Skip to main content
Journal of Cellular and Molecular Medicine logoLink to Journal of Cellular and Molecular Medicine
. 2014 Feb 18;18(3):371–390. doi: 10.1111/jcmm.12236

Methodological challenges in utilizing miRNAs as circulating biomarkers

Leni Moldovan a, Kara E Batte b, Joanne Trgovcich c, Jon Wisler c, Clay B Marsh a,*, Melissa Piper a
PMCID: PMC3943687  NIHMSID: NIHMS553042  PMID: 24533657

Abstract

MicroRNAs (miRNAs) have emerged as important regulators in the post-transcriptional control of gene expression. The discovery of their presence not only in tissues but also in extratissular fluids, including blood, urine and cerebro-spinal fluid, together with their changes in expression in various pathological conditions, has implicated these extracellular miRNAs as informative biomarkers of disease. However, exploiting miRNAs in this capacity requires methodological rigour. Here, we report several key procedural aspects of miRNA isolation from plasma and serum, as exemplified by research in cardiovascular and pulmonary diseases. We also highlight the advantages and disadvantages of various profiling methods to determine the expression levels of plasma-and serum-derived miRNAs. Attention to such methodological details is critical, as circulating miRNAs become diagnostic tools for various human diseases.

Keywords: biomarkers, microRNA, circulating microRNAs, plasma, serum, real-time PCR, profiling

Introduction

MicroRNAs (miRNAs) are small non-coding RNA sequences of about 22 nucleotides that are important post-transcriptional regulators of gene expression. They impact many if not most developmental [1] and homeostatic processes such as the immune response and metabolism [2, 3] and epigenetic processes as well [1, 4].

A plethora of studies in humans, complemented by animal models, demonstrate the importance of tissue miRNAs during developmental processes [511] and various disease pathologies [1217]. Thus, significant changes of tissue miRNA ‘signatures' occur in various diseases, such as cancers [18, 19], cardiovascular disease (CVD)[2126], diseases of the lung [2731], kidney [16] and nervous system [32, 33]. However, routine biopsies from any organ for miRNA profiling are not a practical option; therefore, investigators are turning towards less invasive procedures, involving circulating miRNA biomarkers (for a review of practical issues, see [34]).

The presence of miRNA molecules in human plasma was concomitantly reported by several groups [3538]. Once it became clear that some miRNAs were released from their cells of origin and could be captured in various extracellular fluids, numerous studies began investigating whether tissue-and disease-specific miRNA signatures were also reflected in fluids such as blood, urine, spinal fluid or saliva [39]. Extracellular miRNAs are remarkably stable in the circulation [3537, 40]. As a result of cellular damage/tissue injury, such as in acute myocardial infarction (MI) [41, 42], atherosclerosis [44], non-small cell lung cancer [45], neurodegenerative diseases [4649], skin fibrosis [50] and osteoarthritis [51] (Table 1), miRNA expression can change in the blood. Notably, this list is increasingly expanding. Furthermore, miRNA expression can change not only in the blood but in other body fluids as well (Table 1). Thus, circulating miRNAs are attractive candidates for disease monitoring to serve as valuable prognostic indicators of disease progression or resolution. It is predicted that changes in miRNA expression in body fluids occur earlier than conventional biomarkers. For instance, the markers of inflammation and repair in the cardiovascular and lung systems include troponin, C-reactive protein, chemokines and cytokines [52, 53]. However, by the time these proteins are detectable in the circulation, much of the tissue damage has already occurred, which makes it crucial that better biomarkers be uncovered for the early detection of diseases.

Table 1.

Examples of circulating miRNAs detected in various human pathologies

MiRNAs differentially expressed Condition Sample type Anticoagulant Volume (ml) Isolation method* Spike-in/endogenous controls Detection method No. of miRNAs tested Refs.
miRs-21,-155,-210 Diffuse large B-cell lymphoma Serum N/A 2 Trizol miR-16 qRT-PCR 3 [71]
N/A Healthy donors Plasma Unknown 40 Trizol None qRT-PCR pre-amp 420 [36]§
let-7a, miRs-16,-92a,-122,-142-3p,-150 Healthy donors Plasma (p), EV, Serum (s) EDTA 30 (p) 10 (s) miRNeasy; 5 vol Qiazol cel-miRs Microarray qRT-PCR 375 [68]**
miRs-15b,-16 Prostate cancer Plasma Serum EDTA 10 mirVana PARIS cel-miRs qRT-PCR pre-amp 368 [37]††
miRs-16,-21,-24,-34b Healthy donors Plasma EDTA 0.4 Tri-Reagent LS miRNeasy cel-miR-39 qRT-PCR pre-amp 667 [72]‡‡
miRs-15b,-16,-24,-122 Healthy donors Plasma Serum EDTA 0.4 mirVana PARIS cel-miRs qRT-PCR 3 [73]§§
let-7b, miR-223 NSCLC Plasma Serum Heparin 0.1 Total RNA purification kit cel-miRs qRT-PCR 30 [74]§§
N/A N/A Plasma Serum U 0.4 miRNeasy; 10 vol Qiazol cel-miRs qRT-PCR N/A [75]¶¶
miR-499 ACS; CHF Plasma EDTA 0.5 mirVana PARIS Novel Synthetic miR qRT-PCR 1 [41]
miRs-1,-133a,-133b,-208a,-208b,-499 ACS Plasma U 0.05 Master Pure RNA Purification Kit cel-miR-54 qRT-PCR 6 [43]
miRs-1,-133 AMI Plasma Citrate 1 miRNeasy U6 snRNA qRT-PCR 2 [42]
miR-1 AMI‡‡ Serum N/A 0.2 miRNAs Isolation Kit None (absolute quantification) qRT-PCR 1 [76]
miRs-1,-16, 133a,-208a,-499 AMI‡‡ Plasma EDTA 0.1 TRI Reagent BD TR126 cel-miR-39 Microarray qRT-PCR 5 [77]
miRs-1,-122, 133a,-133b,-375,-499-5p AMI‡‡ Plasma Serum EDTA; thrombin 2.4 mirVana PARIS miR-17-5p; median normalization qRT-PCR 667 [52]
miRs-208b,-499 AMI, VM, DD; AHF Plasma Citrate; EDTA 0.1 mirVana PARIS cel-miRs qRT-PCR 6–12 [78]
miRs-1,-133a AMI‡‡ Serum N/A 0.35 Trizol LS N/A qRT-PCR 3 [79]
miRs-133,-328 AMI Plasma EDTA 0.5-1 Trizol LS U6 snRNA qRT-PCR 3 [80]
miRs-17,-92a,-126,-145,-155,-199a CAD Plasma Serum EDTA 0.25 TRI Reagent BD miRNAeasy cel-miR-39 Microarray qRT-PCR >1000 [81]
miRs-126,-133a,-499 CAD Plasma EDTA U TRI Reagent BD cel-miR-39 qRT-PCR 7 [82]
miR-423-5p HF Plasma Citrate 0.5–4.4 mirVana PARIS miR-1249 Sequencing***; qRT-PCR >1000 [83]
miRs-122,-126,-499 CHF Plasma EDTA 0.5 mirVana PARIS Novel Synthetic miR qRT-PCR pre-amp 3 [84]
hcmv-miR-UL112, let-7e, miR-296–5p HT Plasma EDTA 0.25 TRI Reagent BD Median normalization; U6 snRNA Microarray qRT-PCR >1000 [67]
miRs-27b,-130a,-210 ASO Serum N/A U QIAamp Circulating Nucleic Acid Kit U6 snRNA qRT-PCR 13 [85]
miR-21 AS Plasma EDTA U TRIzol cel-miR-39 qRT-PCR 1 [86]
miRs-105,-106a,-223 FH††† Plasma (HDL, EV) U U miRNeasy RNU6; Array Specific Microarray qRT-PCR pre-amp >1000 [69]‡‡‡
miRs-20a,-21,-133a, 146a,-221,-222,-210,-328 Exercise –ind. Cardiovasc. Adaptation Plasma EDTA 0.1 MicroRNA Extraction Kit miR-422b qRT-PCR 8 [87]
miRs-122,-370 HL Plasma EDTA 0.4 mirVana PARIS cel-miR-39 qRT-PCR 4 [88]
miRs-1,-21,-29a,-133a,-208 MI Plasma U 0.4 mirVana PARIS U6 snRNA qRT-PCR 5 [89]
miRs-208b,-499, MI Plasma Citrate U mirVana PARIS cel-miRs qRT-PCR 2 [90]
miR-133a MI Serum N/A 0.1 miRNeasy cel-miR-39 qRT-PCR 1 [91]
miRs-126,-197,-223 MI Plasma U U miRNeasy U6 snRNA qRT-PCR 19 [92]
miRs-15b,-27b NSCLC Serum N/A 0.5 mirVana PARIS§§§ ΔCt matrix qRT-PCR  181 [93]
miRs-1,-30d,-486,-499 NSCLC Serum N/A 50 TRIzol¶¶¶ Normalization to total RNA qRT-PCR; sequencing 11 by qRT-PCR [94]
let7f, miRs-20b,-30e-3p NSCLC Plama EV U 3 Dynabeads + mirVana PARIS†† miR-142-3p and miR-30b qRT-PCR 365 [95]
let7b, miRs-20b,-33a, 199b, 200a,-518b,-635,-662,-935 COPD, lung cancer Whole blood PAXgene 5 miRNeasy Quantile normalization Microarray; qRT-PCR 863 [96]
miRs-1, 7,-21,-30a,-126,-200b,-210,-219,-324,-451,-486 Lung cancer Plasma EDTA 0.2 mirVana PARIS RNU-6B Microarray; qRT-PCR  >200 [61]
miR-150 PAH Plasma Citrate U miRNeasy cel-miR-39 Microarray; qRT-PCR  1223 [59]
miRs-21,-210, 486-5p SPN Plasma EDTA U mirVana PARIS miR-16 qRT-PCR 5 [97]
miR-483-5p HCC Plasma U 0.25 miRNeasy U6 snRNA; cel-miR-39 qRT-PCR TLDA cards A & B 5 [98]
let-7d-5p, let-7 g-5p, miRs-15b-5p,-142-3p,-191-5p,-301a-3p,-545 Alzheimer Plasma U 1 mirVana PARIS modified Ath-159a; NegA nCounter**** ˜800 [99]
miR-210 AKI Plasma U U MasterPure RNA Purification Kit cel-miR-54 Microarray >1000; 3 by qRT-PCR [100]
*

Reagents and supplier information: Master Pure RNA Purification Kit (Epicentre Biotechnologies, Madison, WI, USA); MicroRNA Extraction Kit (BenevBio, Mission Viejo, CA, USA); miRNA Isolation Kit (RNA Bioscience, Salt Lake City, UT, USA); miRNeasy, Qiazol, QIAamp Circulating Nucleic Acid Kit (Qiagen, Valencia, CA, USA); mirVana PARIS, Trizol and Trizol LS [Ambion/Invitrogen/Applied Biosytems Inc. (ABI), Life Technologies, Carlsbad, CA, USA]; TRI Reagent BD and Tri-Reagent LS (Sigma-Aldrich, St. Louis MO, USA); Total RNA Purification Kit (Norgen Biotek Corp, Thorold, Canada).

Quantitative real-time-PCR (qRT-PCR) technologies include TaqMan, TLDA Cards and SYBR Green methodologies using a wide range of products from ABI, Qiagen, Exiqon (Woburn, MA, USA), Roche (Indianapolis, IN, USA) and Promega (Madison, WI, USA).

First study to show miRNA presence in plasma.

§

First study to analyse miRNA in plasma microvesicles in healthy individuals.

Various array technologies are from Affymetrix (Santa Clara, CA, USA), Agilent (Santa Clara, CA, USA), Exiqon miRCURY LNA Array (Woburn, MA, USA), Genom Biochip MPEA (Febit biomed GmbH, Heidelberg, Germany), Fluidigm (San Francisco, CA, USA).

**

Reports that some circulating miRNA species may be associated with Argonaute2.

††

Describes the use of Caenorhabditis elegans miRNAs spike-ins as exogenous controls for the efficiency of extraction procedure.

‡‡

Useful methods paper.

§§

Useful methods paper; compares plasma and serum.

¶¶

Comprehensive methods review paper.

***

Sequencing-based profiling from Illumnia (San Diego, CA, USA).

†††

In addition to humans, study included rodent models.

‡‡‡

Report that circulating miRNAs are associated with HDL.

§§§

Reported yield 300–500 ng/ml of serum.

¶¶¶

Reported yield 5–10 μg/50 ml of serum.

****

nCounter miRNA expression assay v1 (Nanostring Technology, Seattle, WA, USA).

ACS: acute coronary syndrome; AHF: acute heart failure; AKI: acute kidney injury; AMI: acute myocardial infarction; AS: atherosclerosis; ASO: atherosclerosis obliterans; CAD: coronary artery disease; cc: cell culture; cel-miRs: C. elegans synthetic miR-39, miR-54 and/or miR-238; CHF: congestive heart failure; COPD: congestive obstructive pulmonary disease; DD: diastolic dysfunction; EDTA: ethylenediaminetetraacetic acid; EV: extracellular vesicles (including exosomes); FH: familial hypercholesterolaemia; H: human; HCC: hepatocellular carcinoma; HF: heart failure; HL: hyperlipidaemia; HT: hypertension; pre-amp.: pre-amplification; MI: myocardial infarction; NSCLC: non-small cell lung carcinoma; PAH: pulmonary artery hypertension; SPN: solitary pulmonary nodules; U: unknown (not mentioned in paper); VM: viral myocarditis.

However, miRNA fingerprinting as a novel and valid diagnostic, prognostic and disease surveillance tool is still in the descriptive stages. Currently, information is being gathered and compared in various disease states. The necessity of more thorough studies, based on much larger patient cohorts, is required for their utility. This is underscored by several studies using conventional biomarkers to assess tissue damage from organs that would also require invasive biopsies. For instance, comparison of circulating miRNA profile with at least one protein marker of acute coronary syndromes, namely troponin, led to conflicting results [55, 56]. At the same time, there are only few studies that analyse miRNA expression outside lung tissue in chronic obstructive pulmonary disease [57, 58], pulmonary artery hypertension [59], idiopathic pulmonary fibrosis [60] or lung cancer [61]. Several studies have begun to examine the potential of miRNAs to predict liver damage as well [62, 63].

Another level of complexity is added by the fact that, although many circulating extracellular miRNA species are repeatedly found having differential expression, such as miRs-1,-122,-126 and-223 in MI (Table 1; for comprehensive recent reviews, see Refs. [43, 64]), some of these are also associated with leukaemias and cancers [12, 66], further complicating identification of a disease-specific profile. In addition, extracellular miRNAs derived from viruses such as human cytomegalovirus can be found in patients with CVD. For example, hcmv-miR-UL112 is elevated in the circulation in hypertension [67] (Table 1). Although the role of viral miRNAs in infection is just beginning to be elucidated, little is known about their involvement in the pathogenesis of other human diseases.

Sources of variability in miRNA assessment from body fluids involve both the extraction methodology and the analysis platform employed, which may lead to inconsistent or even contradictory results. A comprehensive literature survey revealed that many methodological details are often overlooked, thus making it difficult to directly compare the specific efficiency and accuracy of various methods. The matter is further complicated by the fact that in the blood, miRNAs are either associated with proteins, such as argonaute [68], lipoproteins [69], or contained within cellular fragments designated as exosomes, microparticles, microvesicles or extracellular vesicles (EVs) [36, 37]. Extraction and analysis of miRNA from any or all of these components may pose specific challenges and yield different results.

Initial studies [36, 37] utilized a large volume of blood to isolate and profile >450 miRNAs. In contrast, advances in the approaches now allow the interrogation of <1 ml of blood to examine the expression of extracellular miRNAs (Table 1). However, in-depth analyses of the miRNA profiles often require more starting biological fluid, which may be a limiting factor. This is especially true when the patients are children, elderly or seriously ill. In addition, the RNA quantity can dictate the platform used for miRNA analysis. As shown in Table 1, current methods vary among laboratories both for miRNA isolation from plasma or serum and for obtaining miRNA profiles from these isolates.

In this review, we will highlight important considerations, such as blood collection methods or the choice of a profiling platform, when embarking on studies to profile circulating miRNAs. Throughout the review, we will refer to circulating extracellular miRNAs, as those found in the serum or plasma and not blood cells, unless otherwise stated. We focus on studies using limited sample volumes, as would be practical for disease profiling in clinical settings, and highlight optimal miRNA isolation protocols for small sample sizes. In this review, we also discuss analytic methods to identify circulating miRNAs, as well as other critical factors when designing these studies.

Current methodologies for extracting miRNAs from plasma and serum

The purposes of detecting extracellular miRNAs are to use them as disease biomarkers, as well as to understand disease pathogenesis. Therefore, much of the effort so far involves profiling miRNA species in various body fluids to uncover differential expression patterns and to correlate them with physiological status or disease progression. The composition of these fluids is quite dissimilar, which implies that isolation methods cannot be directly transposed from one tissue/fluid to another. To date, most RNA and miRNA extraction methods use a phenol:chloroform-based extraction technique, often facilitated by adding guanidinium thiocyanate, whereas newer methods, faster and some automated, include a selective solid phase (silica) adsorption of RNA from the phenol:chloroform extraction onto mini-columns, followed by elution in water or a buffer [101]. MiRNA molecules behave physically and chemically different from the larger RNA molecules, and their quantitative recovery requires optimization of existing total RNA isolation procedures. Various manufacturers have employed different strategies for this purpose.

MiRNA profiling is a multi-step process that includes: (i ) blood collection; (ii ) RNA purification; (iii ) RNA quantification and quality control; and (iv ) RNA profiling (Fig. 1). We will further discuss each of these steps, highlighting methodological aspects that may impact the outcome, as well as possible pitfalls. In addition to limited sample volumes, these involve: (i ) the lack of standard protocols; (ii ) interference of anticoagulants in PCR-based profiling [101]; (iii ) low recovery of miRNAs [64, 81]; and (iv ) the lack of known invariant miRNA species to be used as endogenous controls.

Figure 1.

Figure 1

Summary of the workflow in designing miRNA profiling from plasma or serum. One needs to first decide if plasma or serum will be analysed. If plasma is selected, then the anticoagulant should be carefully chosen, because of interference with downstream applications. Consistency in the time of collection, as well as other phlebotomy parameters, is necessary, concomitant with a similar time frame of miRNA extraction. For the extraction method selection, other factors have to be considered, such as the available initial fluid volume and the expected yield. Finally, quality control is a necessary step for successful downstream applications.

Lastly, we should stress the fact that it is not possible to unequivocally assign the origin of detected miRNAs. While circulating miRNAs are associated with EVs [36, 70], plasma lipoproteins [69] and proteins [68], the levels of each of these components, combined with efficiency of the extraction method used, further contribute to bias in interpreting miRNA expression levels. The sections below discuss newer methods for isolating all circulating extracellular miRNAs, regardless of plasma or serum components with which they may be associated.

Considerations for the collection of plasma and serum

Plasma or serum?

The essential difference between plasma and serum is the presence and absence, respectively, of fibrinogen and clotting factors. More importantly, platelets contain a wide spectrum of miRNAs [36], and these may be released into the serum during coagulation, together with miRNAs from red and white blood cells [103]. In the clinical settings, both plasma and serum are used for extracellular miRNA detection. Plasma is routinely collected in various tubes containing anticoagulants [K2 ethylenediaminetetraacetic acid (EDTA), Na3 citrate, heparin] and serum in tubes that promote coagulation and permit clot separation from serum (serum separators). Many studies comparing these biological fluids side-by-side find little or no difference in extracellular miRNA quantification [35, 37], although higher concentrations were consistently found in sera [103].

Moreover, some important factors to consider include that: (i ) platelets release EVs upon stimulation [104]; (ii ) serum miRNAs are associated with EVs [102]; and (iii ) some circulating miRNAs in both plasma and serum, such as miRs-126,-150,-191 and-223, display differential expression upon platelet activation, which responds to antiplatelet therapy [105]. These variables may contribute to notable differences in reporting. Indeed, in a study dedicated to the methodological analysis of circulating extracellular miRNAs, the concentration of several miRNAs (miRs-15b,-16 and-24) was significantly higher in plasma collected in EDTA-containing tubes compared with serum [73]. Wang et al. found higher total miRNA concentration in serum; however, for specific miRNA, the results were more nuanced: for some of the more abundant miRNA species in blood cells (miR-150, miR-16 and miR-126), the concentrations were the same between serum and plasma, whereas for others, they observed higher concentrations in the serum (e.g. miR-15b and miR-451) [103].

Thus, while plasma and sera generally have similar miRNA expression patterns, in specific instances, significant differences between these biological fluids are apparent [103, 105]. As there is insufficient data to account for all of these situations, a standard approach should be followed to ensure intra-experiment or intra-clinical trial consistency. Additional caution should be used in the case of archived samples, which are mostly stored as sera [103].

The effect of anticoagulants

In the clinical setting, the commonly used anticoagulants for plasma collection are EDTA [36], citrate [42] or heparin [74]. In selecting a tube for blood collection, consideration must be given to the platform that will be used to profile the miRNA. One needs to avoid introducing reagents that inhibit enzymes that will be used later in the process, and that are difficult to be removed prior to subsequent profiling. For example, heparin inhibits the reverse transcriptase and polymerase enzymes used in PCR [106, 107]. Therefore, heparinase I [74] or LiCl [108] treatments of RNA preparations need to be performed prior to quantitative real-time PCR (qRT-PCR)-based profiling. Of note, removal of the heparin may not be complete and may also reduce RNA yield. Similar to heparin, citrate can interfere with PCR from blood samples [109]. Fichtlscherer et al. found that miRNA expression profiling was better in EDTA-anticoagulated blood compared with blood collected in sodium citrate [81]. Recently, we have confirmed that citrate interferes with qRT-PCR-based miRNA assessment [101].

In conclusion, the choice among regularly used anticoagulants should be EDTA. Citrate and heparin should be avoided as a result of their inhibitory effects on qRT-PCR. In addition, heparinized blood should be avoided for next-generation sequencing (NGS) studies as well, as these also utilize reverse transcriptase (reviewed in Section 3.5). The effect of citrate on this platform is unknown at this time. Furthermore, if a patient has several tubes of blood collected for various clinical tests, the EDTA-containing tube must be collected prior to those containing heparin or citrate, to avoid inadvertent contamination with these anticoagulants. All these critical elements need to be outlined in clinical protocols as well as in the study reports.

The effect of fasting and blood draw timing

Some blood components, such as glucose and lipoproteins, vary according to food intake. This is common knowledge; however, much less is known about how this impacts miRNA levels. Fasting may affect miRNA detection in different ways. First, various miRNAs may be transported in circulation as ‘cargo' by different carriers, which might themselves be affected by the fasting state of the individual. For example, HDL particles, which can carry some extracellular miRNAs, vary diurnally in their blood concentrations [69]. Second, higher levels of circulating lipoproteins might interfere with the extraction process, affecting efficiency of miRNA recovery [75]. In this context, one needs to correlate the observed changes in miRNA expression patterns in patients with atherosclerosis with their lipoprotein levels (Table 1). In the clinical setting, the general practice is to draw blood in fasting conditions; however, this is rarely mentioned in reports on circulating plasma or serum miRNAs. Wang et al. are one of the few research groups who have recorded this information [77].

Another factor that is not routinely considered is the diurnal variation in circulating miRNAs levels. Although it is not known whether circulating miRNAs undergo daily variations in concentration in human cases, there is at least one report of miRNA circadian rhythm in mouse serum [110]. Previously, we found no difference in the expression of extracellular miRNAs from plasma collected in the morning or afternoon from healthy individuals [36].

Thus, a good practice should be to maintain consistency in time of day for sample collection to minimize variability because of such unknown factors. Moreover, reports should attempt to capture all this information when describing their protocols in publications. Ultimately, this information will be necessary to facilitate standardization of practices for analysis of circulating miRNAs.

MiRNA stability and interference of blood cells with detection

miRNAs are desirable candidate biomarkers because of their stability over time, even after repeated freeze-thaw cycles in plasma and serum [35, 37, 72, 73]. Their stability is explained by several protective mechanisms, related to the association of extracellular miRNAs with plasma proteins [111], specifically argonaute-2 [68] or lipoproteins [69], or, in other circumstances, their encapsulation in EVs [36, 70]. Circulating EVs, which are phospholipid-enclosed vesicles of various sizes and cellular origins [112], deserve special attention because they are produced by many cell types, and their quantity, quality and composition vary according to the physio-pathological status, including cardiovascular events [113, 114]. One needs to take in account that the EVs are a mixed population in the blood. There are numerous published methods and several excellent reviews for EVs isolation and characterization from plasma or serum [115, 118]. To date, the technology is not adequate or advanced enough to isolate EV subpopulations of sufficient purity for miRNA profiling.

Nevertheless, although the extracellular miRNAs are stable in purified plasma or serum, it is recommended that the blood be processed within 6 hrs in EDTA tubes [119]. It is possible that the cellular components of blood are releasing miRNAs during the storage period that would account for some changes. In fact, extracellular miRNA expression changes in packed red blood cells after long-term storage [120, 121]. Thus, rapid processing, within 2–4 hrs from collection, is optimal. Furthermore, as peripheral blood cells, including erythrocytes, can contribute to extracellular miRNAs found in plasma or serum [122], increased blood cell counts in an individual need to be considered in the total miRNA recovered from blood-based analysis.

Phlebotomy variables such as lysis of erythrocytes and other blood cells, and subsequent miRNA release should be also considered [73, 123]. The importance of these factors also lies in the fact that haemoglobin and lactoferrin may be released in the process, and these have been demonstrated to inhibit subsequent qRT-PCR [107]. As discussed, platelet-derived miRNAs are present in plasma and serum [36, 105]. In some medical conditions, platelet counts increase and may be reflected as changes in plasma or serum miRNA expression patterns [124]. Although it is difficult to determine the contribution of each of these blood cell components to miRNAs found in the serum or plasma, studies should include at least WBC and platelet counts from individuals, and data should be expressed with regard to these variables. In addition, platelets can be inadvertently activated and thus induced to release miRNAs when using small gauge needles to collect blood [118]. Thus, standardization and reporting of these specific details are important when analysing circulating miRNAs as biomarkers.

In conclusion, to reduce the influence of all these factors, the protocols should be consistent in sample acquisition, storage and processing within the first 2–4 hrs from collection. Moreover, samples should be checked for small clots and haemolysis in both plasma and sera, as these could contribute to the variability in miRNAs detection. Another useful practice is to discard the first couple of ml of blood, which would remove possible tissue and cell contaminants derived from the puncture site [118]. Presentation of detailed information in publications will certainly advance our understanding of extracellular miRNAs in the circulation and how to best use them as biomarkers.

Extraction methods

As discussed in the introduction to this section, most of present techniques for miRNA isolation are based on phenol:chloroform extraction. However, until recently, these methods were designed for use with tissues or cell cultures. A major difficulty in RNA/miRNA isolation from body fluids, including plasma and serum, is the generation of large volumes of aqueous phase. To ensure adequate denaturing and removal of the high protein content from samples (albumin, immunoglobulins, coagulation and complement components among others), the lysis reagent-to-specimen ratio has to be increased several-fold. Together with the starting fluid volume, this is the most variable step in the protocols reviewed (Table 1). Unfortunately, very few authors reported the yield of RNA recovered in the specific conditions used, making it difficult to determine the efficiency of these extraction protocols.

The transport conditions in the blood for a particular miRNA species such as argonaute-2 [68], lipoproteins [69] or EVs [36, 102] also need to be considered prior to RNA isolation. As some EVs are produced from the exocytic pathway, their membranes have a different lipid composition from cellular plasma membranes [125, 126] indicating that methods used to extract miRNAs from cells may not be suitable for extracting EVs-derived miRNAs leading to changes in expression. Indeed, comparison of RNA extraction methods from in vitro–generated exosomes reported that phenol-based methods were less efficient than column-based ones for total RNA [101, 127]. Conversely, a better yield of small RNAs was recovered from exosomes using combined phenol and column-based methods [101, 127]. Of particular note, the total RNA extracted from EVs consists of various RNA species, of which miRNAs comprise a small percentage and transfer RNAs are the most abundant [128]. How each method impacts the isolation of various RNA species still needs to be empirically examined. Therefore, this section highlights general protocol considerations for isolating miRNAs from plasma or serum.

Phenol:Chloroform extraction for miRNA

This procedure relies on the differential solubility of cellular components in organic solvents, such as phenol, chloroform or ethanol. The main components of this protocol are phenol and guanidinium thiocyanate, mostly commonly marketed as Trizol®. Because Trizol® denatures proteins, including RNases, samples are safe for long-term storage [129]. After phase separation, RNA is recovered by precipitation with isopropyl alcohol. As miRNAs are small, ample time is needed to recover these RNA species. Thus, it is recommend that at least an overnight precipitation at −20°C or −80°C [36], followed by longer pelleting times be used, such as a 16,000–21,000 × g centrifugation for 1 hr, at 4°C.

While Trizol® is widely employed (Table 1), there are important considerations using this method. Recently, a paper was retracted [130], as the authors realized that there was selective loss of small RNA molecules with low GC content when using Trizol, especially when low number of cells were analysed [131]. The same was found true for some pre-miRNAs, small interfering RNA duplexes and transfer RNAs [131]. Although this study refers to miRNAs extracted from cells rather than body fluids, one should nevertheless be cautious as it is unknown if these factors alter RNA isolation from plasma or serum as well.

Silica-based miRNA recovery methods

Several methodologies are available, of which miRVanaPARIS™ and miRNeasy® mini kits are more widely used (Table 1). As the majority of studies do not report the actual yield and quality of miRNA, it makes direct comparison of these methods difficult.

The miRVanaPARIS™ kit is a commercially available method of separating both nucleic acids and proteins. The method is unique in that it isolates native proteins and small RNAs, using a non-ionic detergent to disrupt cells prior to phenol:chloroform extraction. The miRNAs are isolated over a glass-fibre filter. Specific to the miRVanaPARIS™ method is a two-part, sequential filtration with increasing ethanol concentrations, allowing for the collection of a highly enriched fraction of RNA molecules below 200 nucleotides. This method works very well for the isolation of miRNA from tissues, and greater than a third of the articles cited in this review have used this kit for body fluids as well, as it requires small starting fluid volumes (from <100 μl, up to 625 μl). On the basis of the lack of details in these studies (Table 1), we do not know the efficiency of miRNA recovery from body fluids.

The miRNeasy® Mini kit uses a silica-based column system to recover miRNAs. Some groups reported that this kit leads to a two-to threefold better yield than miRVanaPARIS™ kit [75, 81]; however, only a few studies used it for miRNA extraction from plasma/serum (Table 1) [39, 62]. The kit also utilizes phenol:chloroform to separate the miRNA from other plasma components, by the adsorption on a silica mini-column in the presence of ethanol. Notably, the binding, washing and elution steps can be automated using the Qiagen QIAcube [132], which, when available, decreases both working time and variability.

Several variations of the QIAzol™ and miRNeasy® protocol exist, encompassing () different starting plasma/serum volumes, between 40 and 400 μl [38, 81] and (ii ) different QIAzol:fluid ratios, from 1:1 [35, 37] to 3:1 [77], 3.5:1 [62], 5:1 [68] or 10:1 [134]. As the yields are almost never reported, one should empirically determine the best QIAzol™-to-sample ratio to obtain high-quality RNA from small sample volumes for one's studies. We recently published an optimized protocol in using this approach [101].

Other kits used for miRNA isolation from plasma/serum samples

The MicroRNA Extraction Kit has been used for total and miRNA isolation from 100 μl plasma that was subsequently subjected to qRT-PCR [87]. Specifically, this study examined the expression of only eight miRNAs from the recovered RNA. miRNA is extracted by homogenization in the lysis buffer rather than phenol and chloroform. The sample is then treated with ethanol. Prior to the recovery of miRNAs, the manufacturer suggests an optional in-column DNase treatment using a NucleoPur™ spin column.

For analysis of diagnostic and prognostic values of several circulating miRNAs in acute coronary syndrome, Widera et al. used MasterPureRNA Purification Kit [43]. These investigators started from 50 μl of plasma to examine the expression of a few miRNAs by qRT-PCR. Thus, both these kits recovered sufficient and suitable RNA to examine the expression of limited miRNA species.

Instead of using phenol:chloroform and silica-based columns, the Exiqon miRCURYRNA Isolation kit is based on a proprietary resin as a separation matrix. A recent comparison of several methods for miRNA isolation from exosomes found that this kit gave the highest yields [127]. However, to date, we did not find the miRCURYkit used for miRNA extraction from body fluids, although in its protocol, it specifies that it can be used with as low as 100 μl of whole blood.

Sigma-Aldrich® markets the mir Premier® microRNA Isolation Kit to purify miRNAs and other small RNAs without phenol and chloroform. According to the manufacturer, the proprietary lysis solution achieves three functions: RNA release, inactivation of ribonucleases, and precipitation of large RNA and genomic DNA. Similar to the majority of other kits, RNA is recovered by a silica column. To date, we did not find studies using this kit to extract miRNAs from plasma or serum.

In conclusion, the choice of the specific methodology for miRNA extraction from body fluids depends on many factors, such as (i ) available initial volume, (ii ) number of miRNA species intended to be assessed, (iii ) method of subsequent analysis, (iv ) ease of use and (v ) price per sample. Most of these variables are investigator/study-dependent; therefore, there is no universal solution.

Global quantity and quality assessment of miRNAs

Several methods exist to determine the concentration and quality of purified miRNAs. Spectrophotometric analysis is one of the easiest and most common methods to determine concentration and protein or phenol contamination in RNA preparations [135]. As the absorbance of phenol (270 nm) is in close proximity to that of nucleic acids (260 nm), phenol contamination can lead to overestimation of RNA quantity. Furthermore, proteins absorb light at 280 nm, but the ratio of absorbance at 260 and 280 (260/280) estimates protein contamination in a nucleic acid solution with low sensitivity [135].

It is difficult, using this method, to discern the relative composition of the isolated miRNAs compared with total RNA or other small non-coding molecules and precursor miRNAs. As both total extracellular RNA and miRNAs are recovered, the miRNA concentration is prone to be overestimated if the total RNA is degraded during the isolation process. These issues can be reflected in unreliable profiling results [136]. As extracellular miRNAs are less than 1% of the total RNA recovered [128], their concentration is often under the detection limits of spectrophotometric devices. In this instance, it is recommended to use a fixed volume rather than a fixed miRNA amount for qRT-PCR [37, 75].

When the miRNA amount and concentration are sufficient, quality assessment of the preparation can be performed by capillary electrophoresis using the Small RNA kit in the Agilent Bioanalyzer. This system examines the presence of small RNAs between 6 and 150 nucleotides and using an algorithm based on ribosomal RNA detection, assigns an RNA integrity number (RIN) to demonstrate miRNA quality [137, 138]. Although this system can quantify small RNA molecules <150 nucleotides, it cannot distinguish between precursor and mature miRNA forms. A low RIN sample might indicate the presence of degraded miRNA that is not of sufficient quality to profile large number of miRNAs. However, examination of individual miRNAs by qRT-PCR might still be accomplished [137]. We [36] and others [139, 140] found that RINs >6 are acceptable for miRNA profiling from plasma or serum samples. Of note, when miRNA is extracted from purified EVs, this algorithm is not applicable, as it relies on ribosomal RNAs, which are not consistent in EVs [127]. Thus, if using RIN for quality evaluation, then each investigator must determine and define the acceptable quality levels through their study.

MiRNAs profiling methods

To date, there are several methods to examine miRNA expression profiling including (i ) qRT-PCR, (ii ) microarrays, (iii ) sequence-specific hybridization in solution followed by miRNA molecules counting based on reporter probes and (iv) direct sequencing (Table 1). Each method has advantages and limitations (Fig. 2). In addition, miRNA assessment is also confronted with (i ) the inability to distinguish between precursor and mature forms of miRNAs on some profiling platforms and (ii ) the difficulty to validate and correlate individual miRNA expression between different profiling platforms [18, 103]. In addition, considerations for endogenous controls are dependent on the platform used. Although there are inconsistencies across profiling platforms, qRT-PCR platforms seem to have better sensitivity than array technologies for miRNA profiling from body fluids [141]. These issues and methods are briefly described below.

Figure 2.

Figure 2

Comparison of the common miRNA profiling methods. Workflow: indicates main steps after miRNA isolation for getting the raw expression data. Sensitivity, Specificity, Throughput, Absolute quantification/accuracy and Flexibility are classified as follows: +++ (very high); ++ (moderate); +/++ (moderate to low); + (low); and +/-(low or not applicable). Flexibility: refers to ease of customization. The only technology that allows absolute quantification is qRT-PCR, whereas NextGen sequencing is the only platform that can identify novel miRNAs. Data analysis is classified as: E (relatively easy); M (moderate) with various software applications available; and D (difficult) requiring advanced computational infrastructure. Other: Considerations and challenges for the respective technologies.

Using a qRT-PCR approach is limited in the species as well as the number of detectable miRNA species compared with microarrays. However, medium-and large-scale, high-throughput platforms are commercially available using a qRT-PCR approach, such as microfluidic cards or plates. These are customizable. However, for profiling a large number of miRNA species, the microfluidic cards are at the same time more flexible, but more limited in the number of miRNAs that can be examined compared with microarrays. To date, Qiagen PCR arrays cover human, mouse and rat, with primer assays extending to Rhesus, hamster, chicken, several viruses, etc., whereas Life Technologies assays cover ˜200 species. The design of PCR arrays can accommodate as few or as many miRNA species desired, but the higher their number, the more labour-intensive the study becomes, and less price-efficient compared with microarray methodology.

Real-time PCR

Overview of qRT-PCR platforms

Analysis of miRNA expression by qRT-PCR can be performed by either TaqMan® or SYBR® Green methodologies. However, each requires specific reverse transcription (RT) and PCR reagents. For the Life Technologies miRNA TaqMan® assays, a unique stem-loop RT primer is designed to quantify only mature miRNAs and is miRNA specific. This primer produces a primer/mature miRNA chimera that extends the 3′ end of the miRNA, generating a longer RT product for the specific primers and probe to anneal during the PCR step.

Similar to the TaqMan® assay, the Qiagen SYBR® green–based miScript PCR System enables the profiling of several hundred miRNAs from the same sample. The RT for this system utilizes poly(A) polymerase and a unique universal oligo(dT) primer, which extends the template and converts mature miRNA species, as well as all other RNA species (precursor miRNA, other ncRNAs and mRNAs) into cDNA. The specificity is facilitated by a dual buffer system for the RT step, which selectively converts either miRNA or mRNA to cDNA. For PCR, a miRNA-specific forward primer and a universal reverse primer are used in combination with QuantiTect SYBR Green PCR Master Mix.

Exiqon also offers a SYBR green–based qRT-PCR system, miRCURY LNAPCR Kit. This kit examines the expression of individual miRNAs requiring as little as 1 pg total RNA. To increase sensitivity, miRNA-specific PCR primers generated with their ‘locked nucleic acid' (LNA™) technology [21] are utilized. The ribose rings of nucleotides are ‘locked' by connecting the 2′-O to the 4′-C atom using a methylene bridge. This locked conformation increases thermal stability of the nucleotide and high-affinity base pairing to the complementary DNA strand. Furthermore, the oligonucleotides used for PCR can be shorter and still maintain high melting temperatures.

Pre-amplification for qRT-PCR

To profile hundreds of differentially expressed miRNAs from serum or plasma when RNA quantity is limited poses a technical challenge. Pre-amplification of cDNA can overcome this issue. Applied Biosystems offers a kit for pre-amplification of the cDNA obtained after RT using the Megaplex™ Primer pools [36, 72]. These stem-looped RT primers reduce the number of RT reactions, as well as the amount of total input RNA required to 1 ng per sample. Pre-amplification enhances the detection of low-expressed species and does not appear to introduce a bias in miRNA detection [143]. Furthermore, Megaplex™-generated template is versatile and can be used for individual TaqMan® miRNA Assays or TaqMan® Low Density Arrays (TLDA) Cards. The TaqMan® assay is one of the most specific and sensitive methods for miRNA profiling [141, 144], and is also one of the most widely used in our literature survey. It may be labour-intensive if using a 384-well format, but the TLDA cards and the TaqMan® Open Array® MicroRNA panels simplify the profiling process. However, additional equipment is required to use these platforms.

qRT-PCR normalization controls

Quantification of miRNA expression by qRT-PCR may be performed in two different ways [36, 144]. One method involves calculating the absolute quantification of the number of copies of a gene using a standard curve. This is generated for each miRNA by serial dilutions of the cDNA obtained from a known amount of template. The procedure is advantageous for diagnostic purposes, especially considering that there is no consensus on endogenous controls (as described below). However, it is impractical to use this method when profiling a higher number of miRNA species. It has been used occasionally, for instance for finding miRNA biomarkers for early detection of lung cancer [146].

The second method is the relative quantification: the level of expression is expressed as a ratio of the Ct values of the genes of interest to the Ct values of one or more genes, the ‘endogenous controls', considered invariant in (at least) the conditions tested including the control samples [147]. The method has been widely used for mRNA and miRNA isolated from tissues and cells, as many genes and small miRNA were found relatively constant and thus useful as endogenous controls. However, characterizing miRNA expression in plasma/serum using relative quantification has encountered a serious challenge: since their identification in these fluids, no universally invariant calibrator miRNA or any other small RNA molecule has been found to date. There are a few invariant miRNAs in solid tissues [148], but their presence and/or consistent expression in various pathological instances has not yet been demonstrated for any body fluid. This problem has three solutions so far, neither perfect.

The first solution is to spike-in exogenous, synthetic miRNA mimetics. These allow for normalization, as well as estimating the efficiency of miRNA extraction and the reverse transcription step. Mitchell et al. used three synthetic miRNAs corresponding to Caenorhabditis elegans miRNAs that do not have homologous sequences in humans: cel-miR-39, cel-miR-54 and cel-miR-238 [37]. These miRNAs are spiked-in after the addition of the denaturing agents to avoid their degradation by plasma RNases. Some investigators use all three [70, 73, 74, 78], while others use only one [43, 72, 81, 82]. Finally, some groups designed their own spike-in controls [41, 84] or used synthetic human miR-422b as it is minimally expressed in plasma [87]. However, a serious drawback of the spike-in method is its reliability for normalization when quantification of extracted RNA is not possible. Spike-in miRNAs are a good measure of the extraction efficiency; however, when the starting volume is low and the amount of miRNA cannot be reliably detected, some investigators chose to use the same starting sample volume for each sample, rather than the same miRNA amount [e.g. [42, 74]]. Moreover, using a fixed volume does not imply that RNA content is invariable; therefore, spiking-in of a fixed amount of one or several synthetic miRNAs into samples containing variables amounts of RNA with the aim of normalizing qRT-PCR is not the best choice.

The second solution is to use endogenous controls, such as small nuclear (sn)RNA and small nucleolar (sno)RNAs [63, 67, 80, 85, 149], or specific miRNAs including miR-1249 [83], miR-223 or miR-16 [75]. However, these miRNAs have been shown to be invariant only in specific instances, whereas in other situations, they were found to change with disease [150]. Furthermore, snRNAs/snoRNAs are variable among tissues [151]. We have utilized snRNAs and snoRNAs as well as 18s and 5s rRNAs for normalization of plasma miRNAs [36, 70].

Finally, mean normalization of the miRNA profiling data sets is another possibility to quantify miRNA expression [36, 70]. In this case, one determines which miRNA Ct values are invariant across all (control and disease) samples. Thus, either one or a set or several consistent miRNAs across the samples can be used for normalization. One caveat is that it may be difficult to compare data using this method with other normalization methods or data representing another disease condition. More studies are needed to identify miRNAs that can serve as a true universal miRNA endogenous controls [152]. As extracellular RNA expression data are shared and placed in repositories such as Vesiclepedia [153], the identification of invariant miRNAs in body fluids should emerge.

Some general rules regarding whether to use an endogenous or exogenous control need to be considered. Either control type needs to be amendable for the miRNA assay. They should have sizes similar to miRNAs, such as snRNAs, and the amplification efficiency should be comparable to the tested miRNAs. For an endogenous control, the level of expression should be invariant across tissues and cell types as well as in various physiological and pathological conditions. Exogenous controls need to be stable. Applied Biosystems offers detection assays for several snRNAs/snoRNAs, miRNAs as well as traditional controls (18S rRNA). However, their applicability to plasma or serum samples is reduced because the presence and the variability (or lack thereof) of these small RNA species are not comprehensively established. To date, no universal endogenous control for all experimental conditions exists, neither for RNA nor miRNA, a fact especially true for body fluids including plasma and serum. Thus, selection of several (at least three) exogenous and endogenous controls with the lowest variability among all samples within a specific study should be considered for normalization.

Hybridization-based detection of miRNA expression

A remarkable advantage of microarrays is their comprehensive coverage and, for some, the ability to be customized, thus making them a flexible and versatile tool [154]. However, specific instrumentation and software is required to perform these analyses. Depending on the microarray manufacturer, they differ according to the chemistries involved in miRNA labelling as well as probe design and methods used to immobilize the probes [155]. Moreover, as opposed to qRT-PCR, they cannot be used for absolute quantification and they have a lower sensitivity and specificity than qRT-PCR. For instance, Jensen et al. [141] have found that the GeneChip miRNA 2.0 Array platform from Affymetrix was not reliable at a low input level, similar to that of the miRNAs that can be recovered from small amounts of plasma (250 μl). Another limiting factor for microarrays is that, like qRT-PCR, they can only detect known miRNA species [155].

Furthermore, although both intra-and inter-platform reproducibility of data is obviously desirable, this is not always attainable. Various studies have been designed to understand the reproducibility, sensitivity and specificity of various platforms for mRNA [156] or miRNA [157]. With respect to miRNA, the results are staggering: the intra-platform reproducibility is usually good to very good; however, the agreement between various microarray technologies is low, especially for the lower expressed miRNAs [143, 157159]. This outcome was the same regardless of the profiling technology, and qRT-PCR always proved to be the most sensitive. The inability to validate miRNA expression profiles across platforms may be due not only to differences in technologies but may also be a reflection of the lack of standard methods for (i ) normalization, (ii ) miRNA processing and (iii ) distinguishing mature miRNAs from precursor miRNAs or other ncRNAs.

The specific choice of microarray platform depends on several factors, an important one being availability. Many are highly sophisticated and mostly accessible in core laboratories because of the high cost of instrumentation and technical specialization. Other factors may also influence the platform choice: (i ) As most of the arrays require as low as 100 ng starting material, the amount of biological material available may be a deciding factor. The miRNA needs to be concentrated within a maximum volume of 3–8 μl for many platforms. This can be easily attained when the sources are tissues or cells, but much more difficult from the small sample volumes from plasma. (ii ) The number of miRNAs to be examined and whether the investigator is interested in, for example, viral miRNA expression indicate that a wider coverage is needed (such as Affymetrix). If the objective is to follow a specific set of miRNAs as biomarkers or to better understand their role, then arrays that encompass a lower number of miRNAs are more suitable. Some of the technologies described are customizable (e.g. Agilent or NanoString), whereas for others, the fabrication process makes customization less easy to apply and more expensive (e.g. Affymetrix). (iii) Finally, other important issues are sensitivity, reproducibility and a good understanding of the type of information obtained by the raw data. Below, we highlight several microarray types that are widely used and one unique system involving miRNA hybridization in fluid rather than on a solid surface.

Agilent oligonucleotides microarrays

Samples containing as low as 100 ng of miRNA can be profiled using the Agilent printed DNA oligonucleotide microarrays [160, 161]. The labelling method starts from total RNA, without fractionation or amplification, and consists of RNA dephosphorylation followed by dye (Cyanine 3-pCp, Cy3) ligation using T4 RNA ligase. The dynamic range is reported to be from 0.2 amol to 2 fmol miRNA, and the probe design helps to equalize melting temperatures by adding a G to the 5′ end to complement the 3′ cytosine introduced during labelling and ultimately stabilizes the cognate miRNA [161]. The slide printing allows for customizations, for instance [162], that vary mainly in the oligonucleotide probe design and miRNA labelling, as well as other procedural aspects. A good agreement was found between the detection by Agilent microarray and qRT-PCR, although a few miRNAs were consistently differentially expressed between platforms [163]. As such, this finding calls for caution when qRT-PCR is used to validate for microarray results.

Affymetrix GeneChip© miRNA arrays

Affymetrix is widely known for mRNA GeneChips© and has recently released GeneChip miRNA 3.0 Array. This new version covers 100% of the miRNAs described in miRBase v.17 for 153 organisms as well as viral miRNAs, all on the same array. It has probe sets for small nucleolar snoRNAs and pre-miRNA hairpins. For detection, a poly(A) tail is added to the 3′ end, then a 3DNA® dendrimer (a branched structure of single-and double-stranded DNA conjugated with numerous biotin molecules) is ligated. The biotin-labelled miRNA is then hybridized to the chip and detected by streptavidin-phycoerythrin. The miRNA GeneChip© requires a minimum of 100 ng input total RNA and is not customizable. The procedure is straightforward and has the advantage of detecting human and viral miRNAs at the same time, which may shed light into the possible connections between some diseases and various viral infections. However, several reports found low correlation between Affymetrix miRNA GeneChip results and other types of microarrays or qRT-PCR [163, 164].

Exiqon miRCURY LNA™ microRNA arrays

This array has extensive coverage for human, mouse and rat miRNAs based on miRBase v.19 and contains 3100 capture probes as well as 146 viral miRNAs. Notably, this fluorescent system requires as little as 30 ng of input RNA. The capture probes have been designed using their LNA™ technology [21]. The LNA incorporation into the probes standardizes hybridization conditions, which are optimized against the melting temperature rather than GC content to increase specificity. This system also includes 52 spike-in control miRNAs for normalization. As for the other microarrays, the miRNA is first dephosphorylated and then a fluorophore is attached to the 3′ end. An advantage of these arrays is that they also allow for either one-or two-colour assays. However, although Exiqon claims higher sensitivity and specificity of the LNA technology, especially applicable to plasma or serum [165], we could find only few instances of its utilization for plasma miRNA profiling in humans [e.g. [166]].

NanoString technologies nCounter© miRNA expression assay

Unlike the microarray tools, NanoString Technologies does not utilize printed chips. Rather, it uses fluorescent colour-coded ‘molecular barcode' oligonucleotides that hybridize directly to the target molecules. A biotinylated capture probe and a miRNA-specific fluorescent reporter probe are hybridized to the miRNA in solution. The tripartite-hybridized molecule is affinity-purified and attached to a streptavidin-coated slide, then imaged to count each fluorescent ‘bar code' [167]. Importantly, the nCounter© Analysis System does not rely on enzymes for processing or amplification, but requires 100 ng of concentrated RNA (33 ng/μl). Therefore, when using plasma/serum-derived miRNAs, an additional concentration step is necessary. The company claims a sensitivity close to qRT-PCR; however, the number of miRNA species detectable in a single reaction is limited to ˜800, and the normalization method may affect outcome.

Next-generation sequencing

Perhaps, there is no more promising technology for analysing miRNA profiles than NGS, also known as massively parallel sequencing. There are currently several companies offering high-throughput NGS platforms [Illumina HiSeq200 or GAIIX, ABI SOLiD and the Roche GS FLX+ (454)], although newer smaller scale NGS platforms are becoming available to be used in individual laboratories (Illumina MiSeq, Invitrogen Ion Torrent and Roche GS Junior 454). In the short term, NGS offers important advantages over other technologies, including the possibility to generate comprehensive and definitive analyses of miRNAs in samples, comprising those derived from sera and plasma [168, 169]. The primary advantage of this technology is that it does not require knowledge of target miRNAs nor does it require specific probes or primers, and therefore does not limit studies to known miRNAs. Thus, until such time as unambiguous libraries of miRNAs are generated for humans and other organisms, NGS is currently the best platform for miRNA discovery.

Next-generation sequencing also offers several additional advantages compared to microarray profiling (reviewed by Pritchard and colleagues [155]): (i ) NGS is extremely sensitive. ABI claims that its SOLiD system detects one miRNA copy per cell. (ii ) NGS provides relative expression data for small RNAs in a sample with greater dynamic range than miRNA microarrays. Next-generation sequencing platforms discern relative expression over a six to seven log-fold range, allowing relative quantification of miRNAs that are estimated to vary in abundance by four orders of magnitude. It should be noted, however, that qRT-PCR is still the only platform capable of generating absolute quantification. (iii ) Another specific advantage of NGS relative to microarray technology is that NGS provides sequence data, allowing the investigator to distinguish isomiRs from miRNAs that differ by a single nucleotide, including changes related to RNA editing that may influence miRNA stability [170, 171]. (iv ) Lastly, NGS will generate a profile of all small RNAs in a sample, including ncRNAs, such as short interfering RNA (siRNA), piwi-RNA (piRNA) and repeat-associated siRNA (rasiRNA). Depending on the source and size range of RNA analysed, NGS may also capture other non-coding small RNAs including promoter-associated small RNAs (PASRs), transcription initiation RNAs (tiRNAs), centromere repeat-associated small interacting RNAs (crasiRNAs) and telomere-specific small RNAs (tel-sRNAs) [172].

A detailed overview of NGS platforms for miRNA sequencing and sample preparation parameters is reviewed elsewhere [155]. Sample preparation is similar for all platforms. After RNA isolation and size fractionation of the small RNA population, adaptors are ligated to the ends of the RNA molecules and reverse transcription is used to generate cDNA. PCR amplification of cDNA occurs on an immobilized surface and solid-phase PCR on the Illumina platform. The Roche and ABI platforms immobilize cDNA on beads followed by emulsion PCR.

The ABI SOLiD system requires 10–500 ng of total RNA, feasible for RNA amounts isolated from plasma or serum and can generate 120 million tags per slide or 240 million per run. The Illumina small RNA protocols require 50 ng–1 μg of RNA. Illumina systems can deliver 3.4 million to 3 billion reads per run and offer the advantage of multiplexing, which can generate 1–600 Gb of output. Sequences are then mapped to the appropriate reference genomes.

There are also some disadvantages of NGS platforms that require consideration: First, there is the high cost of profiling miRNA populations using NGS. Second, the data analysis requires an advanced computational infrastructure and bioinformatics support. We previously described a pipeline for bioinformatics analysis based on NGS data from cells infected with human cytomegalovirus including sequence tag alignment, threshold determination, visualization to reference genomes and comparison to sequences in miRBase [173]. Finally, it has been observed that miRNA sequences generated by NGS vary from miRBase reference sequence as a result of RNA editing [174, 175], 3′ end nucleotide additions and clusters of isomiRs showing variable 3′ and 5′ ends relative to the reference genome, which can influence comparison to validated sequences in miRBase.

The most profound advantage to the NGS approach is the potential for discovery of novel small RNAs such as the detection of ncRNAs. Notably, validation of these observations using experimental approaches (PCR, cloning and sequencing and/or northern analysis) to satisfy accepted miRNA criteria including size and the formation of hairpin precursor structures is necessary. Although miRBase provides a comprehensive database of validated miRNAs from multiple organisms, databases of other small RNAs are absent or still evolving. An excellent resource in this regard is RNAdb 2.0—a database of mammalian ncRNAs (http://research.imb.uq.edu.au/rnadb/) [177]. Thus, experimentally validated RNA sequences identified through NGS that do not fulfil standard miRNA criteria may require additional studies to establish classification in other small RNA families. In summary, NGS also offers researchers the opportunity to extend analysis to the exciting and rapidly expanding arena of small regulatory RNAs that may be relevant to disease. Cost and availability of computational infrastructure and bioinformatics support are top considerations when choosing this approach.

Discussion

The search for useful diagnostic and prognostic biomarkers obtained from minimally invasive techniques is at the forefront of many disease-oriented studies. From these studies, circulating extracellular miRNAs from the blood are emerging as ideal candidates. As highlighted throughout this review, using extracellular miRNAs from body fluids as biomarkers is in its infancy. While an abundance of studies report miRNAs differential expression, important procedural information is lacking and needs to be standardized and disseminated to the research community. Importantly, attention to the fasting status of the patient and blood collection procedures, such as anticoagulants and needle size, should be considered when profiling extracellular miRNA from blood. Furthermore, extraction methods vary among reports, while information on the starting fluid volume and especially the efficiency of extraction is often missing. The availability of this information will enable investigators to standardize procedures and perhaps compare profiles across various patient cohorts.

As discussed, there are currently no known endogenous controls for serum or plasma miRNAs [36, 37]. While the miRNA profile of plasma and serum are comparable in general terms [35, 37], subtle changes in expression patterns may be present. Finally, optimization of RNA extraction methodologies from fluid samples should enable the analysis of a few miRNA species up to several hundred miRNAs. Ultimately, the RNA quantity and quality may dictate the platform used to obtain the extracellular miRNA profile.

Conclusions

There is sufficient evidence to support circulating extracellular miRNAs, retrieved either from plasma or serum, as important players and/or promising biomarkers for a variety of diseases. Current studies are improving the ability to retrieve sufficient RNA for analysis and profile numerous miRNAs. However, the field still lacks consistency and standardization, and thus effort is needed to establish common practices. Furthermore, several important considerations are needed before designing studies to analyse circulating miRNAs. As this field evolves, circulating miRNA profiles from plasma or serum are still susceptible to significant technological advancement. The potential to obtain extracellular miRNA signature from body fluids collected by non-invasive means exists. In fact, the knowledge of isolating extracellular RNAs from plasma or serum can be extrapolated to other body fluids, such as saliva or urine [68, 165]. Certainly, this information will lead to the development of diagnostic tests from easily assessable biofluids. Thus, as future experimental and clinical investigations start filling the gaps of our knowledge, the power of using not only extracellular miRNAs but perhaps other RNA species will greatly expand our understanding of diseases and may lead to new exciting therapies.

Key Points

  • Introduction

  • Current methodologies for extracting miRNAs from plasma and serum

    • – Considerations for the collection of plasma and serum

    • – Extraction methods

    • – Global quantity and quality assessment of miRNAs

  • MiRNA Profiling Methods

    • – Real-time PCR

    • – Microarray-based detection of miRNA expression

    • – Exiqon miRCURY LNATM microRNA arrays

    • – NanoString Technologies nCounterc miRNA Expression Assay

    • – Next-generation sequencing

  • Discussion

  • Conclusions

Acknowledgments

Supported by NHLBI R01 HL067176 and R01 HL102464 (PI: Marsh) and American Thoracic Society (PI: Piper). This article is subject to the NIH Public Access Policy.

Conflicts of interest

The authors confirm that there are no conflicts of interest.

Author contribution

LM, KB, JT, JW, CBM and MGP wrote the manuscript. CBM and MGP critically reviewed the manuscript.

References

  • 1.Amiel J, Pontual de L, Henrion-Caude A. miRNA, development and disease. Adv Genet. 2012;80:1–36. doi: 10.1016/B978-0-12-404742-6.00001-6. [DOI] [PubMed] [Google Scholar]
  • 2.Baltimore D, Boldin MP, O'Connell RM, et al. MicroRNAs: new regulators of immune cell development and function. Nat Immunol. 2008;9:839–45. doi: 10.1038/ni.f.209. [DOI] [PubMed] [Google Scholar]
  • 3.Suzuki HI, Miyazono K. Emerging complexity of microRNA generation cascades. J Biochem. 2011;149:15–25. doi: 10.1093/jb/mvq113. [DOI] [PubMed] [Google Scholar]
  • 4.Leonardo TR, Schultheisz HL, Loring JF, et al. The functions of microRNAs in pluripotency and reprogramming. Nat Cell Biol. 2012;14:1114–21. doi: 10.1038/ncb2613. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Thum T, Galuppo P, Wolf C, et al. MicroRNAs in the human heart: a clue to fetal gene reprogramming in heart failure. Circulation. 2007;116:258–67. doi: 10.1161/CIRCULATIONAHA.107.687947. [DOI] [PubMed] [Google Scholar]
  • 6.Ohtani K, Dimmeler S. Control of cardiovascular differentiation by microRNAs. Basic Res Cardiol. 2011;106:5–11. doi: 10.1007/s00395-010-0139-7. [DOI] [PubMed] [Google Scholar]
  • 7.Dumortier O, Obberghen Van E. MicroRNAs in pancreas development. Diabetes Obes Metab. 2012;14:22–8. doi: 10.1111/j.1463-1326.2012.01656.x. [DOI] [PubMed] [Google Scholar]
  • 8.Motti D, Bixby JL, Lemmon VP. MicroRNAs and neuronal development. Semin Fetal Neonatal Med. 2012;17:347–52. doi: 10.1016/j.siny.2012.07.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Khoshgoo N, Kholdebarin R, Iwasiow BM, et al. MicroRNAs and lung development. Pediatr Pulmonol. 2013;48:317–23. doi: 10.1002/ppul.22739. [DOI] [PubMed] [Google Scholar]
  • 10.Marrone AK, Ho J. MicroRNAs: potential regulators of renal development genes that contribute to CAKUT. Pediatr Nephrol. 2013 doi: 10.1007/s00467-013-2599-0. doi: 10.1007/s00467-013-259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Tran T, Andersen R, Sherman SP, et al. Insights into skeletal muscle development and applications in regenerative medicine. Int Rev Cell Mol Biol. 2013;300:51–83. doi: 10.1016/B978-0-12-405210-9.00002-3. [DOI] [PubMed] [Google Scholar]
  • 12.Nana-Sinkam SP, Karsies T, Riscili B, et al. Lung microRNA: from development to disease. Expert Rev Respir Med. 2009;3:373–85. doi: 10.1586/ers.09.30. [DOI] [PubMed] [Google Scholar]
  • 13.Bauersachs J. Regulation of myocardial fibrosis by MicroRNAs. J Cardiovasc Pharmacol. 2010;56:454–9. doi: 10.1097/FJC.0b013e3181ee81df. [DOI] [PubMed] [Google Scholar]
  • 14.Yu S, Li G. MicroRNA expression and function in cardiac ischemic injury. J Cardiovasc Transl Res. 2010;3:241–5. doi: 10.1007/s12265-010-9168-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.D'Alessandra Y, Pompilio G, Capogrossi MC. MicroRNAs and myocardial infarction. Curr Opin Cardiol. 2012;27:228–35. doi: 10.1097/HCO.0b013e3283522052. [DOI] [PubMed] [Google Scholar]
  • 16.Chung AC, Yu X, Lan HY. MicroRNA and nephropathy: emerging concepts. Int J Nephrol Renovasc Dis. 2013;6:169–79. doi: 10.2147/IJNRD.S37885. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Leva Di G, Garofalo M, Croce CM. MicroRNAs in cancer. Annu Rev Pathol. 2013;9 doi: 10.1146/annurev-pathol-012513-104715. doi: 10.1146/annurev-pathol-012513-104715. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Callari M, Dugo M, Musella V, et al. Comparison of microarray platforms for measuring differential microRNA expression in paired normal/cancer colon tissues. PLoS ONE. 2012;7:e45105. doi: 10.1371/journal.pone.0045105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Cuk K, Zucknick M, Madhavan D, et al. Plasma microRNA panel for minimally invasive detection of breast cancer. PLoS ONE. 2013;8:e76729. doi: 10.1371/journal.pone.0076729. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Pizzamiglio S, Bottelli S, Ciniselli CM, et al. A normalization strategy for the analysis of plasma microRNA qPCR data in colorectal cancer. Int J Cancer. 2013 doi: 10.1002/ijc.28530. doi: 10.1002/ijc.28530. [DOI] [PubMed] [Google Scholar]
  • 21.Vester B, Wengel J. LNA (locked nucleic acid): high-affinity targeting of complementary RNA and DNA. Biochemistry. 2004;43:13233–41. doi: 10.1021/bi0485732. [DOI] [PubMed] [Google Scholar]
  • 22.Ikeda S, Kong SW, Lu J, et al. Altered microRNA expression in human heart disease. Physiol Genomics. 2007;31:367–73. doi: 10.1152/physiolgenomics.00144.2007. [DOI] [PubMed] [Google Scholar]
  • 23.Condorelli G, Latronico MV, Dorn GW. microRNAs in heart disease: putative novel therapeutic targets? Eur Heart J. 2010;31:649–58. doi: 10.1093/eurheartj/ehp573. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Dorn GW. MicroRNAs: redefining mechanisms in cardiac disease. J Cardiovasc Pharmacol. 2010;56:589–95. doi: 10.1097/FJC.0b013e3181f8d173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Creemers EE, Tijsen AJ, Pinto YM. Circulating microRNAs: novel biomarkers and extracellular communicators in cardiovascular disease? Circ Res. 2012;110:483–95. doi: 10.1161/CIRCRESAHA.111.247452. [DOI] [PubMed] [Google Scholar]
  • 26.Empel van VP, Windt De LJ, Costa Martins da PA. Circulating miRNAs: reflecting or affecting cardiovascular disease? Curr Hypertens Rep. 2012;14:498–509. doi: 10.1007/s11906-012-0310-7. [DOI] [PubMed] [Google Scholar]
  • 27.Nana-Sinkam SP, Hunter MG, Nuovo GJ, et al. Integrating the MicroRNome into the study of lung disease. Am J Respir Crit Care Med. 2009;179:4–10. doi: 10.1164/rccm.200807-1042PP. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Ezzie ME, Crawford M, Cho JH, et al. Gene expression networks in COPD: microRNA and mRNA regulation. Thorax. 2011;67:122–31. doi: 10.1136/thoraxjnl-2011-200089. [DOI] [PubMed] [Google Scholar]
  • 29.Skrzypski M, Dziadziuszko R, Jassem J. MicroRNA in lung cancer diagnostics and treatment. Mutat Res. 2011;717:25–31. doi: 10.1016/j.mrfmmm.2011.04.002. [DOI] [PubMed] [Google Scholar]
  • 30.Yendamuri S, Kratzke R. MicroRNA biomarkers in lung cancer: MiRacle or quagMiRe? Transl Res. 2011;157:209–15. doi: 10.1016/j.trsl.2010.12.015. [DOI] [PubMed] [Google Scholar]
  • 31.Zandberga E, Kozirovskis V, Abols A, et al. Cell-free microRNAs as diagnostic, prognostic, and predictive biomarkers for lung cancer. Genes Chromosom Cancer. 2013;52:356–69. doi: 10.1002/gcc.22032. [DOI] [PubMed] [Google Scholar]
  • 32.Fenoglio C, Ridolfi E, Galimberti D, et al. MicroRNAs as active players in the pathogenesis of multiple sclerosis. Int J Mol Sci. 2012;13:13227–39. doi: 10.3390/ijms131013227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Dogini DB, Avansini SH, Vieira AS, et al. MicroRNA regulation and dysregulation in epilepsy. Front Cell Neurosci. 2013;7:172. doi: 10.3389/fncel.2013.00172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Bernardo BC, Charchar FJ, Lin RC, et al. A microRNA guide for clinicians and basic scientists: background and experimental techniques. Heart Lung Circ. 2012;21:131–42. doi: 10.1016/j.hlc.2011.11.002. [DOI] [PubMed] [Google Scholar]
  • 35.Chen X, Ba Y, Ma L, et al. Characterization of microRNAs in serum: a novel class of biomarkers for diagnosis of cancer and other diseases. Cell Res. 2008;18:997–1006. doi: 10.1038/cr.2008.282. [DOI] [PubMed] [Google Scholar]
  • 36.Hunter MP, Ismail N, Zhang X, et al. Detection of microRNA expression in human peripheral blood microvesicles. PLoS ONE. 2008;3:e3694. doi: 10.1371/journal.pone.0003694. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Mitchell PS, Parkin RK, Kroh EM, et al. Circulating microRNAs as stable blood-based markers for cancer detection. Proc Natl Acad Sci USA. 2008;105:10513–8. doi: 10.1073/pnas.0804549105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Cortez MA, Calin GA. MicroRNA identification in plasma and serum: a new tool to diagnose and monitor diseases. Expert Opin Biol Ther. 2009;9:703–11. doi: 10.1517/14712590902932889. [DOI] [PubMed] [Google Scholar]
  • 39.Weber JA, Baxter DH, Zhang S, et al. The microRNA spectrum in 12 body fluids. Clin Chem. 2010;56:1733–41. doi: 10.1373/clinchem.2010.147405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Gilad S, Meiri E, Yogev Y, et al. Serum microRNAs are promising novel biomarkers. PLoS ONE. 2008;3:e3148. doi: 10.1371/journal.pone.0003148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Adachi T, Nakanishi M, Otsuka Y, et al. Plasma microRNA 499 as a biomarker of acute myocardial infarction. Clin Chem. 2010;56:1183–5. doi: 10.1373/clinchem.2010.144121. [DOI] [PubMed] [Google Scholar]
  • 42.Ai J, Zhang R, Li Y, et al. Circulating microRNA-1 as a potential novel biomarker for acute myocardial infarction. Biochem Biophys Res Commun. 2010;391:73–7. doi: 10.1016/j.bbrc.2009.11.005. [DOI] [PubMed] [Google Scholar]
  • 43.Widera C, Gupta SK, Lorenzen JM, et al. Diagnostic and prognostic impact of six circulating microRNAs in acute coronary syndrome. J Mol Cell Cardiol. 2011;51:872–5. doi: 10.1016/j.yjmcc.2011.07.011. [DOI] [PubMed] [Google Scholar]
  • 44.Menghini R, Casagrande V, Federici M. MicroRNAs in endothelial senescence and atherosclerosis. J Cardiovasc Transl Res. 2013;6:924–30. doi: 10.1007/s12265-013-9487-7. [DOI] [PubMed] [Google Scholar]
  • 45.Huang Y, Hu Q, Deng Z, et al. MicroRNAs in body fluids as biomarkers for non-small cell lung cancer: a systematic review. Technol Cancer Res Treat. 2013 doi: 10.7785/tcrt.2012.500377. doi: 10.7785/tcrt.2012.500377. [DOI] [PubMed] [Google Scholar]
  • 46.Cheng L, Quek CY, Sun X, et al. The detection of microRNA associated with Alzheimer's disease in biological fluids using next-generation sequencing technologies. Front Genet. 2013;4:150. doi: 10.3389/fgene.2013.00150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Dorval V, Nelson PT, Hebert SS. Circulating microRNAs in Alzheimer's disease: the search for novel biomarkers. Front Mol Neurosci. 2013;6:24. doi: 10.3389/fnmol.2013.00024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Jin XF, Wu N, Wang L, et al. Circulating microRNAs: a novel class of potential biomarkers for diagnosing and prognosing central nervous system diseases. Cell Mol Neurobiol. 2013;33:601–13. doi: 10.1007/s10571-013-9940-9. [DOI] [PubMed] [Google Scholar]
  • 49.Sheinerman KS, Umansky SR. Circulating cell-free microRNA as biomarkers for screening, diagnosis and monitoring of neurodegenerative diseases and other neurologic pathologies. Front Cell Neurosci. 2013;7:150. doi: 10.3389/fncel.2013.00150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Babalola O, Mamalis A, Lev-Tov H, et al. The role of microRNAs in skin fibrosis. Arch Dermatol Res. 2013;305:763–6. doi: 10.1007/s00403-013-1410-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Gonzalez A. Osteoarthritis year 2013 in review: genetics and genomics. Osteoarthritis Cartilage. 2013;21:1443–51. doi: 10.1016/j.joca.2013.07.001. [DOI] [PubMed] [Google Scholar]
  • 52.D'Alessandra Y, Devanna P, Limana F, et al. Circulating microRNAs are new and sensitive biomarkers of myocardial infarction. Eur Heart J. 2010;31:2765–73. doi: 10.1093/eurheartj/ehq167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Eeden Van S, Leipsic J, Paul Man SF, et al. The relationship between lung inflammation and cardiovascular disease. Am J Respir Crit Care Med. 2012;186:11–6. doi: 10.1164/rccm.201203-0455PP. [DOI] [PubMed] [Google Scholar]
  • 54.Zhang Y, Kaminski N. Biomarkers in idiopathic pulmonary fibrosis. Curr Opin Pulm Med. 2012;18:441–6. doi: 10.1097/MCP.0b013e328356d03c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Oerlemans MI, Mosterd A, Dekker MS, et al. Early assessment of acute coronary syndromes in the emergency department: the potential diagnostic value of circulating microRNAs. EMBO Mol Med. 2012;4:1176–85. doi: 10.1002/emmm.201201749. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Li YQ, Zhang MF, Wen HY, et al. Comparing the diagnostic values of circulating microRNAs and cardiac troponin T in patients with acute myocardial infarction. Clinics (Sao Paulo) 2013;68:75–80. doi: 10.6061/clinics/2013(01)OA12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Pottelberge GR, Mestdagh P, Bracke KR, et al. MicroRNA expression in induced sputum of smokers and patients with chronic obstructive pulmonary disease. Am J Respir Crit Care Med. 2011;183:898–906. doi: 10.1164/rccm.201002-0304OC. [DOI] [PubMed] [Google Scholar]
  • 58.Soeda S, Ohyashiki JH, Ohtsuki K, et al. Clinical relevance of plasma miR-106b levels in patients with chronic obstructive pulmonary disease. Int J Mol Med. 2013;31:533–9. doi: 10.3892/ijmm.2013.1251. [DOI] [PubMed] [Google Scholar]
  • 59.Rhodes CJ, Wharton J, Boon RA, et al. Reduced microRNA-150 is associated with poor survival in pulmonary arterial hypertension. Am J Respir Crit Care Med. 2013;187:294–302. doi: 10.1164/rccm.201205-0839OC. [DOI] [PubMed] [Google Scholar]
  • 60.Molina-Pinelo S, Suarez R, Pastor MD, et al. Association between the miRNA signatures in plasma and bronchoalveolar fluid in respiratory pathologies. Dis Markers. 2012;32:221–30. doi: 10.3233/DMA-2011-0882. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Boeri M, Verri C, Conte D, et al. MicroRNA signatures in tissues and plasma predict development and prognosis of computed tomography detected lung cancer. Proc Natl Acad Sci USA. 2011;108:3713–8. doi: 10.1073/pnas.1100048108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Wang K, Zhang S, Marzolf B, et al. Circulating microRNAs, potential biomarkers for drug-induced liver injury. Proc Natl Acad Sci USA. 2009;106:4402–7. doi: 10.1073/pnas.0813371106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Starkey Lewis PJ, Dear J, Platt V, et al. Circulating microRNAs as potential markers of human drug-induced liver injury. Hepatology. 2011;54:1767–76. doi: 10.1002/hep.24538. [DOI] [PubMed] [Google Scholar]
  • 64.Fichtlscherer S, Zeiher AM, Dimmeler S. Circulating microRNAs: biomarkers or mediators of cardiovascular diseases? Arterioscler Thromb Vasc Biol. 2011;31:2383–90. doi: 10.1161/ATVBAHA.111.226696. [DOI] [PubMed] [Google Scholar]
  • 65.Zampetaki A, Willeit P, Drozdov I, et al. Profiling of circulating micrornas: from single biomarkers to re-wired networks. Cardiovasc Res. 2011;93:555–62. doi: 10.1093/cvr/cvr266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Garzon R, Calin GA, Croce CM. MicroRNAs in cancer. Annu Rev Med. 2009;60:167–79. doi: 10.1146/annurev.med.59.053006.104707. [DOI] [PubMed] [Google Scholar]
  • 67.Li S, Zhu J, Zhang W, et al. Signature microRNA expression profile of essential hypertension and its novel link to human cytomegalovirus infection. Circulation. 2011;124:175–84. doi: 10.1161/CIRCULATIONAHA.110.012237. [DOI] [PubMed] [Google Scholar]
  • 68.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. 2011;108:5003–8. doi: 10.1073/pnas.1019055108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Vickers KC, Palmisano BT, Shoucri BM, et al. MicroRNAs are transported in plasma and delivered to recipient cells by high-density lipoproteins. Nat Cell Biol. 2011;13:423–33. doi: 10.1038/ncb2210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Ismail N, Wang Y, Dakhlallah D, et al. Macrophage microvesicles induce macrophage differentiation and miR-223 transfer. Blood. 2013;121:984–95. doi: 10.1182/blood-2011-08-374793. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Lawrie CH, Gal S, Dunlop HM, et al. Detection of elevated levels of tumour-associated microRNAs in serum of patients with diffuse large B-cell lymphoma. Br J Haematol. 2008;141:672–5. doi: 10.1111/j.1365-2141.2008.07077.x. [DOI] [PubMed] [Google Scholar]
  • 72.Turchinovich A, Weiz L, Langheinz A, et al. Characterization of extracellular circulating microRNA. Nucleic Acids Res. 2011;39:7223–33. doi: 10.1093/nar/gkr254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.McDonald JS, Milosevic D, Reddi HV, et al. Analysis of circulating microRNA: preanalytical and analytical challenges. Clin Chem. 2011;57:833–40. doi: 10.1373/clinchem.2010.157198. [DOI] [PubMed] [Google Scholar]
  • 74.Heegaard NH, Schetter AJ, Welsh JA, et al. Circulating microRNA expression profiles in early stage non-small cell lung cancer. Int J Cancer. 2011;130:1378–86. doi: 10.1002/ijc.26153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Kroh EM, Parkin RK, Mitchell PS, et al. Analysis of circulating microRNA biomarkers in plasma and serum using quantitative reverse transcription-PCR (qRT-PCR) Methods. 2010;50:298–301. doi: 10.1016/j.ymeth.2010.01.032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Cheng Y, Tan N, Yang J, et al. A translational study of circulating cell-free microRNA-1 in acute myocardial infarction. Clin Sci. 2010;119:87–95. doi: 10.1042/CS20090645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Wang GK, Zhu JQ, Zhang JT, et al. Circulating microRNA: a novel potential biomarker for early diagnosis of acute myocardial infarction in humans. Eur Heart J. 2010;31:659–66. doi: 10.1093/eurheartj/ehq013. [DOI] [PubMed] [Google Scholar]
  • 78.Corsten MF, Dennert R, Jochems S, et al. Circulating MicroRNA-208b and MicroRNA-499 reflect myocardial damage in cardiovascular disease. Circ Cardiovasc Genet. 2010;3:499–506. doi: 10.1161/CIRCGENETICS.110.957415. [DOI] [PubMed] [Google Scholar]
  • 79.Kuwabara Y, Ono K, Horie T, et al. Increased microRNA-1 and microRNA-133a levels in serum of patients with cardiovascular disease indicate the existence of myocardial damage. Circ Cardiovasc Genet. 2011;4:446–54. doi: 10.1161/CIRCGENETICS.110.958975. [DOI] [PubMed] [Google Scholar]
  • 80.Wang R, Li N, Zhang Y, et al. Circulating MicroRNAs are promising novel biomarkers of acute myocardial infarction. Intern Med. 2011;50:1789–95. doi: 10.2169/internalmedicine.50.5129. [DOI] [PubMed] [Google Scholar]
  • 81.Fichtlscherer S, Rosa De S, Fox H, et al. Circulating microRNAs in patients with coronary artery disease. Circ Res. 2010;107:677–84. doi: 10.1161/CIRCRESAHA.109.215566. [DOI] [PubMed] [Google Scholar]
  • 82.Rosa De S, Fichtlscherer S, Lehmann R, et al. Transcoronary concentration gradients of circulating microRNAs. Circulation. 2011;124:1936–44. doi: 10.1161/CIRCULATIONAHA.111.037572. [DOI] [PubMed] [Google Scholar]
  • 83.Tijsen AJ, Creemers EE, Moerland PD, et al. MiR423-5p as a circulating biomarker for heart failure. Circ Res. 2010;106:1035–9. doi: 10.1161/CIRCRESAHA.110.218297. [DOI] [PubMed] [Google Scholar]
  • 84.Fukushima Y, Nakanishi M, Nonogi H, et al. Assessment of plasma miRNAs in congestive heart failure. Circ J. 2011;75:336–40. doi: 10.1253/circj.cj-10-0457. [DOI] [PubMed] [Google Scholar]
  • 85.Li T, Cao H, Zhuang J, et al. Identification of miR-130a, miR-27b and miR-210 as serum biomarkers for atherosclerosis obliterans. Clin Chim Acta. 2011;412:66–70. doi: 10.1016/j.cca.2010.09.029. [DOI] [PubMed] [Google Scholar]
  • 86.Villar AV, Garcia R, Merino D, et al. Myocardial and circulating levels of microRNA-21 reflect left ventricular fibrosis in aortic stenosis patients. Int J Cardiol. 2012;167:2875–81. doi: 10.1016/j.ijcard.2012.07.021. [DOI] [PubMed] [Google Scholar]
  • 87.Baggish AL, Hale A, Weiner RB, et al. Dynamic regulation of circulating microRNA during acute exhaustive exercise and sustained aerobic exercise training. J Physiol. 2011;589:3983–94. doi: 10.1113/jphysiol.2011.213363. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Gao W, He HW, Wang ZM, et al. Plasma levels of lipometabolism-related miR-122 and miR-370 are increased in patients with hyperlipidemia and associated with coronary artery disease. Lipids Health Dis. 2012;11:55. doi: 10.1186/1476-511X-11-55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Zile MR, Mehurg SM, Arroyo JE, et al. Relationship between the temporal profile of plasma microRNA and left ventricular remodeling in patients after myocardial infarction. Circ Cardiovasc Genet. 2011;4:614–9. doi: 10.1161/CIRCGENETICS.111.959841. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Devaux Y, Vausort M, Goretti E, et al. Use of circulating microRNAs to diagnose acute myocardial infarction. Clin Chem. 2012;58:559–67. doi: 10.1373/clinchem.2011.173823. [DOI] [PubMed] [Google Scholar]
  • 91.Eitel I, Adams V, Dieterich P, et al. Relation of circulating MicroRNA-133a concentrations with myocardial damage and clinical prognosis in ST-elevation myocardial infarction. Am Heart J. 2012;164:706–14. doi: 10.1016/j.ahj.2012.08.004. [DOI] [PubMed] [Google Scholar]
  • 92.Zampetaki A, Willeit P, Tilling L, et al. Prospective study on circulating MicroRNAs and risk of myocardial infarction. J Am Coll Cardiol. 2012;60:290–9. doi: 10.1016/j.jacc.2012.03.056. [DOI] [PubMed] [Google Scholar]
  • 93.Hennessey PT, Sanford T, Choudhary A, et al. Serum microRNA biomarkers for detection of non-small cell lung cancer. PLoS ONE. 2012;7:e32307. doi: 10.1371/journal.pone.0032307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Hu Z, Chen X, Zhao Y, et al. Serum microRNA signatures identified in a genome-wide serum microRNA expression profiling predict survival of non-small-cell lung cancer. J Clin Oncol. 2010;28:1721–6. doi: 10.1200/JCO.2009.24.9342. [DOI] [PubMed] [Google Scholar]
  • 95.Silva J, Garcia V, Zaballos A, et al. Vesicle-related microRNAs in plasma of nonsmall cell lung cancer patients and correlation with survival. Eur Respir J. 2011;37:617–23. doi: 10.1183/09031936.00029610. [DOI] [PubMed] [Google Scholar]
  • 96.Leidinger P, Keller A, Borries A, et al. Specific peripheral miRNA profiles for distinguishing lung cancer from COPD. Lung Cancer. 2011;74:41–7. doi: 10.1016/j.lungcan.2011.02.003. [DOI] [PubMed] [Google Scholar]
  • 97.Shen J, Liu Z, Todd NW, et al. Diagnosis of lung cancer in individuals with solitary pulmonary nodules by plasma microRNA biomarkers. BMC Cancer. 2011;11:374. doi: 10.1186/1471-2407-11-374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Shen J, Wang A, Wang Q, et al. Exploration of genome-wide circulating microRNA in hepatocellular carcinoma (HCC): MiR-483-5p as a potential biomarker. Cancer Epidemiol Biomarkers Prev. 2013;22:2364–73. doi: 10.1158/1055-9965.EPI-13-0237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Kumar P, Dezso Z, MacKenzie C, et al. Circulating miRNA biomarkers for Alzheimer's disease. PLoS ONE. 2013;8:e69807. doi: 10.1371/journal.pone.0069807. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Lorenzen JM, Kielstein JT, Hafer C, et al. Circulating miR-210 predicts survival in critically ill patients with acute kidney injury. Clin J Am Soc Nephrol. 2011;6:1540–6. doi: 10.2215/CJN.00430111. [DOI] [PubMed] [Google Scholar]
  • 101.Moldovan L, Batte K, Wang Y, et al. Analyzing the circulating microRNAs in exosomes/extracellular vesicles from serum or plasma by qRT-PCR. Methods Mol Biol. 2013;1024:129–45. doi: 10.1007/978-1-62703-453-1_10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Gallo A, Tandon M, Alevizos I, et al. The majority of microRNAs detectable in serum and saliva is concentrated in exosomes. PLoS ONE. 2012;7:e30679. doi: 10.1371/journal.pone.0030679. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Wang K, Yuan Y, Cho JH, et al. Comparing the microRNA spectrum between serum and plasma. PLoS ONE. 2012;7:e41561. doi: 10.1371/journal.pone.0041561. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.George JN, Thoi LL, McManus LM, et al. Isolation of human platelet membrane microparticles from plasma and serum. Blood. 1982;60:834–40. [PubMed] [Google Scholar]
  • 105.Willeit P, Zampetaki A, Dudek K, et al. Circulating microRNAs as novel biomarkers for platelet activation. Circ Res. 2013;112:595–600. doi: 10.1161/CIRCRESAHA.111.300539. [DOI] [PubMed] [Google Scholar]
  • 106.Willems M, Moshage H, Nevens F, et al. Plasma collected from heparinized blood is not suitable for HCV-RNA detection by conventional RT-PCR assay. J Virol Methods. 1993;42:127–30. doi: 10.1016/0166-0934(93)90184-s. [DOI] [PubMed] [Google Scholar]
  • 107.Al Soud WA, Radstrom P. Purification and characterization of PCR-inhibitory components in blood cells. J Clin Microbiol. 2001;39:485–93. doi: 10.1128/JCM.39.2.485-493.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Wang J, Chen J, Chang P, et al. MicroRNAs in plasma of pancreatic ductal adenocarcinoma patients as novel blood-based biomarkers of disease. Cancer Prev Res. 2009;2:807–13. doi: 10.1158/1940-6207.CAPR-09-0094. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Garcia ME, Blanco JL, Caballero J, et al. Anticoagulants interfere with PCR used to diagnose invasive aspergillosis. J Clin Microbiol. 2002;40:1567–8. doi: 10.1128/JCM.40.4.1567-1568.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Shende VR, Goldrick MM, Ramani S, et al. Expression and rhythmic modulation of circulating microRNAs targeting the clock gene Bmal1 in mice. PLoS ONE. 2011;6:e22586. doi: 10.1371/journal.pone.0022586. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Wang K, Zhang S, Weber J, et al. Export of microRNAs and microRNA-protective protein by mammalian cells. Nucleic Acids Res. 2010;38:7248–59. doi: 10.1093/nar/gkq601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Pol van der E, Boing AN, Harrison P, et al. Classification, functions, and clinical relevance of extracellular vesicles. Pharmacol Rev. 2012;64:676–705. doi: 10.1124/pr.112.005983. [DOI] [PubMed] [Google Scholar]
  • 113.Lee ST, Chu K, Jung KH, et al. Circulating CD62E+ microparticles and cardiovascular outcomes. PLoS ONE. 2012;7:e35713. doi: 10.1371/journal.pone.0035713. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Shantsila E, Kamphuisen PW, Lip GY. Circulating microparticles in cardiovascular disease: implications for atherogenesis and atherothrombosis. J Thromb Haemost. 2010;8:2358–68. doi: 10.1111/j.1538-7836.2010.04007.x. [DOI] [PubMed] [Google Scholar]
  • 115.Orozco AF, Lewis DE. Flow cytometric analysis of circulating microparticles in plasma. Cytometry A. 2010;77:502–14. doi: 10.1002/cyto.a.20886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Yuana Y, Bertina RM, Osanto S. Pre-analytical and analytical issues in the analysis of blood microparticles. Thromb Haemost. 2010;105:396–408. doi: 10.1160/TH10-09-0595. [DOI] [PubMed] [Google Scholar]
  • 117.Grant R, Ansa-Addo E, Stratton D, et al. A filtration-based protocol to isolate human Plasma Membrane-derived Vesicles and exosomes from blood plasma. J Immunol Methods. 2011;371:143–51. doi: 10.1016/j.jim.2011.06.024. [DOI] [PubMed] [Google Scholar]
  • 118.Witwer K, Buzas EI, Bemis LT, et al. Standardization of sample collection, isolation and analysis methods in extracellular vesicle research. J Extracell Vesicles. 2013;2:20360. doi: 10.3402/jev.v2i0.20360. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Tsui NB, Ng EK, Lo YM. Stability of endogenous and added RNA in blood specimens, serum, and plasma. Clin Chem. 2002;48:1647–53. [PubMed] [Google Scholar]
  • 120.Kannan M, Mohan KV, Kulkarni S, et al. Membrane array-based differential profiling of platelets during storage for 52 miRNAs associated with apoptosis. Transfusion. 2009;49:1443–50. doi: 10.1111/j.1537-2995.2009.02140.x. [DOI] [PubMed] [Google Scholar]
  • 121.Kannan M, Atreya C. Differential profiling of human red blood cells during storage for 52 selected microRNAs. Transfusion. 2010;50:1581–8. doi: 10.1111/j.1537-2995.2010.02585.x. [DOI] [PubMed] [Google Scholar]
  • 122.Duttagupta R, Jiang R, Gollub J, et al. Impact of cellular miRNAs on circulating miRNA biomarker signatures. PLoS ONE. 2011;6:e20769. doi: 10.1371/journal.pone.0020769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Kirschner MB, Kao SC, Edelman JJ, et al. Haemolysis during sample preparation alters microRNA content of plasma. PLoS ONE. 2011;6:e24145. doi: 10.1371/journal.pone.0024145. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Schafer AI. Thrombocytosis and thrombocythemia. Blood Rev. 2001;15:159–66. doi: 10.1054/blre.2001.0162. [DOI] [PubMed] [Google Scholar]
  • 125.Laulagnier K, Motta C, Hamdi S, et al. Mast cell-and dendritic cell-derived exosomes display a specific lipid composition and an unusual membrane organization. Biochem J. 2004;380:161–71. doi: 10.1042/BJ20031594. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Llorente A, Skotland T, Sylvanne T, et al. Molecular lipidomics of exosomes released by PC-3 prostate cancer cells. Biochim Biophys Acta. 2013;1831:1302–9. doi: 10.1016/j.bbalip.2013.04.011. [DOI] [PubMed] [Google Scholar]
  • 127.Eldh M, Lotvall J, Malmhall C, et al. Importance of RNA isolation methods for analysis of exosomal RNA: evaluation of different methods. Mol Immunol. 2012;50:278–86. doi: 10.1016/j.molimm.2012.02.001. [DOI] [PubMed] [Google Scholar]
  • 128.Bellingham SA, Coleman BM, Hill AF. Small RNA deep sequencing reveals a distinct miRNA signature released in exosomes from prion-infected neuronal cells. Nucleic Acids Res. 2012;40:10937–49. doi: 10.1093/nar/gks832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Ma W, Wang M, Wang ZQ, et al. Effect of long-term storage in TRIzol on microarray-based gene expression profiling. Cancer Epidemiol Biomarkers Prev. 2010;19:2445–52. doi: 10.1158/1055-9965.EPI-10-0565. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Kim YK, Yeo J, Ha M, et al. Cell adhesion-dependent control of microRNA decay. Mol Cell. 2011;43:1005–14. doi: 10.1016/j.molcel.2011.07.031. [DOI] [PubMed] [Google Scholar]
  • 131.Kim YK, Yeo J, Kim B, et al. Short structured RNAs with low GC content are selectively lost during extraction from a small number of cells. Mol Cell. 2012;46:893–5. doi: 10.1016/j.molcel.2012.05.036. [DOI] [PubMed] [Google Scholar]
  • 132.Qiagen. 2011. Purification of total RNA, including small RNAs, from serum or plasma using the miRNeasy Mini Kit; available at: http://www.qiagen.com/products/rnastabilizationpurification/microrna/mirneasyminikit aspx#Tabs=t2.
  • 133.Rayner KJ, Esau CC, Hussain FN, et al. Inhibition of miR-33a/b in non-human primates raises plasma HDL and lowers VLDL triglycerides. Nature. 2011;478:404–7. doi: 10.1038/nature10486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Laterza OF, Lim L, Garrett-Engele PW, et al. Plasma MicroRNAs as sensitive and specific biomarkers of tissue injury. Clin Chem. 2009;55:1977–83. doi: 10.1373/clinchem.2009.131797. [DOI] [PubMed] [Google Scholar]
  • 135.Glasel JA. Validity of nucleic acid purities monitored by 260 nm/280 nm absorbance ratios. Biotechniques. 1995;18:62–3. [PubMed] [Google Scholar]
  • 136.Becker C, Hammerle-Fickinger A, Riedmaier I, et al. mRNA and microRNA quality control for RT-qPCR analysis. Methods. 2010;50:237–43. doi: 10.1016/j.ymeth.2010.01.010. [DOI] [PubMed] [Google Scholar]
  • 137.Mueller O, Lightfoot S, Schroeder A. RNA Integrity Number (RIN) – standardization of RNA quality control. Waldbronn, Germany: Agilent Technologies; 2004. Agilent Application Note, Publication Number 5989-1165EN. [Google Scholar]
  • 138.Tissot C. Analysis of miRNA content in total RNA preparations using the Agilent 2100 bioanalyzer. Waldbronn, Germany: Agilent Technologies; 2008. Agilent Application Note, Publication Number 5989-7870EN. [Google Scholar]
  • 139.Vasilescu C, Rossi S, Shimizu M, et al. MicroRNA fingerprints identify miR-150 as a plasma prognostic marker in patients with sepsis. PLoS ONE. 2009;4:e7405. doi: 10.1371/journal.pone.0007405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Shen J, Todd NW, Zhang H, et al. Plasma microRNAs as potential biomarkers for non-small-cell lung cancer. Lab Invest. 2011;91:579–87. doi: 10.1038/labinvest.2010.194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.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. 2011;12:435. doi: 10.1186/1471-2164-12-435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Zampetaki A, Kiechl S, Drozdov I, et al. Plasma microRNA profiling reveals loss of endothelial miR-126 and other microRNAs in type 2 diabetes. Circ Res. 2010;107:810–7. doi: 10.1161/CIRCRESAHA.110.226357. [DOI] [PubMed] [Google Scholar]
  • 143.Chen Y, Gelfond JA, McManus LM, et al. Reproducibility of quantitative RT-PCR array in miRNA expression profiling and comparison with microarray analysis. BMC Genomics. 2009;10:407. doi: 10.1186/1471-2164-10-407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Schmittgen TD, Lee EJ, Jiang J, et al. Real-time PCR quantification of precursor and mature microRNA. Methods. 2008;44:31–8. doi: 10.1016/j.ymeth.2007.09.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods. 2001;25:402–8. doi: 10.1006/meth.2001.1262. [DOI] [PubMed] [Google Scholar]
  • 146.Zheng D, Haddadin S, Wang Y, et al. Plasma microRNAs as novel biomarkers for early detection of lung cancer. Int J Clin Exp Pathol. 2011;4:575–86. [PMC free article] [PubMed] [Google Scholar]
  • 147.Schmittgen TD, Livak KJ. Analyzing real-time PCR data by the comparative C(T) method. Nat Protoc. 2008;3:1101–8. doi: 10.1038/nprot.2008.73. [DOI] [PubMed] [Google Scholar]
  • 148.Peltier HJ, Latham GJ. Normalization of microRNA expression levels in quantitative RT-PCR assays: identification of suitable reference RNA targets in normal and cancerous human solid tissues. RNA. 2008;14:844–52. doi: 10.1261/rna.939908. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Zhu HT, Dong QZ, Wang G, et al. Identification of suitable reference genes for qRT-PCR analysis of circulating microRNAs in hepatitis B virus-infected patients. Mol Biotechnol. 2011;50:49–56. doi: 10.1007/s12033-011-9414-6. [DOI] [PubMed] [Google Scholar]
  • 150.Filkova M, Aradi B, Senolt L, et al. Association of circulating miR-223 and miR-16 with disease activity in patients with early rheumatoid arthritis. Ann Rheum Dis. 2013 doi: 10.1136/annrheumdis-2012-202815. ; doi: 10.1136/annrheumdis-2012-202815. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Castle JC, Armour CD, Lower M, et al. Digital genome-wide ncRNA expression, including SnoRNAs, across 11 human tissues using polyA-neutral amplification. PLoS ONE. 2010;5:e11779. doi: 10.1371/journal.pone.0011779. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Davoren PA, McNeill RE, Lowery AJ, et al. Identification of suitable endogenous control genes for microRNA gene expression analysis in human breast cancer. BMC Mol Biol. 2008;9:76. doi: 10.1186/1471-2199-9-76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Kalra H, Simpson RJ, Ji H, et al. Vesiclepedia: a compendium for extracellular vesicles with continuous community annotation. PLoS Biol. 2012;10:e1001450. doi: 10.1371/journal.pbio.1001450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Liu CG, Calin GA, Meloon B, et al. An oligonucleotide microchip for genome-wide microRNA profiling in human and mouse tissues. Proc Natl Acad Sci USA. 2004;101:9740–4. doi: 10.1073/pnas.0403293101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Pritchard CC, Cheng HH, Tewari M. MicroRNA profiling: approaches and considerations. Nat Rev Genet. 2012;13:358–69. doi: 10.1038/nrg3198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Shi L, Jones WD, Jensen RV, et al. The balance of reproducibility, sensitivity, and specificity of lists of differentially expressed genes in microarray studies. BMC Bioinformatics. 2008;9:S10. doi: 10.1186/1471-2105-9-S9-S10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Sato F, Tsuchiya S, Terasawa K, et al. Intra-platform repeatability and inter-platform comparability of microRNA microarray technology. PLoS ONE. 2009;4:e5540. doi: 10.1371/journal.pone.0005540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Git A, Dvinge H, Salmon-Divon M, et al. Systematic comparison of microarray profiling, real-time PCR, and next-generation sequencing technologies for measuring differential microRNA expression. RNA. 2010;16:991–1006. doi: 10.1261/rna.1947110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Wang B, Howel P, Bruheim S, et al. Systematic evaluation of three microRNA profiling platforms: microarray, beads array, and quantitative real-time PCR array. PLoS ONE. 2011;6:e17167. doi: 10.1371/journal.pone.0017167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Wolber PK, Collins PJ, Lucas AB, et al. The Agilent in situ-synthesized microarray platform. Methods Enzymol. 2006;410:28–57. doi: 10.1016/S0076-6879(06)10002-6. [DOI] [PubMed] [Google Scholar]
  • 161.Wang H, Ach RA, Curry B. Direct and sensitive miRNA profiling from low-input total RNA. RNA. 2007;13:151–9. doi: 10.1261/rna.234507. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Liu CG, Calin GA, Volinia S, et al. MicroRNA expression profiling using microarrays. Nat Protoc. 2008;3:563–78. doi: 10.1038/nprot.2008.14. [DOI] [PubMed] [Google Scholar]
  • 163.Ach RA, Wang H, Curry B. Measuring microRNAs: comparisons of microarray and quantitative PCR measurements, and of different total RNA prep methods. BMC Biotechnol. 2008;8:69. doi: 10.1186/1472-6750-8-69. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Pradervand S, Weber J, Lemoine F, et al. Concordance among digital gene expression, microarrays, and qPCR when measuring differential expression of microRNAs. Biotechniques. 2010;48:219–22. doi: 10.2144/000113367. [DOI] [PubMed] [Google Scholar]
  • 165.Exiqon. 2011. Profiling of microRNA in Blood Serum/Plasma. Guidelines for the miRCURY LNA TM Universal RT microRNA PCR System; available at: http://www.exiqon.com/ls/Documents/Scientific/microRNA-serum-plasma-guidelines.pdf.
  • 166.Fu Y, Yi Z, Wu X, et al. Circulating microRNAs in patients with active pulmonary tuberculosis. J Clin Microbiol. 2011;49:4246–51. doi: 10.1128/JCM.05459-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Geiss GK, Bumgarner RE, Birditt B, et al. Direct multiplexed measurement of gene expression with color-coded probe pairs. Nat Biotechnol. 2008;26:317–25. doi: 10.1038/nbt1385. [DOI] [PubMed] [Google Scholar]
  • 168.Diehl P, Fricke A, Sander L, et al. Microparticles: major transport vehicles for distinct miRNAs in circulation. Cardiovasc Res. 2012;93:633–44. doi: 10.1093/cvr/cvs007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Pan M, Ge Q, Li H, et al. Sequencing the miRNAs in maternal plasma from women before and after parturition. J Nanosci Nanotechnol. 2012;12:4035–43. doi: 10.1166/jnn.2012.6196. [DOI] [PubMed] [Google Scholar]
  • 170.Lee LW, Zhang S, Etheridge A, et al. Complexity of the microRNA repertoire revealed by next-generation sequencing. RNA. 2010;16:2170–80. doi: 10.1261/rna.2225110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Wulff BE, Nishikura K. Modulation of microRNA expression and function by ADARs. Curr Top Microbiol Immunol. 2012;353:91–109. doi: 10.1007/82_2011_151. [DOI] [PubMed] [Google Scholar]
  • 172.Sana J, Faltejskova P, Svoboda M, et al. Novel classes of non-coding RNAs and cancer. J Transl Med. 2012;10:103. doi: 10.1186/1479-5876-10-103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Camerlengo T, Ozer GH, Zhang G, et al. Ohio Collaborative Conference on Bioinformatics. Los Alamitos, CA: IEEE Computer Society; 2009. Computational challenges and solutions to the analysis of micro RNA profiles in virally-infected cells derived by massively parallel sequencing; pp. 32–6. [Google Scholar]
  • 174.Kuchenbauer F, Morin RD, Argiropoulos B, et al. In-depth characterization of the microRNA transcriptome in a leukemia progression model. Genome Res. 2008;18:1787–97. doi: 10.1101/gr.077578.108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Morin RD, O'Connor MD, Griffith M, et al. Application of massively parallel sequencing to microRNA profiling and discovery in human embryonic stem cells. Genome Res. 2008;18:610–21. doi: 10.1101/gr.7179508. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Buermans HP, Ariyurek Y, Ommen van G, et al. New methods for next generation sequencing based microRNA expression profiling. BMC Genomics. 2010;11:716. doi: 10.1186/1471-2164-11-716. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Pang KC, Stephen S, Dinger ME, et al. RNAdb 2.0–an expanded database of mammalian non-coding RNAs. Nucleic Acids Res. 2007;35:D178–82. doi: 10.1093/nar/gkl926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Kawashita Y, Jinnin M, Makino T, et al. Circulating miR-29a levels in patients with scleroderma spectrum disorder. J Dermatol Sci. 2011;61:67–9. doi: 10.1016/j.jdermsci.2010.11.007. [DOI] [PubMed] [Google Scholar]

Articles from Journal of Cellular and Molecular Medicine are provided here courtesy of Blackwell Publishing

RESOURCES