [8 8].
ncRNAs in tissue samples & body fluids
As already mentioned earlier in this article, one of the character-
istics of a good biomarker is the accessibility through noninvasive
methods. Thus far, clinical analysis of tumors is mainly based on
molecular analysis of the tumor tissue itself, and these diseased tis-
sues can usually only be obtained through invasive methods such
as surgery and biopsies. This makes biomarkers detected in tumor
tissue samples, although widely used, not the best options for diag-
nostic and prognostic purposes. Recently, circulating RNAs have
been discovered in different kinds of body fluids including blood
(plasma and serum), urine and saliva, of which diagnostic samples
are easy to collect. Therefore, circulating RNAs, and miRNAs in
particular, might be extremely suitable as noninvasive biomarkers
for a plethora of human diseases. However, some level of caution
should be taken into consideration when assessing the usefulness
of circulating RNAs as biomarkers, as recent studies report on
the importance of the origin of biomarkers and their impact on
biomarker specificity. For example, a significant proportion of
miRNAs derived from red and white blood cells have been found
to be present as contaminants in plasma preparation [89] , and differ-
ences in blood cell counts and hemolysis can alter plasma miRNA
biomarker levels by up to 50-fold [9 0]. In addition, inherent dif-
ferences in biological samples and the methods of collecting and
analyzing them can dramatically affect the detection and quanti-
fication of miRNAs and other ncRNAs [91, 92] . Therefore, in order
Van Roosbroeck, Pollet & Calin
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
187
www.expert-reviews.com
Review
to identify true disease-specific circulating RNAs, the approaches
used for quantification of these RNAs should be optimized and
validated for accurate quantification of circulating RNAs [91,92] .
The presence of cancer-specific miRNAs in plasma, serum, spu-
tum and urine of cancer patients is well documented [3, 4,93], despite
the fact that high amounts of ribonucleases circulate in the blood of
cancer patients [88]. These findings suggest that cell-free RNAs are
protected from RNase-mediated degradation, probably either due to
their packaging into microparticles, such as exosomes, microvesicles,
apoptotic bodies and apoptotic microparticles [94], or due to their
stabilization by binding to protein complexes [95–97]. Recently, high-
density lipoproteins were described to transport circulating miRNAs
in plasma [9 8]. The microvesicles and exosomes contain a variety of
proteins and nucleic acids including cytoplasmic and membrane
proteins, mRNAs of protein-coding genes, miRNAs, ncRNAs and
both genomic DNAs and cDNAs [99,10 0]. However, with the excep-
tion of prostate cancer 3 (PCA3) in exosomes isolated from urine
of prostate cancer patients [101] , long ncRNAs have not yet been
observed in microparticles extracted from body fluids. Nevertheless,
long ncRNAs are expressed in different types of body fluids includ-
ing blood [10 2] . For example, Panzitt et al. were able to detect the
highly upregulated in liver cancer ncRNA HULC in the periph-
eral blood cells of hepatocellular carcinoma (HCC) patients and
demonstrated that expression of this long ncRNA is upregulated in
these patients [8 5]. Since the extraction was done on peripheral blood
mononuclear cells and not on exosomes or microvesicles, protection
of the HULC long ncRNA from the RNases in the blood most
likely occurs through other mechanisms, such as binding to protein
complexes, as many long ncRNAs have protein-binding potential.
Disease-associated circulating cell-free RNAs are not restricted
to body fluids of cancer patients, but can also be detected in other
human diseases, including but not limited to CVDs (reviewed by
[5, 6]), kidney disease [1 03 ] , autoimmune diseases such as inflam-
matory bowel disease [10 4 ] , Crohn’s disease [15] , systemic lupus
erythematosus (SLE) [105 ] and multiple sclerosis [1 0 6] , neurological
disorders including Alzheimer’s disease [11] , preeclampsia [107 ] ,
tuberculosis [10 8] , chronic hepatitis C and nonalcoholic fatty liver
disease [1 09 ] , and hepatitis B infection [110 ] .
Despite the increasing number of reports on the presence of
ncRNAs in body fluids, the functions of these cell-free RNAs
remain poorly understood. One of the hypotheses ascribes a
hormone-like effect to circulating miRNAs, in which miRNAs
are released by a donor cell, spread to cells located in other parts
of the organism and affect acceptor cells at these distant sites [4] .
Methods of detecting ncRNAs
To be able to use ncRNAs as biomarkers for human disease, it
is essential to develop accurate determination methods. Several
established techniques are commonly used for the discovery, iden-
tification and detection of ncRNAs (TABle 1). The ideal method
is sensitive enough to provide a quantitative expression analysis,
even with low amounts of starting material; can process multiple
samples in parallel; is specific enough to detect only the ncRNA
of interest without detecting closely related ncRNAs; is easy to
carry out; and does not need expensive reagents or equipment [111].
Northern blotting is one of the gold standards for the detection
of ncRNAs, but is not very suitable for clinical diagnostics as it is
time consuming and low throughput, necessitates relatively large
amounts of RNA and uses radioactivity for detection (TABle 1).
Custom ncRNA microarrays and quantitative reverse transcrip-
tion PCR (qRT-PCR) can be higher throughput, are more sensi-
tive, cheaper, easier to perform and already have some commercial
applications (TABle 1 & Box 1). These techniques, however, all use
ncRNAs extracted from the entire patient sample. It is generally
known that many tumors comprise a mix of normal and malig-
nant cells, and the amount of malignant cells can sometimes be
very low. For example, in Hodgkin’s lymphoma (HL), the number
of malignant Hodgkin’s and Reed Sternberg cells typically com-
prise less than 1% of the tumor mass [1 12 ] . RNA extraction of the
total tumor sample will thus result in a huge ‘contamination’ with
normal, nonmalignant cells and the biomarker that one wants to
measure will be below the detection limit. In situ hybridization
(ISH) overcomes this problem, as the ncRNA is detected within
the tissue, and if combined with immunostaining to identify the
cells of interest, ncRNA expression can be directly assessed in
the neoplastic cells (TABl e 1). In addition, ISH gives information
on the subcellular localization of the ncRNA. The most recently
established method to detect ncRNAs is RNA sequencing. This
technique is highly sensitive, highly specific, can be used for high-
throughput analysis and also discovers ‘de novo’ ncRNAs (TABle 1).
RNA sequencing generates a massive amount of complex data
that need to be analyzed by a well-trained bioinformatician. This,
together with the high costs of a single-RNA-seq run, makes this
method not yet appropriate for diagnostic purposes. However, the
rapidly reducing costs of deep sequencing and the announcement
of the ‘$1,000 genome’, the possibility to sequence a human-sized
human genome in 2 h at a cost of US$1000 [113 ] opens the door for
RNA-seq to be used for biomarker detection in patient samples.
There are many different platforms available to detect and quan-
tify ncRNAs (TABles 1 & 2); however, there are some technical chal-
lenges that make it difficult to correlate measurements performed
on different platforms. An important factor is the different primer
design for the measurements by qRT-PCR and microarrays, the
two most commonly used detection methods [8]. In addition,
many research groups use different protocols for sample prepara-
tion, which makes comparisons between studies, even if the same
platform is used, difficult. Some studies use an enriched small RNA
fraction, whereas others use total RNA; however, the influence of
the fractionation procedure has not yet been systematically inves-
tigated [114 ] . Furthermore, assessment of quantity and quality is
more challenging for miRNAs than for mRNAs of protein-coding
genes, for which sizes and relative abundance of ribosomal RNAs
can be used to check RNA integrity. Finally, technical and biologi-
cal variability has been assessed by a range of different normaliza-
tion processes, mostly based on the expression of reference genes.
Especially, for cell-free miRNA analysis, no known extracellular
reference RNA is currently suitable for a proper normalization [8].
As all established techniques have their own strengths and
weaknesses (TABle 1), there is a continuous search for novel detec-
tion methods that can be used to detect and quantify ncRNAs in
miRNAs & long noncoding RNAs as biomarkers in human diseases
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
Expert Rev. Mol. Diagn. 13(2 ), (2013)
188
Review
patient samples. A selection of emerging new methods, developed
over the last 5–6 years, is given in TABl e 2. This table explains the
mechanism of action of the detection techniques, as well as the
advantages and limitations.
Clinical use of ncRNAs as biomarkers
An illustrative list of long ncRNAs that can be used as biomarkers,
and which will be discussed in the following sections, is shown
in TABle 3.
ncRNAs & disease susceptibility
It has been proposed that variations in the expression and/or
sequences of miRNAs and long ncRNAs, if inherited in the
germline, could be important in disease predisposition [10, 35,115] .
Genome-wide association studies often identify disease-associated
mutations (frequently SNPs) in noncoding, but transcribed, por-
tions of the human genome [116] . These sequence variations rarely
occur within the functional seed sequences of miRNAs [1 17] , but
are observed in primary miRNAs (pri-miRs) and precursor miR-
NAs (pre-miRs), the longer progenitors of the small miRNAs, in
a number of human cancers [118 –1 2 3] , as well as other human dis-
eases [124–127]. SNPs located within or near miRNA binding sites
(e.g., 3 untranslated regions (UTRs) of protein-coding genes)
may also influence tumor susceptibility and serve as potential
biomarkers of disease risk (reviewed by [13, 128]). In addition, SNPs
within genes involved in miRNA biogenesis modify cancer risk
[12 9] and affect tumorigenesis and tumor progression [130 ] .
In this regard, two useful databases have recently been released.
The publicly available miRdSNP database combines PubMed
data on 630 disease-associated SNPs and their association with
miRNA binding sites on 3 UTRs of human genes in 204 diseases
[131] . A catalog of miRNA seed region polymorphisms (miR-seed-
SNPs) consisting of 149 SNPs in six species was generated, and an
online tool for the detection of miRNA polymorphisms (miRNA
SNiPer) in vertebrates was developed [76].
Last, an increasing number of reports describe SNPs within or
near genomic sites where long ncRNAs are located. Two SNPs,
located within the boundaries of UCRs, were identified [132], which
were significantly associated with familial breast cancer, whereas
polymorphisms in the antisense ncRNA ANRIL locus were found
Table 1. Established methods to detect and quantify noncoding RNAs.
Study (year) Method Advantages Limitations Ref.
Lau et al. (2001); Lee et al. (2008) Northern blot Gold standard for ncRNA
detection
Specific
Limited sensitivity (nM)
Low throughput
Time consuming
Limited for quantification
Relatively large amounts of RNA
needed
Radioactivity generally used for
detection
[26 6,2 67]
Gupta and Mo (2011); Hanna
et al. (2012); Kloosterman et al.
(2006)
In situ hybridization Locates miRNA in tissue and
cell compartments
Low throughput
Invasive sample collection
Limited sensitivity
Very limited quantification
[26 8–270 ]
Calin et al. (2004); Lodes et al.
(2009); Roderburg et al. (2012);
Schrauder et al. (2012); Zhao et al.
(2012); Tang et al. (2007)
Microarray Very high throughput In some cases, fair specificity
Medium sensitivity (pM)
Sensitivity and specificity can be
improved by LNA modification of the
probes
Limited for quantification
Results are often validated with
qRT-PCR
[156 , 271–275]
Benes and Castoldi (2010); Chugh
et al. (2010); Jensen et al. (2011);
Bianchi et al. (2012); Heneghan
et al. (2010); Heneghan et al.
(2010); Hu et al. (2012)
qRT-PCR Semi-high throughput
Good quantification
Amplification enables
superior sensitivity (fM)
Difficult to distinguish single-
nucleotide differences
Not for ncRNA discovery
[276 –282]
Hu et al. (2012); Brase et al. (2011) RNA sequencing Most next-generation
platforms offer possibilities
for high throughput
High sensitivity (<fM)
High specificity
Can be used for discovery of
novel ncRNAs
Large amounts of complex data that
need to be analyzed
High costs
[282,2 83]
LNA: Locked nucleic acid; ncRNA: Noncoding RNA; qRT-PCR: Quantitative reverse transcription PCR.
Van Roosbroeck, Pollet & Calin
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
189
www.expert-reviews.com
Review
to be associated with CVDs such as coronary artery disease, diabe-
tes and multiple cancers, including Philadelphia chromosome posi-
tive acute lymphocytic leukemia [133–136]. In addition, expression of
a non-poly(A) circular ncRNA emanating from the ANRIL locus
correlates with enhanced risk of atherosclerosis [13 4] . Very recently,
a genetic variant in HULC has been described that contributes to
risk of hepatitis B virus-related HCC [13 7] . A ‘gene-desert’ region
on chromosome 8q24 upstream of the MYC oncogene harbors a
number of SNPs associated with prostate cancer [138 ,13 9]. In this
region, two long ncRNAs have been identified: PC AT-1 [140 ] and
PRNCR1 [141] . Although the relationship between both long ncR-
NAs and the 8q24 SNPs is not clear at present time, these findings
suggest that so-called ‘gene deserts’ may harbor long ncRNA genes,
and that SNPs in these regions may affect uncovered aspects of
biology [35]. Overexpression of PRINS has been found to correlate
with psoriasis susceptibility [1 4 2] , whereas a SNP in the promoter
of a B-cell specific antisense transcript (S AS- ZFAT ) determines
susceptibility to autoimmune thyroid disease [143].
ncRNA biomarkers in cancer
Cancer diagnosis
Many research groups sought diagnostic miRNA biomarkers that
distinguish cancer patients from cancer-free individuals and that
enable early detection. They typically established expression pro-
files to develop miRNA signatures of various cancer types. For
example, miRNAs signatures have been established that distinguish
between normal subjects and patients with solid tumors such as
lung cancer, breast cancer, gastric cancer, prostate cancer, colorectal
cancer, pancreatic cancer, HCC and glioblastoma, or hematological
malignancies such as chronic lymphocytic leukemia (CLL) and
acute lymphocytic leukemia, chronic myeloid leukemia and acute
myeloid leukemia, follicular lymphoma, diffuse large B-cell lym-
phoma and HL (reviewed by [9,10,115] ). The number of reports on
long ncRNAs in cancer diagnosis is much more restricted, but has
progressively grown over recent years ([1 02 ] ; reviewed by [88]). The
main causes for this rising interest are the increasing realization
that long ncRNAs may play important roles in physiological and
pathological processes in the cell [88] and the development of effec-
tive high-throughput expression analysis technologies (TABles 1 &
2). The most eminent long ncRNA used as a diagnostic biomarker
in cancer is undoubtedly PCA3, previously known as DD3, whose
expression was found to be highly specific in prostate tissue and
which is highly overexpressed in prostate cancer [14 4 ] . These fea-
tures, together with the detectability in urine and the expression
in early-stage tumors, resulted in the use of PCA3 as a diagnostic
biomarker already exploited in the clinic (Box 1). Two other long
ncRNAs, HULC and HOTAIR, are highly expressed in HCC [85]
and colorectal cancer that metastasizes to the liver [84 ], and in meta-
static breast cancer [81] , respectively, compared with noncancerous
tissues. In contrast, HOTAIR levels were found to be higher in
colorectal cancer tissues as compared with adjacent uninvolved
tissue [82] . Furthermore, T-UCR expression profiles can be used
to differentiate human cancers as genome-wide profiling revealed
that T-UCRs have distinct signatures in human leukemias and
carcinomas compared with the corresponding normal tissues [30] .
Fina lly, UCA1, also known as CUDR, was identified as a highly
specific (91.8%) and very sensitive (80.9%) biomarker detected in
the urine of bladder cancer patients [14 5] .
Tumor classification
miRNA expression signatures can also be used to differentiate
between different tumor subtypes and to classify tumors of uncer-
tain cellular origin [1 46 ,147] . For instance, in breast cancer [148] ,
gastric cancer [149], acute myeloid leukemia [15 0] and HL [151] , it
has been described that miRNA signatures differentiate between
different molecular and histological subtypes. A four miRNA-
based molecular test that differentiates between specific cancer
(sub)types has been developed (Box 1). In addition, cancer subtypes
Box 1. Commercially available noncoding RNA-based diagnostic assays.
An increasing number of biotech companies started commercializing ncRNA-based diagnostic tests, primarily for certain neoplastic
diseases. For instance, Rosetta Genomics [402] offers four miRNA-based molecular tests and has a fifth assay in the pipeline. The miRview®
mets² assay can differentiate between 42 different tumor types using a custom array platform containing 64 miRNAs and has an overall
sensitivity of 85% and a specificity of 99.3% [259] . The miRview lung test uses a qRT-PCR assay based on eight miRNAs (overall sensitivity
93.7%; overall specificity 97.9%) that differentiates primary lung cancers into four types: squamous cell carcinoma, nonsquamous NSCLC,
carcinoid and small-cell carcinoma [260], while miRview squamous is based on the expression level of hsa-miR-205 measured by qRT-PCR
and distinguishes squamous from nonsquamous NSCLC with an overall sensitivity and specificity of 100% [2 61,262]. The miRview meso
molecular assay is able to differentiate malignant pleural mesothelioma from carcinomas in the lung and pleura based on a three-miRNA
qRT-PCR assay that has an overall sensitivity of 100% and a specificity of 94% [263]. Finally, the miRview kidney diagnostic test, which
is still under development, is a two-step decision-tree classifier that uses expression levels of six miRNAs: the first step uses expression
levels of hsa-miR-210 and hsa-miR-221 to distinguish between the two pairs of subtypes; the second step uses either hsa-miR-200c with
hsa-miR-139-5p to identify oncocytoma from chromophobe or hsa-miR-31 with hsa-miR-126 to identify conventional from papillary
tumors [2 64] .
One US FDA-approved diagnostic assay for prostate cancer is based on the detection of the long ncRNA PCA3 in patient urine samples.
The ProgensaPCA3 test from Gen-Probe Inc. [403] uses the transcription-mediated amplification (TMA™) technology to determine a
PCA3 score from male urine. TMA is an isothermal nucleic acid-based method that can amplify RNA or DNA targets a billion-fold in less
than 1 h [2 65]. The PCA3 assay has a sensitivity of 60–70%, a specificity of 70–80% and a negative-predictive value of 90% [403]. PCA3 is
the first long ncRNA to be used in clinical diagnostic assays, but with the recent developments in the ncRNA world, many more will most
likely follow soon.
ncRNA: Noncoding RNA; NSCLC: Non-small-cell lung cancer; qRT-PCR: Quantitative reverse transcription PCR.
miRNAs & long noncoding RNAs as biomarkers in human diseases
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
Expert Rev. Mol. Diagn. 13(2 ), (2013)
190
Review
Table 2. Selection of emerging new methods to detect and quantify noncoding RNAs.
Study (year) Method ncRNA detection mechanism Advantages and limitations Ref.
Wang et al. (2011);
Wanunu et al. (2010 )
Nanopore-based
RNA detection
An applied voltage pulls ncRNAs either
through biological or solid state
nanopores. As the molecules
translocate, they partially block the ion
flow through the pore, which is
detected as a drop in the measured
current. The detection of specific
ncRNAs can be carried out by adding a
complementary DNA probe to look for
a target-specific hybridization signature
signal
+
-
Single-molecule detection without the
need for labels or amplification
Can quantify subpicomolar levels of
cancer-associated miRNAs and
distinguish single-nucleotide
differences between miRNA family
members
Can be made into a small instrument
Possibilities for high throughput
Extremely rapid
Requires sophisticated detection
instruments
Complex data analysis
[28 4,2 85]
Chan et al. (2010) TIRFM Complementary DNA probes are
hybridized to ncRNAs and labeled with
fluorescent YOYO-1. The fluorescent
hybrids are imaged by an electron-
multiplying charge-coupled device-
coupled TIRFM system and quantified
by single-molecule counting
+
-
Single-molecule sensitivity
Very accurate quantification
Low throughput
Very advanced optical setup
[28 6]
Neely et al. (2006) Single-molecule
detection based
on dual
fluorescent
labeling
Homogeneous assay using two probes,
each with a specific dye. If both probes
bind to the ncRNA, the detector will
see two fluorescent peaks. Unbound
probes will appear as a single peak
+
-
Single-molecule sensitivity
Expensive labels
Very advanced optical setup
[28 7]
Gao and Yang (2006);
Gao and Yu (2007);
Peng and Gao (2011);
Cissell and Deo (2009)
Electrochemical
detection
ncRNAs are labeled with
electrocatalytic nanoparticles. The
labeled ncRNAs can hybridize with
complementary DNA immobilized on a
chip. The nanoparticles of the
hybridized ncRNA are able to catalyze
an oxidation reaction, which will
increase the current
+
-
Cost-effective sensitive sensors (high
fM)
Background signal has to be verified in
real samples
Need for special nanoparticle labels
[28 8–291]
Zhang et al. (2009) SiNWs Label-free and direct hybridization
assay for ultrasensitive detection of
miRNAs using silicon nanowires.
Peptide nucleic acids serve as a
complementary receptor. Resistance
change is directly corrected to the
concentrations of hybridized ncRNA
+
-
Label free, small device
Very sensitive (1 fM)
Very specific (single nucleotide)
Background signal has to be verified in
real samples
[292]
Sioss et al. (2012) Nanoresonator
chip
Detection of ncRNAs after
hybridization with complementary
DNA bound on metallic or silica
nanowires. Mass-amplifying gold
nanoparticles bind to the ncRNA–DNA
sandwich and will induce a large shift
in the resonance frequency of the
nanoresonator, which can be measured
optically
+
-
Promising technology for excellent
quantitative sensitivity and specificity
At this moment, still limited
reproducibility and sensitivity
Complex production process for
nanoresonators
[293]
Robertson and Vora
(2012)
Flow cytometry-
FISH
In situ hybridization of cells, induced
fluorescence measured with flow
cytometer
+
-
High throughput detection of in situ
hybridization
No quantification
No location of the ncRNA
[294 ]
LOD: Limit of detection; ncRNA: Noncoding RNA; SERS: Surface-enhanced Raman spectroscopy; SiNW: Silicon nanowire; SNA: Spherical nucleic acid;
SPR: Surface plasmon resonance; TIRFM: Total internal reflection fluorescence microscopy.
Van Roosbroeck, Pollet & Calin
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
191
www.expert-reviews.com
Review
can be further subcategorized by miRNA expression profiles
unique to specific genotypes [55] . For example in lung cancer, KRAS
mutation-positive tumors are associated with miR-495 upregula-
tion and EGFR mutation-positive tumors show upregulated miR-
21 and miR-25, whereas miR-155 is exclusively upregulated in lung
cancer tumors without KRAS or EGFR mutation [152 ,153 ] .
That also long ncRNAs can also be used to classify tumors was
illustrated for pediatric tumors [5 6] . The authors demonstrated
the presence and overexpression of a 250-kb stretch of noncod-
ing transcript, which they call a putative very long ncRNA, in
Ewing’s family of tumors, but not in the other cancer tumor
types analyzed (PAX-FKHR fusion-positive and -negative rhabdo-
myosarcoma, osteosarcoma, neuroblastoma and Wilms’ tumors).
However, further validation of these findings is necessary as this
study only considered a small number of patients for each subtype.
Furthermore, a novel ncRNA in rhabdomyosarcoma, NCRMS,
Table 2. Selection of emerging new methods to detect and quantify noncoding RNAs (cont.).
Study (year) Method ncRNA detection mechanism Advantages and limitations Ref.
Driskell et al. (2009) SERS Absorption of ncRNA to silver or gold
nanoparticles can be detected through
Raman scattering spectroscopy
+
-
High sensitivity (fM)
Provides a very specific molecular
fingerprint
No labeling required
Sophisticated instrument needed
Low reproducibility
Difficulty of spectra interpretation,
especially in more complex matrices
[295]
Cissell et al. (2008) Bioluminescence Heterogeneous or homogeneous
competitive hybridization assays with
bioluminescent labeled DNA probes
+
-
Good sensitivity (LOD = 1 fM)
Expensive consumables
Difficult to standardize
[296 ]
Fang et al. (2006);
Nasheri et al. (2011);
Sípová et al. (2010);
Wark et al. (2008)
SPR SPR is an optical detection method that
is sensitive for small changes of the
local refractive index that are, for
example, induced by the hybridization
of ncRNA to immobilized capture DNA.
The signal can be amplified by using
hefty labels. SPR imaging enables SPR
in a microarray format
+
-
Good sensitivity (10 fM)
Expensive read-out system
Risk for high background signal as
result of nonspecific binding
interactions
Limited throughput
[297–300]
Dong et al. (2012) Fluorescence
quenching on
graphene oxide
ssDNA strongly binds to graphene
oxide, which also quenches fluorescent
dyes. Fluorescent probes hybridized
with ncRNA are protected from
adsorption to the graphene. The
sensitivity can be enhanced by
isothermal strand-displacement
amplification
+
-
Amplification process
High sensitivity (LOD = 2.1 fM)
Fluorescent labels are expensive
[3 01]
Alhasan et al. (2012) Scanometric
miRNA array
Isolated miRNAs are enzymatically
ligated to a universal linker followed by
hybridization onto miRNA microarray.
Universal SNA-functionalized gold
nanoparticle conjugates are
subsequently hybridized to detect
captured miRNA targets. The signal
intensity is amplified by depositing
gold with gold-enhancing solution and
before imaging
+
-
Sensitivity (1 fM)
Nonspecific adsorption of gold
nanoparticles may induce high
background signal
[30 2]
Jiang et al. (2012) Base stacking
hybridization
coupling with
time-resolved
fluorescence
technology
The ncRNA and fluorescent DNA tags
can hybridize together with an
immobilized DNA capture probe. If no
ncRNA is present, the short duplex
between the fluorescent tag and
capture probe will be disrupted when
washing the surface, because the
binding is energetically unfavorable
+
-
Rapid, universal label
Sensitive (20 fM)
Needs fluorescent tag
[303]
LOD: Limit of detection; ncRNA: Noncoding RNA; SERS: Surface-enhanced Raman spectroscopy; SiNW: Silicon nanowire; SNA: Spherical nucleic acid;
SPR: Surface plasmon resonance; TIRFM: Total internal reflection fluorescence microscopy.
miRNAs & long noncoding RNAs as biomarkers in human diseases
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
Expert Rev. Mol. Diagn. 13(2 ), (2013)
192
Review
Table 3. Illustrative list of long noncoding RNAs (potentially) used as biomarkers in human diseases.
Disease Involved ncRNA Name/description Application
Prostate cancer PCA3 or DD3 Prostate cancer 3 or differential
display code 3 long ncRNA
Diagnostic biomarker detected in exosomes
isolated from urine of prostate cancer patients
Liver cancer, CRC that
metastasizes to the liver
HULC Highly upregulated in liver long
ncRNA
Potential diagnostic biomarker highly
upregulated in blood of liver and hepatic
metastatic CRC patients
Generic variant contributes to risk of HBV-
related hepatocellular carcinoma
Coronary artery disease,
atherosclerosis, stroke, intracranial
aneurism, diabetes, ALL,
melanoma, glioma, lung cancer,
prostate cancer, NF type 1
ANRIL Antisense ncRNA in the INK4
locus
Polymorphisms in ANRIL as biomarker for
disease susceptibility
Prostate cancer PC AT-1 Prostate cancer-associated
ncRNA transcript 1
Polymorphisms in 8q24 region (region of
PC AT-1) associated with risk for prostate
cancer
Potential biomarker for prostate cancer
subtype stratification
Prostate cancer PRNCR1 Prostate cancer ncRNA 1 Polymorphisms in 8q24 region (region of
PRNC1) associated with risk of prostate cancer
Psoriasis PRINS Psoriasis susceptibility-related
RNA gene induced by stress
Overexpression correlated with disease
susceptibility
Autoimmune thyroid disease SAS-ZFAT Small antisense transcript of
zinc-finger gene in AITD
susceptibility region
SNP in promoter of SAS-ZFAT determines
susceptibility to autoimmune thyroid disease
Metastatic breast cancer,
colorectal cancer, HCC, GIST,
pancreatic cancer
HOTA IR HOX antisense intergenic RNA Prognostic biomarker in several types of
cancer, as high expression is associated with
poor prognosis, metastasis and tumor
recurrence
Rhabdomyosarcoma NCRMS ncRNA in rhabdomyosarcoma Potential biomarker in tumor classification due
to differential expression between
rhabdomyosarcoma subtypes
NSCLC, HCC, breast carcinoma,
pancreatic cancer, CRC, prostate
cancer, ESS of the uterus,
osteosarcoma, bladder cancer
MALAT1 Metastasis-associated lung
adenocarcinoma transcript 1
Prognostic marker in many human carcinomas,
as high expression is associated with poor
prognosis, poor response to therapy,
metastasis and tumor recurrence
HNSCC, bladder cancer, CRC,
cervical cancer
CUDR = UCA1 Cancer-upregulated drug-
resistant gene or urothelial
carcinoma associated 1
Overexpression associated with drug
resistance in HNSCC cell lines
Diagnostic biomarker detected in urine of
bladder cancer patients
HCC, ovarian cancer, bladder
cancer, HNSCC
H19 Imprinted (at IGF2 locus)
maternally expressed
untranslated mRNA
Potential diagnostic marker for HCC and
ovarian cancer
Potential prognostic marker in bladder cancer
and HNSCC, with higher expression associated
with higher grade disease and indicative for
tumor recurrence
May induce MDR1-associated drug resistance
in HCC cells
Myocardial infarction MIAT Myocardial infarction-
associated transcript
Polymorphisms in MIAT confer risk of
myocardial infarction
AITD: Autoimmune thyroid disease; ALL: Acute lymphocy tic leukemia; CRC: Colorectal cancer; ESS: Endometrial stromal sarcoma; GIST: Gastrointestinal stromal
tumor; HBV: Hepatitis B virus; HCC: Hepatocellular carcinoma; HNSCC: Head and neck squamous cell carcinoma; ncRNA: Noncoding RNA; NF: Neurofibromatosis;
NSCLC: Non-small-cell lung cancer.
Van Roosbroeck, Pollet & Calin
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
193
www.expert-reviews.com
Review
shows differential expression between rhabdomyosarcoma
subtypes [15 4] .
Cancer prognosis
To date, the amount of prognostic ncRNAs outnumbers the
ncRNAs used for diagnostic purposes, mainly because diagnos-
tic biomarkers preferably enable early detection. One of the first
miRNA signatures associated with prognosis was established in
CLL: a 13-miRNA expression signature has been identified that is
able to differentiate between CLL cases with low versus high levels
of ZAP-70, between unmutated and mutated IgV(H), and that
was associated with the presence or absence of disease progression
[155 ,156 ] . After these initial reports, many other prognostic miRNA
markers have been identified in a variety of cancers, including
non-small-cell lung cancer (NSCLC), breast cancer, gastric can-
cer, prostate cancer, HCC, glioblastoma, acute and chronic leuke-
mias, diffuse large B-cell lymphoma and HL (reviewed by [9,10]).
One well-known example of a prognostic long ncRNA is metas-
tasis-associated lung adenocarcinoma transcript 1 (MAL AT1),
which is highly expressed in many solid tumors [157,158] , predicts
poor prognosis in NSCLC [15 9,16 0] and is an independent prognos-
tic marker for predicting HCC recurrence after liver transplan-
tation [1 61] . Signatures for transcribed UCRs could differentiate
CLL patients with poor prognosis (high expression of ZAP-70)
from CLL patients with good prognosis (low ZAP-70 expression)
[30] , and were able to significantly distinguish between short-term
and long-term neuroblastoma survivors [16 2] .
One major cause of cancer-related death is tumor metastasis,
a multistage process that includes dissemination of tumor cells
to and proliferation at distant sites [163]. Therefore, identification
of biomarkers associated with disease progression is extremely
significant as these markers may be used as a basis for the devel-
opment of new therapeutic strategies reducing the mortality
and morbidity, and improving the survival of metastatic cancer
patients [164 ] . miRNAs have been shown to act as either meta-
static activators or as metastatic repressors in a variety of human
cancers, including breast cancer, testicular germ cell tumors,
colon cancer, gliomas, pancreatic cancer, nasopharyngeal car-
cinoma and prostate cancer (reviewed by [9,16 3]). Several long
ncRNAs have also been implicated in metastasis. For instance,
HOTAIR has a unique association with patient prognosis and
correlates with metastasis in breast cancer patients [81] , colorectal
carcinoma [82], HCC [1 65] and gastrointestinal stromal tumors
[16 6] . MALAT1 was first identified in metastatic NSCLC [159 ] ,
and its high expression was subsequently found to be related to
colorectal cancer metastasis [167] and to predict tumor recurrence
of HCC after liver transplantation [1 61] . In a recent report, the
relative abundance of a set of long ncRNAs was studied in pan-
creatic tissue samples and sets of intronic long ncRNAs whose
abundance is correlated with pancreatic ductal adenocarcinoma
or metastasis [168]. Finally, a unique 26-kb intergenic ncRNA
transcript, located on chromosome 2, was found to be highly
expressed in a primary Ewing’s family of tumors that did not
metastasize, whereas low expression was observed in primary
tumors that eventually metastasized [5 6 ] .
Recent reports indicate that sequence variations of ncRNAs can
also act as prognostic biomarkers. SNPs located in pre-miRNAs
[12 0, 169], in the miRNA binding site located in the 3 UTR of
protein-coding genes [1 70 ,171] and even in the miRNA flanking
region [1 72] have been found to have prognostic value in a variety
of human cancers.
These data show the potential of miRNAs and long ncRNAs
as biomarkers in cancer prognosis as well as to stratify metastatic
from nonmetastatic cancers. In addition, sequence variations of
ncRNAs may be potential prognostic biomarkers in human cancer.
Prediction of therapeutic responses
ncRNAs can be employed to predict response to therapy as their
expression is known to play a role in drug sensitivity and resist-
ance to therapy [1 0,173,174] . An 11 miRNA signature was estab-
lished in ovarian cancer cell lines, which differentiates between
cisplatin-resistant versus cisplatin-sensitive cells [17 5] . With respect
to single-miRNA biomarkers, low levels of miR-10b were found
to be associated with improved response to neoadjuvant therapy
and outcome in pancreatic ductal adenocarcinoma [176 ] , whereas
high expression in colorectal cancer confers resistance to chemo-
therapeutic agents [177 ] . Moreover, increased miR-21 expression
is correlated with chemoresistance in pancreatic cancer cells
[17 8– 181] , glioblastoma multiforme [18 2] , bladder cancer cells [1 83 ] ,
and head and neck squamous cell carcinoma cells [1 84 ] . In breast
cancer, miR-210 levels correlate with sensitivity to trastuzumab
[185 ] and miR-125b is predictive of chemoresistance [186 ,187] . In
addition, miR-125b is also associated with resistance to therapy
in ovarian cancer cells [188], glioblastoma stem cells [18 9] and pedi-
atric acute promyelocytic leukemia [19 0] . In NSCLC, miR-128b is
demonstrated to be a predictive factor of response to therapy [191].
To the best of our knowledge, only two long ncRNAs have been
associated with the prediction of chemotherapeutic response:
MALAT1, which is highly expressed in a poor responder group of
osteosarcoma [192] , and CUDR, whose overexpression is associated
with drug resistance in head and neck squamous cell carcinoma
cell lines [193] . In addition, it is suggested that the long ncRNA
H19 induces MDR1-associated drug resistance in HCC cells [19 4] .
Modulation of the ncRNAs involved in resistance to therapy may
sensitize cancer cells to chemotherapeutic agents [19 5,19 6] .
Sequence variations in ncRNAs may also be predictive of thera-
peutic responses. Among others, SNPs in miRNA binding sites
of protein-coding genes have been found to be correlated with
radiotherapy outcome in bladder cancer [19 7] , and with methotrex-
ate [198] and cisplatin resistance [19 9] . In addition, polymorphisms
in pri-miRNA and in components of the miRNA biogenesis
machinery have been found to be related to treatment outcome
in metastatic colon cancer patients treated with 5-fluorouracil
and irinotecan [2 00] .
ncRNA biomarkers in CVD
According to the WHO, CVDs are the leading cause of death
worldwide [20 1] . In 2008, more than 17 million people died
from CVD, and although the mortality rates due to CVD have
reduced, the disease burden remains high [20 2]. Therefore, better
miRNAs & long noncoding RNAs as biomarkers in human diseases
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
Expert Rev. Mol. Diagn. 13(2 ), (2013)
194
Review
diagnostic and therapeutic strategies are required to address these
high morbidity and mortality rates.
miRNAs play an essential role in cardiovascular biology and
are often deregulated in several heart diseases including myo-
cardial hypertrophy, heart failure and arrhythmia (reviewed by
[203] ). These deregulations can be caused by aberrant expression
of genes from the miRNA biogenesis machinery, as decreased
Dicer expression was observed in patients with end-stage dilated
cardiomyopathy and heart failure [204], or by aberrant expres-
sion of miRNAs themselves, for example deregulation of miR-1
and miR-133 in cardiac hypertrophy and arrhythmia [203] . Since
heart and vascular tissues are almost impossible to obtain from
patients with CVDs, most research has focused on the discovery
of circulating miRNA biomarkers for either diagnostic and/or
prognostic purposes ([205 –210] ; reviewed by [6 ,211] ).
A few long ncRNAs are reported to have a function in CVDs,
but their use as biomarkers still need to be validated. ANRIL was
found to be associated with atherosclerosis [21 2] and genetic vari-
ants at the 9p21 locus of ANRIL correlate with CVD susceptibil-
ity [13 3,13 4,213–215]. Similarly, altered expression of the long ncRNA
MIAT by SNPs confers risk of myocardial infarction [216 ].
ncRNA biomarkers in autoimmune diseases
Autoimmune diseases form a very heterogeneous group of more
than 100 distinctive diseases in which the patient’s immune system
in mistakenly activated to attack substances and tissues normally
present in the body. miRNAs are essential for normal immune
functions, immune cell development and prevention of autoim-
munity (reviewed by [217, 218]). Although most studies were per-
formed in cultured cells and animals, there is evidence that some
miRNAs play a critical role in the pathogenesis of auto immune
diseases such as SLE (e.g., miR-21, miR-125a, miR-146a and
miR-148a), rheumatoid arthritis (e.g., miR-124a, miR-146a and
miR-155) and multiple sclerosis (e.g., miR-17-5p, miR-20a, miR-
34a, miR-155 and miR-326), reviewed by [219] . Understanding the
involvement of long ncRNAs in autoimmunity is only in its infancy
and, to the best of our knowledge, no long ncRNA has been found
to be directly implicated in the development of the major types of
autoimmune diseases. Some preliminary data in a murine model
system point to a link between the long ncRNA GAS5 and disease
susceptibility to SLE [22 0]. In addition, the chromosomal locus of
GAS5, 1q25, has been associated with human SLE development
in genetic studies [221–223]. The identification of abnormal ncRNA
expression in autoimmune diseases is still emerging, hence the
potential value of miRNAs and long ncRNAs as biomarkers is
still largely unexplored and warrants further investigation.
ncRNA biomarkers in neurological diseases
The nervous system has the broadest spectrum of miR NA expres-
sion of all human tissues, with approximately 70% of experi-
mentally detectable miRNAs being expressed in the brain [2 24] .
Thus, it is not surprising that miRNA deregulation has been
described in almost all neurological diseases studied [32], ranging
from neuro developmental, neurovascular and neurodegenerative
disorders to tumors of the central nervous system, nervous system
trauma and psychiatric disorders (reviewed by [22 5]). Although
differential miRNA expression profiles have been generated that
could potentially be used to diagnose, classify and predict the
prognosis of some neurological disorders, to date, there is no
diagnostic laboratory test available [111].
The majority of long ncRNAs are expressed in the nervous
system [2 26] and several are deregulated in neurological disorders
(reviewed by [51 ,22 7–230] ). Similar to CVDs, biological samples of
diseased tissues are very difficult, if not impossible, to obtain.
For example, in Alzheimer’s disease, up until now, the diagnosis
relies principally on clinical criteria and a pathological confirma-
tion can only be made by postmortem analysis of plaques and
tangles in the brain [231] . In the last few years, the search for useful
biomarkers in neurological disorders has begun to concentrate
on the detection in body fluids, such as cerebrospinal fluid and
plasma, although the main focus has been on proteins and pep-
tides [2 31,23 2]. Nevertheless, miRNAs begin to emerge as biological
body fluid markers of neurological diseases [2 07,233–235]. Whether
long ncRNAs in body fluids can also be useful as neurological
biomarkers remains to be elucidated.
ncRNA biomarkers in infectious diseases
To study the role of ncRNAs in viral infections, two aspects have
to be taken into account: the viral ncRNAs and host ncRNAs.
Some viruses encode their own ncRNAs, which target host
mRNAs or viral RNAs. These viral ncRNAs may be used as
diagnostic markers for viral infection or stage of infection [23 6].
On the other hand, parts of the viral life cycle, such as viral trans-
lation, may be regulated by host ncRNAs [237] . Finally, the virus
may alter the expression of host ncRNAs and, in this way, change
the regulation of host or viral ncRNA targets [2 37] .
The involvement of miRNAs in viral infection is reviewed else-
where [23 7,23 8]. An intron-less viral long noncoding poly adenylated
nuclear RNA was found to be highly expressed during the lytic
phase of infection of the Kaposi’s sarcoma-associated herpes
virus [239,240], the causative agent of various human cancers and
immuno proliferative disorders. Two recent studies used next-
generation whole transcriptome sequencing to identify ncRNAs
differentially expressed after virus expression. A first study showed
differential expression of approximately 500 annotated, long
ncRNAs and 1000 nonannotated genomic regions after severe
acute respiratory syndrome corona virus infection in mice [ 24 1] .
In addition, the authors demonstrated that over 40% of ncR-
NAs were similarly regulated in vitro in response to influenza
virus infection, suggesting the existence of a signature profile
associated with pathogenicity. In the second study, a software
tool ‘RandA’ (ncRNA Read-and-Analyze) that performs com-
prehensive ncRNA profiling and differential expression analysis
on data generated by deep sequencing while running on a local
computer was developed [2 42] . They used RandA to identify and
validate ncRNA shifts in samples infected with mycoplasma, HIV
and Epstein–Barr virus. Overall, the knowledge on the roles of
miRNAs and long ncRNAs in infectious diseases is limited and
further research is necessary in order to exploit them as useful
biomarkers.
Van Roosbroeck, Pollet & Calin
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
195
www.expert-reviews.com
Review
ncRNAs as a new strategy for therapeutic
interventions
Their widespread role in physiological processes and deregula-
tion in human diseases makes ncRNAs very attractive targets for
novel therapeutics. Several strategies are currently being inves-
tigated that either silence overexpressed ncRNAs or reactivate
down regulated ncRNAs (reviewed by [32 ,203,243]). To date, the
main focus has been on miRNAs, because long ncRNA-based
research is only in premature stages. However, similar strategies
to restore deregulated miRNAs could potentially also be useful
for long ncRNAs.
Inactivation of overexpressed miRNAs can be achieved by
using antisense oligonucleotides, which are complementary to the
miRNA and inhibit miRNAs based on this base pair complemen-
tarity [3 2]. To increase the stability of antisense oligonucleotides,
different chemical modifications have been incorporated, leading
to the development of locked nucleic acid anti-miRs [24 4], anti-
sense miRNA oligonucleotides [245] and antagomirs [2 46] . Another
approach is the use of ‘miRNA sponges’, transcripts expressed from
strong promoters and containing multiple artificial miRNA bind-
ing sites that compete with natural miRNA targets for miRNA
binding [2 47] . The search for small-molecule inhibitors of miRNAs
is also ongoing [248] , although it is still under debate whether small
molecules are indeed able to inhibit miRNA function. On the
other hand, the ‘microRNA replacement strategy’ can be used
to re-activate downregulated miRNAs. These miRNA mimics
are synthetic dsRNAs that are designed to mirror the action of
mature endogenous miRNAs (reviewed by [249]). Although these
therapeutic strategies are very promising, some major hurdles
remain, including the problem of proper in vivo delivery and off-
target specificity [250 ]. Many biotech companies are trying to over-
come these obstacles to advance miRNA-based therapeutics to the
clinic. In fact, one locked nucleic acid-modified oligonucleotide,
SPC3649 or miravirsen, has gone forward from preclinical trials
in primates [ 251] to Phase I and IIa clinical trials in humans with
chronic hepatitis C virus infection [40 1] . SPC3649 was demon-
strated to block the liver-specific host miRNA miR-122, which
is critical to hepatitis C virus accumulation in the liver [ 251] . The
Phase IIa clinical trial recently finished and no official results have
been published yet. However, a late-breaking oral presentation
on the 2011 AASLD meeting reported that the drug was safe
and well tolerated, and resulted in a significant dose-dependent
decrease in hepatitis C RNA levels in patients [252] . Miravirsen is
the first miRNA-targeted drug to enter clinical trials, but most
likely, others will follow soon.
For more information on the use of ncRNAs as targets of
therapeutic interventions, the reader is redirected to some other
reviews [253–257].
Expert commentary
The discovery of informative biomarkers is not only important to
understand physiopathological processes of diseases, but is also
essential for improved diagnostics and therapeutics [8] . Variations
in expression and/or sequences of ncRNAs have been associated
with disease predisposition, which drives the development of
biomarkers for genetic disease susceptibility. Although in many
instances genetic screening might be very useful, it raises various
ethical questions that need to be answered [258] .
A critical issue, particularly in cancers, is early detection. Most
current diagnostic biomarkers are based on molecular analysis of
the tumor tissue itself, which is usually very small in early stages.
Consequently, many tumor types can only be detected in later
stages, when patients tend to be refractory to therapy, which leads
to high mortality rates. In this regard, the discovery of tumor-
specific ncRNAs in body fluids such as serum, saliva and urine is of
key interest, as even in early stages, significant changes in ncRNA
concentration levels could be detected [4] , and these samples can
be obtained through noninvasive methods. More research is, how-
ever, needed in order to establish circulating ncRNAs as diagnostic
and prognostic biomarkers, as the precise functions of these cell-
free RNAs remain poorly understood [4]. To effectively screen
for these ncRNA biomarkers, there is a need for accurate and
extremely sensitive determination methods. The field of ncRNA
diagnostics is rapidly evolving. Several established techniques,
such as microarrays, ISH, real-time PCR and RNA sequencing,
are routinely used nowadays (TABle 1) . Meanwhile, novel methods
are being developed that are more sensitive and selective, and can
be used at high throughput and at low cost (TABle 2).
Even though the knowledge on ncRNAs has significantly
advanced over recent years, much more research is needed to
elucidate the full extent to which ncRNAs exert their patho-
logical effect, before they can be widely used in diagnostics and
prognostics of human diseases.
Five-year view
In the last decade, the involvement of ncRNAs in human diseases
has become increasingly apparent, and many research groups have
started to explore the molecular use of ncRNAs in diagnostics and
therapeutics. In particular, miRNAs have been extensively inves-
tigated, although more and more research starts to focus on other
ncRNAs as well. As their importance in human diseases becomes
increasingly apparent, the molecular use of ncRNAs in diagnos-
tics and therapeutics is now beginning to be explored. The pres-
ence of ncRNAs in body fluids makes them extremely suitable as
noninvasive biomarkers, and it is like they will become the main
source of biomarkers in the coming years. The first diagnostic tests
employing miRNAs and long ncRNAs are commercially available
(Box 1) and many others are in the pipeline. ncRNAs can also be
used as ‘druggable’ targets, and although some hurdles associated
with in vivo delivery and off-target specificity still need to be over-
come before ncRNAs can be widely used as therapeutic targets, an
increasing number of biotech companies are trying to overcome
these obstacles, and their first reports are promising for the near
future. In fact, the first miRNA antagonist is in clinical trials for
patients with chronic hepatitis C infection [4 0 1] , and we expect many
more to follow soon.
miRNAs & long noncoding RNAs as biomarkers in human diseases
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
Expert Rev. Mol. Diagn. 13(2 ), (2013)
196
Review
Financial & competing interests disclosure
K Van Roosbroeck is supported by a fellowship of the Henri Benedictus Fund/
Belgian American Education Foundation. J Pollet is supported by fellowships
of the Fulbright Association and the Belgian American Education Foundation.
GA Calin is an Alan M Gewirtz Leukemia & Lymphoma Society Scholar. He
is also supported as a Fellow at The University of Texas MD Anderson Research
Trust, as a University of Texas System Regents Research Scholar, and by the
CLL Global Research Foundation. Work in GA Calin’s laboratory is supported
in part by the NIH/NCI (C A135444), a Department of De fense Breast Cancer
Idea Award, Developmental Research Awards in Breast Cancer, Ovarian
Cancer, Brain Cancer, Prostate Cancer, Multiple Myeloma, Leukemia
(P50 CA100632) and Head and Neck (P50 CA097007) SPOREs, a SINF
MDACC_DKFZ grant in CLL, the Laura and John Arnold Foundation,
the RGK Foundation and the Estate of CG Johnson Jr. The authors have no
other relevant affiliations or financial involvement with any organization or
entity with a financial interest in or financial conflict with the subject matter
or materials discussed in the manuscript apart from those disclosed.
No writing assistance was utilized in the production of this manuscript.
Key issues
miRNAs have more complex functions than initially assumed: they do not only negatively regulate expression of protein-coding genes,
but they can also upregulate expression of target genes, act as competing endogenous RNAs (ceRNAs) that divert miRNAs away from
their mRNA targets, function as RNA decoy to modulate the function of regulatory proteins, and interact with and regulate other
noncoding RNAs (ncRNAs).
Most ncRNAs are good biomarkers as they are stable (even in body fluids), their expression is specific to tissues, organs or biological
stages, they show increased or decreased expression levels in disease, and they can be easily measured by methods such as quantitative
PCR and microarrays.
Variations in expression and/or sequences of miRNAs and long ncRNAs are involved in disease predisposition.
ncRNAs are very useful cancer biomarkers as they can be used to establish a diagnosis, classify tumors, determine disease stage and
predict disease outcome and evolution.
The use of ncRNAs as biomarkers is not only restricted to cancer, but extends to a plethora of other human diseases, including
cardiovascular diseases, autoimmune diseases, neurological disorders and infectious diseases.
The clinical use of ncRNAs as biomarkers for cancer is translated into four commercially available diagnostic tests that distinguish
between different cancer subtypes.
ncRNAs are very attractive targets for therapeutic interventions, and several strategies are currently being explored that either silence
overexpressed ncRNAs or re-activate downregulated ncRNAs.
A first miRNA antagonist, miravirsen, is in Phase IIa clinical trial for patients with chronic hepatitis C virus infection.
References
1 Carninci P, Kasukawa T, Katayama S et al.;
FANTOM Consortium; RIKEN Genome
Exploration Research Group and Genome
Science Group (Genome Network Project
Core Group). The transcriptional
landscape of the mammalian genome.
Science 309(5740), 1559–1563 (2005).
2 Birney E, Stamatoyannopoulos JA, Dutta
A et al. Identification and analysis of
functional elements in 1% of the human
genome by the ENCODE pilot project.
Nature 447(7146), 799–816 (2007).
3 Gibb EA, Brown CJ, Lam WL. The
functional role of long non-coding RNA in
human carcinomas. Mol. Cancer 10, 38
(2 011) .
4 Cortez MA, Bueso-Ramos C, Ferdin J,
Lopez-Berestein G, Sood AK, Calin GA.
MicroRNAs in body fluids – the mix of
hormones and biomarkers. Nat. Rev. Clin.
Oncol. 8(8), 467–477 (2011).
5 Creemers EE, Tijsen AJ, Pinto YM.
Circulating microRNAs: novel biomarkers
and extracellular communicators in
cardiovascular disease ? Circ. Res. 110(3),
483 –495 (2012).
6 Di Stefano V, Zaccagnini G, Capogrossi
MC, Martelli F. microRNAs as peripheral
blood biomarkers of cardiovascular
disease. Vascul. Pharmacol. 55 (4 ), 111–118
(2 011) .
7 Edwards JK, Pasqualini R, Arap W, Calin
GA. MicroRNAs and ultraconserved genes
as diagnostic markers and therapeutic
targets in cancer and cardiovascular
diseases. J. Cardiovasc. Transl. Res. 3(3),
271–279 (2010).
8 Etheridge A, Lee I, Hood L, Galas D,
Wang K. Extracellular microRNA: a new
source of biomarkers. Mutat. Res. 717(1–2 ),
85 –9 0 (2011).
9 Fabbri M. miRNAs as molecular biomark-
ers of cancer. Expert Rev. Mol. Diagn.
10(4), 435–444 (2010).
10 Ferracin M, Veronese A, Negrini M.
Micromarkers: miRNAs in cancer
diagnosis and prognosis. Expert Rev. Mol.
Diagn. 10(3), 297–308 (2010).
11 Geekiyanage H, Jicha GA, Nelson PT,
Chan C. Blood serum miRNA:
non-invasive biomarkers for Alzheimer’s
disease. Exp. Neurol. 235(2), 491–496
(2012).
12 Lai CY, Yu SL, Hsieh MH et al. Micro-
RNA expression aberration as potential
peripheral blood biomarkers for
schizophrenia. PLoS ONE 6(6), e21635
(2 011) .
13 Pelletier C, Weidhaas JB. MicroRNA
binding site polymorphisms as biomarkers
of cancer risk. Expert Rev. Mol. Diagn.
10(6), 817–829 (2010).
14 Wang Y, Liang Y, Lu Q. MicroRNA
epigenetic alterations: predicting
biomarkers and therapeutic targets in
human diseases. Clin. Genet. 74(4 ),
307–315 (2008).
15 Zahm AM, Thayu M, Hand NJ, Horner
A, Leonard MB, Friedman JR. Circulating
microRNA is a biomarker of pediatric
Crohn disease. J. Pediatr. Gastroenterol.
Nutr. 53(1), 26–33 (2011).
16 Lee RC, Feinbaum RL, Ambros V. The
C. elegans heterochronic gene lin-4
encodes small RNAs with antisense
complementarity to lin-14. Cell 75(5),
843–854 (1993).
17 Wightman B, Ha I, Ruvkun G.
Posttranscriptional regulation of the
heterochronic gene lin-14 by lin-4 mediates
Van Roosbroeck, Pollet & Calin
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
197
www.expert-reviews.com
Review
temporal pattern formation in C. elegans.
Cell 75(5), 855–862 (1993).
18 Sumazin P, Yang X, Chiu HS et al. An
extensive microRNA-mediated network of
RNA–RNA interactions regulates
established oncogenic pathways in
glioblastoma. Cell 147(2), 370–381 (2011).
19 Vasudevan S. Posttranscriptional
upregulation by microRNAs. Wiley
Interdiscip. Rev. RNA 3(3), 311–330
(2012).
20 Salmena L, Poliseno L, Tay Y, Kats L,
Pandolfi PP. A ceRNA hypothesis: the
Rosetta Stone of a hidden RNA language?
Cell 146(3), 353–358 (2011).
21 Poliseno L, Salmena L, Zhang J, Carver B,
Haveman WJ, Pandolfi PP. A coding-inde-
pendent function of gene and pseudogene
mRNAs regulates tumour biology. Nature
465(7301), 1033–1038 (2010).
22 Karreth FA, Tay Y, Perna D et al. In vivo
identification of tumor-suppressive PTEN
ceRNAs in an oncogenic BRAF-induced
mouse model of melanoma. Cell 147(2),
382–395 (2011).
23 Tay Y, Kats L, Salmena L et al. Coding-
independent regulation of the tumor
suppressor PTEN by competing endog-
enous mRNAs. Cell 147(2), 344–357
(2 011) .
24 Cesana M, Cacchiarelli D, Legnini I et al.
A long noncoding RNA controls muscle
differentiation by functioning as a
competing endogenous RNA. Cell 147(2),
358 –369 (2011).
25 Wang J, Liu X, Wu H et al. CREB
up-regulates long non-coding RNA, HULC
expression through interaction with
microRNA-372 in liver cancer. Nucleic
Acids Res. 38(16), 5366–5383 (2010).
26 Eiring AM, Harb JG, Neviani P et al.
miR-328 functions as an RNA decoy to
modulate hnRNP E2 regulation of mRNA
translation in leukemic blasts. Cell 140(5),
652– 665 (2010).
27 Pan YZ, Morris ME, Yu AM. Micro-
RNA-328 negatively regulates the
expression of breast cancer resistance
protein (BCRP/ABCG2) in human cancer
cells. Mol. Pharmacol. 75(6), 1374–1379
(2009).
28 Bhattacharyya SN, Habermacher R,
Martine U, Closs EI, Filipowicz W. Relief
of microRNA-mediated translational
repression in human cells subjected to
stress. Cell 125(6 ), 1111–1124 (2006 ).
29 Kedde M, Strasser MJ, Boldajipour B et al.
RNA-binding protein Dnd1 inhibits
microRNA access to target mRNA. Cell
131(7), 1273–1286 (2007).
30 Calin GA, Liu CG, Ferracin M et al.
Ultraconserved regions encoding ncRNAs
are altered in human leukemias and
carcinomas. Cancer Cell 12(3), 215–229
(2007).
31 Hansen TB, Wiklund ED, Bramsen JB
et al. miRNA-dependent gene silencing
involving Ago2-mediated cleavage of a
circular antisense RNA. EMBO J. 30(21),
4414–4422 (2011).
32 Esteller M. Non-coding RNAs in human
disease. Nat. Rev. Genet. 12 (12), 861–874
(2 011) .
33 Mercer TR, Dinger ME, Mattick JS. Long
non-coding RNAs: insights into functions.
Nat. Rev. Genet. 10 (3), 155 –159 ( 20 09).
34 Kapranov P, Cheng J, Dike S et al. RNA
maps reveal new RNA classes and a
possible function for pervasive transcrip-
tion. Science 316(5830 ), 1484–1488
(2007).
35 Prensner JR, Chinnaiyan AM. The
emergence of lncRNAs in cancer biology.
Cancer Discov. 1(5), 391–407 (2011).
36 Khalil AM, Guttman M, Huarte M et al.
Many human large intergenic noncoding
RNAs associate with chromatin-modifying
complexes and affect gene expression. Proc.
Natl Acad. Sci. USA 10 6(28) , 116 67–11672
(2009).
37 Kotake Y, Nakagawa T, Kitagawa K et al.
Long non-coding RNA ANRIL is
required for the PRC2 recruitment to and
silencing of p15(INK4B) tumor suppres-
sor gene. Oncogene 30 (16), 1956 –1962
(2 011) .
38 Rinn JL, Kertesz M, Wang JK et al.
Functional demarcation of active and silent
chromatin domains in human HOX loci by
noncoding RNAs. Cell 129(7), 1311–1323
(2007).
39 Li T, Hu JF, Qiu X et al. CTCF regulates
allelic expression of Igf2 by orchestrating a
promoter-polycomb repressive complex 2
intrachromosomal loop. Mol. Cell. Biol.
28(20), 6473–6482 (2008).
40 Pandey RR, Mondal T, Mohammad F et al.
Kcnq1ot1 antisense noncoding RNA
mediates lineage-specific transcriptional
silencing through chromatin-level
regulation. Mol. Cell 32 (2), 232–246
(2008).
41 Plath K, Fang J, Mlynarczyk-Evans SK
et al. Role of histone H3 lysine 27
methylation in X inactivation. Science
300 (5616), 131–135 (2003).
42 Wang KC, Yang YW, Liu B et al. A long
noncoding RNA maintains active
chromatin to coordinate homeotic gene
expression. Nature 472(7341), 120–124
(2 011) .
43 Wang D, Garcia-Bassets I, Benner C et al.
Reprogramming transcription by distinct
classes of enhancers functionally defined by
eRNA. Nature 474(7351), 390–394 (2011).
44 Ørom UA, Derrien T, Beringer M et al.
Long noncoding RNAs with enhancer-like
function in human cells. Cell 143(1),
46 –58 (2010).
45 Pennacchio LA, Ahituv N, Moses AM et al.
In vivo enhancer analysis of human
conserved non-coding sequences. Nature
444(7118), 499–502 (2006).
46 Visel A, Prabhakar S, Akiyama JA et al.
Ultraconservation identifies a small subset
of extremely constrained developmental
enhancers. Nat. Genet. 40(2), 158–160
(2008).
47 Bernard D, Prasanth KV, Tripathi V et al.
A long nuclear-retained non-coding RNA
regulates synaptogenesis by modulating
gene expression. EMBO J. 29 (18),
3082–3093 (2010).
48 Tripathi V, Ellis JD, Shen Z et al. The
nuclear-retained noncoding R NA MALAT1
regulates alternative splicing by modulating
SR splicing factor phosphorylation. Mol.
Cell 39(6), 925–938 (2010).
49 Faghihi MA, Modarresi F, Khalil AM et al.
Expression of a noncoding RNA is elevated
in Alzheimer’s disease and drives rapid
feed-forward regulation of β-secretase. Nat.
Med. 14(7), 723–730 (2008).
50 Wang KC, Chang HY. Molecular
mechanisms of long noncoding RNAs.
Mol. Cell 43(6), 904–914 (2011).
51 Niland CN, Merry CR, Khalil AM.
Emerging roles for long non-coding RNAs
in cancer and neurological disorders. Front.
Genet. 3, 25 (2012).
52 Mendell JT, Olson EN. MicroRNAs in
stress signaling and human disease. Cell
148 (6), 1172–1187 (2012).
53 Small EM, Olson EN. Pervasive roles of
microRNAs in cardiovascular biology.
Nature 469(7330), 336–342 (2011).
54 Esquela-Kerscher A, Slack FJ. Oncomirs
– microRNAs with a role in cancer. Nat.
Rev. Cancer 6(4), 259–269 (2006).
55 Enfield KS, Pikor LA, Martinez VD, Lam
WL. Mechanistic roles of noncoding RNAs
in lung cancer biology and their clinical
implications. Genet. Res. Int. 2012, 737416
(2012).
miRNAs & long noncoding RNAs as biomarkers in human diseases
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
Expert Rev. Mol. Diagn. 13(2 ), (2013)
198
Review
56 Mitra SA, Mitra AP, Triche TJ. A central
role for long non-coding RNA in cancer.
Front. Genet. 3, 17 (2012).
57 Wapinski O, Chang HY. Long noncoding
RNAs and human disease. Trends Cell Biol.
21(6), 354–361 (2011).
58 Calin GA, Sevignani C, Dumitru CD et al.
Human microRNA genes are frequently
located at fragile sites and genomic regions
involved in cancers. Proc. Natl Acad. Sci.
USA 101(9), 2999–3004 (2004).
59 Rossi S, Sevignani C, Nnadi SC, Siracusa
LD, Calin GA. Cancer-associated genomic
regions (CAGRs) and noncoding RNAs:
bioinformatics and therapeutic
implications. Mamm. Genome 19(7–8 ),
526–540 (2008).
60 Zhang L, Huang J, Yang N et al.
microRNAs exhibit high frequency
genomic alterations in human cancer. Proc.
Natl Acad. Sci. USA 103 (24), 9136–9141
(2006).
61 Esteller M. Epigenetics in cancer. N. Engl.
J. Med. 358(11), 1148–1159 ( 20 08).
62 Feinberg AP, Vogelstein B. Hypomethyla-
tion distinguishes genes of some human
cancers from their normal counterparts.
Nature 301(5895), 89–92 (1983).
63 Lujambio A, Portela A, Liz J et al. CpG
island hypermethylation-associated
silencing of non-coding RNAs transcribed
from ultraconserved regions in human
cancer. Oncogene 29(48), 6390–6401
(2010).
64 Lujambio A, Ropero S, Ballestar E et al.
Genetic unmasking of an epigenetically
silenced microRNA in human cancer cells.
Cancer Res. 67(4), 1424–1429 (2007).
65 Saito Y, Liang G, Egger G et al. Specific
activation of microRNA-127 with
downregulation of the proto-oncogene
BCL6 by chromatin-modifying drugs in
human cancer cells. Cancer Cell 9(6),
435–443 (2006).
66 Brueckner B, Stresemann C, Kuner R et al.
The human let-7a-3 locus contains an
epigenetically regulated microRNA gene
with oncogenic function. Cancer Res. 67(4),
1419–1423 (2007).
67 Pan W, Zhu S, Yuan M et al. Micro-
RNA-21 and microRNA-148a contribute
to DNA hypomethylation in lupus CD4+ T
cells by directly and indirectly targeting
DNA methyltransferase 1. J. Immunol.
184(12), 6773– 6781 (2010).
68 Zhao S, Wang Y, Liang Y et al. Micro-
RNA-126 regulates DNA methylation in
CD4(+) T cells and contributes to systemic
lupus erythematosus by targeting DNA
methyltransferase 1. Arthritis Rheum.
63(5), 1376–1386 (2011).
69 He L, He X, Lim LP et al. A microRNA
component of the p53 tumour suppressor
network. Nature 447(7148), 1130–1134
(2007).
70 Huarte M, Guttman M, Feldser D et al. A
large intergenic noncoding RNA induced
by p53 mediates global gene repression in
the p53 response. Cell 142(3), 409–419
(2010).
71 Barsyte-Lovejoy D, Lau SK, Boutros PC
et al. The c-Myc oncogene directly induces
the H19 noncoding RNA by allele-specific
binding to potentiate tumorigenesis. Cancer
Res. 66(10), 5330–5337 (2006).
72 Chang TC, Yu D, Lee YS et al. Widespread
microRNA repression by Myc contributes
to tumorigenesis. Nat. Genet. 40(1), 43–50
(2008).
73 Wu H, Tao J, Chen PJ et al. Genome-wide
analysis reveals methyl-CpG-binding
protein 2-dependent regulation of
microRNAs in a mouse model of Rett
syndrome. Proc. Natl Acad. Sci. USA
107(42 ), 18161–18166 (2010).
74 Manolio TA, Brooks LD, Collins FS.
A HapMap harvest of insights into the
genetics of common disease. J. Clin. Invest.
118(5), 1590–1605 (2008).
75 Wojcik SE, Rossi S, Shimizu M et al.
Non-coding RNA sequence variations in
human chronic lymphocytic leukemia and
colorectal cancer. Carcinogenesis 31(2),
208–215 (2010).
76 Zorc M, Skok DJ, Godnic I et al. Catalog of
microRNA seed polymorphisms in
vertebrates. PLoS ONE 7(1), e30737 (2012).
77 Melo SA, Esteller M. A precursor
microRNA in a cancer cell nucleus: get me
out of here! Cell Cycle 10(6), 922–925
(2 011) .
78 Bernards R. It’s diagnostics, stupid. Cell
141(1), 13–17 (2010).
79 Clark MB, Johnston RL, Inostroza-Ponta
M et al. Genome-wide analysis of long
noncoding RNA stability. Genome Res.
22(5), 885–898 (2012).
80 Tani H, Mizutani R, Salam KA et al.
Genome-wide determination of RNA
stability reveals hundreds of short-lived
noncoding transcripts in mammals.
Genome Res. 22(5), 947–956 (2012).
81 Gupta R A, Shah N, Wang KC et al. Long
non-coding RNA HOTAIR reprograms
chromatin state to promote cancer
metastasis. Nature 464(7291), 1071–1076
(2010).
82 Kogo R, Shimamura T, Mimori K et al.
Long noncoding RNA HOTAIR regulates
polycomb-dependent chromatin modifica-
tion and is associated with poor prognosis
in colorectal cancers. Cancer Res. 71(20),
6320–6326 (2011).
83 Yang Z, Zhou L, Wu LM et al. Over-
expression of long non-coding RNA
HOTAIR predicts tumor recurrence in
hepatocellular carcinoma patients following
liver transplantation. Ann. Surg. Oncol.
18(5), 1243–1250 (2011).
84 Matouk IJ, Abbasi I, Hochberg A, Galun
E, Dweik H, Akkawi M. Highly upregu-
lated in liver cancer noncoding RNA is
overexpressed in hepatic colorectal
metastasis. Eur. J. Gastroenterol. Hepatol.
21(6), 688–692 (2009).
85 Panzitt K, Tschernatsch MM, Guelly C
et al. Characterization of HULC, a novel
gene with striking up-regulation in
hepatocellular carcinoma, as noncoding
RNA. Gastroenterology 132(1), 330–342
(2007).
86 Tinzl M, Marberger M, Horvath S, Chypre
C. DD3PCA3 RNA analysis in urine –
a new perspective for detecting prostate
cancer. Eur. Urol. 46(2), 182–186;
discussion 187 (2004).
87 Spizzo R, Almeida MI, Colombatti A,
Calin GA. Long non-coding RNAs and
cancer: a new frontier of translational
research? Oncogene 31(43), 4577– 4587
(2012).
88 Reis EM, Verjovski-Almeida S. Perspectives
of long non-coding RNAs in cancer
diagnostics. Front. Genet. 3, 32 (2012).
89 Duttagupta R, Jiang R, Gollub J, Getts
RC, Jones KW. Impact of cellular miRNAs
on circulating miRNA biomarker
signatures. PLoS ONE 6(6), e20769 (2011).
90 Pritchard CC, Kroh E, Wood B et al.
Blood cell origin of circulating
microRNAs: a cautionary note for cancer
biomarker studies. Cancer Prev. Res.
(Phila.) 5(3), 492–497 (2012).
91 Kim DJ, Linnstaedt S, Palma J et al.
Plasma components affect accuracy of
circulating cancer-related microRNA
quantitation. J. Mol. Diagn. 14(1), 71–80
(2012).
92 McDonald JS, Milosevic D, Reddi HV,
Grebe SK, Algeciras-Schimnich A. Analysis
of circulating microRNA: preanalytical
and analytical challenges. Clin. Chem.
57(6), 833–840 (2011).
93 Weber JA, Baxter DH, Zhang S et al. The
microRNA spectrum in 12 body fluids.
Clin. Chem. 56(11), 1733–1741 (2010).
Van Roosbroeck, Pollet & Calin
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
199
www.expert-reviews.com
Review
94 Orozco AF, Lewis DE. Flow cytometric
analysis of circulating microparticles in
plasma. Cytometry. A 77(6), 502–514
(2010).
95 Arroyo JD, Chevillet JR, Kroh EM et al.
Argonaute2 complexes carry a population
of circulating microRNAs independent of
vesicles in human plasma. Proc. Natl
Acad. Sci. USA 108(12), 5003–5008
(2 011) .
96 Turchinovich A, Weiz L, Langheinz A,
Burwinkel B. Characterization of
extracellular circulating microRNA.
Nucleic Acids Res. 39(16), 7223–7233
(2 011) .
97 Wang K, Zhang S, Weber J, Baxter D,
Galas DJ. Export of microRNAs and
microRNA-protective protein by mamma-
lian cells. Nucleic Acids Res. 38(20),
7248 –7259 (2010).
98 Vickers KC, Palmisano BT, Shoucri BM,
Shamburek RD, Remaley AT. Micro RNAs
are transported in plasma and delivered to
recipient cells by high-density lipopro-
teins. Nat. Cell Biol. 13(4), 423–433
(2 011) .
99 Valadi H, Ekström K, Bossios A, Sjöstrand
M, Lee JJ, Lötvall JO. Exosome-mediated
transfer of mRNAs and microRNAs is a
novel mechanism of genetic exchange
between cells. Nat. Cell Biol. 9(6),
654659 (2007).
100 van der Vos KE, Balaj L, Skog J,
Breakefield XO. Brain tumor microvesicles:
insights into intercellular communication
in the nervous system. Cell. Mol. Neurobiol.
31(6), 949–959 (2011).
101 Nilsson J, Skog J, Nordstrand A et al.
Prostate cancer-derived urine exosomes: a
novel approach to biomarkers for prostate
cancer. Br. J. Cancer 100 (10), 1603–1607
(2009).
102 Gibb EA, Vucic EA, Enfield KS et al.
Human cancer long non-coding RNA
transcriptomes. PLoS ONE 6(10), e25915
(2 011) .
103 Lorenzen JM, Thum T. Circulating and
urinary microRNAs in kidney disease.
Clin. J. Am. Soc. Nephrol. 7(9) , 1528–1533
(2012).
104 Paraskevi A, Theodoropoulos G,
Papaconstantinou I, Mantzaris G, Nikiteas
N, Gazouli M. Circulating MicroRNA in
inflammatory bowel disease. J. Crohns.
Colitis 6(9), 900–904 (2012).
105 Wang G, Tam LS, Li EK et al. Serum and
urinary free microRNA level in patients
with systemic lupus erythematosus. Lupus
20(5), 493–500 (2011).
106 Siegel SR, Mackenzie J, Chaplin G,
Jablonski NG, Griffiths L. Circulating
microRNAs involved in multiple sclerosis.
Mol. Biol. Rep. 39(5), 6219–6225 (2012).
107 Wu L, Zhou H, Lin H et al. Circulating
microRNAs are elevated in plasma from
severe preeclamptic pregnancies.
Reproduction 143(3), 389–397 (2012).
108 Fu Y, Yi Z, Wu X, Li J, Xu F. Circulating
microRNAs in patients with active
pulmonary tuberculosis. J. Clin. Microbiol.
49(12), 4246– 4251 (2011).
109 Cermelli S, Ruggieri A, Marrero JA,
Ioannou GN, Beretta L. Circulating
microRNAs in patients with chronic
hepatitis C and non-alcoholic fatty liver
disease. PLoS ONE 6(8), e23937 (2011).
110 Ji F, Yang B, Peng X, Ding H, You H, Tien
P. Circulating microRNAs in hepatitis B
virus-infected patients. J. Viral Hepat.
18(7), e242–e251 (2011).
111 de Planell-Saguer M, Rodicio MC.
Analytical aspects of microRNA in
diagnostics: a review. Anal. Chim. Acta
699(2), 134–152 (2011).
112 Küppers R. The biology of Hodgkin’s
lymphoma. Nat. Rev. Cancer 9(1), 15–27
(2009).
113 Defrancesco L. Life Technologies promises
$1,000 genome. Nat. Biotechnol. 30(2),
126 (2012).
114 Salminen WF, Yang X, Shi Q, Mendrick
D. Using microRNA as biomarkers of
drug-induced liver injury. J. Mol. Biomark.
Diagn. 2 (6 ), 119 (2011).
115 Calin GA, Croce CM. MicroRNA
signatures in human cancers. Nat. Rev.
Cancer 6(11), 857–866 (2006).
116 Halvorsen M, Martin JS, Broadaway S,
Laederach A. Disease-associated mutations
that alter the RNA structural ensemble.
PLoS Genet. 6(8), e1001074 (2010).
117 Richardson K, Lai CQ, Parnell LD, Lee
YC, Ordovas JM. A genome-wide survey
for SNPs altering microRNA seed sites
identifies functional candidates in GWAS.
BMC Genomics 12, 504 (2011).
118 Shen J, Ambrosone CB, Zhao H. Novel
genetic variants in microRNA genes and
familial breast cancer. Int. J. Cancer 124(5),
1178 –1182 (2009) .
119 Srivastava K, Srivastava A, Mittal B.
Common genetic variants in
pre-microRNAs and risk of gallbladder
cancer in North Indian population.
J. Hum. Genet. 55(8), 495 499 (2010).
120 Xing J, Wan S, Zhou F et al. Genetic
polymorphisms in pre-microRNA genes as
prognostic markers of colorectal cancer.
Cancer Epidemiol. Biomarkers Prev. 21(1),
217–227 (2012).
121 Xu B, Feng NH, Li PC et al. A functional
polymorphism in pre-miR-146a gene is
associated with prostate cancer risk and
mature miR-146a expression in vivo.
Prostate 70(5), 467–472 (2010).
122 Yang R, Schlehe B, Hemminki K et al.
A genetic variant in the pre-miR-27a
oncogene is associated with a reduced
familial breast cancer risk. Breast Cancer
Res. Trea t. 121(3), 693–702 (2010).
123 Zeng Y, Sun QM, Liu NN et al. Correla-
tion between pre-miR-146a C/G polymor-
phism and gastric cancer risk in Chinese
population. World J. Gastroenterol. 16(28),
3578–3583 (2010).
124 Luo X, Yang W, Ye DQ et al. A functional
variant in microRNA-146a promoter
modulates its expression and confers
disease risk for systemic lupus
erythematosus. PLoS Genet. 7(6),
e1002128 (2011).
125 Xu J, Hu Z, Xu Z et al. Functional variant
in microRNA-196a2 contributes to the
susceptibility of congenital heart disease in
a Chinese population. Hum. Mutat. 30(8),
1231–1236 (2009).
126 Xu Y, Li F, Zhang B et al. MicroRNAs and
target site screening reveals a pre-
microRNA-30e variant associated with
schizophrenia. Schizophr. Res. 119(1–3),
219–227 (2010).
127 Zhou B, Rao L, Peng Y et al. Common
genetic polymorphisms in pre-microRNAs
were associated with increased risk of
dilated cardiomyopathy. Clin. Chim. Acta
411(17–18), 1287–1290 (2010).
128 Blitzblau RC, Weidhaas JB. MicroRNA
binding-site polymorphisms as potential
biomarkers of cancer risk. Mol. Diagn.
Ther. 14(6), 335–342 (2010).
129 Horikawa Y, Wood CG, Yang H et al.
Single nucleotide polymorphisms of
microRNA machinery genes modify the
risk of renal cell carcinoma. Clin. Cancer
Res. 14(23), 7956–7962 (2008).
130 Rotunno M, Zhao Y, Bergen AW et al.
Inherited polymorphisms in the RNA-
mediated interference machinery affect
microRNA expression and lung cancer
survival. Br. J. Cancer 103(12), 1870–1874
(2010).
131 Bruno AE, Li L, Kalabus JL, Pan Y, Yu A,
Hu Z. miRdSNP: a database of disease-
associated SNPs and microRNA target sites
on 3UTRs of human genes. BMC
Genomics 13, 44 (2012).
miRNAs & long noncoding RNAs as biomarkers in human diseases
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
Expert Rev. Mol. Diagn. 13(2 ), (2013)
200
Review
132 Yang R, Frank B, Hemminki K et al. SNPs
in ultraconserved elements and familial
breast cancer risk. Carcinogenesis 29(2),
351–355 (2008).
133 Broadbent HM, Peden JF, Lorkowski S
et al.; PROCARDIS Consortium.
Susceptibility to coronary artery disease
and diabetes is encoded by distinct, tightly
linked SNPs in the ANRIL locus on
chromosome 9p. Hum. Mol. Genet. 17(6),
806–814 (2008).
134 Burd CE, Jeck WR, Liu Y, Sanoff HK,
Wang Z, Sharpless NE. Expression of
linear and novel circular forms of an INK4/
ARF-associated non-coding RNA
correlates with atherosclerosis risk. PLoS
Genet. 6(12), e1001233 (2010) .
135 Cunnington MS, Santibanez Koref M,
Mayosi BM, Burn J, Keavney B. Chromo-
some 9p21 SNPs associated with multiple
disease phenotypes correlate with ANRIL
expression. PLoS Genet. 6(4), e1000899
(2010).
136 Iacobucci I, Sazzini M, Garagnani P et al.
A polymorphism in the chromosome 9p21
ANRIL locus is associated to Philadelphia
positive acute lymphoblastic leukemia.
Leuk. Res. 35(8), 1052–1059 (2011).
137 Liu Y, Pan S, Liu L et al. A genetic variant
in long non-coding RNA HULC
contributes to risk of HBV-related
hepatocellular carcinoma in a Chinese
population. PLoS ONE 7(4), e35145
(2012).
138 Pal P, Xi H, Guha S et al. Common
variants in 8q24 are associated with risk for
prostate cancer and tumor aggressiveness in
men of European ancestry. Prostate 69(14),
1548 –1556 (2 009 ).
139 Salinas CA, Kwon E, Carlson CS et al.
Multiple independent genetic variants in
the 8q24 region are associated with prostate
cancer risk. Cancer Epidemiol. Biomarkers
Prev. 17(5), 1203–1213 (2008).
140 Prensner JR, Iyer MK, Balbin OA et al.
Transcriptome sequencing across a prostate
cancer cohort identifies PCAT-1, an
unannotated lincRNA implicated in
disease progression. Nat. Biotechnol. 29(8),
742 –749 (2011) .
141 Chung S, Nakagawa H, Uemura M et al.
Association of a novel long non-coding RNA
in 8q24 with prostate cancer susceptibility.
Cancer Sci. 102(1), 245–252 (2011).
142 Sonkoly E, Bata-Csorgo Z, Pivarcsi A et al.
Identification and characterization of a
novel, psoriasis susceptibility-related
noncoding RNA gene, PRINS. J. Biol.
Chem. 280(25), 24159–24167 (2005).
143 Shirasawa S, Harada H, Furugaki K et al.
SNPs in the promoter of a B cell-specific
antisense transcript, SAS-ZFAT, determine
susceptibility to autoimmune thyroid
disease. Hum. Mol. Genet. 13(19),
2221–2231 (2004).
144 Bussemakers MJ, van Bokhoven A,
Verhaegh GW et al. DD3: a new prostate-
specific gene, highly overexpressed in
prostate cancer. Cancer Res. 59(23),
5975–5979 (1999).
145 Wang XS, Zhang Z, Wang HC et al. Rapid
identification of UCA1 as a very sensitive
and specific unique marker for human
bladder carcinoma. Clin. Cancer Res.
12(16), 4851–4858 (2006).
146 Lu J, Getz G, Miska EA et al. MicroRNA
expression profiles classify human cancers.
Nature 435(7043), 834–838 (2005).
147 Rosenfeld N, Aharonov R, Meiri E et al.
MicroRNAs accurately identify cancer
tissue origin. Nat. Biotechnol. 26(4),
462–469 (2008).
148 Blenkiron C, Goldstein LD, Thorne NP
et al. MicroRNA expression profiling of
human breast cancer identifies new markers
of tumor subtype. Genome Biol. 8(10),
R214 (2007) .
149 Ueda T, Volinia S, Okumura H et al.
Relation between microRNA expression
and progression and prognosis of gastric
cancer: a microRNA expression analysis.
Lancet Oncol. 11(2), 136–146 (2010).
150 Garzon R, Garofalo M, Martelli MP
et al. Distinctive microRNA signature of
acute myeloid leukemia bearing cytoplas-
mic mutated nucleophosmin. Proc. Natl
Acad. Sci. USA 105 (10), 3945–3950
(2008).
151 Navarro A, Gaya A, Martinez A et al.
MicroRNA expression profiling in classic
Hodgkin lymphoma. Blood 111(5) ,
2825–2832 (2008).
152 Dacic S, Kelly L, Shuai Y, Nikiforova MN.
miRNA expression profiling of lung
adenocarcinomas: correlation with
mutational status. Mod. Pathol. 23(12),
157 7–1582 (2010).
153 Seike M, Goto A, Okano T et al. MiR-21 is
an EGFR-regulated anti-apoptotic factor in
lung cancer in never-smokers. Proc. Natl
Acad. Sci. USA 106(29), 12085–12090
(2009).
154 Chan AS, Thorner PS, Squire JA, Zielenska
M. Identification of a novel gene NCRMS
on chromosome 12q21 with differential
expression between rhabdomyosarcoma
subtypes. Oncogene 21(19), 3029–3037
(2002).
155 Calin GA, Ferracin M, Cimmino A et al.
A MicroRNA signature associated with
prognosis and progression in chronic
lymphocytic leukemia. N. Engl. J. Med.
353(17), 1793–1801 (2005).
156 Calin GA, Liu CG, Sevignani C et al.
MicroRNA profiling reveals distinct
signatures in B cell chronic lymphocytic
leukemias. Proc. Natl Acad. Sci. USA
101(32 ), 11755–11760 (2004 ).
157 Lin R, Maeda S, Liu C, Karin M,
Edgington TS. A large noncoding RNA is
a marker for murine hepatocellular
carcinomas and a spectrum of human
carcinomas. Oncogene 26(6), 851–858
(2007).
158 Yamada K, Kano J, Tsunoda H et al.
Phenotypic characterization of endometrial
stromal sarcoma of the uterus. Cancer Sci.
97(2), 106–112 (2006).
159 Ji P, Diederichs S, Wang W et al.
MALAT-1, a novel noncoding RNA, and
thymosin β4 predict metastasis and survival
in early-stage non-small cell lung cancer.
Oncogene 22(39), 8031–8041 (2003).
160 Schmidt LH, Spieker T, Koschmieder S
et al. The long noncoding MALAT-1 RNA
indicates a poor prognosis in non-small cell
lung cancer and induces migration and
tumor growth. J. Thorac. Oncol. 6(12),
1984–1992 (2011) .
161 Lai MC, Yang Z, Zhou L et al. Long
non-coding RNA MALAT-1 overexpres-
sion predicts tumor recurrence of
hepatocellular carcinoma after liver
transplantation. Med. Oncol. 29(3),
1810–1816 (2012).
162 Scaruffi P, Stigliani S, Moretti S et al.
Transcribed-ultra conserved region
expression is associated with outcome in
high-risk neuroblastoma. BMC Cancer 9,
441 (2009).
163 Nicoloso MS, Spizzo R, Shimizu M, Rossi
S, Calin GA. MicroRNAs – the micro
steering wheel of tumour metastases. Nat.
Rev. Cancer 9(4), 293–302 (2009).
164 Eccles SA, Welch DR. Metastasis: recent
discoveries and novel treatment strategies.
Lancet 369(9574), 1742–1757 (2007).
165 Geng YJ, Xie SL, Li Q, Ma J, Wang GY.
Large intervening non-coding RNA
HOTAIR is associated with hepatocellular
carcinoma progression. J. Int. Med. Res.
39 (6), 2119–2128 (2011).
166 Niinuma T, Suzuki H, Nojima M et al.
Upregulation of miR-196a and HOTAIR
drive malignant character in
gastrointestinal stromal tumors. Cancer
Res. 72 (5), 112 6–113 6 ( 2012).
Van Roosbroeck, Pollet & Calin
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
201
www.expert-reviews.com
Review
167 Xu C, Yang M, Tian J, Wang X, Li Z.
MALAT-1: a long non-coding RNA and its
important 3 end functional motif in
colorectal cancer metastasis. Int. J. Oncol.
39(1), 169–175 (2011).
168 Tahira AC, Kubrusly MS, Faria MF et al.
Long noncoding intronic RNAs are
differentially expressed in primary and
metastatic pancreatic cancer. Mol. Cancer
10, 141 (2011).
169 Permuth-Wey J, Thompson RC, Burton
Nabors L et al. A functional polymorphism
in the pre-miR-146a gene is associated with
risk and prognosis in adult glioma.
J. Neurooncol. 105(3), 639–646 (2011).
170 Luo J, Cai Q, Wang W et al. A micro-
RNA-7 binding site polymorphism in
HOXB5 leads to differential gene
expression in bladder cancer. PLoS ONE
7(6), e40127 (2012).
171 Smits KM, Paranjape T, Nallur S et al.
A let-7 microRNA SNP in the KRAS
3UTR is prognostic in early-stage
colorectal cancer. Clin. Cancer Res. 17(24),
7723–7731 (2011).
172 Hu Z, Shu Y, Chen Y et al. Genetic
polymorphisms in the precursor microRNA
flanking region and non-small cell lung
cancer survival. Am. J. Respir. Crit. Care
Med. 183(5), 641–648 (2011).
173 Fojo T. Multiple paths to a drug resistance
phenotype: mutations, translocations,
deletions and amplification of coding genes
or promoter regions, epigenetic changes
and microRNAs. Drug Resist. Updat.
10(1–2), 59–67 (2007).
174 Ma J, Dong C, Ji C. MicroRNA and drug
resistance. Cancer Gene Ther. 17(8),
523 –531 (2010).
175 Kumar S, Kumar A, Shah PP, Rai SN,
Panguluri SK, Kakar SS. MicroRNA
signature of cis-platin resistant vs. cis-platin
sensitive ovarian cancer cell lines.
J. Ovarian Res. 4(1), 17 (2011).
176 Preis M, Gardner TB, Gordon SR et al.
MicroRNA-10b expression correlates with
response to neoadjuvant therapy and
survival in pancreatic ductal
adenocarcinoma. Clin. Cancer Res. 17(17),
5812–5821 (2011).
177 Nishida N, Yamashita S, Mimori K et al.
MicroRNA-10b is a prognostic indicator in
colorectal cancer and confers resistance to
the chemotherapeutic agent 5-fluorouracil
in colorectal cancer cells. Ann. Surg. Oncol.
19(9), 3065–3071 (2012).
178 Giovannetti E, Funel N, Peters GJ et al.
MicroRNA-21 in pancreatic cancer:
correlation with clinical outcome and
pharmacologic aspects underlying its role
in the modulation of gemcitabine activity.
Cancer Res. 70(11), 4528 –4538 (2010).
179 Hwang JH, Voortman J, Giovannetti E
et al. Identification of microRNA-21 as a
biomarker for chemoresistance and clinical
outcome following adjuvant therapy in
resectable pancreatic cancer. PLoS ONE
5(5), e10630 (2010).
180 Moriyama T, Ohuchida K, Mizumoto K
et al. MicroRNA-21 modulates biological
functions of pancreatic cancer cells
including their proliferation, invasion, and
chemoresistance. Mol. Cancer Ther. 8(5),
1067–1074 (2009).
181 Shi GH, Ye DW, Yao XD et al. Involve-
ment of microRNA-21 in mediating
chemo-resistance to docetaxel in androgen-
independent prostate cancer PC3 cells. Acta
Pharmacol. Sin. 31(7), 867–873 (2010).
182 Li Y, Li W, Yang Y et al. MicroRNA-21
targets LRRFIP1 and contributes to
VM-26 resistance in glioblastoma
multiforme. Brain Res. 1286, 13–18
(2009).
183 Tao J, Lu Q, Wu D et al. microRNA-21
modulates cell proliferation and sensitivity
to doxorubicin in bladder cancer cells.
Oncol. Rep. 25(6), 1721–1729 ( 2011).
184 Bourguignon LY, Earle C, Wong G, Spevak
CC, Krueger K. Stem cell marker (Nanog)
and Stat-3 signaling promote micro-
RNA-21 expression and chemoresistance in
hyaluronan/CD44-activated head and neck
squamous cell carcinoma cells. Oncogene
31(2), 149–160 (2012).
185 Jung EJ, Santarpia L, Kim J et al. Plasma
microRNA 210 levels correlate with
sensitivity to trastuzumab and tumor
presence in breast cancer patients. Cancer
118(10), 2603–2614 (2012).
186 Wang H, Tan G, Dong L et al. Circulating
miR-125b as a marker predicting
chemoresistance in breast cancer. PLoS
ONE 7(4), e34210 (2012).
187 Zhou M, Liu Z, Zhao Y et al. MicroRNA-
125b confers the resistance of breast cancer
cells to paclitaxel through suppression of
pro-apoptotic Bcl-2 antagonist killer 1
(Bak1) expression. J. Biol. Chem. 285(28),
21496 –21507 (2010).
188 Kong F, Sun C, Wang Z et al. miR-125b
confers resistance of ovarian cancer cells to
cisplatin by targeting pro-apoptotic Bcl-2
antagonist killer 1. J. Huazhong Univ. Sci.
Technol. Med. Sci. 31(4), 543–549 (2011).
189 Shi L, Zhang S, Feng K et al. MicroRNA-
125b-2 confers human glioblastoma stem
cells resistance to temozolomide through
the mitochondrial pathway of apoptosis.
Int. J. Oncol. 40(1), 119–129 (2012).
190 Zhang H, Luo XQ, Feng DD et al.
Upregulation of microRNA-125b
contributes to leukemogenesis and increases
drug resistance in pediatric acute
promyelocytic leukemia. Mol. Cancer 10,
108 (2011).
191 Weiss GJ, Bemis LT, Nakajima E et al.
EGFR regulation by microRNA in lung
cancer: correlation with clinical response
and survival to gefitinib and EGFR
expression in cell lines. Ann. Oncol. 19(6 ),
1053–1059 (2008).
192 Fellenberg J, Bernd L, Delling G, Witte D,
Zahlten-Hinguranage A. Prognostic
significance of drug-regulated genes in
high-grade osteosarcoma. Mod. Pathol.
20 (10), 1085 –1094 (2007).
193 Tsang WP, Wong TW, Cheung AH, Co
CN, Kwok TT. Induction of drug
resistance and transformation in human
cancer cells by the noncoding RNA
CUDR. RNA 13(6), 890–898 (2007).
194 Tsang WP, Kwok TT. Riboregulator H19
induction of MDR1-associated drug
resistance in human hepatocellular
carcinoma cells. Oncogene 26(33),
4877–4881 (2007).
195 Ren Y, Zhou X, Mei M et al. Micro-
RNA-21 inhibitor sensitizes human
glioblastoma cells U251 (PTEN-mutant)
and LN229 (PTEN-wild type) to taxol.
BMC Cancer 10, 27 (2010).
196 Zhang S, Wan Y, Pan T et al. Micro-
RNA-21 inhibitor sensitizes human
glioblastoma U251 stem cells to chemo-
therapeutic drug temozolomide. J. Mol.
Neurosci. 47(2), 346–356 (2012).
197 Teo MT, Landi D, Taylor CF et al. The
role of microRNA-binding site polymor-
phisms in DNA repair genes as risk
factors for bladder cancer and breast
cancer and their impact on radiotherapy
outcomes. Carcinogenesis 33(3), 581–586
(2012).
198 Mishra PJ, Humeniuk R, Mishra PJ,
Longo-Sorbello GS, Banerjee D, Bertino
JR. A miR-24 microRNA binding-site
polymorphism in dihydrofolate reductase
gene leads to methotrexate resistance. Proc.
Natl Acad. Sci. USA 104(33), 13513–13518
(2007).
199 Wu Y, Xiao Y, Ding X et al. A
miR-200b/200c/429-binding site
polymorphism in the 3 untranslated region
of the AP-2a gene is associated with
cisplatin resistance. PLoS ONE 6(12),
e29043 (2011).
miRNAs & long noncoding RNAs as biomarkers in human diseases
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
Expert Rev. Mol. Diagn. 13(2 ), (2013)
202
Review
200 Boni V, Zarate R, Villa JC et al. Role of
primary miRNA polymorphic variants in
metastatic colon cancer patients treated
with 5-fluorouracil and irinotecan.
Pharmacogenomics J. 11(6), 429–436
(2 011) .
201 Global Atlas on Cardiovascular Disease
Prevention and Control. Mendis S, Puska P,
Norrving B (Eds). WHO, Geneva,
Switzerland (2011).
202 Roger VL, Go AS, Lloyd-Jones DM et al.;
American Heart Association Statistics
Committee and Stroke Statistics
Subcommittee. Executive summary: heart
disease and stroke statistics – 2012
update: a report from the American Heart
Association. Circulation 125(1), 188–197
(2012).
203 Catalucci D, Gallo P, Condorelli G.
MicroRNAs in cardiovascular biology and
heart disease. Circ. Cardiovasc. Genet. 2(4),
402–408 (2009).
204 Chen JF, Murchison EP, Tang R et al.
Targeted deletion of Dicer in the heart
leads to dilated cardiomyopathy and heart
failure. Proc. Natl Acad. Sci. USA 105(6),
2111–2116 (2008) .
205 Hulsmans M, Sinnaeve P,
Van der Schueren B, Mathieu C, Janssens
S, Holvoet P. Decreased miR-181a
expression in monocytes of obese patients is
associated with the occurrence of metabolic
syndrome and coronary artery disease.
J. Clin. Endocrinol. Metab. 97(7),
E1213–E1218 (2012).
206 Satoh M, Minami Y, Takahashi Y, Tabuchi
T, Nakamura M. A cellular microRNA,
let-7i, is a novel biomarker for clinical
outcome in patients with dilated cardio-
myopathy. J. Card. Fail. 17(11), 923 –929
(2 011) .
207 Weng H, Shen C, Hirokawa G et al.
Plasma miR-124 as a biomarker for cerebral
infarction. Biomed. Res. 32(2), 135–141
(2 011) .
208 Widera C, Gupta SK, Lorenzen JM et al.
Diagnostic and prognostic impact of six
circulating microRNAs in acute coronary
syndrome. J. Mol. Cell. Cardiol. 51(5) ,
872–875 (2011).
209 Xiao J, Jing ZC, Ellinor PT et al.
MicroRNA-134 as a potential plasma
biomarker for the diagnosis of acute pulmo-
nary embolism. J. Transl. Med. 9, 159
(2 011) .
210 Zeng L, Liu J, Wang Y et al.
MicroRNA-210 as a novel blood biomarker
in acute cerebral ischemia. Front. Biosci.
(Elite Ed.) 3, 1265–1272 (2011).
211 Kawashima T, Shioi T. MicroRNA,
emerging role as a biomarker of heart
failure. Circ. J. 75(2), 268–269 (2011).
212 Holdt LM, Teupser D. Recent studies of
the human chromosome 9p21 locus, which
is associated with atherosclerosis in human
populations. Arterioscler. Thromb. Vasc.
Biol. 32(2), 196–206 (2012).
213 Holdt LM, Beutner F, Scholz M et al.
ANRIL expression is associated with
atherosclerosis risk at chromosome 9p21.
Arterioscler. Thromb. Vasc. Biol. 30(3),
620– 627 (2010).
214 Liu Y, Sanoff HK, Cho H et al. INK4/ARF
transcript expression is associated with
chromosome 9p21 variants linked to
atherosclerosis. PLoS ONE 4(4), e5027
(2009).
215 Zhang W, Chen Y, Liu P et al. Variants on
chromosome 9p21.3 correlated with
ANRIL expression contribute to stroke risk
and recurrence in a large prospective stroke
population. Stroke 43(1), 14–21 (2012).
216 Ishii N, Ozaki K, Sato H et al. Identifica-
tion of a novel non-coding RNA, MIAT,
that confers risk of myocardial infarction.
J. Hum. Genet. 51(12), 1087–1099 (2006 ).
217 Furer V, Greenberg JD, Attur M,
Abramson SB, Pillinger MH. The role of
microRNA in rheumatoid arthritis and
other autoimmune diseases. Clin. Immunol.
136(1), 1–15 (2010).
218 Pauley KM, Cha S, Chan EK. MicroRNA
in autoimmunity and autoimmune
diseases. J. Autoimmun. 32(3–4), 189–194
(2009).
219 Dai R, Ahmed SA. MicroRNA, a
new paradigm for understanding
immunoregulation, inflammation, and
autoimmune diseases. Tr ans l. Res. 157 (4 ),
163–179 (2011).
220 Haywood ME, Rose SJ, Horswell S et al.
Overlapping BXSB congenic intervals, in
combination with microarray gene
expression, reveal novel lupus candidate
genes. Genes Immun. 7(3), 250–263
(2006).
221 Johanneson B, Lima G, von Salomé J,
Alarcón-Segovia D, Alarcón-Riquelme
ME; Collaborative Group on the Genetics
of SLE, The BIOMED II Collaboration on
the Genetics of SLE and Sjögrens
syndrome. A major susceptibility locus for
systemic lupus erythematosus maps to
chromosome 1q31. Am. J. Hum. Genet.
71(5), 1060–1071 (2002).
222 Tsao BP. The genetics of human systemic
lupus erythematosus. Trends Immunol.
24(11), 595–602 (2003).
223 Tsao BP. Update on human systemic lupus
erythematosus genetics. Curr. Opin.
Rheumatol. 16(5), 513–521 (2004).
224 Cao X, Yeo G, Muotri AR, Kuwabara T,
Gage FH. Noncoding RNAs in the
mammalian central nervous system. Annu.
Rev. Neurosci. 29, 77–103 (2006).
225 Omahen DA. MicroRNA and diseases of
the nervous system. Neurosurgery 69(2),
440–454 (2011).
226 Mercer TR, Dinger ME, Sunkin SM,
Mehler MF, Mattick JS. Specific expression
of long noncoding RNAs in the mouse
brain. Proc. Natl Acad. Sci. USA 105(2),
716–721 (2008).
227 Johnson R. Long non-coding RNAs in
Huntington’s disease neurodegeneration.
Neurobiol. Dis. 46(2), 245–254 (2012).
228 Kohtz JD, Berghoff EG. Regulatory long
non-coding RNAs and neuronal disorders.
Physiol. Behav. 100(3), 250 –254 (2010).
229 Qureshi IA, Mattick JS, Mehler MF. Long
non-coding RNAs in nervous system
function and disease. Brain Res. 1338,
20–35 (2010).
230 Tan H, Xu Z, Jin P. Role of noncoding
RNAs in trinucleotide repeat neurodegen-
erative disorders. Exp. Neurol. 235(2),
469 –475 (2012 ).
231 Decourt B, Sabbagh MN. BACE1 as a
potential biomarker for Alzheimer’s disease.
J. Alzheimers Dis. 24(Suppl. 2), 53–59
(2 011) .
232 Blennow K, Hampel H, Weiner M,
Zetterberg H. Cerebrospinal fluid and
plasma biomarkers in Alzheimer disease.
Nat. Rev. Neurol. 6(3), 131–144 (2010).
233 Cogswell JP, Ward J, Taylor IA et al.
Identification of miRNA changes in
Alzheimer’s disease brain and CSF yields
putative biomarkers and insights into
disease pathways. J. Alzheimers Dis. 14(1),
27–41 (2008).
234 Ghidoni R, Benussi L, Paterlini A,
Albertini V, Binetti G, Emanuele E.
Cerebrospinal fluid biomarkers for
Alzheimer’s disease: the present and the
future. Neurodegener. Dis. 8(6), 413–420
(2 011) .
235 Schipper HM, Maes OC, Chertkow HM,
Wang E. MicroRNA expression in
Alzheimer blood mononuclear cells. Gene
Regul. Syst. Bio. 1, 263–274 (2007).
236 Schütz S, Sarnow P. Interaction of viruses
with the mammalian RNA interference
pathway. Virology 344(1), 151–157 (2006).
237 Roberts AP, Lewis AP, Jopling CL. The
role of microRNAs in viral infection. Prog.
Mol. Biol. Transl. Sci. 102, 101–139 (2011).
Van Roosbroeck, Pollet & Calin
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
203
www.expert-reviews.com
Review
238 Sun G, Rossi JJ. MicroRNAs and their
potential involvement in HIV infection.
Trends Pharmacol. Sci. 32(11), 675–681
(2 011) .
239 Sun R, Lin SF, Gradoville L, Miller G.
Polyadenylated nuclear RNA encoded by
Kaposi sarcoma-associated herpesvirus.
Proc. Natl Acad. Sci. USA 93(21),
11883–11888 (1996).
240 Zhong W, Ganem D. Characterization of
ribonucleoprotein complexes containing an
abundant polyadenylated nuclear RNA
encoded by Kaposi’s sarcoma-associated
herpesvirus (human herpesvirus 8).
J. Virol. 71(2), 1207–1212 (1997).
241 Peng X, Gralinski L, Armour CD et al.
Unique signatures of long noncoding RNA
expression in response to virus infection
and altered innate immune signaling.
MBio 1(5), pii: e00206-10 (2010).
242 Isakov O, Ronen R, Kovarsky J et al. Novel
insight into the non-coding repertoire
through deep sequencing analysis. Nucleic
Acids Res. 40(11), e86 (2012).
243 Galasso M, Sana ME, Volinia S. Non-
coding RNAs: a key to future personalized
molecular therapy? Genome Med. 2(2), 12
(2010).
244 Elmén J, Lindow M, Schütz S et al.
LNA-mediated microRNA silencing in
non-human primates. Nature 452(7189),
896–899 (2008).
245 Esau CC. Inhibition of microRNA with
antisense oligonucleotides. Methods 44(1),
55–60 (2008).
246 Krützfeldt J, Rajewsky N, Braich R et al.
Silencing of microRNAs in vivo with
‘a nt agomi rs ’. Nature 438(7068), 685–689
(2005).
247 Ebert MS, Neilson JR, Sharp PA.
MicroRNA sponges: competitive inhibitors
of small R NAs in mammalian cells. Nat.
Methods 4(9), 721–726 (2007).
248 Deiters A. Small molecule modifiers of the
microRNA and RNA interference pathway.
AAPS J. 12(1), 51–60 (2010).
249 Bader AG, Brown D, Winkler M. The
promise of microRNA replacement
therapy. Cancer Res. 70(18), 7027–7030
(2010).
250 Hébert SS, De Strooper B. Alterations of
the microRNA network cause neurodegen-
erative disease. Trends Neurosci. 32(4),
199–206 (2009).
251 Lanford RE, Hildebrandt-Eriksen ES, Petri
A et al. Therapeutic silencing of micro-
RNA-122 in primates with chronic
hepatitis C virus infection. Science
327(5962), 198–201 (2010).
252 Janssen HL, Reesink HW, Zeuzem S et al.
A randomized, double-blind, placebo
(PLB) controlled safety and anti-viral proof
of concept study of miravirsen (MIR), an
oligonucleotide targeting miR-122, in
treatment naive patients with genotype 1
(GT1) chronic HCV infection. Hepatology
54(S1), 1430A (2011).
253 Fiedler J, Gupta SK, Thum T. MicroRNA-
based therapeutic approaches in the
cardiovascular system. Cardiovasc. Ther.
30(1), e9–e15 (2012).
254 Garzon R, Marcucci G, Croce CM.
Targeting microRNAs in cancer: rationale,
strategies and challenges. Nat. Rev. Drug
Discov. 9(10), 775–789 (2010).
255 Ruberti F, Barbato C, Cogoni C. Targeting
microRNAs in neurons: tools and
perspectives. Exp. Neurol. 235(2), 419–426
(2012).
256 Stenvang J, Kauppinen S. MicroRNAs as
targets for antisense-based therapeutics.
Expert Opin. Biol. Ther. 8(1), 59 –81
(2008).
257 Thum T. MicroRNA therapeutics in
cardiovascular medicine. EMBO Mol. Med.
4(1), 3–14 (2012).
258 Goldstein DB. Growth of genome
screening needs debate. Nature 476(7358),
27–28 (2 011) .
259 Aharonov RT, Rosenwald S, Edmonston
TB et al. A second-generation microRNA-
based assay for diagnosing tumor tissue
origin. Presented at: ASCO Meeting.
29(Suppl. 15), Abstract 10575 (2011).
260 Edmonston TB, Gibori H, Kushnir M
et al. New microRNA-based diagnostic test
for lung cancer classification. Presented at:
ASCO Meeting. 29(Suppl. 15), Abstract
10531 (2011).
261 Bishop JA, Benjamin H, Cholakh H,
Chajut A, Clark DP, Westra WH. Accurate
classification of non-small cell lung
carcinoma using a novel microRNA-based
approach. Clin. Cancer Res. 16(2 ), 610–619
(2010).
262 Lebanony D, Benjamin H, Gilad S et al.
Diagnostic assay based on hsa-miR-205
expression distinguishes squamous from
nonsquamous non-small-cell lung
carcinoma. J. Clin. Oncol. 27(12),
2030–2037 (2009).
263 Benjamin H, Lebanony D, Rosenwald S
et al. A diagnostic assay based on micro-
RNA expression accurately identifies
malignant pleural mesothelioma. J. Mol.
Diagn. 12(6), 771–779 (2010).
264 Fridman E, Dotan Z, Barshack I et al.
Accurate molecular classification of renal
tumors using microRNA expression.
J. Mol. Diagn. 12(5), 687–696 (2010).
265 De La Taille A. Progensa PCA3 test for
prostate cancer detection. Expert Rev. Mol.
Diagn. 7(5), 491–497 (2007).
266 Lau NC, Lim LP, Weinstein EG, Bartel DP.
An abundant class of tiny RNAs with
probable regulatory roles in Caenorhabditis
elegans. Science 294(5543), 858–862 (2001).
267 Lee EJ, Baek M, Gusev Y, Brackett DJ,
Nuovo GJ, Schmittgen TD. Systematic
evaluation of microRNA processing
patterns in tissues, cell lines, and tumors.
RNA 14(1), 35–42 (2008).
268 Gupta A, Mo YY. Detection of microRNAs
in cultured cells and paraffin-embedded
tissue specimens by in situ hybridization.
Methods Mol. Biol. 676, 73–83 (2011).
269 Hanna JA, Wimberly H, Kumar S, Slack F,
Agarwal S, Rimm DL. Quantitative
analysis of microRNAs in tissue microar-
rays by in situ hybridization. BioTechniques
52(4), 235–245 (2012).
270 Kloosterman WP, Wienholds E, de Bruijn
E, Kauppinen S, Plasterk RH. In situ
detection of miRNAs in animal embryos
using LNA-modified oligonucleotide
probes. Nat. Methods 3(1), 27–29 (2006).
271 Lodes MJ, Caraballo M, Suciu D, Munro
S, Kumar A, Anderson B. Detection of
cancer with serum miRNAs on an
oligonucleotide microarray. PLoS ONE
4(7), e6229 (2009).
272 Roderburg C, Mollnow T, Bongaerts B
et al. Micro-RNA profiling in human
serum reveals compartment-specific roles of
miR-571 and miR-652 in liver cirrhosis.
PLoS ONE 7(3), e32999 (2012).
273 Schrauder MG, Strick R,
Schulz-Wendtland R et al. Circulating
micro-RNAs as potential blood-based
markers for early stage breast cancer
detection. PLoS ONE 7(1), e29770 (2012).
274 Zhao B, Jin L, Wei J et al. A simple and fast
method for profiling microRNA expression
from low-input total RNA by microarray.
IUBMB Life 64(7), 612– 616 (2 012 ).
275 Tang X, Gal J, Zhuang X, Wang W, Zhu
H, Tang G. A simple array platform for
microRNA analysis and its application in
mouse tissues. RNA 13(10), 1803 –1822
(2007).
276 Benes V, Castoldi M. Expression profiling
of microRNA using real-time quantitative
PCR, how to use it and what is available.
Methods 50(4), 244–249 (2010).
277 Chugh P, Tamburro K, Dittmer DP.
Profiling of pre-micro RNAs and
microRNAs using quantitative real-time
miRNAs & long noncoding RNAs as biomarkers in human diseases
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
Expert Rev. Mol. Diagn. 13(2 ), (2013)
204
Review
PCR (qPCR) arrays. J. Vis. Exp. (46),
pii :2210 (2010).
278 Jensen SG, Lamy P, Rasmussen MH et al.
Evaluation of two commercial global
miRNA expression profiling platforms for
detection of less abundant miRNAs. BMC
Genomics 12, 435 (2011).
279 Bianchi N, Zuccato C, Finotti A,
Lampronti I, Borgatti M, Gambari R.
Involvement of miRNA in erythroid
differentiation. Epigenomics 4(1), 51–65
(2012).
280 Heneghan HM, Miller N, Kerin MJ.
Circulating miRNA signatures: promis-
ing prognostic tools for cancer. J. Clin.
Oncol. 28(29), e573–e574; author reply
e575 (2010 ).
281 Heneghan HM, Miller N, Kerin MJ.
MiRNAs as biomarkers and therapeutic
targets in cancer. Curr. Opin. Pharmacol.
10(5), 543–550 (2010).
282 Hu Z, Dong J, Wang LE et al. Serum
microRNA profiling and breast cancer risk:
the use of miR-484/191 as endogenous
controls. Carcinogenesis 33(4), 828–834
(2012).
283 Brase JC, Johannes M, Schlomm T et al.
Circulating miRNAs are correlated with
tumor progression in prostate cancer. Int. J.
Cancer 128(3), 608–616 (2011).
284 Wang Y, Zheng D, Tan Q, Wang MX, Gu
LQ. Nanopore-based detection of circulat-
ing microRNAs in lung cancer patients.
Nat. Nanotechnol. 6(10), 668–674 (2011).
285 Wanunu M, Dadosh T, Ray V, Jin J,
McReynolds L, Drndic M. Rapid
electronic detection of probe-specific
microRNAs using thin nanopore sensors.
Nat. Nanotechnol. 5(11), 807–814 (2010).
286 Chan HM, Chan LS, Wong RN, Li HW.
Direct quantification of single-molecules of
microRNA by total internal reflection
fluorescence microscopy. Anal. Chem.
82(16), 6911– 6918 (2010).
287 Neely LA, Patel S, Garver J et al. A
single-molecule method for the quantita-
tion of microRNA gene expression. Nat.
Methods 3(1), 41–46 (2006).
288 Gao Z, Yang Z. Detection of microRNAs
using electrocatalytic nanoparticle tags.
Anal. Chem. 78 (5), 1470–1477 (2006).
289 Gao Z, Yu YH. Direct labeling microRNA
with an electrocatalytic moiety and its
application in ultrasensitive microRNA
assays. Biosens. Bioelectron. 22(6), 933–940
(2007).
290 Peng Y, Gao Z. Amplified detection of
microRNA based on ruthenium oxide
nanoparticle-initiated deposition of an
insulating film. Anal. Chem. 83 (3),
820–827 (2011).
291 Cissell KA, Deo SK. Trends in microRNA
detection. Anal. Bioanal. Chem. 394(4),
1109–1116 (2009 ).
292 Zhang GJ, Chua JH, Chee RE, Agarwal A,
Wong SM. Label-free direct detection of
MiRNAs with silicon nanowire biosensors.
Biosens. Bioelectron. 24 (8), 2504 –2508
(2009).
293 Sioss JA, Bhiladvala RB, Pan W et al.
Nanoresonator chip-based RNA sensor
strategy for detection of circulating tumor
cells: response using PCA3 as a prostate
cancer marker. Nanomedicine 8(6),
1017–1025 (2012).
294 Robertson KL, Vora GJ. Locked nucleic
acid flow cytometry-fluorescence in situ
hybridization (LNA flow-FISH): a method
for bacterial small RNA detection. J. Vis.
Exp. 59, e3655 (2012).
295 Driskell JD, Primera-Pedrozo OM, Dluhy
RA, Zhao Y, Tripp RA. Quantitative
surface-enhanced Raman spectroscopy
based analysis of microRNA mixtures.
Appl. Spectrosc. 63(10), 1107–1114
(2009).
296 Cissell KA, Rahimi Y, Shrestha S, Hunt
EA, Deo SK. Bioluminescence-based
detection of microRNA, miR21 in breast
cancer cells. Anal. Chem. 80(7), 2319–2325
(2008).
297 Fang S, Lee HJ, Wark AW, Corn RM.
Attomole microarray detection of
microRNAs by nanoparticle-amplified SPR
imaging measurements of surface
polyadenylation reactions. J. Am. Chem.
Soc. 128(43), 14044–14046 (2006).
298 Nasheri N, Cheng J, Singaravelu R, Wu P,
McDermott MT, Pezacki JP. An
enzyme-linked assay for the rapid
quantification of microRNAs based on
the viral suppressor of RNA silencing
protein p19. Anal. Biochem. 412(2),
165–172 (2011).
299 Sípová H, Zhang S, Dudley AM, Galas D,
Wang K, Homola J. Surface plasmon
resonance biosensor for rapid label-free
detection of microribonucleic acid at
subfemtomole level. Anal. Chem. 82(24),
10110–10115 (2010).
300 Wark AW, Lee HJ, Corn RM. Multiplexed
detection methods for profiling microRNA
expression in biological samples. Angew.
Chem. Int. Ed. Engl. 47(4), 644–652
(2008).
301 Dong H, Zhang J, Ju H et al. Highly
sensitive multiple microRNA detection
based on fluorescence quenching of
graphene oxide and isothermal strand-
displacement polymerase reaction. Anal.
Chem. 84 (10), 4587–4593 (2012).
302 Alhasan AH, Kim DY, Daniel WL et al.
Scanometric microRNA array profiling
of prostate cancer markers using
spherical nucleic acid-gold nanoparticle
conjugates. Anal. Chem. 84(9), 4153–
4160 (2012).
303 Jiang L, Duan D, Shen Y, Li J. Direct
microRNA detection with universal tagged
probe and time-resolved fluorescence
technology. Biosens. Bioelectron. 34(1),
291–295 (2012).
Websites
401 ClinicalTrials.gov.
www.clinicaltrials.gov
402 Rosetta Genomics. Rosetta Genomics
Leading the microRNA Revolution.
www.rosettagenomics.com
403 Hologic Gen-Probe. Hologic Gen-Probe
Nucleic Acid Test.
www.gen-probe.com.
Van Roosbroeck, Pollet & Calin
Expert Review of Molecular Diagnostics Downloaded from informahealthcare.com by UTMD Anderson Cancer Center on 07/06/14
For personal use only.
... Therefore, the present study chose to use the RNA-sequencing technique, a highthroughput assay which provides a complete overview for miRNA expression profiling. Although commonly used for screening purposes (where the objective usually is to identify miRNAs that are differentially expressed between biofluids, tissues, or subgroups of patients [27]), this technique can also be used to select the most adequate candidate normalizer miRNAs (as was the case in this study, where the objective was to identify the miRNAs whose expression pattern was equivalent across all samples in both groups of ...
... Therefore, the present study chose to use the RNA-sequencing technique, a highthroughput assay which provides a complete overview for miRNA expression profiling. Although commonly used for screening purposes (where the objective usually is to identify miRNAs that are differentially expressed between biofluids, tissues, or subgroups of patients [27]), this technique can also be used to select the most adequate candidate normalizer miRNAs (as was the case in this study, where the objective was to identify the miRNAs whose expression pattern was equivalent across all samples in both groups of patients evaluated, those who did and those who did not develop cisplatin-induced nephrotoxicity). ...
Article
Full-text available
The microRNA (miRNA) expression profile by qRT-PCR depends directly on the most appropriate normalization strategy adopted; however, currently there is no universally adequate reference gene. Therefore, this study aimed to determine, considering RNA-Seq results, the most adequate endogenous normalizer for use in the relative quantification of urine miRNAs from head and neck cancer patients, treated with cisplatin chemoradiotherapy. The massive sequencing was performed to identify the miRNAs differentially expressed between the group with cisplatin nephrotoxicity (n = 6) and the one without (n = 6). The candidate endogen normalizer was chosen according to four criteria: (1) the miRNA must be expressed in most samples; (2) the miRNA must have a fold change value between 0.99 and 1.01; (3) the miRNA must have a p-value ≥ 0.98; and (4) the miRNA must not be commented on by the final GeneGlobe (Qiagen, Hilden, Germany) analysis. Four miRNAs met all the criteria (hsa-miR-363-5p, hsa-miR-875-5p, hsa-miR-4302, and hsa-miR-6749-5p) and were selected for validation by qRT-PCR in a cohort of 49 patients (including the 12 sequencing participants). Only hsa-miR-875-5p was shown to be an adequate normalizer for the experimental condition under investigation, as it exhibited invariant expression between the two groups.
... of ncRNAs along with the recent advancements in cutting edge sequence 'omics' technologies and single cell analysis techniques in understanding the molecular mechanisms of brain in physiological and pathological state, make these (ncRNAs) suitable candidates to be used both as ideal biomarkers and the therapeutic agents. 24,25 ...
Article
Full-text available
The human genome encrypts around 20,000 protein coding genes, constituting around 1% of the total human genome sequence. The rest of it initially labeled as a “junk DNA” is transcribed to a distinct class of non-coding RNAs (ncRNAs) which do not code for any protein in the cell and their presence was quite intriguing to the researchers. The recent studies, however, have surprisingly revealed the vital roles of these ncRNAs in regulating an array of diverse cellular and biological processes in different organs including brain. The dysfunction of these regulatory ncRNAs in human brain causes certain neurological disorders and brain tumors which earlier have been widely linked to various risk factors such as oxidative stress, genetic mutations, aberrant protein degradation and dysfunctional neural network. This review provides an overview of different types of ncRNAs, their regulatory roles in brain functions and neurological disorders along with their prospects to be used as potential biomarkers and therapeutics.
... Furthermore, post-transcriptionally modified miRNAs are thought to control more than a third of all the known genes coding for protein [7]. Following the discovery of evidence connecting carcinogenesis in human chronic lymphocytic leukemia (shown) to miRNAs [8,9], the potential of miRNAs in the diagnosis, prognosis, and therapy in a wide range of cancers has sparked considerable interest [10,11]. miRNAs have become intriguing tools and targets for novel therapeutic approaches thanks to new insights into their role in cancer. ...
Article
Full-text available
The dynamic spectrum of microRNA (miRNA) has grown significantly over the years with its identification and exploration in cancer therapeutics and is currently identified as an important resource for innovative strategies due to its functional behavior for gene regulation and modulation of complex biological networks. The progression of cancer is the consequence of uncontrolled, nonsynchronous procedural faults in the biological system. Diversified and variable cellular response of cancerous cells has always raised challenges in effective cancer therapy. miRNAs, a class of non-coding RNAs (ncRNAs), are the natural genetic gift, responsible to preserve the homeostasis of cell to nurture. The unprecedented significance of endogenous miRNAs has exhibited promising therapeutic potential in cancer therapeutics. Currently, miRNA mimic miR-34, and an antimiR aimed against miR-122 has entered the clinical trials for cancer treatments. This review, highlights the recent breakthroughs, challenges, clinical trials, and advanced delivery vehicles in the administration of miRNA therapies for precision oncology.
... These features make lncRNAs very suitable as non-invasive biomarkers. Unlike micro RNAs, lncRNAs show greater complexity of their functions and have wider biological activities [42]. ...
Article
Full-text available
Background The role of the long non-coding RNAs (lncRNAs) in the pathogenesis of systemic lupus erythematosus (SLE) is mostly unknown, despite increasing evidence that lncRNAs extensively participate in physiological and pathological conditions. Aim To detect the level of lncRNA-Cox2, HOTAIR, IL-6, and MMP-9 in the serum of SLE patients and to correlate these levels with disease activity and patients’ clinical and laboratory data to evaluate the value of these biomarkers for SLE diagnosis and assessment of disease activity. Methods Blood samples from 58 SLE patients, and 60 healthy controls (HCs) were used for detection of lncRNAs-Cox2 and HOTAIR expression levels by real-time polymerase chain reaction. Both IL-6 and MMP-9 serum levels were assayed by enzyme-linked immunosorbent assay. Lupus activity was assessed with the Systemic Lupus Erythematosus Disease Activity Index (SLEDAI). Results The serum expression levels of lncRNA-Cox2 and HOTAIR were significantly up-regulated in SLE patients vs HCs (fold change [median (IQR) was 1.29(0.81–1.71, P <0.0001) and 2.68(0.95–3.67), P = 0.038) for lncRNA-Cox2 and HOTAIR, respectively. Serum levels of both IL-6 and MMP-9 were significantly high in SLE patients compared with HCs ( P≤ 0.001 for each). The up-regulated lncRNA-Cox2 was positively associated with the presence of neurological manifestations in SLE patients ( P = 0.007). Furthermore, HOTAIR expression level had significantly positive correlation with IL-6 ( r = 0.578, P <0.0001), MMP-9 level (r = 0.762, P <0.0001), nephritis grades ( r = 0.296, P = 0.024) and proteinuria ( r = 0.287, P = 0.035). LncRNA-Cox2 showed sensitivity and specificity 72.4%, and 100.0% respectively. HOTAIR sensitivity was 60.3%, and specificity was 100.0%. By multiple logistic regression analysis, lncRNA-Cox2 and HOTAIR were found as SLE independent predictors. Conclusion LncRNA-COX2 and HOTAIR can be used as new non-invasive biomarkers for the diagnosis of SLE.
... Wu et al., 2018; Zhu et al., 2017). MicroRNAs such as miR-96, miR-17-92 cluster, miR-21, miR-155, and miR-221 has been found profoundly upregulated in fewer cancers (M.Li et al., 2010; Y. Li et al., 2017;Sundarbose et al., 2013; implying their oncogenic role in promoting uncontrolled cell growth and division(Simmons et al., 2016;van Roosbroeck et al., 2013). The poor sensitivity and specificity of available biomarkers for cancer detection, ...
Article
MicroRNAs are small non-protein coding RNAs of 19–24 nucleotides that operate as regulators of gene expression in the eukaryotic system at both the transcriptional and post-transcriptional levels. Promising evidences suggest that most of the miRNAs and the related components of the RNA interference pathway are decisively significant in the regulation of nearly all cellular processes of both animals and plants. MicroRNA has crucial role in different metabolic processes like cell division, growth, proliferation, differentiation as well as apoptosis of the cell. Scientists are given more insight into this novel idea of cross kingdom regulation system by miRNA after discovery of exogenous plant MIR168a influencing the transport system of cholesterol in mice by down regulating the receptor adopter protein 1 expression of law density lipoprotein. These types of influential exploration about cross kingdom regulation of miRNAs have taken more intention towards its role in animal as well as microbial system. In human health, the animal sources, plant, microbes derived miRNAs and diet based miRNA plays significant role. Dysregulation of miRNAs is correlated with development of various pathological events such as neurological disorders, cardiovascular diseases, cancer, AMR, which are the most threatened field for human society. Here we detailed about the role of miRNA in human health with diverse source of miRNAs and its role in therapeutic and diagnosis of disease like cancer in human.
... 10 Therefore, the further study of miRNAs is helpful for the prediction, diagnosis, and prognosis of cancer in the future. 11 MiR-379-5p is differentially expressed in various cancer tissues and cells, such as hepatocellular carcinoma, prostate cancer, gastric cancer, lung cancer, and osteosarcoma. [12][13][14][15][16] Studies have shown that the expression level of miR-379 was significantly decreased in BC tissues compared with normal breast tissues and significantly negatively correlated with cyclin B1. 17 KIF4A, a kinesin superfamily member (KIFs), belongs to the KIFA subfamily. ...
Article
Full-text available
Background: Many studies have shown that microRNAs (miRNAs) play an essential role in gene regulation and tumor development. This study aimed to explore the expression of miR-379-5p and its mechanisms of affecting proliferation, migration, and invasion in breast cancer (BC). Methods: MiRNAs and mRNAs expression data of BC and normal breast tissue samples were downloaded from the TCGA and GEO databases. qRT-PCR was used to detect the expression of miR-379-5p in human normal breast epithelial cell lines and human BC cell lines. The proliferation ability of transfected cells was detected by colony formation and EdU assays. The mobility and invasion ability of transfected cells was measured by wound healing and transwell assays. The relative protein expression of transfected cells was detected by western blot. Dual luciferase reporter assay was performed to identify the targeted binding of miR-379-5p and KIF4A. Results: MiR-379-5p was lowly expressed in BC tissue samples and BC cell lines. The target genes of miR-379-5p were involved in many cancer-related signaling pathways. PPI analysis and the cytoHubba algorithm of Cytoscape identified 10 genes as the hub genes. Survival analysis showed that only KIF4A expression in 10 hub genes was significantly associated with the prognosis of BC patients and was significantly upregulated in BC. Overexpression of miR-379-5p inhibited proliferation, migration, and invasion in the BC cell line MDA-MB-231, which could be reversed by KIF4A. Conclusions: MiR-379-5p inhibits proliferation, migration, and invasion of BC by targeting KIF4A.
... Towards that direction, the pleiotropic action of non-coding RNAs (ncRNAs) in the regulation of gene expression makes them as putative diagnostic, prognostic or predictive molecular markers not only in CRC, but also in a wider spectrum of human diseases [8][9][10][11]. Small ncRNAs (sncRNAs) constitute a class of ncRNAs, which are characterized by a length shorter than 200 nucleotides (nt) and are heavily implicated in the regulation of gene expression. Among these small ncRNAs, an emerging frontier is represented by microRNAs (miRNAs) that have obtained a clinical relevance only after much research effort [12][13][14][15]. ...
Article
Full-text available
Colorectal cancer (CRC) is the second most lethal cause of cancer-related deaths in Europe. Fragments of tRNAPro are conserved among vertebrates, characterized by pleiotropic regulatory functions and have been found to discriminate colorectal tumors from normal colorectal mucosa. In the current study, we investigated the prognostic utility of 5′-tiRNA-ProTGG levels in CRC. For this purpose, total RNA was extracted from 155 malignant colorectal tumors and 74 adjacent non-cancerous tissue specimens, polyadenylated and reverse-transcribed using an oligo-dT adapter as primer. Real-time quantitative PCR (qPCR) was used to assess the levels of 5′-tiRNA-ProTGG. Kaplan-Meier survival analysis demonstrated that high 5′-tiRNA-ProTGG levels predict both poor disease-free survival (DFS) and overall survival (OS) of CRC patients. Of note, high 5′-tiRNA-ProTGG levels retain their unfavorable prognostic value in patients with rectal cancer and/or moderately differentiated CRC (grade II). More importantly, multivariate cox regression analysis highlighted that the overexpression of 5′-tiRNA-ProTGG constitutes an adverse prognostic factor predicting short-term relapse of CRC patients independently of the established prognosticators in CRC. Finally, bioinformatics analysis unveiled a potentially critical role of 5′-tiRNA-ProTGG regarding the maintenance of cellular homeostasis, signaling, cell communication, and cellular motility.
Article
Background: MicroRNA-155 has been discussed as a biomarker in cancer diagnosis and prognosis. Although relevant studies have been published, the role of microRNA-155 remains uncertain because of insufficient data. Methods: We conducted a literature search in PubMed, Embase, and Web of Science databases to obtain relevant articles and extract data to evaluate the role of microRNA-155 in cancer diagnosis and prognosis. Results: The pooled results showed that microRNA-155 presented a remarkable diagnostic value in cancers (area under the curve = 0.90, 95% confidence interval (CI 0.87-0.92; sensitivity = 0.83, 95% CI 0.79-0.87; specificity = 0.83, 95% CI 0.80-0.86), which was maintained in the subgroups stratified by ethnicity (Asian and Caucasian), cancer types (breast cancer, lung cancer, hepatocellular carcinoma, leukemia, and pancreatic ductal adenocarcinoma), sample types (plasma, serum, tissue), and sample size (n >100 and n <100). In prognosis, a combined hazard ratio (HR) showed that microRNA-155 was significantly associated with poor overall survival (HR = 1.38, 95% CI 1.25-1.54) and recurrence-free survival (HR = 2.13, 95% CI 1.65-2.76), and was boundary significant with poor progression-free survival (HR = 1.20, 95% CI 1.00-1.44), but not significant with disease-free survival (HR = 1.14, 95% CI 0.70-1.85). Subgroup analyses in overall survival showed that microRNA-155 was associated with poor overall survival in the subgroups stratified by ethnicity and sample size. However, the significant association was maintained in cancer types subgroups of leukemia, lung cancer, and oral squamous cell carcinoma, but not in colorectal cancer, hepatocellular carcinoma, and breast cancer, and was maintained in sample types subgroups of bone marrow and tissue, but not in plasma and serum. Conclusions: Results from this meta-analysis demonstrated that microRNA-155 was a valuable biomarker in cancer diagnosis and prognosis.
Article
Full-text available
Exosomes are vesicles of endocytic origin released by many cells. These vesicles can mediate communication between cells, facilitating processes such as antigen presentation. Here, we show that exosomes from a mouse and a human mast cell line (MC/9 and HMC-1, respectively), as well as primary bone marrow-derived mouse mast cells, contain RNA. Microarray assessments revealed the presence of mRNA from approximately 1300 genes, many of which are not present in the cytoplasm of the donor cell. In vitro translation proved that the exosome mRNAs were functional. Quality control RNA analysis of total RNA derived from exosomes also revealed presence of small RNAs, including microRNAs. The RNA from mast cell exosomes is transferable to other mouse and human mast cells. After transfer of mouse exosomal RNA to human mast cells, new mouse proteins were found in the recipient cells, indicating that transferred exosomal mRNA can be translated after entering another cell. In summary, we show that exosomes contain both mRNA and microRNA, which can be delivered to another cell, and can be functional in this new location. We propose that this RNA is called " exosomal shuttle RNA " (esRNA). Exosomes are small (50–90 nm) membrane vesicles of endocytic origin that are released into the extracellular environment on fusion of multivesicular bodies (MVB) with the plasma membrane 1. Many cells have the capacity to release exosomes, including reticulo-cytes 2 , dendritic cells 3 , B cells 4 , T cells 5 , mast cells 6 , epithelial cells 7 and tumour cells 8. The functions of exosomes are not completely understood, although it has been shown that exosomes can participate in the signalling events contributing to antigen presentation to T cells 4 and the development of tolerance 9. Several mechanisms have been hypothesized describing the interactions of exosomes and recipient cells. Exosomes can bind to cells through recep-tor–ligand interactions, similar to cell–cell communication mediating , for example, antigen presentation 4. Alternatively, exosomes could putatively attach or fuse with the target-cell membrane, delivering exosomal surface proteins and perhaps cytoplasm to the recipient cell 10,11. Finally, exosomes may also be internalized by the recipient cells by mechanisms such as endocytosis 12. Exosomes were isolated from a mast-cell line (MC/9), primary bone marrow-derived mast cells (BMMC) and a human mast-cell line (HMC-1) through a series of microfiltration and ultracentrifugation steps modified from what has been previously described 4. To confirm that the structures studied indeed are exosomes, they were examined by electron microscopy (Fig. 1a), flow cytometric analysis (FACS; Fig. 1b), and proteomic analysis (see Supplementary Information, Table S1). The electron micrographs of the exosomes revealed rounded structures with a size of approximately 50–80 nm, similar to previously described exo-somes 4,13–15. The identity of the studied vesicles was further confirmed as exosomes by FACS analysis (Fig. 1b), which show the presence of the surface protein CD63 — a commonly used marker of exosomes. Finally, extensive protein analysis of the MC/9 exosomes was performed on multiple samples using LC-MS/MS technology. A total of 271 proteins were identified (see Supplementary Information, Table S1) from three preparations of the isolated vesicles, of which 47 proteins were present in all three samples. More importantly, a large number of the proteins found in the preparations were the same as proteins previously identified in exosomes produced by other cells (that is, exosomes derived from intestinal epithelial cells, urine, dendritic cells, microglia, melanoma, T-cells and B-cells). In particular, 60% of the 47 proteins found in all samples of mast-cell exosomes have been previously found in other types of exosomes. Moreover, 39% of the 271 total proteins found in the analysed exosome samples have also been previously found in other types of exosomes. Thus, the electron microscopy, the FACS, and the detailed protein analyses each provided significant evidence in favour of the identification of the isolated vesicles as exosomes. The presence of nucleic acids was examined in exosomes derived from MC/9, BMMC and HMC-1 cells to define a potential mechanism by which exosomes may mediate cell–cell communication. These assessments showed that isolated exosomes contain no DNA (see Supplementary Information, Fig. S1). However, substantial amounts of RNA were detected by agarose gel electrophoresis, spectrophotometry
Article
MicroRNAs have essential functional roles in brain development and neuronal specification but their roles in neurode- generative diseases such as Alzheimer's disease (AD) is unknown. Using a sensitive qRT-PCR platform we identified regional and stage-specific deregulation of miRNA expression in AD patient brains. We used experimental validation in addition to literature to reveal how the deregulated brain microRNAs are biomarkers for known and novel pathways in AD pathogenesis related to amyloid processing, neurogenesis, insulin resistance, and innate immunity. We additionally recovered miRNAs from cerebrospinal fluid and discovered AD-specific miRNA changes consistent with their role as potential biomarkers of disease.
Article
10575 Background: Cancer of unknown primary (CUP) constitutes 3%-5% of all newly diagnosed cancer cases. It presents a major diagnostic challenge as knowing the tumor tissue of origin (ToO) of the cancer is crucial for choosing the optimal treatment. MicroRNAs are a family of non-coding, regulatory RNA genes involved in development and differentiation that were shown to be involved in carcinogenesis. MicroRNAs, highly stable in clinical samples and tissue-specific, serve as ideal biomarkers for cancer diagnosis. Our first generation assay identifies the ToO using a set of 48 microRNAs measured on a qRT-PCR platform and differentiates 25 different tumor types. We present here the development and validation of a second generation clinical assay that can identify 42 different tumor types using a custom array platform. Methods: Over 1300 primary and metastatic tumor formalin-fixed and paraffin-embedded (FFPE) tissue samples were used for the training and development of the assay. High-purity RNA was extracted...