Skip to main content
  • Research article
  • Open access
  • Published:

Assessing diversity of the female urine microbiota by high throughput sequencing of 16S rDNA amplicons

Abstract

Background

Urine within the urinary tract is commonly regarded as "sterile" in cultivation terms. Here, we present a comprehensive in-depth study of bacterial 16S rDNA sequences associated with urine from healthy females by means of culture-independent high-throughput sequencing techniques.

Results

Sequencing of the V1V2 and V6 regions of the 16S ribosomal RNA gene using the 454 GS FLX system was performed to characterize the possible bacterial composition in 8 culture-negative (<100,000 CFU/ml) healthy female urine specimens. Sequences were compared to 16S rRNA databases and showed significant diversity, with the predominant genera detected being Lactobacillus, Prevotella and Gardnerella. The bacterial profiles in the female urine samples studied were complex; considerable variation between individuals was observed and a common microbial signature was not evident. Notably, a significant amount of sequences belonging to bacteria with a known pathogenic potential was observed. The number of operational taxonomic units (OTUs) for individual samples varied substantially and was in the range of 20 - 500.

Conclusions

Normal female urine displays a noticeable and variable bacterial 16S rDNA sequence richness, which includes fastidious and anaerobic bacteria previously shown to be associated with female urogenital pathology.

Background

Microbes, including bacteria, viruses and protists, reside both on the surface and deep within numerous sites in the human body. It is estimated that trillions of microorganisms inhabit the average healthy human and that microbial cell counts in and on the human body outnumber the human cells by a factor of 10 [1, 2]. Studies confirm that humans live in a symbiosis with most of these microbes, whose roles span from harmless to important to life and health [1, 3, 4]. However, microorganisms can also be detrimental to their host and cause diseases such as digestive disorders, obesity, skin diseases, oral disease, bacterial vaginosis (BV), sexual transmitted diseases and urinary tract infections (UTI) [2, 5–9].

Urine within the urinary tract has in general been considered sterile [10, 11], based upon a lack of culturable microbial cells present in urine specimens obtained by the clean-catch method and by catheterization [12–15]. Confirmation of a UTI relies on demonstrating significant bacteriuria (or funguria) in a voided midstream urine sample. Traditionally, 105 colony-forming units per ml (CFU/ml) is the threshold for defining a positive (significant) culture result [16, 17]. Conventional culturing techniques favor the fast growing and modest bacteria, whereas fastidious bacteria can evade the standard culture conditions [18]. The presence of intracellular bacteria in uroepithelial cells [19], and even biofilm formation in the urinary tract has been suggested [20, 21]. Investigation of healthy urine specimens has demonstrated the presence of non-culturable bacterial cells [22]. These findings stress that bacteria present in urine specimens can escape detection by culture-dependent methods, and that the current view of bacterial diversity in urine thus may be incomplete. This leaves a cryptic fraction of bacteria that may be explored by other means.

Culture-independent, 16S ribosomal DNA (rDNA) sequencing has been widely utilized in the past two decades to study bacterial diversity from various habitats since sequencing of PCR-amplified 16S rDNA overcomes the limitations of culture-based bacterial detection [23]. However, often the search for microbial agents is performed only after a disease state has been diagnosed. Only a few investigations including urine from healthy persons using 16S rDNA PCR have been reported [12, 24–26]. These studies had a variable success rate in actually obtaining sequences, resulting in a limited overview of the healthy urine bacterial flora. However, two recent 16S rDNA studies by Nelson et al. (2010) and Dong et al. (2011) [27, 28] have shown that the male urine contains multiple bacterial genera.

Advances in sequencing technology, such as massively parallel pyrosequencing as developed by 454 Life Sciences [29], allow for extensive characterization of microbial populations in a high throughput and cost effective manner [30, 31]. Amplicons of partial 16S rRNA genes are sequenced on microscopic beads placed separately in picoliter-sized wells, bypassing previously needed cloning and cultivation procedures. Such sequencing has revealed an unexpectedly high diversity within various human-associated microbial communities, e.g. oral-, vaginal-, intestinal- and male first catch urine microbiota [4, 28, 32, 33], but female urine microbial diversity has so far not been studied using high throughput sequencing (HTS) methods.

Here, we have investigated the bacterial diversity in urine microbiota from healthy females by means of 16S rDNA amplicon 454 pyrosequencing. This study demonstrates the use of this methodology for investigating bacterial sequence diversity in female urine samples. Our results indicate a diverse spectrum of bacterial profiles associated with healthy, culture negative female urine and provide a resource for further studies in the field of molecular diagnostics of urine specimens.

Methods

Urine sampling

Urine was collected by the clean catch method in which healthy adult female volunteers (n = 8), collected midstream urine into a sterile container. Specimens were initially kept at 4°C, and within an hour transported to the laboratory for DNA isolation. All specimens were culture negative, as tested by the Urological Clinic at the University Hospital HF Aker-Oslo. Samples were taken with informed consent and the study was approved by the Regional Committee for Medical Research Ethics East-Norway (REK Øst Prosjekt 110-08141c 1.2008.367).

DNA isolation

30 ml urine volume was pelleted by centrifugation at 14000 RCF for 10 min at 4°C. 25 ml of the supernatant was decanted and the pellet was resuspended in the remaining volume. 5 ml of the sample was again pelleted by centrifugation for 10 min at 16000 × g (4°C). The pellet and some supernatant (up to 100 μl) were processed further. DNA was isolated from the urine pellets with DNeasy Blood & Tissue kit (QIAGEN, Germany), following the tissue spin-column protocol with minor modifications. Briefly, cell lysis was initiated by adding 100 μl POWERlyse lysis buffer (NorDiag ASA, Oslo, Norway) followed by incubation at 80°C for 10 min. Finally, 200 μl of Qiagen buffer AL was added. Samples were mixed by pulse-vortexing for 15 sec. From this point onward, purification was carried out as per manufacturer's instructions. Finally, the DNA was eluted in 100 μl of AE buffer from the kit. The DNA concentrations in the samples were measured by using the Quant-iT PicoGreen dsDNA assay kit (Molecular Probes, Invitrogen USA) and ranged from 0.33 ng/μl to 1.59 ng/μl.

16S rDNA PCR

DNA (10 μl of 1:9 dilution) was amplified by PCR using the broad range 16S rDNA primers described in Table 1. The composite primers each comprised a 17-20 bases target specific region at their 3' end and a 19 bases region of the Primer A (forward primer) or the Primer B (reverse primer) sequences needed for GS FLX amplicon sequencing (454 Life Sciences, USA) at their 5'end. PCR reactions were performed using 25 μl (final volume) mixtures containing 1× GeneAmp PCR Gold Buffer Applied Biosystems, 3.5 mM MgCl2, 0.2 mM GeneAmp dNTP, 10 pmol of each primer and 0.025 U/μl AmpliTaq Gold DNA Polymerase, LD (Applied Biosystems, USA). The amplification protocol for the V1V2 amplicon primers was: 95°C for 10 min, followed by 35 cycles of 95°C for 30 s, 50°C for 30 s and 72°C for 25 s, and a final elongation step at 72°C for 7 min. The protocol for the V6 amplicon primers was: 95°C for 10 min, followed by 35 cycles of 95°C for 30 s, 50°C for 25 s and 72°C for 25 s, and a final elongation step at 72°C for 7 min. Replicate PCRs were performed for each sample. A positive control (with previously amplified bacterial DNA) as template was run for every PCR.

Table 1 PCR primers used

PCR amplicons were detected and confirmed for DNA from all eight subjects by agarose gel electrophoresis prior to pyrosequencing (data not shown).

All crucial steps during DNA isolation and the entire PCR set up were performed in a laminar air flow (LAF)-bench, illuminated with a UV lamp prior to use in order to avoid possible contaminants. In addition, negative DNA extraction controls (lysis buffer and kit reagents only) were amplified and sequenced as contamination controls.

Additionally, negative PCR controls (sterile Molecular Biology Grade Water from 5PRIME (VWR, Norway) as template) were run for every PCR protocol, resulting in no PCR product.

454 pyrosequencing

Replicate PCR products were pooled and purified using Agencourt AMPure PCR purification (Beckman Coulter, USA). DNA concentration and quality were assessed on a Bioanalyzer 2100 (Agilent, USA). Equal amounts of both amplicons (V1V2 and V6) for a single subject or contamination control were pooled and sequenced using GS FLX chemistry in the same lane of a PicoTiterPlate divided into 16 lanes. Each of the amplicons was pyrosequenced together, except for samples F1 and F3.

454 pyrosequencing was performed by the Norwegian Sequencing Centre (NSC) at the Department of Biology, University of Oslo, Norway.

Sequence read analysis

A total of 190 287 reads were produced (female urine 165 041 raw reads and contamination control 25 246 raw reads). The initial sequence reads were split into two pools using the V1V2 and V6 primer sequences via the sfffile program from 454 Life Sciences, thus reducing the sequences to 152 413 urine reads (Table 2) due to the program splitting on exact match to primer.

Table 2 Sampling depth and biodiversity found by amplicon 454 pyrosequencing V1V2 and V6 regions from eight culture negative female urine samples

The 454 pyrosequencing method has a characteristic error rate in the form of insertion/deletion errors at homopolymer runs. To correct for this phenomenon, the raw reads were processed with PyroNoise [34] with a minimum length cutoff of 218 and 235 nt for the V1V2 and V6 regions, respectively. The PyroNoise program clusters all reads whose flowgrams indicate that they could stem from the same sequence, while also considering read abundance. After denoising, one sequence per cluster together with the number of reads mapping to that cluster is reported. Next, the sequences (at this stage one sequence per denoised cluster) that did not have an exact match to the primer were removed, and the forward primer sequence itself was also trimmed. Finally, the urine sample sequence sets were stripped for sequences that could be from the same source as those in the contamination control dataset. This was done by using the program ESPRIT http://www.biotech.ufl.edu/people/sun/esprit.html[35] to do a complete linkage clustering at 1% genetic difference of each sample together with its respective control. Before clustering, the control sequences were weighed so that there were the same number of reads stemming from both the sample and the control going into the process. Within each cluster the frequency of sample vs control sequence was calculated, and any sample sequences found in clusters where 50% or more of the sequences belonged to the control were removed.

For taxonomic grouping we used MEGAN V3.4 http://www-ab.informatik.uni-tuebingen.de/software/megan/welcome.html[36, 37], which uses blast hits to place reads onto a taxonomy by assigning each read to a taxonomic group at a level in the NCBI taxonomy. The sequence reads (one read per denoised cluster from the pyronoise step) that passed the filtering steps were compared to a curated version of the SSUrdp database [38] using blastn with parameters set to a maximum expectation value (E) of 10-5. The 25 best hits were kept. To reflect abundance behind each denoised sequence cluster, prior to taxonomic classification each entry in the blast output file was replicated as many times as there were reads mapping to its query sequence. MEGAN analysis of these blast records was performed using a minimum alignment bit score threshold of 100, and the minimum support filter was set to a threshold of 5 (the minimum number of sequences that must be assigned to a taxon for it to be reported). These parameters were consistently used throughout this analysis. When comparing the individual datasets using MEGAN, the number of reads were normalized to 100 000 for each dataset using the compare tool in MEGAN.

Sequences generated in this study have been submitted to the Sequence Read Archive with the study accession number ERP000957. It can be accessed directly through http://www.ebi.ac.uk/ena/data/view/ERP000957.

Clustering of reads into OTUs

Numbers of operational taxonomic units (OTUs), rarefaction curves, Chao1 richness estimations and Shannon diversities were calculated using MOTHUR v1.17.0 [39], both on each separate sample and on pooled V1V2 and V6 sequences, after replicating each sequence to reflect the amount of reads mapping to its denoised cluster. Each sequence set was first reduced to unique sequences, before a single linkage preclustering step as described by Huse et al., 2010 [40] was performed. In this step, shorter and less abundant sequences were merged with longer and more abundant sequences with a maximum of two differing nucleotides. OTUs were calculated using average clustering at 3%, using a pairwise distance matrix. Distances were calculated using Needleman-Wunsch, discounting endgaps while counting internal gaps separately.

Considering that the Shannon index is sensitive to the original number of sequences generated from a given sample [41] we calculated the Shannon index for normalized numbers of sequences for each separate sample. A random number of reads, corresponding to the lowest number of sequences in a sample group, i.e. 2720 for V1V2 and 2988 for V6, were picked 100 times from each sequence set. These new sequence sets were processed through MOTHUR in the same fashion as the full sequence sets and the average of the resulting Shannon values are shown in Table 2.

Results

454 pyrosequencing data

In our study a total of 78 346 sequences for the V1V2 region and 74 067 sequences for the V6 region were obtained (Table 2). The quality filtering approach as described in Methods eliminated 40% of the sequenced reads. Additionally, since the bacterial identification technique (broad range 16S rDNA PCR) utilized in this study was highly sensitive and susceptible to environmental contamination, we included negative control extractions, followed by PCR and sequencing, to determine the contamination resulting from the chemicals and consumables used. The read datasets were stripped for sequences found to cluster predominantly with contamination control sequences. This resulted in removal of an additional 1% of the reads, showing that background contamination levels were low (Table 2).

Identity of the bacterial DNA found in female human urine

An analysis using MEGAN of all pooled reads from the two different amplicon libraries of the 16S rRNA gene (i.e. V1V2 and V6 regions) revealed a total of eleven phyla in female urine, with the bacterial DNA sequences predominantly found in Firmicutes (65%), Bacteroidetes (18%), Actinobacteria (12%), Fusobacteria (3%), and Proteobacteria (2%) (Figure 1A). The other 6 phyla were represented by less than 1% of the total sequence reads. The phylum Chloroflexi was identified by only the V6 sequence dataset; similarly, the phyla Spirochaetes, Synergistetes and Fibrobacteres were only identified by the V1V2 sequence dataset.

Figure 1
figure 1

Summary of the microbial phyla and orders detected in human female urine. A: An overview of the taxonomy at the phylum level as computed using MEGAN V3.4, using normalized counts by pooling together the V1V2 and V6 16S rDNA reads. The size of the circles is scaled logarithmically to the number of reads assigned to the taxon. Nodes denoted as "Not assigned" and "No hits" are the number of reads that were assigned to a taxon with fewer than 5 hits, or did not match to any sequence when compared to the SSUrdp database, respectively. B and C: Comparison of taxonomic assignments for human female urine sequences at the order level. Reads obtained using the V1V2 hypervariable 16S rDNA region were predominantly assigned to Lactobacillales, and identified in total 18 different orders where Desulfuromonadales and Spirochaetales are unique to this V1V2 dataset. V6 reads revealed a slightly higher diversity with 20 different orders; Bdellovibrionales, Myxococcales, Rhizobiales and Enterobacteriales are only identified by this V6 method.

When examining the two sequence sets separately, 22 different orders were identified in total. The 4 most abundant bacterial orders were the same for both regions sequenced; Lactobacillales (53% for V1V2 and 55% for V6), Bacteroidales (20% for V1V2 and 16% for V6), Clostridiales (10% for V1V2 and 11% for V6), and Bifidobacteriales (9% for V1V2 and 13% for V6) (Figure 1B and 1C). Additionally, 18 other orders were detected in both the V1V2 and V6 datasets. Further, Bdellovibrionales, Myxococcales, Rhizobiales and Enterobacteriales were only identified in the V6 sequence dataset, while Desulfuromonadales and Spirochaetales were only observed in the V1V2 dataset (Figure 1B and 1C).

Analyzing the data at the genus level revealed 45 different genera. 88% and 87% of the reads in the V1V2 and V6 sequence datasets, respectively, were assigned to Lactobacillus, Prevotella and Gardnerella (Figure 2A). These three major genera found in female human urine belong to the three most predominantly detected phyla: Firmicutes, Bacteroidetes and Actinobacteria (Figure 1A). Out of the 45 different genera, 17 genera were unique for the V1V2 sequence reads, whereas a total of 10 genera were uniquely found with V6 sequence reads.

Figure 2
figure 2

Bacterial genera detected in healthy female urine. A: Comparison of healthy female urine bacterial genera abundance determined by sequencing 2 different hypervariable 16S rDNA regions, V1V2 and V6. Relative abundance of 18 major bacterial genera found in the sequence pool of eight different urine samples are shown for the two 16S rDNA regions. Groups denoted "other" represent minor groups classified. Y-axis represents relative abundance. B: Heat map showing the relative abundance of bacterial genera across urine samples of eight healthy females. Genera denoted as phylum_genus, samples denoted as samplenumber_V1V2 or V6. Taxa marked with asterisk (*) could not be assigned to any genera, and are shown at the lowest common taxon: family and order. Color intensity of the heat map is directly proportional to log 10 scale of the abundance normalized sequence data as done by MEGAN.

Keeping the same parameters as for the analysis at higher taxonomic levels, a small number of bacterial reads from the V1V2 and V6 dataset were assigned to species level, see Additional file 1: Table S1. When comparing to previous reports from literature [9, 17, 37, 42–81], nine out of the 45 species listed are associated with UTI. Twenty of the species listed represent uncultured bacteria, many of them with an unknown pathogenic potential (Additional file 1: Table S1).

Variation between urine samples from different individuals

The distribution of the different taxa differed markedly among the urine specimens. 16S rDNA sequences from the phyla Firmicutes and Bacteroidetes were found in all urine samples. Sequences from Proteobacteria and Actinobacteria were observed in 6/8 and 5/8 urine samples respectively, while sequences from Fusobacteria were identified in only 2 samples. The remaining six phyla defined in our pooled urine sequence dataset were only detected once among the urine samples; Spirochaetes, Chloroflexi, Fibrobacteres and Acidobacteria in sample F7, Tenericutes in sample F4 and Synergistetes in sample F2. These results indicate that there is a noticeable intra-individual variation in urine 16S rDNA sequences even at the phylum level.

The interpersonal microbial sequence diversity and the distribution of bacterial DNA at the genus level in each individual are shown in the heat map in Figure 2B. In the majority of the urine specimens (6 out of 8) one genus was dominant, i.e. represented by at least 75% of the reads, while in two specimens (sample F7 and F8) there was a more even distribution among the represented genera (Figure 2B). A polymicrobial state is suggested for all but a single urine specimen based on both of the 16S rDNA sequence datasets. The exception was sample F3, which showed only the presence of Lactobacillus based on the V1V2 reads, while the V6 amplicon sequence data identified seven additional bacterial genera, though at a low frequency. The most frequently identified genus was Prevotella, with sequences present in 7 out of 8 urine samples. Sequences assigned to Lactobacillus, Peptoniphilus and Dialister were also frequently detected (6/8), followed by Finegoldia (5/8), Anaerococcus, Allisonella, Streptococcus, Staphylococcus (all 4/8). Interestingly, reads assigned to Gardnerella were only identified in 3/8 urine samples, even though this genus was the 3rd most abundant group in the pooled sequence dataset for both the V1V2 and V6 regions (Figure 2A). Three other genera and a group of 5 genera were identified by reads belonging to 3 or 2 urine samples, respectively. 24 genera were only detected in 1 out of the 8 samples.

Species richness and diversity estimates of the female urine microbiota

Bacterial taxonomic richness and diversity varied greatly among urine samples investigated in this study. Community richness and diversity were determined using rarefaction plots, Chao1 and Shannon index estimations (Figure 3 and Table 2).

Figure 3
figure 3

Number of OTUs as function of the total number of sequences. A and B: Rarefaction curves of individual samples for the V1V2 (A) and the V6 datasets (B). Curves were generated at 3% genetic difference using MOTHUR v1.17.0 [39]. C and D: Rarefaction curves of the pooled dataset for both V1V2 reads (C) and V6 reads (D). OTUs with ≤3%, ≤6% and ≤10% pairwise sequence difference generated using MOTHUR v1.17.0 [39] are assumed to belong to the same species, genus and family, respectively.

Rarefaction curves were generated for 3% genetic difference level (e.g., at the species level). The number of OTUs calculated for the eight individual samples ranged from 20-504 and 63-499 OTUs for the V1V2 and V6 regions, respectively (Figure 3A, B and Table 2). OTU numbers of the total bacterial community in the female urine at 3% difference for the V1V2 sequence pool was calculated to 1209 OTUs and to 1435 OTUs for the V6 sequence pool (Figure 3C, D and Table 2). Furthermore, total unique OTUs for the V1V2 pooled reads were 1354 and for the V6 pooled reads 2069 (Table 2).

To compare the diversity between the eight different urine samples, the Shannon diversity index was determined both with the original, and with normalized numbers of sequences (Table 2). There was no substantial difference between the two Shannon indices calculated for the same sample.

Discussion

In this work we sequenced two different variable regions of 16S rDNA isolated from eight culture-negative urine samples. Urine samples are at risk of contamination by the bacterial flora of the female urogenital system [82, 83], therefore sampling of mid-stream urine was performed by the clean catch method, under guidance of an experienced urotherapy nurse. To avoid further bacterial growth, which could skew the results, the samples were kept on ice and analyzed within an hour. Amplicon lengths used here exceed the typical fragment size (150-200 bp) of circulating cell-free DNA in urine [84], thus reducing the frequency of such DNA in our analyses.

Bacterial profile of female urine

The sequences found in the samples were mainly assigned to the Firmicutes phylum (65%) with Bacteriodetes, Actinobacteria, Fusobacteria and Proteobacteria members accounting for most of the remaining sequences (Figure 1A). This overall composition of phyla is comparable to prior 16S rDNA sequencing studies of the human urogenital tract (vaginal microbiota [79] and male urogenital tract [27, 28, 85]). However, we also found sequences from Fibrobacteres, a phylum not previously associated with human microbiota as described by the Human Microbiome Project catalog (HMP) [69, 86], the Human Oral Microbiome Database (HOMD) [70, 87] and in studies on the gastrointestinal tract, vaginal and male urine bacterial flora [27, 28, 79, 88, 89].

Our analysis revealed that the bacterial composition in human female urine specimens is polymicrobial and that there is considerable variation between urine samples (Figure 2B). Lactobacillus, Prevotella and Gardnerella were the dominant genera (Figure 2A), however, not every urine sample exhibited 16S rDNA from these genera (Figure 2B), indicating that a single characteristic microbial community for female urine cannot be established. Similar results were also seen in Nelson et al. (2010) [27] and Dong et al. (2011) [28] in their studies on male urine composition. While Lactobacillus and Prevotella were not among the dominant genera in the first study [27], rDNA sequences belonging to these genera were dominant in the latter study [28], as it is in our data. Lactobacillus was, however, considerably more abundant in female than in male urine. The two studies on male urine did not display the genus Gardnerella (typically associated with the female vagina), as a major bacterium, while this genus is one of three dominating genera in our study. In contrast, Sneathia, another vaginal bacterium - only present at low abundance in female urine, was reported as a dominant genus in male urine.

Comparison of V1V2 and V6 primer sets

Two different primer sets previously used for investigating human microbial communities [32, 33] covering different parts of the hypervariable regions were used in this study. The V1V2 region is noted for its robustness for taxonomic classification, while the V6 region is more appropriate for measuring microbial diversity due to high variability [32, 90, 91]. These differences were also reflected in our study where V1V2 uncovered a wider taxonomical range (Figure 2 and Table 2). Both rDNA regions detected approximately the same groups at phylum and order level, however, a larger difference was evident at the genus level. The V1V2 method detected 35 different genera in total, 16 of which were not found in the V6 dataset. The V6 method detected 28 genera in total, where 10 genera were unique to this dataset. Thus, using a combination of these two primer sets clearly maximized the bacterial diversity that could be detected.

Estimated species richness in female urine microbiota

Our OTU calculations on female urine displayed richness levels that were in the same range as reported for commensal vaginal microbiota (1584 OTUs) [79], but lower than those reported for oral (3011 to 5669 OTUs) [4, 92] and fecal samples (up to 5200 OTUs) [90]. For all but one sample, the Chao1 minimum richness estimates for the V1V2 dataset are in close agreement with the observed number of OTUs (Table 2). In addition, the rarefaction curves approached saturation, demonstrating that the OTU diversity was almost completely covered by the V1V2 variable region (Figure 3A and 3C). In contrast, the Chao1 estimates and the rarefaction curves for all but one of the V6 samples indicated that the current sequencing effort for the V6 variable region was not exhaustive (Table 2 and Figure 3B, D).

Clinical significance of the bacterial DNA identified in human female urine

The anaerobe microbial profile of urine specimens is not routinely investigated in microbiological laboratories since fastidious bacteria often evade standard culture conditions. The present work shows that, besides bacterial species associated with vaginal, fecal and skin bacterial flora, unsurprising considering the anatomy of the female urogenital tract, several types of bacteria previously not seen in female urine were identified. Interestingly, some species detected have earlier been described as causing UTI and bacterial vaginosis (BV), but here we also detect these potentially pathogenic species in asymptomatic healthy female urine samples. For example, most of the fastidious (opportunistic), mostly anaerobic pathogenic bacteria identified by 16S rDNA PCR and sequencing in a study of UTI samples [9], were also detected in our study. On the other hand, uropathogenic E.coli (UPEC), a common cause of UTI [93], was not detected in any of our urine samples.

Lactobacillus was dominant in the urine microbiota (see Figure 2A), as it is in the human vaginal microbiota, and all of the other genera previously found in vaginal microbiota were also identified in our samples [64, 79]. BV is in a majority of cases characterized by a shift in composition of the vaginal microbial community that results in decreased number of lactic producing bacteria and increased numbers of other facultative or anaerobic species in relation to normal bacterial flora [79]. A similar shift in bacterial composition as seen in BV was found in 4 of our eight urine samples: Lactobacillus was either present at a low abundance or not detected at all, and the other genera present were mostly anaerobes. One of these, the anaerobe Prevotella disiens is also typically found in females with genital tract infections. Furthermore, the genus Gardnerella, comprising only the species G. vaginalis, is involved in BV, as well as associated with preterm delivery [94, 95], and also reported as an uropathogen [9, 96].

Both the species Aerococcus urinae and the genus Ureaplasma, examples of "difficult-to-culture pathogens" commonly not detectable by conventional culture methods [52], were detected in our samples. A. urinae is generally associated with bladder infection in elderly people, but can also cause serious complications, such as infective endocarditis when not detected and treated during UTI diagnosis [97, 98]. Ureaplasma spp occurs more commonly in patients with symptoms of UTI than previously thought [99], and the species Ureaplasma urealyticum has also been associated with chronic urinary symptoms in women [100]. Whether or not these potentially pathogenic bacteria represent non-pathogenetic variants or are simply not causing any disease in this setting remains to be investigated.

Conclusion

Our finding of sequences of these potentially disease-causing species and genera in healthy female urine is an example of the enhanced resolution that can be obtained by high-throughput sequencing. This study also shows that the urine medium of asymptomatic females is harboring a surprisingly wide range of bacteria, including many potentially associated with pathogenic conditions. Apparently, such bacteria are part of the healthy urine microbiota.

References

  1. Backhed F, Ley RE, Sonnenburg JL, Peterson DA, Gordon JI: Host-bacterial mutualism in the human intestine. Science (New York, NY). 2005, 307 (5717): 1915-1920. 10.1126/science.1104816.

    Article  Google Scholar 

  2. Turnbaugh PJ, Hamady M, Yatsunenko T, Cantarel BL, Duncan A, Ley RE, Sogin ML, Jones WJ, Roe BA, Affourtit JP, et al: A core gut microbiome in obese and lean twins. Nature. 2009, 457 (7228): 480-484. 10.1038/nature07540.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  3. Hooper LV, Midtvedt T, Gordon JI: How host-microbial interactions shape the nutrient environment of the mammalian intestine. Annual review of nutrition. 2002, 22: 283-307. 10.1146/annurev.nutr.22.011602.092259.

    Article  PubMed  CAS  Google Scholar 

  4. Keijser BJ, Zaura E, Huse SM, van der Vossen JM, Schuren FH, Montijn RC, ten Cate JM, Crielaard W: Pyrosequencing analysis of the oral microflora of healthy adults. Journal of dental research. 2008, 87 (11): 1016-1020. 10.1177/154405910808701104.

    Article  PubMed  CAS  Google Scholar 

  5. Sanz Y, Santacruz A, Gauffin P: Gut microbiota in obesity and metabolic disorders. The Proceedings of the Nutrition Society. 2010, 1-8.

    Google Scholar 

  6. Weisenseel P, Prinz JC: Incidental detection of S. pyogenes-DNA in psoriatic skin by PCR. Archives of dermatological research. 2005, 296 (12): 573-576. 10.1007/s00403-005-0559-7.

    Article  PubMed  CAS  Google Scholar 

  7. Aas JA, Griffen AL, Dardis SR, Lee AM, Olsen I, Dewhirst FE, Leys EJ, Paster BJ: Bacteria of dental caries in primary and permanent teeth in children and young adults. J Clin Microbiol. 2008, 46 (4): 1407-1417. 10.1128/JCM.01410-07.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  8. Oakley BB, Fiedler TL, Marrazzo JM, Fredricks DN: Diversity of human vaginal bacterial communities and associations with clinically defined bacterial vaginosis. Appl Environ Microbiol. 2008, 74 (15): 4898-4909. 10.1128/AEM.02884-07.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  9. Imirzalioglu C, Hain T, Chakraborty T, Domann E: Hidden pathogens uncovered: metagenomic analysis of urinary tract infections. Andrologia. 2008, 40 (2): 66-71. 10.1111/j.1439-0272.2007.00830.x.

    Article  PubMed  CAS  Google Scholar 

  10. Dukes CE: Urine examination and clinical interpretation. 1939, New York: Oxford Medical Publications

    Google Scholar 

  11. Osborne NG: Acute Urinary-Tract Infection: A Condition Overdiagnosed in Women?. Journal of Gynecologic Surgery. 2008, 24 (1): 51-54. 10.1089/gyn.2007.999993.

    Article  Google Scholar 

  12. Haarala M, Jalava J, Laato M, Kiilholma P, Nurmi M, Alanen A: Absence of bacterial DNA in the bladder of patients with interstitial cystitis. J Urol. 1996, 156 (5): 1843-1845. 10.1016/S0022-5347(01)65549-6.

    Article  PubMed  CAS  Google Scholar 

  13. Keay S, Schwalbe RS, Trifillis AL, Lovchik JC, Jacobs S, Warren JW: A prospective study of microorganisms in urine and bladder biopsies from interstitial cystitis patients and controls. Urology. 1995, 45 (2): 223-229. 10.1016/0090-4295(95)80009-3.

    Article  PubMed  CAS  Google Scholar 

  14. Keay S, Zhang CO, Baldwin BR, Jacobs SC, Warren JW: Polymerase chain reaction amplification of bacterial 16S rRNA genes in interstitial cystitis and control patient bladder biopsies. J Urol. 1998, 159 (1): 280-283. 10.1016/S0022-5347(01)64082-5.

    Article  PubMed  CAS  Google Scholar 

  15. Domingue GJ, Ghoniem GM, Bost KL, Fermin C, Human LG: Dormant microbes in interstitial cystitis. J Urol. 1995, 153 (4): 1321-1326. 10.1016/S0022-5347(01)67594-3.

    Article  PubMed  CAS  Google Scholar 

  16. Barnett BJ, Stephens DS: Urinary tract infection: an overview. Am J Med Sci. 1997, 314 (4): 245-249. 10.1097/00000441-199710000-00007.

    PubMed  CAS  Google Scholar 

  17. Murray PR, Baron EJ, Jorgensen JH, Landry ML, Pfaller MA: Manual of Clinical Microbiology. 2007, ASM Press, 1: 9

    Google Scholar 

  18. Pace NR: A molecular view of microbial diversity and the biosphere. Science (New York, NY). 1997, 276 (5313): 734-740. 10.1126/science.276.5313.734.

    Article  CAS  Google Scholar 

  19. Rosen DA, Hooton TM, Stamm WE, Humphrey PA, Hultgren SJ: Detection of intracellular bacterial communities in human urinary tract infection. PLoS medicine. 2007, 4 (12): e329-10.1371/journal.pmed.0040329.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Hancock V, Ferrieres L, Klemm P: Biofilm formation by asymptomatic and virulent urinary tract infectious Escherichia coli strains. FEMS microbiology letters. 2007, 267 (1): 30-37. 10.1111/j.1574-6968.2006.00507.x.

    Article  PubMed  CAS  Google Scholar 

  21. Salo J, Sevander JJ, Tapiainen T, Ikaheimo I, Pokka T, Koskela M, Uhari M: Biofilm formation by Escherichia coli isolated from patients with urinary tract infections. Clinical nephrology. 2009, 71 (5): 501-507.

    Article  PubMed  CAS  Google Scholar 

  22. Anderson M, Bollinger D, Hagler A, Hartwell H, Rivers B, Ward K, Steck TR: Viable but nonculturable bacteria are present in mouse and human urine specimens. J Clin Microbiol. 2004, 42 (2): 753-758. 10.1128/JCM.42.2.753-758.2004.

    Article  PubMed  PubMed Central  Google Scholar 

  23. Woo PC, Lau SK, Teng JL, Tse H, Yuen KY: Then and now: use of 16S rDNA gene sequencing for bacterial identification and discovery of novel bacteria in clinical microbiology laboratories. Clin Microbiol Infect. 2008, 14 (10): 908-934. 10.1111/j.1469-0691.2008.02070.x.

    Article  PubMed  CAS  Google Scholar 

  24. Heritz DM, Lacroix JM, Batra SD, Jarvi KA, Beheshti B, Mittelman MW: Detection of eubacteria in interstitial cystitis by 16S rDNA amplification. J Urol. 1997, 158 (6): 2291-2295. 10.1016/S0022-5347(01)68237-5.

    Article  PubMed  CAS  Google Scholar 

  25. Lacroix JM, Jarvic K, Batrab SD, Heritze DM, Mittelman MW: PCR-based technique for the detection of bacteria in semen and urine. J Microbiol Methods. 1996, 26 (1-2): 61-71. 10.1016/0167-7012(96)00844-5.

    Article  CAS  Google Scholar 

  26. Riemersma WA, van der Schee CJ, van der Meijden WI, Verbrugh HA, van Belkum A: Microbial population diversity in the urethras of healthy males and males suffering from nonchlamydial, nongonococcal urethritis. J Clin Microbiol. 2003, 41 (5): 1977-1986. 10.1128/JCM.41.5.1977-1986.2003.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  27. Nelson DE, Van Der Pol B, Dong Q, Revanna KV, Fan B, Easwaran S, Sodergren E, Weinstock GM, Diao L, Fortenberry JD: Characteristic male urine microbiomes associate with asymptomatic sexually transmitted infection. PLoS ONE. 2010, 5 (11): e14116-10.1371/journal.pone.0014116.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Dong Q, Nelson DE, Toh E, Diao L, Gao X, Fortenberry JD, Van Der Pol B: The microbial communities in male first catch urine are highly similar to those in paired urethral swab specimens. PLoS ONE. 2011, 6 (5): e19709-10.1371/journal.pone.0019709.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  29. Margulies M, Egholm M, Altman WE, Attiya S, Bader JS, Bemben LA, Berka J, Braverman MS, Chen YJ, Chen Z, et al: Genome sequencing in microfabricated high-density picolitre reactors. Nature. 2005, 437 (7057): 376-380.

    PubMed  CAS  PubMed Central  Google Scholar 

  30. Sogin ML, Morrison HG, Huber JA, Mark Welch D, Huse SM, Neal PR, Arrieta JM, Herndl GJ: Microbial diversity in the deep sea and the underexplored "rare biosphere". Proc Natl Acad Sci USA. 2006, 103 (32): 12115-12120. 10.1073/pnas.0605127103.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  31. McKenna P, Hoffmann C, Minkah N, Aye PP, Lackner A, Liu Z, Lozupone CA, Hamady M, Knight R, Bushman FD: The macaque gut microbiome in health, lentiviral infection, and chronic enterocolitis. PLoS Pathog. 2008, 4 (2): e20-10.1371/journal.ppat.0040020.

    Article  PubMed  PubMed Central  Google Scholar 

  32. Sundquist A, Bigdeli S, Jalili R, Druzin ML, Waller S, Pullen KM, El-Sayed YY, Taslimi MM, Batzoglou S, Ronaghi M: Bacterial flora typing with deep, targeted, chip-based Pyrosequencing. BMC Microbiol. 2007, 7 (1): 108-10.1186/1471-2180-7-108.

    Article  PubMed  PubMed Central  Google Scholar 

  33. Andersson AF, Lindberg M, Jakobsson H, Backhed F, Nyren P, Engstrand L: Comparative analysis of human gut microbiota by barcoded pyrosequencing. PLoS ONE. 2008, 3 (7): e2836-10.1371/journal.pone.0002836.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Quince C, Lanzen A, Curtis TP, Davenport RJ, Hall N, Head IM, Read LF, Sloan WT: Accurate determination of microbial diversity from 454 pyrosequencing data. Nature methods. 2009, 6 (9): 639-641. 10.1038/nmeth.1361.

    Article  PubMed  CAS  Google Scholar 

  35. ESPRIT. [http://www.biotech.ufl.edu/people/sun/esprit.html]

  36. MEtaGenome ANalyzer. [http://www-ab.informatik.uni-tuebingen.de/software/megan/welcome.html]

  37. Huson DH, Auch AF, Qi J, Schuster SC: MEGAN analysis of metagenomic data. Genome Res. 2007, 17 (3): 377-386. 10.1101/gr.5969107. software freely available for academic purposes from, [http://www-ab.informatik.uni-tuebingen.de/software/megan]

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  38. Urich T, Lanzen A, Qi J, Huson DH, Schleper C, Schuster SC: Simultaneous assessment of soil microbial community structure and function through analysis of the meta-transcriptome. PLoS ONE. 2008, 3 (6): e2527-10.1371/journal.pone.0002527.

    Article  PubMed  PubMed Central  Google Scholar 

  39. Schloss PD, Westcott SL, Ryabin T, Hall JR, Hartmann M, Hollister EB, Lesniewski RA, Oakley BB, Parks DH, Robinson CJ, et al: Introducing mothur: open-source, platform-independent, community-supported software for describing and comparing microbial communities. Appl Environ Microbiol. 2009, 75 (23): 7537-7541. 10.1128/AEM.01541-09.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  40. Huse SM, Welch DM, Morrison HG, Sogin ML: Ironing out the wrinkles in the rare biosphere through improved OTU clustering. Environmental microbiology. 2010, 12 (7): 1889-1898. 10.1111/j.1462-2920.2010.02193.x.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  41. Lemos LN, Fulthorpe RR, Triplett EW, Roesch LF: Rethinking microbial diversity analysis in the high throughput sequencing era. J Microbiol Methods. 2011, 86 (1): 42-51. 10.1016/j.mimet.2011.03.014.

    Article  PubMed  CAS  Google Scholar 

  42. Collins MD, Jovita MR, Hutson RA, Ohlen M, Falsen E: Aerococcus christensenii sp. nov., from the human vagina. Int J Syst Bacteriol. 1999, 49 (Pt 3): 1125-1128.

    Article  PubMed  Google Scholar 

  43. Ezaki T, Kawamura Y, Li N, Li ZY, Zhao L, Shu S: Proposal of the genera Anaerococcus gen. nov., Peptoniphilus gen. nov. and Gallicola gen. nov. for members of the genus Peptostreptococcus. Int J Syst Evol Microbiol. 2001, 51 (Pt 4): 1521-1528.

    Article  PubMed  CAS  Google Scholar 

  44. Greub G, Raoult D: "Actinobaculum massiliae," a new species causing chronic urinary tract infection. J Clin Microbiol. 2002, 40 (11): 3938-3941. 10.1128/JCM.40.11.3938-3941.2002.

    Article  PubMed  PubMed Central  Google Scholar 

  45. Hitti J, Hillier SL, Agnew KJ, Krohn MA, Reisner DP, Eschenbach DA: Vaginal indicators of amniotic fluid infection in preterm labor. Obstet Gynecol. 2001, 97 (2): 211-219. 10.1016/S0029-7844(00)01146-7.

    Article  PubMed  CAS  Google Scholar 

  46. Ibler K, Truberg Jensen K, Ostergaard C, Sonksen UW, Bruun B, Schonheyder HC, Kemp M, Dargis R, Andresen K, Christensen JJ: Six cases of Aerococcus sanguinicola infection: clinical relevance and bacterial identification. Scand J Infect Dis. 2008, 40 (9): 761-765. 10.1080/00365540802078059.

    Article  PubMed  CAS  Google Scholar 

  47. Malinen E, Krogius-Kurikka L, Lyra A, Nikkila J, Jaaskelainen A, Rinttila T, Vilpponen-Salmela T, von Wright AJ, Palva A: Association of symptoms with gastrointestinal microbiota in irritable bowel syndrome. World J Gastroenterol. 2010, 16 (36): 4532-4540. 10.3748/wjg.v16.i36.4532.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  48. Nielsen HL, Soby KM, Christensen JJ, Prag J: Actinobaculum schaalii: a common cause of urinary tract infection in the elderly population. Bacteriological and clinical characteristics. Scand J Infect Dis. 2010, 42 (1): 43-47. 10.3109/00365540903289662.

    Article  PubMed  CAS  Google Scholar 

  49. Svenungsson B, Lagergren A, Ekwall E, Evengard B, Hedlund KO, Karnell A, Lofdahl S, Svensson L, Weintraub A: Enteropathogens in adult patients with diarrhea and healthy control subjects: a 1-year prospective study in a Swedish clinic for infectious diseases. Clin Infect Dis. 2000, 30 (5): 770-778. 10.1086/313770.

    Article  PubMed  CAS  Google Scholar 

  50. Vedel G, Toussaint G, Riegel P, Fouilladieu JL, Billoet A, Poyart C: Corynebacterium pseudogenitalium urinary tract infection. Emerg Infect Dis. 2006, 12 (2): 355-356.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Wildeboer-Veloo AC, Harmsen HJ, Welling GW, Degener JE: Development of 16S rRNA-based probes for the identification of Gram-positive anaerobic cocci isolated from human clinical specimens. Clin Microbiol Infect. 2007, 13 (10): 985-992. 10.1111/j.1469-0691.2007.01803.x.

    Article  PubMed  CAS  Google Scholar 

  52. Zhang Q, Kwoh C, Attorri S, Clarridge JE: Aerococcus urinae in urinary tract infections. J Clin Microbiol. 2000, 38 (4): 1703-1705.

    PubMed  CAS  PubMed Central  Google Scholar 

  53. Eubacterium sp. oral clone BU061. [http://www.ncbi.nlm.nih.gov/nuccore/AF385567]

  54. Bjornsson L, Hugenholtz P, Tyson GW, Blackall LL: Filamentous Chloroflexi (green non-sulfur bacteria) are abundant in wastewater treatment processes with biological nutrient removal. Microbiology. 2002, 148 (Pt 8): 2309-2318.

    Article  PubMed  CAS  Google Scholar 

  55. Collins MD, Falsen E, Lemozy J, Akervall E, Sjoden B, Lawson PA: Phenotypic and phylogenetic characterization of some Globicatella-like organisms from human sources: description of Facklamia hominis gen. nov., sp. nov. Int J Syst Bacteriol. 1997, 47 (3): 880-882. 10.1099/00207713-47-3-880.

    Article  PubMed  CAS  Google Scholar 

  56. Gao Z, Tseng CH, Pei Z, Blaser MJ: Molecular analysis of human forearm superficial skin bacterial biota. Proc Natl Acad Sci USA. 2007, 104 (8): 2927-2932. 10.1073/pnas.0607077104.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  57. Hansen J, Gulati A, Sartor RB: The role of mucosal immunity and host genetics in defining intestinal commensal bacteria. Curr Opin Gastroenterol. 2010, 26 (6): 564-571. 10.1097/MOG.0b013e32833f1195.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  58. Healy B, Beukenholt RW, Tuthill D, Ribeiro CD: Facklamia hominis causing chorioamnionitis and puerperal bacteraemia. The Journal of infection. 2005, 50 (4): 353-355. 10.1016/j.jinf.2004.05.013.

    Article  PubMed  CAS  Google Scholar 

  59. Kalyuzhnaya MG, Bowerman S, Lara JC, Lidstrom ME, Chistoserdova L: Methylotenera mobilis gen. nov., sp. nov., an obligately methylamine-utilizing bacterium within the family Methylophilaceae. Int J Syst Evol Microbiol. 2006, 56 (Pt 12): 2819-2823.

    Article  PubMed  CAS  Google Scholar 

  60. Karlsson C, Morgelin M, Collin M, Lood R, Andersson ML, Schmidtchen A, Bjorck L, Frick IM: SufA - a bacterial enzyme that cleaves fibrinogen and blocks fibrin network formation. Microbiology. 2009, 155 (Pt 1): 238-248.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  61. Munson MA, Pitt-Ford T, Chong B, Weightman A, Wade WG: Molecular and cultural analysis of the microflora associated with endodontic infections. Journal of dental research. 2002, 81 (11): 761-766. 10.1177/154405910208101108.

    Article  PubMed  CAS  Google Scholar 

  62. Nikolaitchouk N, Andersch B, Falsen E, Strombeck L, Mattsby-Baltzer I: The lower genital tract microbiota in relation to cytokine-, SLPI- and endotoxin levels: application of checkerboard DNA-DNA hybridization (CDH). APMIS. 2008, 116 (4): 263-277. 10.1111/j.1600-0463.2008.00808.x.

    Article  PubMed  CAS  Google Scholar 

  63. Nikolaitchouk N, Wacher C, Falsen E, Andersch B, Collins MD, Lawson PA: Lactobacillus coleohominis sp. nov., isolated from human sources. Int J Syst Evol Microbiol. 2001, 51 (Pt 6): 2081-2085.

    Article  PubMed  CAS  Google Scholar 

  64. Ravel J, Gajer P, Abdo Z, Schneider GM, Koenig SS, McCulle SL, Karlebach S, Gorle R, Russell J, Tacket CO, et al: Vaginal microbiome of reproductive-age women. Proc Natl Acad Sci USA. 2011, 108 (Suppl 1): 4680-4687.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  65. Riggio MP, Aga H, Murray CA, Jackson MS, Lennon A, Hammersley N, Bagg J: Identification of bacteria associated with spreading odontogenic infections by 16S rRNA gene sequencing. Oral Surg Oral Med Oral Pathol Oral Radiol Endod. 2007, 103 (5): 610-617. 10.1016/j.tripleo.2006.08.009.

    Article  PubMed  Google Scholar 

  66. Shah HN, Collins DM: Prevotella, a new genus to include Bacteroides melaninogenicus and related species formerly classified in the genus Bacteroides. Int J Syst Bacteriol. 1990, 40 (2): 205-208. 10.1099/00207713-40-2-205.

    Article  PubMed  CAS  Google Scholar 

  67. Stackebrandt E, Schumann P, Schaal KP, Weiss N: Propionimicrobium gen. nov., a new genus to accommodate Propionibacterium lymphophilum (Torrey 1916) Johnson and Cummins 1972, 1057AL as Propionimicrobium lymphophilum comb. nov. Int J Syst Evol Microbiol. 2002, 52 (Pt 6): 1925-1927.

    PubMed  CAS  Google Scholar 

  68. Summanen PH, Durmaz B, Vaisanen ML, Liu C, Molitoris D, Eerola E, Helander IM, Finegold SM: Porphyromonas somerae sp. nov., a pathogen isolated from humans and distinct from porphyromonas levii. J Clin Microbiol. 2005, 43 (9): 4455-4459. 10.1128/JCM.43.9.4455-4459.2005.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  69. Human Microbiome Project catalog. [http://www.hmpdacc.org]

  70. Human Oral Microbiome Database. [http://www.homd.org]

  71. European urinalysis guidelines. Scand J Clin Lab Invest Suppl. 2000, 231: 1-86.

  72. Butler JE, Young ND, Lovley DR: Evolution from a respiratory ancestor to fill syntrophic and fermentative niches: comparative fenomics of six Geobacteraceae species. BMC genomics. 2009, 10: 103-10.1186/1471-2164-10-103.

    Article  PubMed  PubMed Central  Google Scholar 

  73. Dekio I, Hayashi H, Sakamoto M, Kitahara M, Nishikawa T, Suematsu M, Benno Y: Detection of potentially novel bacterial components of the human skin microbiota using culture-independent molecular profiling. J Med Microbiol. 2005, 54 (Pt 12): 1231-1238.

    Article  PubMed  CAS  Google Scholar 

  74. Gloux K, Berteau O, El Oumami H, Beguet F, Leclerc M, Dore J: A metagenomic beta-glucuronidase uncovers a core adaptive function of the human intestinal microbiome. Proc Natl Acad Sci USA. 2011, 108 (Suppl 1): 4539-4546.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  75. Haffajee AD, Teles RP, Patel MR, Song X, Veiga N, Socransky SS: Factors affecting human supragingival biofilm composition. I. Plaque mass. J Periodontal Res. 2009, 44 (4): 511-519. 10.1111/j.1600-0765.2008.01154.x.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  76. Hayashi H, Sakamoto M, Benno Y: Fecal microbial diversity in a strict vegetarian as determined by molecular analysis and cultivation. Microbiol Immunol. 2002, 46 (12): 819-831.

    Article  PubMed  CAS  Google Scholar 

  77. Kang JG, Kim SH, Ahn TY: Bacterial diversity in the human saliva from different ages. J Microbiol. 2006, 44 (5): 572-576.

    PubMed  CAS  Google Scholar 

  78. Knapp JS, Koumans EH: Neisseria and Branhamella. Manual of Clinical Microbiology. Edited by: Murray PR, Baron EJ, Pfaller MAea. 1999, Washington, DC, USA: ASM press, 586-603. 7

    Google Scholar 

  79. Ling Z, Kong J, Liu F, Zhu H, Chen X, Wang Y, Li L, Nelson KE, Xia Y, Xiang C: Molecular analysis of the diversity of vaginal microbiota associated with bacterial vaginosis. BMC genomics. 2010, 11: 488-10.1186/1471-2164-11-488.

    Article  PubMed  PubMed Central  Google Scholar 

  80. Silva DG, Tinoco EM, Rocha GA, Rocha AM, Guerra JB, Saraiva IE, Queiroz DM: Helicobacter pylori transiently in the mouth may participate in the transmission of infection. Mem Inst Oswaldo Cruz. 2010, 105 (5): 657-660. 10.1590/S0074-02762010000500009.

    Article  PubMed  CAS  Google Scholar 

  81. Will C, Thurmer A, Wollherr A, Nacke H, Herold N, Schrumpf M, Gutknecht J, Wubet T, Buscot F, Daniel R: Horizon-specific bacterial community composition of German grassland soils, as revealed by pyrosequencing-based analysis of 16S rRNA genes. Appl Environ Microbiol. 2010, 76 (20): 6751-6759. 10.1128/AEM.01063-10.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  82. Jackson SR, Dryden M, Gillett P, Kearney P, Weatherall R: A novel midstream urine-collection device reduces contamination rates in urine cultures amongst women. BJU international. 2005, 96 (3): 360-364. 10.1111/j.1464-410X.2005.05631.x.

    Article  PubMed  Google Scholar 

  83. Bekeris LG, Jones BA, Walsh MK, Wagar EA: Urine culture contamination: a College of American Pathologists Q-Probes study of 127 laboratories. Arch Pathol Lab Med. 2008, 132 (6): 913-917.

    PubMed  Google Scholar 

  84. Umansky SR, Tomei LD: Transrenal DNA testing: progress and perspectives. Expert review of molecular diagnostics. 2006, 6 (2): 153-163. 10.1586/14737159.6.2.153.

    Article  PubMed  CAS  Google Scholar 

  85. Price LB, Liu CM, Johnson KE, Aziz M, Lau MK, Bowers J, Ravel J, Keim PS, Serwadda D, Wawer MJ, et al: The effects of circumcision on the penis microbiome. PLoS ONE. 2010, 5 (1): e8422-10.1371/journal.pone.0008422.

    Article  PubMed  PubMed Central  Google Scholar 

  86. Nelson KE, Weinstock GM, Highlander SK, Worley KC, Creasy HH, Wortman JR, Rusch DB, Mitreva M, Sodergren E, Chinwalla AT, et al: A catalog of reference genomes from the human microbiome. Science (New York, NY). 2010, 328 (5981): 994-999.

    Article  CAS  Google Scholar 

  87. Chen T, Yu WH, Izard J, Baranova OV, Lakshmanan A, Dewhirst FE: The Human Oral Microbiome Database: a web accessible resource for investigating oral microbe taxonomic and genomic information. Database (Oxford). 2010, 2010: baq013-

    Article  Google Scholar 

  88. Carroll IM, Threadgill DW, Threadgill DS: The gastrointestinal microbiome: a malleable, third genome of mammals. Mamm Genome. 2009, 20 (7): 395-403. 10.1007/s00335-009-9204-7.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  89. Arumugam M, Raes J, Pelletier E, Le Paslier D, Yamada T, Mende DR, Fernandes GR, Tap J, Bruls T, Batto JM, et al: Enterotypes of the human gut microbiome. Nature. 2011

    Google Scholar 

  90. Claesson MJ, O'Sullivan O, Wang Q, Nikkila J, Marchesi JR, Smidt H, de Vos WM, Ross RP, O'Toole PW: Comparative analysis of pyrosequencing and a phylogenetic microarray for exploring microbial community structures in the human distal intestine. PLoS ONE. 2009, 4 (8): e6669-10.1371/journal.pone.0006669.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Liu Z, DeSantis TZ, Andersen GL, Knight R: Accurate taxonomy assignments from 16S rRNA sequences produced by highly parallel pyrosequencers. Nucleic Acids Res. 2008, 36 (18): e120-10.1093/nar/gkn491.

    Article  PubMed  PubMed Central  Google Scholar 

  92. Lazarevic V, Whiteson K, Hernandez D, Francois P, Schrenzel J: Study of inter- and intra-individual variations in the salivary microbiota. BMC genomics. 2010, 11: 523-10.1186/1471-2164-11-523.

    Article  PubMed  PubMed Central  Google Scholar 

  93. Foxman B: The epidemiology of urinary tract infection. Nat Rev Urol. 2010, 7 (12): 653-660. 10.1038/nrurol.2010.190.

    Article  PubMed  Google Scholar 

  94. Fredricks DN, Fiedler TL, Marrazzo JM: Molecular identification of bacteria associated with bacterial vaginosis. N Engl J Med. 2005, 353 (18): 1899-1911. 10.1056/NEJMoa043802.

    Article  PubMed  CAS  Google Scholar 

  95. Menard JP, Mazouni C, Salem-Cherif I, Fenollar F, Raoult D, Boubli L, Gamerre M, Bretelle F: High vaginal concentrations of Atopobium vaginae and Gardnerella vaginalis in women undergoing preterm labor. Obstet Gynecol. 2010, 115 (1): 134-140. 10.1097/AOG.0b013e3181c391d7.

    Article  PubMed  Google Scholar 

  96. Domann E, Hong G, Imirzalioglu C, Turschner S, Kuhle J, Watzel C, Hain T, Hossain H, Chakraborty T: Culture-independent identification of pathogenic bacteria and polymicrobial infections in the genitourinary tract of renal transplant recipients. J Clin Microbiol. 2003, 41 (12): 5500-5510. 10.1128/JCM.41.12.5500-5510.2003.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  97. Schuur PM, Sabbe L, van der Wouw AJ, Montagne GJ, Buiting AG: Three cases of serious infection caused by Aerococcus urinae. Eur J Clin Microbiol Infect Dis. 1999, 18 (5): 368-371. 10.1007/PL00015022.

    Article  PubMed  CAS  Google Scholar 

  98. Slany M, Freiberger T, Pavlik P, Cerny J: Culture-negative infective endocarditis caused by Aerococcus urinae. J Heart Valve Dis. 2007, 16 (2): 203-205.

    PubMed  Google Scholar 

  99. Pedraza Aviles AG, Ortiz Zaragoza MC: Symptomatic bacteriuria due to Ureaplasma and Mycoplasma in adults. Rev Latinoam Microbiol. 1998, 40 (1-2): 9-13.

    PubMed  CAS  Google Scholar 

  100. Baka S, Kouskouni E, Antonopoulou S, Sioutis D, Papakonstantinou M, Hassiakos D, Logothetis E, Liapis A: Prevalence of Ureaplasma urealyticum and Mycoplasma hominis in women with chronic urinary symptoms. Urology. 2009, 74 (1): 62-66. 10.1016/j.urology.2009.02.014.

    Article  PubMed  Google Scholar 

  101. Guide To Amplicon Sequencing. [http://www.my454.com/downloads/protocols/Guide_To_Amplicon_Sequencing.pdf]

Download references

Acknowledgements

The authors would like to thank Hege Junita Gaup for technical assistance, and the Norwegian Sequencing Centre (NSC), Department of Biology, University of Oslo, for sequencing services. A special thanks to Professor Lars Magne Eri and urotherapist Turid H Hoel at Aker University Hospital HF, Urological Clinic, for specimen collection. Financial support for this research was provided by grants from the Research Council of Norway to KSJ and from CEES to HS.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Kjetill S Jakobsen.

Additional information

Authors' contributions

HS, AJN, SLJ and KSJ have contributed to the design of this study; HS processed the samples and carried out laboratory procedures. KL, AJN and HS performed the bioinformatics and taxonomic analyses. HS authored the manuscript and all authors edited and commented on the paper. All authors read and approved the final manuscript.

Electronic supplementary material

12866_2011_1533_MOESM1_ESM.DOC

Additional file 1: Table S1: Bacteria species identified in female urine by 16S rDNA amplicon 454 pyrosequencing and their general pathogenic potential. (DOC 218 KB)

Authors’ original submitted files for images

Rights and permissions

Open Access This article is published under license to BioMed Central Ltd. This is an Open Access article is distributed under the terms of the Creative Commons Attribution License ( https://creativecommons.org/licenses/by/2.0 ), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

Siddiqui, H., Nederbragt, A.J., Lagesen, K. et al. Assessing diversity of the female urine microbiota by high throughput sequencing of 16S rDNA amplicons. BMC Microbiol 11, 244 (2011). https://doi.org/10.1186/1471-2180-11-244

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2180-11-244

Keywords