Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2014 Sep 22.
Published in final edited form as: Bioessays. 2014 Jan 16;36(4):394–406. doi: 10.1002/bies.201300150

Transfer and functional consequences of dietary microRNAs in vertebrates: Concepts in search of corroboration

Negative results challenge the hypothesis that dietary xenomiRs cross the gut and regulate genes in ingesting vertebrates, but important questions persist

Kenneth W Witwer 1,*, Kendal D Hirschi 2
PMCID: PMC4109825  NIHMSID: NIHMS608060  PMID: 24436255

Abstract

If validated, diet-derived foreign microRNA absorption and function in consuming vertebrates would drastically alter our understanding of nutrition and ecology. RNA interference mechanisms of C. elegans are enhanced by uptake of environmental RNA and amplification and systemic distribution of RNA interference effectors. Therapeutic exploitation of RNAi in treating human disease is difficult because these accessory processes are absent or diminished in most animals. A recent report challenged multiple paradigms, suggesting that ingested microRNAs are transferred to blood, accumulate in tissues, and exert canonical regulation of endogenous transcripts. Independent replication of these findings has been elusive, and multiple disconfirmatory findings have been published. In the face of mounting negative results, any additional positive reports must provide the proverbial “extraordinary proof” to support such claims. In this article, we review the evidence for and against a significant role for dietary microRNAs in influencing gene expression, and make recommendations for future studies.

Keywords: cholesterol, diet, epigenetics, exosome, microRNA, nutrition, RNAi

Introduction

Eating is an engagement with the world [1] that transforms our environment into our bodies. While epidemiology studies have shown that plant-based diets help lower the risk of many diseases, synthetic supplements of vitamins and phytochemicals often do not have the same potency as complex plant materials. The lack of mechanistic understanding of the demonstrated health effects of plant-rich diets has hampered effective exploitation of this benefit. Conflicting nutritional information abounds, and popular communication of scientific studies often misinforms dietary perceptions and practices. A large branch of nutrition is based on observational studies where startling correlations can make media headlines; however, the claims often fail to be repeatable, and the published corrections appear only in the scientific literature [2]. The confused public is thus left remembering the story of women who eat additional breakfast cereal giving birth to more boys [2], for example.

A recent experimental study reported that microRNAs (miRNAs) from plants may control target genes in the consumer [3], making the same sort of media “splash” that has traditionally been reserved for observational nutritional studies. The idea that nutrition may now encompass the ingestion of genetic information is fascinating as part of the broader concept of “social RNA” presented by Sarkies and Miska in a recent Science editorial [4]. However, the theme of that article, much like the theme here, is that careful replication of these striking results is necessary. Emerging evidence now suggests that the initial claims of delivery and effect of foreign dietary genetic information in mammals may prove to be overstated. Nonetheless, the scientific community must be careful not to dismiss prematurely the concept of dietary transfer of genetic information. Entry of ingested miRNA into the body's pre-existing RNA-mediated regulatory pathways may prove to be specific to the genotype, dietary practices and health status of the consumer and to the specific amount and varieties of foods ingested.

RNA interference: Nematodes to humans

Realizing the tantalizing prospect of therapeutic RNA interference (RNAi) in humans has been a challenge in part because the confluence of three biological characteristics that facilitate RNA silencing in the nematode worm, Caenorhabditis elegans, is largely absent in other organisms, including closely related nematodes [5]. RNAi itself was discovered in C. elegans [6], and the existence of well-conserved eukaryotic silencing pathways was quickly established. Similar yet distinct systems are known in archaea and bacteria [7]. Eukaryotic RNAi is mediated by several RNA classes: miRNA, double stranded RNA (dsRNA), and, in the germ line, PIWI-interacting RNA (piRNA). At their mechanistic heart, different types of RNAi share common themes: processing of double stranded or structured, single-stranded RNA into short, single strands (usually less than 30 nucleotides in length); incorporation into protein machinery; binding to sequences in target messenger RNAs; and degradation, translational blocking, or sequestering of the target [8].

How prevalent is RNAi spread in animal species?

In C. elegans, long dsRNA injected into one part of the organism exert silencing effects in remote locations [9]. A transmembrane protein channel for dsRNA (Fig. 1A), dubbed SID-1 after a “systemic RNA interference-deficient” genomic locus identified in screens, is responsible [10-12]. Conservation of this channel indicated the possibility of systemic dissemination of dsRNA in many other organisms [11]. Some form of systemic RNAi exists in organisms ranging from the honeybee [13] and the flour beetle [14] to planarian flatworms [15]. However, long-range RNA transfer does not occur in the same way in vertebrates, which have developed mechanisms to recognize both single- and double-stranded RNA and its features, thereby triggering powerful innate immune responses [16-21]. Even in C. elegans, dsRNA distribution may not apply equally to RNA molecules of different size or type, as dsRNA of 100 nucleotides or longer are the best silencing effectors [22, 23]. This is due in part to preferential uptake of longer double-stranded molecules by SID-1 [24]. Transfer of double-stranded precursors by SID-1 may, however, result in cytoplasmic presence of single-stranded RNA species [12].

Figure 1.

Figure 1

RNAi-facilitating features of C. elegans. A: Uptake of dsRNA from the environment is allowed in part by the action of transmembrane protein SID-2, promoting endocytosis of bound RNA. B: RNAi amplification. dsRNA molecules are processed to short “primary” siRNA (left). When an siRNA molecules binds to a target, additional target-complementary siRNAs can be generated after extension by RNA-dependent RNA polymerase and processing. C: Transmembrane channel SID-1 grants access to the cytoplasm for long dsRNA, in some cases in conjunction with presentation by SID-2 (not shown). Note that cartoons are for illustration only and are not meant to represent relative size or molecular topology.

Does RNA spread occur in eukaryotes with relatively complex body plans?

A different sort of extracellular RNA (exRNA)-mediated communication may exist in some eukaryotes [25-30]. RNA molecules associate with several extracellular carriers, including lipoproteins [31] and protein complexes [32-34], but extracellular vesicles (EV, encompassing the various definitions of exosome, microvesicle, etc.) have been studied most closely. Much exRNA research has focused on miRNAs. Other RNA molecules are also found outside the cell. The similarities between enveloped viruses and EV have been examined previously [35-37], from biogenesis to cargo to information transfer capabilities.

Nevertheless, the details of how exRNA is transferred in vertebrates between extracellular carrier and recipient cell remain largely unclear. One possible mediator is a human ortholog of C. elegans SID-1, SIDT1 [38]. This transmembrane protein has been reported to facilitate uptake of cholesterol-modified siRNA [39] or contact-dependent transfer of human miR-21 between cultured cancer cells [40]. We expect that ongoing SIDT1 research will identify substrate specificities and its contribution to general RNA uptake or cell contact-independent transfer of RNA. A distinction should of course be made between transfer of exRNA into a recipient cell—with retention of function such as miRNA-mediated regulation—and stimulation of signaling from outside the cell by extracellular RNA. In addition to recognition of dsRNA molecules by TLR3, toll-like receptors can sense short single-stranded RNA molecules [41]. Activation of TLR7 and TLR8 by specific miRNAs has been reported [42, 43]. In summary, many questions remain about systemic RNAi in eukaryotes with different body plan complexity. The various proposed systems, while of considerable potential importance, are not as robust as the dsRNA-based system in C. elegans.

RNAi amplification occurs in few animals

A second C. elegans characteristic contributing to the remarkable implementation of RNAi is amplification of RNAi effectors [44]. Secondary RNA is produced when primary siRNA, processed from long dsRNA, primes RNA-dependent RNA polymerase-mediated copying of target RNA molecules. The resulting dsRNA is then processed into secondary siRNA (Fig. 2B). Successful amplification in C. elegans is generally dependent on the products of the rde-1 and rde-4 genes [45]. Although amplification activity has been reported in some insects, the underlying mechanisms are not well understood, as insects lack RNA-dependent RNA polymerases [46,47]. We are unaware of evidence that amplification of silencing RNA signals occurs in humans.

Figure 2.

Figure 2

Variability of apparent xenomiR detection in human sera: reanalysis of supplemental data from Zhang et al. Four plant miRNAs were detected in each of ten human serum pools examined in the original dietary miRNA study [3]. Pools were composed of 10-11 individuals each, with no assignment criteria specified (other than gender for the two indicated pools). There was substantial variability between pools. Shown are reads per million animal miRNAs for each pool, followed on the right of each panel by a 95% confidence interval for all miRNA-specific data.

C. elegans samples environmental RNA robustly

Efficient uptake of dsRNA molecules from environmental sources (Fig. 1C) also enhances RNAi in C. elegans. Whether soaked in a dsRNA solution [48] or fed with bacteria expressing dsRNA [49], C. elegans assimilate dsRNA from their surroundings, often rendering experimental injection unnecessary. In contrast with SID-1-mediated spreading of dsRNA within the organism, a second protein, SID-2, allows entry of ingested RNA [50,51]. Expressing the C. elegans SID-2 gene in uptake-refractory organisms, including other nematodes, permitted uptake of ingested dsRNA [50, 52]. SID-2 has RNA size requirements, and RNA uptake depends on an acidic extracellular environment such as that of the gut in many organisms [51].

Environmental sampling: Not ubiquitous, but not restricted to C. elegans

C. elegans is manifestly not the only organism that can “sample” environmental RNA, although studies in organisms without amplification mechanisms necessarily involve low copy numbers and are thus fraught with the possibility of contamination artifacts [4, 53]. The exact mechanistic bases of environmental sampling in different species are incompletely understood. Unlike SID-1, the exogenous RNA transporter SID-2 is poorly conserved. In concert with a functional SID-1, SID-2 preferentially allows import of dsRNA of 50 nt or longer [51]. There is some evidence that at least one SID-2-independent RNA uptake pathway may exist in C. elegans [54], and it is clear that other organisms can internalize ingested dsRNA by employing endocytosis that does not require SID-2 [55]. For example, scavenger receptor-mediated endocytosis may allow uptake by specific Drosophila cells [56, 57]. (Mammalian cells may also internalize dsRNA by this means, albeit for presentation to the innate immune system [58].) Uptake may also be enhanced by plant substances that modify the gut barrier [59]. The list of organisms displaying environmental RNA uptake has lengthened considerably to encompass additional nematodes [5, 60, 61], insects [62-65], even shrimp [66] and two species of sponge [67]. dsRNA can act as a species-specific pesticide in insects [64, 68-71]. A “once-removed” environmental transfer was recently reported, in which dietary dsRNA, taken up by Apis (bee), was transferred to an ectoparasite with functional consequences [72].

The challenges of oral RNA delivery to mammals

Even so, simple feeding of RNA is insufficient for uptake by some insect species, including Drosophila [64], which lacks SID-1 transport [10, 73]; this is certainly the case in mammals. It was discovered early during mammalian work that siRNA must be chemically stabilized and conjugated to ensure activity following delivery [74]. Formulation of RNAi therapeutics has since busied many investigators, and the consensus is that engineered modifications and/or “supramolecular assemblies” are essential [75] for stability and uptake, even for delivery by injection. Oral delivery remains a “holy grail.” Researchers have reported dietary delivery of RNA or other molecules, encapsulated for example in various polysaccharide shells [76, 77] or in tumor cell-derived exosomes [78] or other vesicles [79, 80]. Although uptake of orally delivered RNA may be mediated by surveilling immune cells in the gut—raising the possibility that egressing cells could transfer their exogenous cargo to distant recipient cells—RNA delivery to the gut may allow mostly local rather than systemic dissemination [81]. As a result, systemic delivery of RNAi mediators has been achieved almost exclusively by injection in mammalian studies.

Dietary RNA transfer to vertebrates: The evidence

Because of the failure—after more than a decade of intensive inquest—to find robust mammalian homologs of C. elegans RNAi accessory mechanisms or to develop facile oral delivery methods for RNAi [82], a report published online in late 2011 took the scientific community by surprise. Writing in Cell Research, a team led by Chen-Yu Zhang described a series of striking findings on “cross-kingdom” functional RNA transfer. In this study, L. Zhang, et al. used sensitive high-throughput sequencing (HTS) methods to detect high levels of plant miRNAs, or “xenomiRs” [53], in the serum and/or tissues of humans, mice, and calves with plant-based diets.

Abundance and function of xenomiRs—comparable to those of endogenous RNAs?

At least four central findings of this work seemed poised to overturn prevailing RNAi paradigms: 1) Food-derived plant miRNAs entered mammalian circulation naturally and achieved levels comparable to those of abundant endogenous extracellular miRNAs. 2) Plant miRNAs were detected not only in circulation, but also in all murine tissue types examined, and at copy numbers rivaling those of endogenous miRNAs. In liver, for example, some plant miRNAs were more abundant than let-7a. 3) In the mouse, plant miRNAs were said to downregulate at least one endogenous target in liver, the low-density lipoprotein receptor adapter protein 1 (LDLRAP1), within hours of dietary intake. Interestingly, LDLRAP1 downregulation was also reported to increase circulating LDL. 4) Perhaps most unexpectedly, uptake from the diet was independent of chemical modifications, engineered delivery vehicles, or even vesicles or other protecting carriers in the food, as chemically synthesized oligonucleotides with the same sequence as plant MIR168a were reportedly taken up with similar results.

An attractive hypothesis lacks independently corroborating data

Dietary miRNA transfer thus appeared to be an attractive candidate to explain several unsolved mysteries in nutritional science. Observational studies have suggested that specific nutrients contribute to the observed benefits of plant food consumption, but randomized controlled trials have failed to support these conclusions [83]. Could a relationship have been missed because these studies failed to consider miRNAs as a nutrient? There is also indirect support for the concept. For example, although absorption of miRNA is not well-characterized in mammals [84], bacterial factors may facilitate uptake of RNA in the gut [85]. The possibility has also been raised that RNA in a mother's breast milk may survive the infant stomach and intestinal tract to help with gene regulation while the immune and metabolic systems are immature [86-89]. The existence of RNA in vesicles within breast milk is well-established, as is the uptake of such vesicles by cells [89]. The potential involvement of breast milk miRNAs in regulatory pathways of relevance to human infants, such as metabolism of the mother's high fat diet, has been inferred from pathway analyses [86]. However, direct evidence of functional transfer of miRNA between vertebrates—even individuals of the same species—is lacking.

Response to the Cell Research publication ranged from enthusiastic optimism to incredulity. Coverage in journals and the popular press included a “you are what you eat” theme and explored implications of interspecies and interkingdom information exchange. Another leitmotif was dietary miRNA as a new type of nutrient [90], one that might explain the effects of certain folk remedies [91]. Debate about the findings also became a surrogate for genetically modified organism (GMO) controversies. There were suggestions that Zhang's results underscored concerns about unintended RNA-mediated effects of GMOs [91]. We note, however, that these latter two issues were not directly addressed in Zhang's studies or, as far as we know, in any subsequent publications. Indeed, little evidence to support the findings has emerged in the two years, whether from the Zhang lab or from independent groups.

Questions and counterevidence

Amidst intense specialist and lay interest in the Zhang, et al. findings, questions have been raised about systemic distribution of xenomiRs following dietary intake and about the likelihood of post-transcriptional regulation. These questions have stemmed from critical review of the initial publication and examinations of related studies. Importantly, numerous negative experimental studies, including several initiated expressly to confirm the findings of Zhang, et al. [3], have now been published [82, 92, 93].

Critique of published positive findings

Several features of the initial, positive publication have been critiqued [53, 94]. Only four plant miRNAs were consistently detected in humans, and with puzzling variability. Baseline and post-feeding levels of two xenomiRs in vertebrate blood and tissues reportedly approximated the abundance of moderately expressed endogenous miRNAs. If endogenous miRNAs regulate transcripts, as is widely assumed, exogenous RNAs could achieve similar effects. However, the minimal apparent changes in xenomiR levels following feeding are difficult to reconcile with rapid, detectable downregulation of mammalian targets. Finally, the levels of reported exposure in the original experiments would be difficult to replicate in humans. These three critiques are presented in greater depth in Text Box 1.

Public data: Consistent with artifact or scant uptake of dietary RNA

Of 83 public HTS datasets analyzed for evidence of plant miRNAs in animal fluids and tissues [95], 53 yielded reads from fewer than three apparent plant miRNA species. MIR156a was detected among the top three plant miRNA sequences in 18 of 83 profiling studies (22%), or 13 of 76 non-insect studies (17%), albeit with median counts of less than six per million animal miRNA reads (range: 0.4-1089). MIR168a was registered somewhat more often, at a median of 181 counts per million animal miRNA reads in 37 of 76 non-insect studies (49%). In contrast, both MIR156a and MIR168a read counts in the Zhang study were above 2800 per million in some sample pools [3]. In several reviewed datasets, a MIR168 variant was detected in animals that did not receive food containing that miRNA [95]. This observation and a general discrepancy between apparent miRNA abundance in food source and ingesting animal led the authors to conclude that MIR168 appears to be a special case: reported sequencing reads for MIR168 in vertebrate studies might be largely artifactual.

Negative studies

Independent studies have now presented data that are difficult to reconcile with general uptake of exogenous miRNAs at the levels reported by Zhang, et al., levels that might be consistent with regulatory effects.

Negative results following feeding in bees, mice, and humans

In 2013, Snow, et al. reported in RNA Biology that plant miRNAs were undetectable or indistinguishable from background signal in experiments conducted with three organisms: humans, mice, and bees [92]. In plasma of human athletes with diets that included fresh fruits, plant miRNAs could not be detected. Bees also did not appear to take up plant miRNAs.

Wild type mice with diets containing very different levels of specific miRNAs did not have above-background levels of plant miRNAs in tissues [92]. Although slightly higher signal for plant miRNA MIR156a was observed in plasma of mice with plant- versus lard-based diet, levels in the latter were at the limit of detection, and there were no significant differences. A fresh avocado diet did not result in significant increases of plant miRNA. The authors concluded that plant miRNAs—if present at all in the ingesting animal—were found at less than one copy per cell even in presumed target organs such as liver and kidney. In a particularly elegant and definitive experiment, miR-21 knockout mice that have no detectable miR-21 expression [100] were fed for four weeks with an animal lard diet containing high levels of miR-21. At the end of four weeks, no miR-21 was detected above the background level of the qPCR assay in plasma or in any of four tissues (lung, liver, kidney, and stomach) [92].

Non-human primate time-course: Negative

A smaller study, also published in RNA Biology, arrived at similar conclusions [93]. Blood was drawn from two non-human primates before and at one, four, and twelve hours following ingestion of a miRNA-rich soy and fruit mixture. At- or near-background levels of several plant miRNAs were recorded by qPCR assay, but there was no indication of response to feeding. Importantly, from the results of digital droplet PCR experiments, the authors inferred that much or all of the low-level signal they observed was due to non-specific amplification.

Negative results of a direct replication study

The RNA Biology studies did not repeat the Zhang, et al. mouse experiments or examine LDLRAP1, the putative target of miR168a, but a comprehensive attempt by Monsanto and Miragen researchers to replicate the findings was published several months later in Nature Biotechnology [82]. In this study, groups of mice received one of three dietary formulations: standard processed chow, a nutritionally sufficient diet containing 41% rice, or raw rice with a binding agent. This was an improvement on the original study design, which could not address the question of whether dietary insufficiency (raw rice) or LDLRAP1 downregulation by miRNA precipitated the reported LDL increase (see Text Box 1). Dickinson, et al. showed by qPCR that even the 41% rice diet contained more MIR168a than was administered by Zhang, et al., likely due to rice strain differences. LDL levels actually fell in this group, but rose in the all-rice group. The parsimonious explanation is that nutritional intake, not xenomiRs, increases LDL as the body mobilizes energy reserves. Importantly, levels of LDLRAP1 were unchanged across groups, as measured with a sensitive and quantitative ELISA assay. Little or no plant miRNA was found in blood or organs of mice fed with any of these diets, as measured by both HTS and qPCR [82].

Negative results in humans and mice with liver damage

In late 2012, detection of numerous exogenous RNA sequences in human circulation was reported [101]. However, these included only one plant miRNA, MIR168, the same xenomiR that is curiously and perhaps artifactually overrepresented in public high-throughput sequencing datasets [95]. Furthermore, this study found MIR168 at only single digit copies per million reads [101], orders of magnitude below the originally reported levels [3]. This detection level and the identity of the plant miRNA should rightly be viewed as disconfirmatory of Zhang, et al.

In a second study by some of the same researchers, blood samples were taken from mice that did or did not receive liver-damaging doses of acetaminophen [102]. Exogenous RNAs of various types were reportedly detected in blood, including sequences from plants to which the animals do not seem to have been exposed. Only one plant miRNA was found consistently, but, again, it was MIR168a, mapped at fractional reads per million. XenomiRs from various insect and worm species were also detected, albeit in only one or two samples and at similarly low levels. Several of these sequences have been found in organisms to which the mice probably would not have been exposed, are found at low abundance endogenously, or mapped to rare processing variants, not the canonical miRbase sequence of the identified miRNA (Witwer, unpublished analysis of data from [102]). While interesting and actually presented as confirmatory of Zhang, et al., these findings are likely the result of noise in sequencing and analysis, not low-level biological uptake of trace food contaminants. In contrast with the HTS data analyzed in BMC Genomics [95] or provided by Dickinson, et al. [82], but similar to the original work [3], the data underlying these studies do not seem to have been made available in public archives.

Paradigms and hypotheses

It does not appear that transfer of miRNAs from the diet is a reliable occurrence in vertebrates, or that transferred molecules are normally sufficient in number to exert regulatory pressure. Nevertheless, it remains possible that transfer happens under certain circumstances. There may be room for additional research in this area to address several key questions, including those outlined below. Along the way, as also highlighted here, optimized procedures are needed to balance sensitivity and specificity, and careful control experiments are important to avoid overinterpretation of positive results.

How would xenomiRs cross the intestinal barrier?

Transcytosis and other mechanisms

Given the instability of naked RNA, transcytosis—vesicular uptake of RNA carriers and their cargo (Fig. 3A) on one side and release on the other side of a biological barrier—is an attractive option (Fig. 3B). Alternatively, uptake into the cytoplasm by transmembrane RNA transporters is hypothetically possible (Fig. 3C) but would require close contact with a cell or carrier that releases the RNA in the immediate vicinity of the transporter. Additionally, immune cells could take up RNA/RNA carriers and release them on the other side of the barrier (Fig. 3D).

Figure 3.

Figure 3

Extracellular carriers and trans-barrier transport of RNA. A: Extracellular RNA is carried by and protected by incorporation into (left to right): extracellular vesicles (EV) including exosomes, microvesicles, some viruses, and other membrane-bound entities; high- and low-density lipoprotein particles; and protein complexes such as Argonaute-containing structures. The number of RNA molecules within such carriers is not well known. B: Transcytosis may deliver RNA from one side of a cell barrier to the other. Although uncomplexed RNA is shown here, uptake of RNA carriers is more likely. C: Transfer of RNA molecules via transmembrane channels. Here, a double-stranded species is transferred. D: Immune cells in biological barriers may sample nucleic acids and other molecules on one side and release them on the other, with or without movement of the immune cell to a new location.

Sequence specificity?

Degradation-resistant RNA molecules in dietary materials would be packaged into protein complexes and/or vesicles. Since transport mechanisms would recognize the carrier, not its cargo, it is difficult to envision how hypothetical sequence specificity could be achieved at the level of uptake. Specific sequences could be incorporated into protective carriers in the source material (say, by RNA binding proteins), but it is not clear that this occurs.

Permeability and mass action

Every biological barrier has some degree of leakiness. With sufficient dietary intake, some foreign RNA might survive the trip to the intestine. A small number of molecules could traverse the gut and enter the bloodstream. In the absence of sequence-specific uptake, which has not been demonstrated, the transfer of foreign RNA molecules should mirror their distribution and abundance in the source material.

As we have stated, one explanation of the reported human donor pool variability would be uptake of plant xenomiRs by only a small number of individuals, perhaps as a result of altered intestinal permeability. Inflammatory states associated with infectious and non-infectious GI tract disease, stress, and malnutrition can lead to a leaky barrier, and the influences of obesity and alcohol use have also been described. Traditional folk remedies have not been well characterized and might include alteration of gut permeability. Admittedly, the work of Wang, et al. [101] does not support this hypothesis: there was no enrichment of plant MIR168a in the blood of colitis patients. Nevertheless, intestinal permeability remains an area to explore.

Are plant-specific protocols needed for detection of plant miRNAs?

2′ O-methylation of plant miRNAs on the ribose of the 3′ nucleotide [103], a modification also associated with animal piRNAs [104], has been reported to interfere with T4-mediated ligation of adapters, as for example during HTS library preparation [105, 106]. Bias could lead to underrepresentation of plant sequences in mixed plant and animal libraries. Longer ligation times, higher enzyme concentration, selection of relatively unselective enzymes, and addition of ligation enhancers such as polyethylene glycol may promote more accurate comparison of plant and animal sequence levels in such libraries [106].

While optimization of plant- and animal-specific library preparation protocols is useful, protocol differences cannot explain the discrepancies between Zhang, et al. and other studies. Plant HTS libraries have demonstrably been prepared successfully using standard protocols, and any plant-specific ligation issues would simply reduce the apparent plant/animal ratios; they would not explain the near or complete absence of plant miRNAs in most datasets. There is no indication in the Zhang article [3] or papers cited in their methods section [107, 108] that extraordinary steps were taken to promote plant-specific ligation. Importantly, a ligation-independent stem-loop reverse transcription qPCR method was used in three recent or in-press studies reporting negative results [82, 92, 93]. Dickinson, et al. also performed a spike-in experiment (adding plant RNA to mouse RNA in a ratio informed by the levels reported by Zhang), and reported that their HTS results were not biased against plant miRNAs [82].

Artifact: The challenge of sensitive methods

Too much sensitivity, rather than too little, is often a problem. The pervasiveness of nucleic acid contamination was elegantly illustrated in a study led by Bushman and Collman [109]. Searching for constituents of the lung microbiome, the authors sequenced a variety of technical controls and replicates. Since many “lung” microbial species hailed from saliva contamination of the bronchoscope, the investigators refined their collection method by inserting a second sterile scope inside the first. Interestingly, microbial sequences were also recovered from sterile buffer washes of bronchoscopes: although “microbiologically sterile,” nucleic acid fragments could persist through manufacturing or sterilization processes. Sequence reads could even be obtained by purification and sequencing of sterile buffer itself. Only through careful consideration of the various sources of contamination could a true picture of the lung microbiome emerge [109]. The “de-discovery” of the xenotropic murine leukemia virus-related virus provides a second cautionary tale of contamination of various sorts [110-112]. More recently, contaminated spin columns were implicated in the “discovery” of an apparently novel parvovirus [113]. Considering the use of the Zhang paper to question the safety of RNAi technologies, it is also worth mentioning that results of transgenic DNA uptake studies have been highly variable and in some cases consistent with contamination. While some groups have reported detection of transgenes in ingesting organisms, others are unable to find even highly abundant natural food source DNA, including chloroplast genes.

Since plant biology laboratories on the thoroughly pollen-dusted surface of our planet contain many potential contaminants—natural and artificial—special care must be taken to rule out such contaminants in the case of positive results. Sequencing and amplifying technical controls to assess sequencing instrument, reagent, supply, and sample contamination may seem wasteful, but it might help to answer the many questions about positive results. In this regard, investigators would do well to borrow a page from the experience of forensic scientists, who have developed particularly rigorous procedures to reduce, if never entirely eliminate, the chance of contamination.

Finally, although the weight of the current evidence may convince many that plant RNA uptake by vertebrate animals is an infrequent event, a consortium approach may be advisable to answer the question definitively à la the XMRV Scientific Research Working Group [114]. Samples generated in or collected by a central laboratory could be blinded and distributed to several testing centers, including labs that have previously reported negative or positive results. Samples could be from a de novo feeding study, specimens with spiked-in controls, and even samples with reported positive or negative readings from one or more contributing labs. This approach would not be able to identify contamination definitively, but it might allow emergence of consensus on experimental protocols and results.

Do low numbers of xenomiRs preclude all forms of small RNA-mediated regulation?

Non-physiologic expression of both target and putative effector RNA in cultured cells is a common approach to establishing canonical miRNA regulation through RISC. However, miRNA copy numbers in the hundreds or more per cell may be helpful but not always necessary to achieve some measure of canonical target silencing by a targeting miRNA. Mullokandov, et al., for example, used a series of reporters with target sites for specific miRNAs, and observed significant silencing of 67 sensors [115]. 57 of the 67 targeting miRNAs were present at 100 copies per million (0.01%) or more. This evidence has been interpreted by some to suggest that “miRNAs that represent <0.1% (sic) of (a) miRNA pool are unlikely to be functionally relevant” [116]. However, in the Mullokandov, et al. study, 10 of 67 miRNAs were present at less than 0.01%. Combined with the observation that most examined miRNAs, abundant or not, did not suppress their cognate sensors [115], it is apparent that factors beyond miRNA copy number contribute to individual suppressive potency. Among others, subcellular localization [115] and total target numbers [117, 118] influence the copy number of miRNA needed to observe an effect on any particular target. In some circumstances, this could, hypothetically, require only a few copies of a miRNA per cell.

Nor is canonical RISC-mediated suppression the only conceivable or known mechanism of small RNA function. Nuclear effects of small RNA (Fig. 4A) have been reported [119], including the interference of a TAR-derived RNA with retroviral transcription [120, 121]. Since only one genomic integration of HIV-1 typically occurs per cell, chromatin-mediated suppressive mechanisms might require fewer copies of suppressive small RNAs than would be needed to silence many transcripts. Indeed, this mechanism would involve, at most, one small RNA per transcriptional event (Fig. 4A). Althaus and colleagues also reported host and viral small RNAs interactions with the HIV-1 transcript that did not necessarily involve RISC [122]. Interestingly, the reported resistance of HIV-1 to RISC-mediated miRNA regulation [116] could theoretically be due in part to effects of small RNA, which may enforce transcript structure and transport [123, 124]. Structural contributions of small RNA could of course be extended to any type of RNA molecule in the cell (Fig. 4B).

Figure 4.

Figure 4

Non-canonical (RISC-independent) actions of small RNA. A: TAR-derived RNA in HIV-1 transcriptional control. Effects on transcription have been reported for miRNA and other small RNA. Cleavage of a 5′ structured element (TAR, 1) in the nascent HIV-1 transcript may interact with the corresponding sequence in the integrated DNA (2) and lead to further degradation of the transcript, enforcing HIV latency. B: Small RNA may influence secondary/tertiary structure by binding complementary elements separated by considerable distance in the primary sequence.

However, although it is tempting to speculate that low copy number miRNA could exert influence through canonical or non-canonical means, homeopathic effects are implausible: at least a few copies of a targeting miRNA would need to be added to a cell to achieve a change in target amount. As Snow, et al. observed, dietary plant miRNAs, even if reliably quantifiable, are present at less than one copy per cell in target organs [92]. Whatever the prevalence of non-canonical miRNA-mediated regulation, a wide gulf separates the theoretical possibility of xenomiR effects at low copy numbers from the non-physiologic conditions of miRNA-target assays that are often used to establish canonical, direct regulatory effects. In one recent example, RNA species detected in blood were assessed for regulatory effects using a cell culture reporter system [101] and 10 nanomolar transfected RNA. In contrast, the overall RNA concentration in plasma was estimated to be less than 300 femtomolar. Since the authors also estimated that the RNA they studied made up less than 0.01% of the total, even cells in direct contact with blood would be exposed to 300 million-fold lower concentrations of the RNA than was used in the in vitro experiments…and without the benefit of transfection reagents. Results of such experiments, while valuable, may not apply to RNA present in vivo in only trace amounts if at all.

Intestinal permeability, intestinal deliver, and extra-dietary transfer

As we previously mentioned, intestinal permeability may be altered by certain substances, allowing measurable uptake of dietary RNA. Disease states could also affect uptake and/or clearance. In these cases, the observed effects would likely be general. Detection should closely follow RNA distribution in the dietary material, and gut bacterial products should also be detected at higher levels. Permeability and clearance aside, the results of plant vesicle administration [79,80] suggest that xenomiR presence in the intestinal epithelium might be a more appropriate subject for future exogenous miRNA studies than abundance in blood or distant organs.

Snow, et al. astutely reviewed non-dietary means of systemic xenomiR transfer [92]. Indeed, organ or stem cell transplants, blood transfusions, pregnancy, and various states of parasitism may involve direct RNA transfer from one organism to another. Inhalation of atmospheric plant material, injection of plant-based drugs, and fluid exchange (e.g. through different sexual practices) could allow RNA transfer. These mechanisms might present better opportunities for study of foreign RNA-mediated regulation than low-level or undetectable systemic transfer of RNA from the diet.

Concluding remarks

The scientific community has had two years to digest the first report that oral consumption of plant-derived miRNAs affects vertebrate metabolism. The popular and scientific press has accorded at least the requisite 15 minutes to this concept. However, as the poet Emily Dickinson wrote, “Fame is a fickle food- Upon a shifting plate.” This adage may be appropriate here, as several well-designed studies have been unable to confirm that gastrointestinal uptake is a general means of miRNA transfer and subsequent gene regulation. As the initial mixture of enthusiasm and skepticism has shifted, we must carefully consider alternative explanations for the original results. If xenomiRs as a field of inquiry are to endure, any future positive studies should rule out contamination, address rigorously the kinetics and mechanisms of dietary miRNA absorption, distribution, and clearance, and include sensitive assay systems to measure functionality in a reproducible and physiologically relevant manner for any putative diet-derived miRNA.

Table 1. Topics for further study.

Requirement Comments and proposals
Proper normalizers for exogenous RNA experiments Endogenous miRNAs that change with the Circadian day, exercise, feeding, or drugs might be inappropriate. Synthetic spike-in(s)?
Development of tissue sensors Extremely sensitive tissue sensors (as available in C. elegans [11]) could identify effects of low-level RNA transfer.
Understanding of uptake mechanisms Additional work on SIDT1, scavenger receptors, and other transmembrane proteins; further exploration of extracellular vesicles and viruses.
Identification of conditions, treatments that facilitate RNA transfer Examine samples from humans with gut permeability conditions: microbial and small RNA transfer? Are there dietary or medicinal substances with unappreciated effects on permeability?
Appreciation of human/model differences Mice and humans: humans utilize a greater proportion of calcium absorbed in the upper small intestine [125]. Might there also be differences in nucleic acid uptake? Use of humanized mouse models [126].

Acknowledgments

The authors thank Adam Gillum for figure preparation and Jian Yang for reviewing the manuscript. K.W.W. is supported in part by NIH grants R21 AI102659 and P30 MHO75673, as well as R01 NS076357 (to J.E. Clements). K.D.H. acknowledges support from USDA/ARS under the Cooperative Agreement 6250-51000-051 and the Dry Bean Health Research Program and the United Soybean Board.

Footnotes

Conflict of interest: The authors have no conflicts of interest to declare.

References

  • 1.Pollan M. The omnivore's dilemma: a natural history of four meals. New York: Penguin Press; 2006. [Google Scholar]
  • 2.Young SS, Karr A. Deming, data and observational studies: A process out of control and needing fixing. Significance. 2011;8:116–20. [Google Scholar]
  • 3.Zhang L, Hou D, Chen X, Li D, et al. Exogenous plant MIR168a specifically targets mammalian LDLRAP1: evidence of cross-kingdom regulation by microRNA. Cell Res. 2012;22:107–26. doi: 10.1038/cr.2011.158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Sarkies P, Miska EA. Molecular biology. Is there social RNA? Science. 2013;341:467–8. doi: 10.1126/science.1243175. [DOI] [PubMed] [Google Scholar]
  • 5.Felix MA. RNA interference in nematodes and the chance that favored Sydney Brenner. J Biol. 2008;7:34. doi: 10.1186/jbiol97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Fire A, Xu S, Montgomery MK, Kostas SA, et al. Potent and specific genetic interference by double-stranded RNA in Caenorhabditis elegans. Nature. 1998;391:806–11. doi: 10.1038/35888. [DOI] [PubMed] [Google Scholar]
  • 7.Wiedenheft B, Sternberg SH, Doudna JA. RNA-guided genetic silencing systems in bacteria and archaea. Nature. 2012;482:331–8. doi: 10.1038/nature10886. [DOI] [PubMed] [Google Scholar]
  • 8.Wilson RC, Doudna JA. Molecular mechanisms of RNA interference. Annu Rev Biophys. 2013;42:217–39. doi: 10.1146/annurev-biophys-083012-130404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.May RC, Plasterk RH. RNA interference spreading in C. elegans. Methods Enzymol. 2005;392:308–15. doi: 10.1016/S0076-6879(04)92018-6. [DOI] [PubMed] [Google Scholar]
  • 10.Winston WM, Molodowitch C, Hunter CP. Systemic RNAi in C. elegans requires the putative transmembrane protein SID-1. Science. 2002;295:2456–9. doi: 10.1126/science.1068836. [DOI] [PubMed] [Google Scholar]
  • 11.Hunter CP, Winston WM, Molodowitch C, Feinberg EH, et al. Systemic RNAi in Caenorhabditis elegans. Cold Spring Harb Symp Quant Biol. 2006;71:95–100. doi: 10.1101/sqb.2006.71.060. [DOI] [PubMed] [Google Scholar]
  • 12.Shih JD, Hunter CP. SID-1 is a dsRNA-selective dsRNA-gated channel. RNA. 2011;17:1057–65. doi: 10.1261/rna.2596511. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Amdam GV, Simoes ZL, Guidugli KR, Norberg K, et al. Disruption of vitellogenin gene function in adult honeybees by intra-abdominal injection of double-stranded RNA. BMC Biotechnol. 2003;3:1. doi: 10.1186/1472-6750-3-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Bucher G, Scholten J, Klingler M. Parental RNAi in Tribolium (Coleoptera) Curr Biol. 2002;12:R85–6. doi: 10.1016/s0960-9822(02)00666-8. [DOI] [PubMed] [Google Scholar]
  • 15.Newmark PA, Reddien PW, Cebria F, Sanchez Alvarado A. Ingestion of bacterially expressed double-stranded RNA inhibits gene expression in planarians. Proc Natl Acad Sci USA. 2003;100(Suppl 1):11861–5. doi: 10.1073/pnas.1834205100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Goubau D, Deddouche S, Reis ESC. Cytosolic sensing of viruses. Immunity. 2013;38:855–69. doi: 10.1016/j.immuni.2013.05.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Dixit E, Kagan JC. Intracellular pathogen detection by RIG-I-like receptors. Adv Immunol. 2013;117:99–125. doi: 10.1016/B978-0-12-410524-9.00004-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Co JG, Witwer KW, Gama L, Zink MC, et al. Induction of innate immune responses by SIV in vivo and in vitro: Differential expression and function of RIG-I and MDA5. J Infect Dis. 2011;204:1104–14. doi: 10.1093/infdis/jir469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Loo YM, Gale M., Jr Immune signaling by RIG-I-like receptors. Immunity. 2011;34:680–92. doi: 10.1016/j.immuni.2011.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Zhang SY, Herman M, Ciancanelli MJ, Perez de Diego R, et al. TLR3 immunity to infection in mice and humans. Curr Opin Immunol. 2013;25:19–33. doi: 10.1016/j.coi.2012.11.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Peisley A, Hur S. Multi-level regulation of cellular recognition of viral dsRNA. Cell Mol Life Sci. 2013;70:1949–63. doi: 10.1007/s00018-012-1149-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Parrish S, Fleenor J, Xu S, Mello C, et al. Functional anatomy of a dsRNA trigger: differential requirement for the two trigger strands in RNA interference. Mol Cell. 2000;6:1077–87. doi: 10.1016/s1097-2765(00)00106-4. [DOI] [PubMed] [Google Scholar]
  • 23.Kumar P, Pandit SS, Baldwin IT. Tobacco rattle virus vector: A rapid and transient means of silencing manduca sexta genes by plant mediated RNA interference. PLoS One. 2012;7:e31347. doi: 10.1371/journal.pone.0031347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Feinberg EH, Hunter CP. Transport of dsRNA into cells by the transmembrane protein SID-1. Science. 2003;301:1545–7. doi: 10.1126/science.1087117. [DOI] [PubMed] [Google Scholar]
  • 25.Simons M, Raposo G. Exosomes--vesicular carriers for intercellular communication. Curr Opin Cell Biol. 2009;21:575–81. doi: 10.1016/j.ceb.2009.03.007. [DOI] [PubMed] [Google Scholar]
  • 26.El Andaloussi S, Mager I, Breakefield XO, Wood MJ. Extracellular vesicles: biology and emerging therapeutic opportunities. Nat Rev Drug Discov. 2013;12:347–57. doi: 10.1038/nrd3978. [DOI] [PubMed] [Google Scholar]
  • 27.Valadi H, Ekstrom K, Bossios A, Sjostrand M, et al. Exosome-mediated transfer of mRNAs and microRNAs is a novel mechanism of genetic exchange between cells. Nat Cell Biol. 2007;9:654–9. doi: 10.1038/ncb1596. [DOI] [PubMed] [Google Scholar]
  • 28.Pegtel DM, Cosmopoulos K, Thorley-Lawson DA, van Eijndhoven MA, et al. Functional delivery of viral miRNAs via exosomes. Proc Natl Acad Sci USA. 2010;107:6328–33. doi: 10.1073/pnas.0914843107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Skog J, Wurdinger T, van Rijn S, Meijer DH, et al. Glioblastoma microvesicles transport RNA and proteins that promote tumour growth and provide diagnostic biomarkers. Nat Cell Biol. 2008;10:1470–6. doi: 10.1038/ncb1800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Witwer KW, Buzas EI, Bemis LT, Bora A, et al. Standardization of sample collection, isolation and analysis methods in extracellular vesicle research. J Extracell Vesicles. 2013;2 doi: 10.3402/jev.v2i0.20360. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Vickers KC, Palmisano BT, Shoucri BM, Shamburek RD, 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]
  • 32.Turchinovich A, Burwinkel B. Distinct AGO1 and AGO2 associated miRNA profiles in human cells and blood plasma. RNA Biol. 2012;9:1066–75. doi: 10.4161/rna.21083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Turchinovich A, Weiz L, Langheinz A, Burwinkel B. Characterization of extracellular circulating microRNA. Nucleic Acids Res. 2011;39:7223–33. doi: 10.1093/nar/gkr254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Arroyo JD, Chevillet JR, Kroh EM, Ruf IK, 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]
  • 35.Fang Y, Wu N, Gan X, Yan W, et al. Higher-order oligomerization targets plasma membrane proteins and HIV gag to exosomes. PLoS Biol. 2007;5:e158. doi: 10.1371/journal.pbio.0050158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Wurdinger T, Gatson NN, Balaj L, Kaur B, et al. Extracellular vesicles and their convergence with viral pathways. Adv Virol. 2012;2012:767694. doi: 10.1155/2012/767694. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Izquierdo-Useros N, Puertas MC, Borras FE, Blanco J, et al. Exosomes and retroviruses: the chicken or the egg? Cell Microbiol. 2011;13:10–7. doi: 10.1111/j.1462-5822.2010.01542.x. [DOI] [PubMed] [Google Scholar]
  • 38.Duxbury MS, Ashley SW, Whang EE. RNA interference: a mammalian SID-1 homologue enhances siRNA uptake and gene silencing efficacy in human cells. Biochem Biophys Res Commun. 2005;331:459–63. doi: 10.1016/j.bbrc.2005.03.199. [DOI] [PubMed] [Google Scholar]
  • 39.Wolfrum C, Shi S, Jayaprakash KN, Jayaraman M, et al. Mechanisms and optimization of in vivo delivery of lipophilic siRNAs. Nat Biotechnol. 2007;25:1149–57. doi: 10.1038/nbt1339. [DOI] [PubMed] [Google Scholar]
  • 40.Elhassan MO, Christie J, Duxbury MS. Homo sapiens systemic RNA interference-defective-1 transmembrane family member 1 (SIDT1) protein mediates contact-dependent small RNA transfer and microRNA-21-driven chemoresistance. J Biol Chem. 2012;287:5267–77. doi: 10.1074/jbc.M111.318865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Gantier MP, Tong S, Behlke MA, Xu D, et al. TLR7 is involved in sequence-specific sensing of single-stranded RNAs in human macrophages. J Immunol. 2008;180:2117–24. doi: 10.4049/jimmunol.180.4.2117. [DOI] [PubMed] [Google Scholar]
  • 42.Lehmann SM, Kruger C, Park B, Derkow K, et al. An unconventional role for miRNA: let-7 activates Toll-like receptor 7 and causes neurodegeneration. Nat Neurosci. 2012;15:827–35. doi: 10.1038/nn.3113. [DOI] [PubMed] [Google Scholar]
  • 43.Fabbri M, Paone A, Calore F, Galli R, et al. MicroRNAs bind to Toll-like receptors to induce prometastatic inflammatory response. Proc Natl Acad Sci USA. 2012;109:E2110–6. doi: 10.1073/pnas.1209414109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Tabara H, Sarkissian M, Kelly WG, Fleenor J, et al. The rde-1 gene, RNA interference, and transposon silencing in C. elegans. Cell. 1999;99:123–32. doi: 10.1016/s0092-8674(00)81644-x. [DOI] [PubMed] [Google Scholar]
  • 45.Parrish S, Fire A. Distinct roles for RDE-1 and RDE-4 during RNA interference in Caenorhabditis elegans. RNA. 2001;7:1397–402. [PMC free article] [PubMed] [Google Scholar]
  • 46.Lipardi C, Wei Q, Paterson BM. RNAi as random degradative PCR: siRNA primers convert mRNA into dsRNAs that are degraded to generate new siRNAs. Cell. 2001;107:297–307. doi: 10.1016/s0092-8674(01)00537-2. [DOI] [PubMed] [Google Scholar]
  • 47.Duan G, Saint RB, Helliwell CA, Behm CA, et al. C. elegans RNA-dependent RNA polymerases rrf-1 and ego-1 silence Drosophila transgenes by differing mechanisms. Cell Mol Life Sci. 2013;70:1469–81. doi: 10.1007/s00018-012-1218-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Tabara H, Grishok A, Mello CC. RNAi in C. elegans: soaking in the genome sequence. Science. 1998;282:430–1. doi: 10.1126/science.282.5388.430. [DOI] [PubMed] [Google Scholar]
  • 49.Timmons L, Fire A. Specific interference by ingested dsRNA. Nature. 1998;395:854. doi: 10.1038/27579. [DOI] [PubMed] [Google Scholar]
  • 50.Winston WM, Sutherlin M, Wright AJ, Feinberg EH, et al. Caenorhabditis elegans SID-2 is required for environmental RNA interference. Proc Natl Acad Sci USA. 2007;104:10565–70. doi: 10.1073/pnas.0611282104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.McEwan DL, Weisman AS, Hunter CP. Uptake of extracellular double-stranded RNA by SID-2. Mol Cell. 2012;47:746–54. doi: 10.1016/j.molcel.2012.07.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Nuez I, Felix MA. Evolution of susceptibility to ingested double-stranded RNAs in Caenorhabditis nematodes. PLoS One. 2012;7:e29811. doi: 10.1371/journal.pone.0029811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Witwer KW. XenomiRs and miRNA homeostasis in health and disease: Evidence that diet and dietary miRNAs directly and indirectly influence circulating miRNA profiles. RNA Biol. 2012;9:1147–54. doi: 10.4161/rna.21619. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Liu H, Wang X, Wang HD, Wu J, et al. Escherichia coli noncoding RNAs can affect gene expression and physiology of Caenorhabditis elegans. Nat Commun. 2012;3:1073. doi: 10.1038/ncomms2071. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Price DR, Gatehouse JA. RNAi-mediated crop protection against insects. Trends Biotechnol. 2008;26:393–400. doi: 10.1016/j.tibtech.2008.04.004. [DOI] [PubMed] [Google Scholar]
  • 56.Saleh MC, van Rij RP, Hekele A, Gillis A, et al. The endocytic pathway mediates cell entry of dsRNA to induce RNAi silencing. Nat Cell Biol. 2006;8:793–802. doi: 10.1038/ncbl439. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Ulvila J, Parikka M, Kleino A, Sormunen R, et al. Double-stranded RNA is internalized by scavenger receptor-mediated endocytosis in Drosophila S2 cells. J Biol Chem. 2006;281:14370–5. doi: 10.1074/jbc.M513868200. [DOI] [PubMed] [Google Scholar]
  • 58.Limmon GV, Arredouani M, McCann KL, Corn Minor RA, et al. Scavenger receptor class-A is a novel cell surface receptor for double-stranded RNA. Faseb J. 2008;22:159–67. doi: 10.1096/fj.07-8348com. [DOI] [PubMed] [Google Scholar]
  • 59.Mao YB, Xue XY, Tao XY, Yang CQ, et al. Cysteine protease enhances plant-mediated bollworm RNA interference. Plant Mol Biol. 2013;83:119–29. doi: 10.1007/s11103-013-0030-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Shannon AJ, Tyson T, Dix I, Boyd J, et al. Systemic RNAi mediated gene silencing in the anhydrobiotic nematode Panagrolaimus superbus. BMC Mol Biol. 2008;9:58. doi: 10.1186/1471-2199-9-58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Bakhetia M, Charlton W, Atkinson HJ, McPherson MJ. RNA interference of dual oxidase in the plant nematode Meloidogyne incognita. Mol Plant Microbe Interact. 2005;18:1099–106. doi: 10.1094/MPMI-18-1099. [DOI] [PubMed] [Google Scholar]
  • 62.Zhu F, Xu J, Palli R, Ferguson J, et al. Ingested RNA interference for managing the populations of the Colorado potato beetle, Leptinotarsa decemlineata. Pest Manag Sci. 2011;67:175–82. doi: 10.1002/ps.2048. [DOI] [PubMed] [Google Scholar]
  • 63.Li J, Chen Q, Lin Y, Jiang T, et al. RNA interference in Nilaparvata lugens (Homoptera: Delphacidae) based on dsRNA ingestion. Pest Manag Sci. 2011;67:852–9. doi: 10.1002/ps.2124. [DOI] [PubMed] [Google Scholar]
  • 64.Whyard S, Singh AD, Wong S. Ingested double-stranded RNAs can act as species-specific insecticides. Insect Biochem Mol Biol. 2009;39:824–32. doi: 10.1016/j.ibmb.2009.09.007. [DOI] [PubMed] [Google Scholar]
  • 65.Sattar S, Addo-Quaye C, Song Y, Anstead JA, et al. Expression of small RNA in Aphis gossypii and its potential role in the resistance interaction with melon. PLoS One. 2012;7:e48579. doi: 10.1371/journal.pone.0048579. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Attasart P, Namramoon O, Kongphom U, Chimwai C, et al. Ingestion of bacteria expressing dsRNA triggers specific RNA silencing in shrimp. Virus Res. 2013;171:252–6. doi: 10.1016/j.virusres.2012.11.012. [DOI] [PubMed] [Google Scholar]
  • 67.Rivera AS, Hammel JU, Haen KM, Danka ES, et al. RNA interference in marine and freshwater sponges: actin knockdown in Tethya wilhelma and Ephydatia muelleri by ingested dsRNA expressing bacteria. BMC Biotechnol. 2011;11:67. doi: 10.1186/1472-6750-11-67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Zhou X, Wheeler MM, Oi FM, Scharf ME. RNA interference in the termite Reticulitermes flavipes through ingestion of double-stranded RNA. Insect Biochem Mol Biol. 2008;38:805–15. doi: 10.1016/j.ibmb.2008.05.005. [DOI] [PubMed] [Google Scholar]
  • 69.Terenius O, Papanicolaou A, Garbutt JS, Eleftherianos I, et al. RNA interference in Lepidoptera: an overview of successful and unsuccessful studies and implications for experimental design. J Insect Physiol. 2011;57:231–45. doi: 10.1016/j.jinsphys.2010.11.006. [DOI] [PubMed] [Google Scholar]
  • 70.Bachman PM, Bolognesi R, Moar WJ, Mueller GM, et al. Characterization of the spectrum of insecticidal activity of a double-stranded RNA with targeted activity against Western Corn Rootworm (Diabrotica virgifera virgifera LeConte) Transgenic Res. 2013;22:1207–22. doi: 10.1007/s11248-013-9716-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Zhang H, Li HC, Miao XX. Feasibility, limitation and possible solutions of RNAi-based technology for insect pest control. Insect Sci. 2013;20:15–30. doi: 10.1111/j.1744-7917.2012.01513.x. [DOI] [PubMed] [Google Scholar]
  • 72.Garbian Y, Maori E, Kalev H, Shafir S, et al. Bidirectional transfer of RNAi between honey bee and Varroa destructor: Varroa gene silencing reduces Varroa population. PLoS Pathog. 2012;8:e1003035. doi: 10.1371/journal.ppat.1003035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Roignant JY, Carre C, Mugat B, Szymczak D, et al. Absence of transitive and systemic pathways allows cell-specific and isoform-specific RNAi in Drosophila. RNA. 2003;9:299–308. doi: 10.1261/rna.2154103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Soutschek J, Akinc A, Bramlage B, Charisse K, et al. Therapeutic silencing of an endogenous gene by systemic administration of modified siRNAs. Nature. 2004;432:173–8. doi: 10.1038/nature03121. [DOI] [PubMed] [Google Scholar]
  • 75.Aliabadi HM, Landry B, Sun C, Tang T, et al. Supramolecular assemblies in functional siRNA delivery: where do we stand? Biomaterials. 2012;33:2546–69. doi: 10.1016/j.biomaterials.2011.11.079. [DOI] [PubMed] [Google Scholar]
  • 76.Aouadi M, Tesz GJ, Nicoloro SM, Wang M, et al. Orally delivered siRNA targeting macrophage Map4k4 suppresses systemic inflammation. Nature. 2009;458:1180–4. doi: 10.1038/nature07774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Ballarin-Gonzalez B, Dagnaes-Hansen F, Fenton RA, Gao S, et al. Protection and systemic translocation of siRNA following oral administration of chitosan/siRNA nanoparticles. Mol Ther Nucleic Acids. 2013;2:e76. doi: 10.1038/mtna.2013.2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Sun D, Zhuang X, Xiang X, Liu Y, et al. A novel nanoparticle drug delivery system: the anti-inflammatory activity of curcumin is enhanced when encapsulated in exosomes. Mol Ther. 2010;18:1606–14. doi: 10.1038/mt.2010.105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Wang B, Zhuang X, Deng ZB, Jiang H, et al. Targeted drug delivery to intestinal macrophages by bioactive nanovesicles released from grapefruit. Mol Ther. 2013 doi: 10.1038/mt.2013.190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Wang Q, Zhuang X, Mu J, Deng ZB, et al. Delivery of therapeutic agents by nanoparticles made of grapefruit-derived lipids. Nat Commun. 2013;4:1867. doi: 10.1038/ncomms2886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Kriegel C, Attarwala H, Amiji M. Multi-compartmental oral delivery systems for nucleic acid therapy in the gastrointestinal tract. Adv Drug Deliv Rev. 2013;65:891–901. doi: 10.1016/j.addr.2012.11.003. [DOI] [PubMed] [Google Scholar]
  • 82.Dickinson B, Zhang Y, Petrick JS, Heck G, et al. Lack of detectable oral bioavailability of plant microRNAs after feeding in mice. Nat Biotechnol. 2013;31:965–7. doi: 10.1038/nbt.2737. [DOI] [PubMed] [Google Scholar]
  • 83.Young SS, Karr A. Deming, Data and observational studies. Significance. 2011;8 [Google Scholar]
  • 84.Lecellier CH, Dunoyer P, Arar K, Lehmann-Che J, et al. A cellular microRNA mediates antiviral defense in human cells. Science. 2005;308:557–60. doi: 10.1126/science.1108784. [DOI] [PubMed] [Google Scholar]
  • 85.Xiang S, Fruehauf J, Li CJ. Short hairpin RNA-expressing bacteria elicit RNA interference in mammals. Nat Biotechnol. 2006;24:697–702. doi: 10.1038/nbt1211. [DOI] [PubMed] [Google Scholar]
  • 86.Munch EM, Harris RA, Mohammad M, Benham AL, et al. Transcriptome profiling of microRNA by Next-Gen deep sequencing reveals known and novel miRNA species in the lipid fraction of human breast milk. PLoS One. 2013;8:e50564. doi: 10.1371/journal.pone.0050564. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Kosaka N, Izumi H, Sekine K, Ochiya T. microRNA as a new immune-regulatory agent in breast milk. Silence. 2010;1:7. doi: 10.1186/1758-907X-1-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Lasser C, Alikhani VS, Ekstrom K, Eldh M, et al. Human saliva, plasma and breast milk exosomes contain RNA: uptake by macrophages. J Transl Med. 2011;9:9. doi: 10.1186/1479-5876-9-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Admyre C, Johansson SM, Qazi KR, Filen JJ, et al. Exosomes with immune modulatory features are present in human breast milk. J Immunol. 2007;179:1969–78. doi: 10.4049/jimmunol.179.3.1969. [DOI] [PubMed] [Google Scholar]
  • 90.Hirschi KD. New foods for thought. Trends Plant Sci. 2012;17:123–5. doi: 10.1016/j.tplants.2012.01.004. [DOI] [PubMed] [Google Scholar]
  • 91.Jiang M, Sang X, Hong Z. Beyond nutrients: Food-derived microRNAs provide cross-kingdom regulation. BioEssays. 2012;34:280–4. doi: 10.1002/bies.201100181. [DOI] [PubMed] [Google Scholar]
  • 92.Snow JW, Hale AE, Isaacs SK, Baggish AL, et al. Ineffective delivery of diet-derived microRNAs to recipient animal organisms. RNA Biol. 2013;10:1107–16. doi: 10.4161/rna.24909. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Witwer KW, McAlexander MA, Queen SE, Adams RJ. Real-time quantitative PCR and droplet digital PCR for plant miRNAs in mammalian blood provide little evidence for general uptake of dietary miRNAs: Limited evidence for general uptake of dietary plant xenomiRs. RNA Biol. 2013;10:1080–6. doi: 10.4161/rna.25246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Petrick JS, Brower-Toland B, Jackson AL, Kier LD. Safety assessment of food and feed from biotechnology-derived crops employing RNA-mediated gene regulation to achieve desired traits: A scientific review. Regul Toxicol Pharmacol. 2013;66:167–76. doi: 10.1016/j.yrtph.2013.03.008. [DOI] [PubMed] [Google Scholar]
  • 95.Zhang Y, Wiggins BE, Lawrence C, Petrick J, et al. Analysis of plant-derived miRNAs in animal small RNA datasets. BMC Genomics. 2012;13:381. doi: 10.1186/1471-2164-13-381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.McAlexander MA, Phillips MJ, Witwer KW. Comparison of methods for miRNA extraction from plasma and quantitative recovery of RNA from cerebrospinal fluid. Front Genet. 2013;4:83. doi: 10.3389/fgene.2013.00083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Hansen KD, Wu Z, Irizarry RA, Leek JT. Sequencing technology does not eliminate biological variability. Nat Biotechnol. 2011;29:572–3. doi: 10.1038/nbt.1910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Wei LQ, Yan LF, Wang T. Deep sequencing on genome-wide scale reveals the unique composition and expression patterns of microRNAs in developing pollen of Oryza sativa. Genome Biol. 2011;12:R53. doi: 10.1186/gb-2011-12-6-r53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Grant-Downton R, Le Trionnaire G, Schmid R, Rodriguez-Enriquez J, et al. MicroRNA and tasiRNA diversity in mature pollen of Arabidopsis thaliana. BMC Genomics. 2009;10:643. doi: 10.1186/1471-2164-10-643. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Parikh VN, Jin RC, Rabello S, Gulbahce N, et al. MicroRNA-21 integrates pathogenic signaling to control pulmonary hypertension: results of a network bioinformatics approach. Circulation. 2012;125:1520–32. doi: 10.1161/CIRCULATIONAHA.111.060269. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Wang K, Li H, Yuan Y, Etheridge A, et al. The complex exogenous RNA spectra in human plasma: an interface with human gut biota? PLoS One. 2012;7:e51009. doi: 10.1371/journal.pone.0051009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Wang K, Yuan Y, Li H, Cho JH, et al. The spectrum of circulating RNA: a window into systems toxicology. Toxicol Sci. 2013;132:478–92. doi: 10.1093/toxsci/kft014. [DOI] [PubMed] [Google Scholar]
  • 103.Yu B, Yang Z, Li J, Minakhina S, et al. Methylation as a crucial step in plant microRNA biogenesis. Science. 2005;307:932–5. doi: 10.1126/science.1107130. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Farazi TA, Juranek SA, Tuschl T. The growing catalog of small RNAs and their association with distinct Argonaute/Piwi family members. Development. 2008;135:1201–14. doi: 10.1242/dev.005629. [DOI] [PubMed] [Google Scholar]
  • 105.Ebhardt HA, Thi EP, Wang MB, Unrau PJ. Extensive 3′ modification of plant small RNAs is modulated by helper component-proteinase expression. Proc Natl Acad Sci USA. 2005;102:13398–403. doi: 10.1073/pnas.0506597102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Munafo DB, Robb GB. Optimization of enzymatic reaction conditions for generating representative pools of cDNA from small RNA. RNA. 2010;16:2537–52. doi: 10.1261/rna.2242610. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Chen X, Li Q, Wang J, Guo X, et al. Identification and characterization of novel amphioxus microRNAs by Solexa sequencing. Genome Biol. 2009;10:R78. doi: 10.1186/gb-2009-10-7-r78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Chen X, Ba Y, Ma L, Cai X, 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]
  • 109.Charlson ES, Bittinger K, Haas AR, Fitzgerald AS, et al. Topographical continuity of bacterial populations in the healthy human respiratory tract. Am J Respir Crit Care Med. 2011;184:957–63. doi: 10.1164/rccm.201104-0655OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Sato E, Furuta RA, Miyazawa T. An endogenous murine leukemia viral genome contaminant in a commercial RT-PCR kit is amplified using standard primers for XMRV. Retrovirology. 2010;7:110. doi: 10.1186/1742-4690-7-110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Erlwein O, Robinson MJ, Dustan S, Weber J, et al. DNA extraction columns contaminated with murine sequences. PLoS One. 2011;6:e23484. doi: 10.1371/journal.pone.0023484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Delviks-Frankenberry K, Cingoz O, Coffin JM, Pathak VK. Recombinant origin, contamination, and de-discovery of XMRV. Curr Opin Virol. 2012;2:499–507. doi: 10.1016/j.coviro.2012.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Naccache SN, Greninger AL, Lee D, Coffey LL, et al. The perils of pathogen discovery: Origin of a novel parvovirus-like hybrid genome traced to nucleic acid extraction spin columns. J Virol. 2013;87:11966–77. doi: 10.1128/JVI.02323-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Simmons G, Glynn SA, Komaroff AL, Mikovits JA, et al. Failure to confirm XMRV/MLVs in the blood of patients with chronic fatigue syndrome: a multi-laboratory study. Science. 2011;334:814–7. doi: 10.1126/science.1213841. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Mullokandov G, Baccarini A, Ruzo A, Jayaprakash AD, et al. High-throughput assessment of microRNA activity and function using microRNA sensor and decoy libraries. Nat Methods. 2012;9:840–6. doi: 10.1038/nmeth.2078. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Whisnant AW, Bogerd HP, Flores O, Ho P, et al. In-depth analysis of the interaction of HIV-1 with cellular microRNA biogenesis and effector mechanisms. MBio. 2013;4:e000193. doi: 10.1128/mBio.00193-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Seitz H. Redefining microRNA targets. Curr Biol. 2009;19:870–3. doi: 10.1016/j.cub.2009.03.059. [DOI] [PubMed] [Google Scholar]
  • 118.Salmena L, Poliseno L, Tay Y, Kats L, et al. A ceRNA hypothesis: the Rosetta Stone of a hidden RNA language? Cell. 2011;146:353–8. doi: 10.1016/j.cell.2011.07.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Younger ST, Corey DR. The puzzle of RNAs that target gene promoters. Chembiochem. 2009;10:1135–9. doi: 10.1002/cbic.200900015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Wagschal A, Rousset E, Basavarajaiah P, Contreras X, et al. Microprocessor, Setx, Xrn2, and Rrp6 co-operate to induce premature termination of transcription by RNAPII. Cell. 2012;150:1147–57. doi: 10.1016/j.cell.2012.08.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Klase Z, Kale P, Winograd R, Gupta MV, et al. HIV-1 TAR element is processed by Dicer to yield a viral micro-RNA involved in chromatin remodeling of the viral LTR. BMC Mol Biol. 2007;8:63. doi: 10.1186/1471-2199-8-63. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Althaus CF, Vongrad V, Niederost B, Joos B, et al. Tailored enrichment strategy detects low abundant small noncoding RNAs in HIV-1 infected cells. Retrovirology. 2012;9:27. doi: 10.1186/1742-4690-9-27. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Ito K, Go S, Komiyama M, Xu Y. Inhibition of translation by small RNA-stabilized mRNA structures in human cells. J Am Chem Soc. 2011;133:19153–9. doi: 10.1021/ja206353c. [DOI] [PubMed] [Google Scholar]
  • 124.Noguchi K, Ishibashi K, Miyokawa K, Hokari M, et al. HIV-1 suppressive sequences are modulated by Rev transport of unspliced RNA and are required for efficient HIV-1 production. PLoS One. 2012;7:e51393. doi: 10.1371/journal.pone.0051393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Coates ME. Gnotobiotic animals in nutrition research. Proc Nutr Soc. 1973;32:53–8. doi: 10.1079/pns19730015. [DOI] [PubMed] [Google Scholar]
  • 126.Bissig KD, Le TT, Woods NB, Verma IM. Repopulation of adult and neonatal mice with human hepatocytes: a chimeric animal model. Proc Natl Acad Sci USA. 2007;104:20507–11. doi: 10.1073/pnas.0710528105. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES