Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Placenta-Derived Fetal Specific mRNA Is More Readily Detectable in Maternal Plasma than in Whole Blood

  • Macy M. S. Heung,

    Affiliations Centre for Research into Circulating Fetal Nucleic Acids, Li Ka Shing Institute of Health Sciences, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China, Department of Chemical Pathology, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China

  • Shengnan Jin,

    Affiliations Centre for Research into Circulating Fetal Nucleic Acids, Li Ka Shing Institute of Health Sciences, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China, Department of Chemical Pathology, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China

  • Nancy B. Y. Tsui,

    Affiliations Centre for Research into Circulating Fetal Nucleic Acids, Li Ka Shing Institute of Health Sciences, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China, Department of Chemical Pathology, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China

  • Chunming Ding,

    Affiliation Stanley Ho Centre for Emerging Infectious Diseases, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China

  • Tak Y. Leung,

    Affiliation Department of Obstetrics and Gynaecology, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China

  • Tze K. Lau,

    Affiliation Department of Obstetrics and Gynaecology, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China

  • Rossa W. K. Chiu,

    Affiliations Centre for Research into Circulating Fetal Nucleic Acids, Li Ka Shing Institute of Health Sciences, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China, Department of Chemical Pathology, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China

  • Y. M. Dennis Lo

    loym@cuhk.edu.hk

    Affiliations Centre for Research into Circulating Fetal Nucleic Acids, Li Ka Shing Institute of Health Sciences, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China, Department of Chemical Pathology, The Chinese University of Hong Kong, Prince of Wales Hospital, Shatin, Hong Kong Special Administrative Region, China

Abstract

Background

Placental mRNA was detected in maternal whole blood, raising the possibility of using maternal blood for noninvasive prenatal diagnosis. We investigated fetal mRNA detection in maternal whole blood and determined if it offered advantages over maternal plasma analysis.

Methodology

The concentrations of placental expressed genes, CSH1, KISS1, PLAC4 and PLAC1 in plasma and whole blood from healthy pregnant and non-pregnant individuals were compared by real-time quantitative reverse-transcriptase polymerase chain reaction analysis. Their fetal specificity was investigated by comparing the transcript concentrations in pre- and post-delivery samples and through SNP genotyping by matrix-assisted laser-desorption and ionization time-of-flight mass spectrometry. The gene expression profiles of pregnant and non-pregnant whole blood were investigated by microarray analysis. Upregulated genes in pregnant whole blood were selected for further quantitative analysis.

Principal Findings

The concentrations of the four transcripts were significantly higher in third trimester maternal whole blood than corresponding plasma without significant correlations. KISS1, PLAC4 and PLAC1 were detected in non-pregnant whole blood but not plasma. The transcripts remained detectable in some postpartum whole blood samples. The PLAC4 mRNA in maternal plasma showed fetal genotype while that in corresponding whole blood indicated both fetal and maternal contributions. Microarray analysis revealed upregulation of genes involved in neutrophil functions in pregnant whole blood including DEFA4, CEACAM8, OLFM4, ORM1, MMP8 and MPO. Though possibly pregnancy-related, they were not pregnancy-specific as suggested by the lack of post-delivery reduction in concentrations.

Conclusions

Maternal plasma is preferred over maternal whole blood for placenta-derived fetal RNA detection. Most studied ‘placental’ mRNA molecules in maternal whole blood were of maternal origin and might be derived from processes such as ‘illegitimate transcription’.

Introduction

The analysis of fetal nucleic acids in maternal plasma holds much promise for noninvasive prenatal diagnosis [1][3]. Many investigations conducted to date were based on the detection of Y chromosomal sequences in maternal plasma [4], thus, limiting the utility of such applications to pregnancies involving male fetuses only. Circulating fetal mRNA analysis, on the other hand, offers a means for noninvasive prenatal assessment that is applicable to pregnancies regardless of the fetal gender [5][7].

Through the detection of the mRNA of placental expressed hormones, namely chorionic somatomammotropin hormone 1 (placental lactogen) (CSH1, NM_001317) and chorionic gonadotropin, beta polypeptide (CGB, NM_000737), the placenta was shown to be a source for fetal mRNA release into maternal plasma [5]. This observation facilitated the development of a microarray-based strategy to systematically identify placental expressed mRNA markers that were detectable in maternal plasma [6]. Furthermore, aberrant concentrations of placental expressed mRNA species have been shown to be associated with pregnancy complications, such as the elevation of corticotropin releasing hormone (CRH, NM_000756) mRNA in maternal plasma of preeclamptic pregnancies [7], [8]. In 2007, Lo et al. demonstrated that fetal chromosomal aneuploidy, such as trisomy 21, can be detected non-invasively from maternal plasma analysis by RNA-single nucleotide polymorphism (RNA-SNP) allelic ratio determination [9]. These studies suggest that circulating placental mRNA detection offers much opportunity for the development of noninvasive prenatal diagnostic or assessment applications.

Besides maternal plasma, maternal whole blood has been reported to be another medium for fetal mRNA detection [10][14]. CSH1, CGB and placenta-specific 4 (PLAC4, NM_182832) mRNA concentrations were reported to be much higher in maternal whole blood than plasma [12], [14]. Indeed, the total RNA content of whole blood is higher than plasma. Hence, in this study, we investigated the detection of fetal mRNA in maternal whole blood and determined if it offered advantages over maternal plasma analysis.

Materials and Methods

Ethics statement

This study was conducted according to the principles expressed in the Declaration of Helsinki. Ethics approval from the Joint Chinese University of Hong Kong-New Territories East Cluster Clinical Research Ethics Committee was obtained (CRE-2005.131). All patients provided written informed consent for the collection of samples and subsequent analysis.

Study design

Using real-time quantitative reverse-transcriptase polymerase chain reaction (QRT-PCR), we first compared the concentrations of four previously studied placental expressed transcripts, CSH1, KiSS-1 metastasis-suppressor (KISS1, NM_002256), placenta-specific 1 (PLAC1, NM_021796) [5] and PLAC4 [9] in maternal plasma and whole blood. Our previous microarray study [6] showed that these transcripts had much higher expression in placental tissues than blood cells. The relative placental-specificity of these transcripts are also supported by data in the public database, SymAtlas v1.2.4, Genomics Institute of the Novartis Research Foundation (Figure S1, S2, S3) [15]. Figure S1, S2, S3 shows that those transcripts are predominantly expressed in placental tissues as compared with the many other human tissues assessed by gene expression microarrays. We then determined the pregnancy-specificity of CSH1, KISS1, PLAC1 and PLAC4 by assessing their disappearance in post-delivery blood samples and detectability in blood samples of non-pregnant females and males. We further confirmed the fetal specificity of PLAC4 and chorionic somatomammotropin hormone-like 1 (CSHL1, NM_022579) mRNA by studying their genotypes in maternal plasma and whole blood using the RNA-SNP approach by matrix-assisted laser-desorption and ionization time-of-flight mass spectrometry. CSHL1 is one of the potentially pregnancy-specific placental expressed transcripts identified in our previous microarray study [6]. Lastly, to exploit if more pregnancy-related circulating mRNA markers could be developed for maternal whole blood analysis, we mined for candidates after performing gene expression microarray comparison of whole blood samples from pregnant and non-pregnant individuals.

Subjects and sample collection

Third trimester healthy women with singleton uncomplicated pregnancies (gestational age range, 38–39 weeks) were recruited with written informed consent from the Prince of Wales Hospital, Hong Kong. 12 mL of peripheral blood was collected into EDTA tubes before and at 24 hours after delivery. Placental tissues were collected immediately after delivery. Age-matched healthy non-pregnant individuals were recruited from the community.

General sample processing

Blood samples were processed within 6 hours. 300 µL whole blood was added into 900 µL Trizol LS reagent (Invitrogen). The remainder of the whole blood sample was processed as previously reported for plasma and buffy coat [5], [9], [16]. All the processed blood samples were stored at −80°C until RNA or DNA extraction. Placental tissue samples were handled as per previous studies [5], [9].

RNA extraction

For every 1.8 mL of plasma-Trizol-LS mixture and 1.2 mL of whole blood-Trizol-LS mixture, 200 µL and 240 µL of chloroform were added, respectively, and centrifuged at 12,000 g for 15 min at 4°C. The aqueous phase collected was mixed with 0.54 volume of 100% ethanol. All mixture from a single sample was applied to a RNeasy minicolumn (Qiagen) and processed following the manufacturer's protocol. Total RNA was eluted with 48 µL of RNase-free water. Placental RNA was extracted as reported [5], [9]. DNase I (Invitrogen) treatment was performed to remove genomic DNA contamination.

DNA extraction

Maternal buffy coat and placental tissue DNA was extracted as previously described [9].

Genes Investigated In This Study

Human genes: CSH1, chorionic somatomammotropin hormone 1 (placental lactogen) (NM_001317); KISS1, KiSS-1 metastasis-suppressor (NM_002256); PLAC1, placenta-specific 1 (NM_021796); PLAC4, placenta-specific 4 (NM_182832); CSHL1, chorionic somatomammotropin hormone-like 1 (NM_022579); DEFA4, defensin, alpha 4, corticostatin (NM_001925); CEACAM8, carcinoembryonic antigen-related cell adhesion molecule 8 (NM_001816); OLFM4, olflactomedin 4 (NM_006418); FLCN, folliculin (NM_144606); ORM1, orosomucoid 1 (NM_000607); MMP8, matrix metallopeptidase 8 (neutrophil collagenase) (NM_002424); MPO, myeloperoxidase (NM_000250).

Real-Time QRT-PCR

All mRNA transcripts were quantified using one-step QRT-PCR. Calibration curves were prepared by serial dilutions of HPLC-purified single-stranded synthetic DNA oligonucleotides (Proligos, Singapore) with the amplicon specified. The sequence information of the primers, probes and calibrators as well as the reaction conditions are summarized in Table S1. QRT-PCR was set up as previously described [6], [9].

RNA-SNP genotyping

Genotyping of PLAC4 SNP, rs8130833, and CSHL1 SNP, rs2246207, was performed involving steps of reverse transcription of RNA, PCR amplification, base extension and mass spectrometric analysis of the extension products [9]. DNA was extracted from placental tissues and maternal buffy coat for determination of the fetal and maternal genomic genotypes. The circulating PLAC4 and CSHL1 mRNA genotypes were then determined in maternal plasma and whole blood samples and further compared with the fetal and maternal genomic genotypes to confirm if they indeed originated from the fetus.

Protocols for the PLAC4 SNP genotyping assay have been described previously [9]. Procedures for CSHL1 were the same as PLAC4 except that 454 ng placental RNA, 2.5 ng whole blood RNA or 48 µL plasma RNA was reverse transcribed at 60°C for 60 min followed by 85°C for 5 min. Placental cDNA of 2.5 µL was added to a total PCR volume of 25 µL. PCR reagents were added to 100 µL for plasma or whole blood cDNA. The PCR reaction mix was the same as that for PLAC4 except 1× HotStar Taq PCR buffer was used for placental cDNA. The PCR was initiated at 95°C for 15 min, followed by 45 cycles of denaturation at 95°C for 20 s, annealing at 65°C for 30 s and extension at 72°C for 1 min with final incubation at 72°C for 3 min. 75 cycles were performed in the primer extension reaction for CSHL1. Primer sequences and molecular weights for the extension primer and extension products of each SNP allele are shown in Tables S2 and S3.

Gene expression microarray analysis

From each of five third trimester pregnant women and five non-pregnant healthy females, 5 mL of peripheral blood was collected into PAXgene™ blood RNA tubes (PreAnalytiX) with further processing and RNA extraction according to manufacturer's instructions. DNase treatment was performed with RNase-Free DNase Set (Qiagen). 7 µg of extracted RNA from each sample was treated following the globin-reduction protocol recommended by Affymetrix [17]. The globin-reduced RNA was labeled and hybridized to the GeneChipH Human Genome U133A and U133B Arrays (Affymetrix, Santa Clara, CA) according to manufacturer's instructions. After hybridization, each array chip was washed, stained and scanned as reported [6]. The data were analyzed with GeneSpring v 7.2 (Agilent Technologies, Palo Alto, CA).

The microarray data in the. CEL format was imported and normalized using the following steps in sequence: (1) raw data processing by Robust Multi-chip Average, with GC-content background correction (GC-RMA); (2) data transformation whereby microarray data with values below 0.001 were set to 0.001; and (3) the signal intensity for each gene was divided by the median of its measurements in all samples. After filtering the gene-list with 70% confidence interval and non-parametric t-test analysis (P = 0.05), 2246 gene transcripts demonstrated differential expression between pregnant and non-pregnant blood samples. The gene list was further filtered by selecting transcripts with 1.5-fold higher expression in the pregnant compared with non-pregnant blood samples, and resulted in 247 gene transcripts. We sorted these genes in descending order of the fold difference.

Statistics

Statistical analyses were performed using Sigma Stat (Systat).

Results

placental mRNA concentrations in third trimester maternal plasma and whole blood

Whole blood and plasma samples from ten third trimester pregnancies were analyzed. The median (range) concentrations for CSH1, KISS1, PLAC4 and PLAC1 mRNA in whole blood were 90 673 (9 324–387 010), 14 712 (0–71 544), 344 330 (88 940–552 686) and 12 498 (3 250–54 328) copies/mL, respectively. The median (range) concentrations for CSH1, KISS1, PLAC4 and PLAC1 mRNA in plasma were 11 338 (1 771–27 754), 535 (0–2 283), 4 317 (2 779–10 737) and 498 (0–1 276) copies/mL, respectively. The median concentrations of CSH1, KISS1, PLAC4 and PLAC1 mRNA were 8-fold (P = 0.001, Mann-Whitney test), 27-fold (P = 0.017, Mann-Whitney test), 80-fold (P<0.001, Mann-Whitney test) and 25-fold (P<0.001, Mann-Whitney test) higher, respectively, in whole blood than plasma. Using the Spearman correlation test, no statistical significant correlation was observed between the whole blood and plasma mRNA signals for CSH1 (R2 = 0.045, P = 0.583), KISS1 (R2 = 0.003, P = 0.365), PLAC4 (R2 = 0.013, P = 0.631), and PLAC1 (R2 = 0.041, P = 0.631).

Post-delivery clearance

Whole blood and plasma samples were also collected 24 hours after delivery from the ten cases discussed above. The four transcripts were no longer detectable in all postpartum plasma samples. The plasma concentrations detected before and after delivery were statistically significant for all transcripts (P<0.001, Wilcoxon test). For the post-delivery whole blood samples, CSH1 and KISS1 mRNA became undetectable in 8 and 9 cases, respectively. However, PLAC4 mRNA remained detectable in all cases at a median concentration of 162 203 copies/mL. PLAC1 mRNA was detected in 7 post-delivery whole blood samples at a median of 10 081 copies/mL (Figure 1). Using the Wilcoxon test, the whole blood concentrations detected before and after delivery were shown to be significantly different for CSH1 and KISS1 (P<0.001) but not for PLAC4 (P = 0.160) and PLAC1 (P = 0.105).

thumbnail
Figure 1. Concentrations of placenta-derived mRNA in pre- and 24-hour-post delivery maternal whole blood.

Scatter plots of (A) CSH1 (B) KISS1 (C) PLAC4 (D) PLAC1 mRNA. Corresponding samples from each individual are connected by a line.

https://doi.org/10.1371/journal.pone.0005858.g001

non-pregnant individuals

CSH1, KISS1, PLAC1 and PLAC4 mRNA were not detectable in the plasma of 10 non-pregnant females and 10 males. Whole blood samples from these 20 individuals showed absence of CSH1 mRNA. KISS1 mRNA was detected in whole blood of 2 each of the non-pregnant females and males at 96 and 1 315 copies/mL in the former group and 264 and 300 copies/mL in the males. PLAC4 mRNA was detected in all 20 whole blood samples with median concentrations of 244 498 copies/mL and 592 372 copies/mL, respectively, in non-pregnant females and males. PLAC1 mRNA could be detected in whole blood of 9 non-pregnant females and 8 males at median concentrations of 12 024 copies/mL and 7 959 copies/mL, respectively. These concentrations were not significantly different from that of the third trimester whole blood samples for PLAC4 (P = 0.065, Kruskal-Wallis one way analysis of variance (Kruskal-Wallis ANOVA) test) and PLAC1 (P = 0.368, Kruskal-Wallis ANOVA test) (Figure 2).

thumbnail
Figure 2. Concentrations of placenta-derived mRNA in third trimester pregnant and non-pregnant female and male whole blood.

Box plots of (A) CSH1 (B) KISS1 (C) PLAC4 (D) PLAC1 mRNA. The line inside each box denotes the median. The lower and upper limits denote the 25th and 75th percentiles, respectively. The lower and upper whiskers denote the 10th and 90th percentiles, respectively. Filled circles denote the outliers. Asterisks denote all the data below detection limit of 8 copies per reaction.

https://doi.org/10.1371/journal.pone.0005858.g002

Genotyping of PLAC4 and CSHL1

Third trimester placental tissues, maternal whole blood, buffy coat and plasma samples were collected. We aimed to compare the genotypes of the placental mRNA transcripts, PLAC4 and CSHL1, in maternal plasma and whole blood with the genomic genotypes of the fetal-maternal pairs to confirm if the circulating transcripts were truly of fetal origin. PLAC4 has a SNP, rs8130833, within the coding region [9]. CSHL1 was targeted because it demonstrated high placental tissue expression at similar levels like CSH1 which was highly homologous with other genes of the growth hormone cluster and that a target-specific RNA-SNP assay could not be developed [6].

The fetal and maternal genotypes for each pregnancy were first determined using placental tissue and maternal buffy coat DNA. Though nucleated fetal cells may be present at a density of 1 to 6 cells per milliliter of maternal blood in the samples [18], [19], such a level of “contaminant” would unlikely affect the predominant maternal genotype detected in maternal blood cells. Cases where the fetal and maternal genotypes differed were considered as informative and included 7 out of 28 cases for PLAC4 and 11 out of 50 cases for CSHL1. The PLAC4 and CSHL1 mRNA genotypes were then determined in maternal plasma and whole blood RNA samples of the informative pregnancies. Representative mass spectra for PLAC4 and CSHL1 RNA-SNP genotyping are shown in Figures S4 and S5. The genotyping data of all informative cases are summarized in Table 1. For PLAC4 mRNA, the genotype in maternal plasma corresponded to that of the fetus while the genotype in whole blood was always heterozygous indicating both fetal and maternal contributions. For CSHL1 mRNA, the genotypes in both the plasma and whole blood samples were identical to the fetus.

thumbnail
Table 1. PLAC4 and CSHL1 RNA-SNP genotyping in third trimester maternal plasma and whole blood.

https://doi.org/10.1371/journal.pone.0005858.t001

Gene expression profile comparison between pregnant and non-pregnant female blood samples

The gene expression profiles of five third trimester and five non-pregnant female whole blood samples were compared using GeneSpring® v 7.2, (Agilent Technologies) software. The data discussed in this publication have been deposited in NCBI's Gene Expression Omnibus [20] and are accessible through GEO Series accession number GSE14771 (http://www.ncbi.nlm.nih.gov/geo/query/acc.cgi?token=hfavtosmyeoygzq&acc=GSE14771). The final gene list was generated by selecting transcripts with 1.5-fold higher expression in the pregnant than non-pregnant blood samples and resulted in 247 gene transcripts. We sorted these genes in descending order of the fold difference (Table 2). Twenty of the top 25 gene transcripts were found to be related to blood cell functions and more specifically the functions of neutrophils (13/25) (Table 2). However, placental expressed transcripts previously identified, namely CSH1, CSHL1, KISS1, PLAC4 and PLAC1 were not present on the list. Defensin, alpha 4, corticostatin (DEFA4, NM_001925), carcinoembryonic antigen-related cell adhesion molecule 8 (CEACAM8, NM_001816), olflactomedin 4 (OLFM4, NM_006418), folliculin (FLCN, NM_144606), orosomucoid 1 (ORM1, NM_000607), matrix metallopeptidase 8 (neutrophil collagenase) (MMP8, NM_002424) and myeloperoxidase (MPO, NM_000250) which represented transcripts with large, medium and small differences in expression between pregnant and non-pregnant blood were further analyzed (Table 2).

thumbnail
Table 2. List of genes more highly expressed in whole blood of third trimester pregnant female than non-pregnant female.

https://doi.org/10.1371/journal.pone.0005858.t002

Quantification of the whole blood mRNA markers in third trimester whole blood and placental tissues.

We assessed if the pregnancy-related whole blood mRNA transcripts were likely to have originated from the placenta. Placental tissue and whole blood samples were collected from ten pregnancies. The median mRNA concentrations for DEFA4, CEACAM8, OLFM4, FLCN, ORM1, MMP8 and MPO were 0, 0, 3, 60, 0, 3 and 2 copies per nanogram placental RNA, respectively. Correspondingly, the median concentrations in whole blood samples were 125, 1181, 781, 207, 338, 386 and 200 copies per nanogram whole blood RNA in third trimester cases.

Concentrations of whole blood mRNA markers in pre- and post-delivery maternal whole blood.

After normalizing with GAPDH to adjust for the leukocyte concentration in whole blood, a statistically significant decreasing trend was observed in the post-delivery maternal blood samples for transcripts DEFA4, CEACAM8, OLFM4 and MMP8 comparing with the corresponding pre-delivery blood samples according to the Wilcoxon test (P<0.05). No significant difference was observed for FLCN (P = 1.00), ORM1 (P = 0.63) and MPO (P = 0.06) (Figure 3).

thumbnail
Figure 3. Quantitative analysis of identified whole blood mRNA markers in pre- and 24-hour-post-delivery maternal whole blood.

(A) DEFA4. (B) CEACAM8. (C) OLFM4. (D) FLCN. (E) ORM1. (F) MMP8. (G) MPO mRNA. Corresponding samples from each individual are connected by a line.

https://doi.org/10.1371/journal.pone.0005858.g003

Concentrations of whole blood mRNA markers in third trimester and non-pregnant whole blood.

When comparing the absolute concentrations of DEFA4, CEACAM8, OLFM4, FLCN, ORM1, MMP8 and MPO mRNA in whole blood samples from third trimester pregnant, non-pregnant female and male, statistical significant differences were observed in the markers (P<0.001, Kruskal-Wallis ANOVA test) except FLCN (P = 0.231, Kruskal-Wallis ANOVA test) (Figure 4).

thumbnail
Figure 4. Concentrations of identified whole blood mRNA in third trimester pregnant and non-pregnant whole blood.

Box plots of (A) DEFA4 (B) CEACAM8 (C) OLFM4 (D) FLCN (E) ORM1 (F) MMP8 (G) MPO mRNA. The line inside each box denotes the median. The lower and upper limits denote the 25th and 75th percentiles, respectively. The lower and upper whiskers denote the 10th and 90th percentiles, respectively. Filled circles denote the outliers. Asterisks indicate the groups with significant differences in mRNA concentrations of the marker (P<0.05, Kruskal-Wallis test followed by pairwise comparison using the Student-Neuman-Keuls test).

https://doi.org/10.1371/journal.pone.0005858.g004

However, after normalization with GAPDH to adjust for the leukocyte concentration in whole blood (data not shown), significant differences among third trimester pregnant, non-pregnant female and male whole blood were only found for transcripts OLFM4 and FLCN (P<0.05, Kruskal-Wallis ANOVA test).

Discussion

For decades, to develop noninvasive prenatal diagnosis, investigators have focused on the search for intact fetal cells in maternal blood [21], [22]. However, the rarity of such cells has hindered their robust identification in maternal blood [23]. The discoveries of cell-free fetal DNA [4] and later placental mRNA [5] in maternal plasma have offered new opportunities for the development of noninvasive prenatal diagnosis [9]. More recently, placental expressed and pregnancy-associated transcripts have been reported to be detectable in maternal whole blood [10][14]. As the RNA yield in whole blood far exceeds that of plasma, we explored in this study if maternal whole blood might represent a better specimen type for the detection of fetal specific placenta-derived mRNA.

We confirmed that placental expressed mRNA concentrations in third trimester whole blood were indeed higher than that in plasma for CSH1, KISS1, PLAC4 and PLAC1. The observation was consistent with previous reports [12], [14]. While the four transcripts were not detectable in all post-delivery maternal plasma samples, there were no significant reductions in PLAC4 and PLAC1 mRNA concentrations in maternal whole blood samples collected 24 h after delivery. In non-pregnant females and males, all four placental transcripts were absent in plasma but PLAC4 and PLAC1 could be detected in the whole blood of both non-pregnant females and males. These data suggested that the PLAC4 and PLAC1 mRNA in whole blood could be derived from tissues or organs other than the placenta. This could also explain the lack of clearance of PLAC4 and PLAC1 mRNA from maternal whole blood after delivery. This is further supported by the lack of statistical significant correlations between the whole blood and plasma signals for CSH1, KISS1, PLAC4 and PLAC1 suggesting that the mRNA species detected in the two specimen types possibly have originated from different tissue sources.

It is believed that fetal or placental RNA is released into the maternal circulation due to apoptosis of the trophoblast cells [5], [24]. If our targeted RNA transcripts were indeed placenta-derived, they should carry the same genotype as the fetus. The concordance between the placental tissue and plasma RNA-SNP genotyping results for PLAC4 and CSHL1 confirmed the fetal origin of such mRNA molecules in maternal plasma. As for maternal whole blood mRNA, the CSHL1 genotype was concordant with that of the fetus while the PLAC4 genotype was always heterozygous even when the fetus was homozygous (cases 1160 and 1359 in Table 1). This latter observation suggests that there is maternal contribution to the detectable PLAC4 mRNA in maternal whole blood. This finding confirms those reported by Go et al [25]. As can be appreciated from Table 1, pregnancies involving a homozygous fetus but a heterozygous mother (cases 1160 and 1359) are more useful for confirming the fetal specificity (i.e. absence of maternal transcript contamination) of the targeted transcript. The presence of a fetal-specific allele in maternal plasma (e.g. G-allele in case 668) merely suggests the existence of fetal contribution to the mRNA pool but does not indicate that there is no maternal transcript contamination. Yet, it is the absence of the maternal specific allele (e.g. G-allele in case 1160) that confirms the fetal-specificity of the transcript.

Concluding from our data, CSH1, CSHL1 and KISS1 appear to be pregnancy-specific in both plasma and whole blood. On the contrary, PLAC4 and PLAC1 are pregnancy-specific only in plasma but not whole blood which has both fetal and maternal contributions. Thus, it is not surprising that Banzola et al reported a lack of quantitative difference in whole blood PLAC4 (located on chromosome 21) mRNA between euploid and trisomy 21 pregnancies [14]. Our data showed that there was no statistical significant difference in whole blood PLAC4 mRNA concentrations even between pregnant and non-pregnant individuals. One possible explanation of the maternal contribution to these apparently “placenta-derived” transcripts in whole blood may be related to the phenomenon of illegitimate expression that lymphocytes had been shown to express transcripts irrelevant to their functions [26][28]. Microarray analyses do not reveal these expression profiles because microarray expression data are often presented in a relative scale where the extremely low abundance signals will be dwarfed by the high abundance signals. These low abundance signals in blood cells may become a significant contaminant in terms of absolute quantities when whole blood is the biological sample analyzed.

Nonetheless, CSH1, CSHL1 and KISS1 appeared to be pregnancy-specific in whole blood. Thus, we systematically investigated if other pregnancy-specific whole blood mRNA markers could be developed based on microarray comparison of pregnant and non-pregnant whole blood samples. We found 247 genes which showed at least 1.5-fold elevation in the former than the latter group. Thirteen of the top 25 transcripts on the list were related to blood cell functions, especially that of neutrophils. This is not surprising as pregnancy has been reported to induce inflammatory change in peripheral blood and neutrophils are major contributors to inflammation [29], [30]. Unexpectedly, none of the previously studied placenta-derived transcripts, CSH1, KISS1, PLAC4 and PLAC1, appeared on the gene list. This was also the case as reported by Maron et al [13]. According to our QRT-PCR data on the pregnancy-related whole blood markers such as DEFA4, the median expression levels of these genes were much higher than that of the placenta-derived transcripts in whole blood (Figures 2 and 4). Therefore, it is possible that CSH1, CSHL1 and KISS1 were ranked much lower than 247. Furthermore, as PLAC4 and PLAC1 were readily detectable in both pregnant and non-pregnant whole blood and possibly without the 1.5-fold threshold difference in expression, they would not have appeared on the sorted gene list.

We have validated the pregnancy-specificity of 7 transcripts with differences in expression between pregnant and non-pregnant whole blood as determined by the microarray analysis. DEFA4 is involved in host defense and found in neutrophils [31]. CEACAM8 is expressed only in neutrophils and eosinophils in humans with undetermined function [32]. OLFM4 is previously recognized as human granulocyte colony stimulating factor stimulated clone-1 with high expression in cancerous tissues [33]. FLCN is speculated to be a tumor suppressor gene with unclear function [34]. ORM1 encodes a key acute phase plasma protein [35]. During pregnancy, ORM1 is suggested to be involved in the maintenance of homeostasis between maternal and fetal systems within the chorioallantoic placenta [36]. MMP8 is expressed in diverse cell types including neutrophils, macrophages, T cells, epithelial cells and endothelial cells in inflammatory conditions [37], [38]. MMP8 was also detected in the chorion during labor and was proposed to be a predictive marker for preterm delivery [39], [40]. MPO is a heme protein primarily hosted in human polymorphonuclear neutrophils [41]. During pregnancy, MPO was observed on the surface of neutrophils [42].

Our QRT-PCR validation of the whole blood markers showed a statistically significant difference between the expression levels among third trimester pregnant, non-pregnant female and male individuals for OLFM4 and FLCN only after normalization with the corresponding GAPDH data. All markers, except FLCN, were related to the hematopoietic system, especially the neutrophils. It has also been reported that the total leukocyte count would rise during pregnancy with neutrophils accounting for most of the increased leukocyte count [43], [44]. Hence, the increased expression level observed from the microarray data was probably due to the augmented leukocyte number instead of the transcript expression level in the cells. This may also explain why we found a significant difference in concentrations of those transcripts only when directly comparing the pregnant and non-pregnant whole blood samples but not after GAPDH normalization.

The whole blood microarray study identified transcripts whose functions were related to the physiological changes of pregnancy. However, their presence in non-pregnant whole blood and lack of clearance after pregnancy suggested that they were not “pregnancy-specific” markers, albeit being “pregnancy-related”. As they were also detectable in blood of non-pregnant individuals, they were therefore not fetal-specific either. Concluding from all the data in this study, it appears that placenta-derived fetal-specific transcripts can be more readily identified from maternal plasma than whole blood. While some transcripts are fetal-specific in maternal whole blood, e.g. CSHL1, due care is needed to validate each new potential whole blood transcript that is meant to be used as a circulating fetal RNA marker. Illegitimate expression by maternal blood cells would need to be excluded before adopting such a marker.

Supporting Information

Figure S1.

Bar-charts adopted from Human GeneAtlas GNF1H showing expression of (A) CSH1 (208356_x_at) and (B) CSHL1 (205958_x_at) in different human tissues. (http://symatlas.gnf.org/SymAtlas)

https://doi.org/10.1371/journal.pone.0005858.s001

(2.28 MB TIF)

Figure S2.

Bar-charts adopted from Human GeneAtlas GNF1H showing expression of KISS1 (205563_at) in different human tissues. (http://symatlas.gnf.org/SymAtlas)

https://doi.org/10.1371/journal.pone.0005858.s002

(0.17 MB TIF)

Figure S3.

Bar-charts adopted from Human GeneAtlas GNF1H showing expression of (A) PLAC4 (214750_at) and (B) PLAC1 (219702_at) in different human tissues. (http://symatlas.gnf.org/SymAtlas)

https://doi.org/10.1371/journal.pone.0005858.s003

(2.60 MB TIF)

Figure S4.

Mass spectra showing PLAC4 RNA-SNP genotypes of placenta, third trimester maternal plasma and whole blood samples. (A) a homozygous placental RNA sample showing a single peak A with unextended primers (UEP). (B) third trimester maternal plasma RNA showing a single peak A as the placental RNA sample. (C) third trimester maternal whole blood RNA showing a peak A and a minor peak G as indicated by the arrow.

https://doi.org/10.1371/journal.pone.0005858.s004

(1.09 MB TIF)

Figure S5.

Mass spectra showing CSHL1 RNA-SNP genotypes of placenta, third trimester maternal plasma and whole blood samples. (A) a heterozygous placental RNA sample showing 2 peaks, allele C and allele T with unextended primers (UEP). (B) third trimester maternal plasma RNA also showing 2 peaks as the placental RNA sample. (C) third trimester maternal whole blood RNA showing both peaks of allele C and allele T.

https://doi.org/10.1371/journal.pone.0005858.s005

(1.07 MB TIF)

Table S1.

Primer, probe and calibrator sequences and reaction conditions for quantitative real-time RT-PCR assays.

https://doi.org/10.1371/journal.pone.0005858.s006

(0.08 MB DOC)

Table S2.

Primer sequences for reverse transcription and PCR amplification of the PLAC4 and CSHL1 SNP.

https://doi.org/10.1371/journal.pone.0005858.s007

(0.03 MB DOC)

Table S3.

Sequences and molecular weights of the extension primer and the expected extension products for each of the alleles of the PLAC4 and CSHL1 SNPs.

https://doi.org/10.1371/journal.pone.0005858.s008

(0.03 MB DOC)

Author Contributions

Conceived and designed the experiments: CD RWKC YMDL. Performed the experiments: MMSH SJ NBYT. Analyzed the data: MMSH SJ NBYT RWKC YMDL. Contributed reagents/materials/analysis tools: TYL TKL. Wrote the paper: MMSH RWKC YMDL. Recruitment of study participants: TYL TKL.

References

  1. 1. Wright CF, Burton H (2009) The use of cell-free fetal nucleic acids in maternal blood for non-invasive prenatal diagnosis. Hum Reprod Update 15: 139–151.
  2. 2. Chiu RWK, Chan KCA, Gao Y, Lau VYM, Zheng W, et al. (2008) Noninvasive prenatal diagnosis of fetal chromosomal aneuploidy by massively parallel genomic sequencing of DNA in maternal plasma. Proc Natl Acad Sci U S A 105: 20458–20463.
  3. 3. Lun FMF, Tsui NBY, Chan KCA, Leung TY, Lau TK, et al. (2008) Noninvasive prenatal diagnosis of monogenic diseases by digital size selection and relative mutation dosage on DNA in maternal plasma. Proc Natl Acad Sci U S A 105: 19920–19925.
  4. 4. Lo YMD, Corbetta N, Chamberlain PF, Rai V, Sargent IL, et al. (1997) Presence of fetal DNA in maternal plasma and serum. Lancet 350: 485–487.
  5. 5. Ng EKO, Tsui NBY, Lau TK, Leung TN, Chiu RWK, et al. (2003) mRNA of placental origin is readily detectable in maternal plasma. Proc Natl Acad Sci U S A 100: 4748–4753.
  6. 6. Tsui NBY, Chim SSC, Chiu RWK, Lau TK, Ng EKO, et al. (2004) Systematic micro-array based identification of placental mRNA in maternal plasma: towards non-invasive prenatal gene expression profiling. J Med Genet 41: 461–467.
  7. 7. Farina A, Chan CWM, Chiu RWK, Tsui NBY, Carinci P, et al. (2004) Circulating corticotropin-releasing hormone mRNA in maternal plasma: relationship with gestational age and severity of preeclampsia. Clin Chem 50: 1851–1854.
  8. 8. Ng EKO, Leung TN, Tsui NBY, Lau TK, Panesar NS, et al. (2003) The concentration of circulating corticotropin-releasing hormone mRNA in maternal plasma is increased in preeclampsia. Clin Chem 49: 727–731.
  9. 9. Lo YMD, Tsui NBY, Chiu RWK, Lau TK, Leung TN, et al. (2007) Plasma placental RNA allelic ratio permits noninvasive prenatal chromosomal aneuploidy detection. Nat Med 13: 218–223.
  10. 10. Concu M, Banzola I, Farina A, Sekizawa A, Rizzo N, et al. (2005) Rapid clearance of mRNA for PLAC1 gene in maternal blood after delivery. Fetal Diagn Ther 20: 27–30.
  11. 11. Farina A, Rizzo N, Concu M, Banzola I, Sekizawa A, et al. (2005) Lower maternal PLAC1 mRNA in pregnancies complicated with vaginal bleeding (threatened abortion <20 weeks) and a surviving fetus. Clin Chem 51: 224–227.
  12. 12. Okazaki S, Sekizawa A, Purwosunu Y, Iwasaki M, Farina A, et al. (2006) Measurement of mRNA of trophoblast-specific genes in cellular and plasma components of maternal blood. J Med Genet 43: e47.
  13. 13. Maron JL, Johnson KL, Slonim D, Lai CQ, Ramoni M, et al. (2007) Gene expression analysis in pregnant women and their infants identifies unique fetal biomarkers that circulate in maternal blood. J Clin Invest 117: 3007–3019.
  14. 14. Banzola I, Rusterholz C, Zannoni L, Rizzo N, Zhong XY, et al. (2008) PLAC4 and beta-HCG mRNA levels are not altered in the maternal circulation of pregnancies with trisomy 21. Prenat Diagn 28: 1262–1267.
  15. 15. Su AI, Wiltshire T, Batalov S, Lapp H, Ching KA, et al. (2004) A gene atlas of the mouse and human protein-encoding transcriptomes. Proc Natl Acad Sci U S A 101: 6062–6067.
  16. 16. Chiu RWK, Poon LLM, Lau TK, Leung TN, Wong EMC, et al. (2001) Effects of blood-processing protocols on fetal and total DNA quantification in maternal plasma. Clin Chem 47: 1607–1613.
  17. 17. Affymetrix (2003) Globin reduction protocol: a method for processing whole blood RNA samples for improved array results. Affymetrix Technical Note 1–9.
  18. 18. Bianchi DW (1999) Fetal cells in the maternal circulation: feasibility for prenatal diagnosis. Br J Haematol 105: 574–583.
  19. 19. Krabchi K, Gros-Louis F, Yan J, Bronsard M, Masse J, et al. (2001) Quantification of all fetal nucleated cells in maternal blood between the 18th and 22nd weeks of pregnancy using molecular cytogenetic techniques. Clin Genet 60: 145–150.
  20. 20. Edgar R, Domrachev M, Lash AE (2002) Gene Expression Omnibus: NCBI gene expression and hybridization array data repository. Nucleic Acids Res 30: 207–210.
  21. 21. Brown ES (1963) Foetal erythrocytes in the maternal circulation. Br Med J 1: 1000–1001.
  22. 22. Mavrou A, Kouvidi E, Antsaklis A, Souka A, Kitsiou Tzeli S, et al. (2007) Identification of nucleated red blood cells in maternal circulation: a second step in screening for fetal aneuploidies and pregnancy complications. Prenat Diagn 27: 150–153.
  23. 23. Bianchi DW, Simpson JL, Jackson LG, Elias S, Holzgreve W, et al. (2002) Fetal gender and aneuploidy detection using fetal cells in maternal blood: analysis of NIFTY I data. National Institute of Child Health and Development Fetal Cell Isolation Study. Prenat Diagn 22: 609–615.
  24. 24. Halicka HD, Bedner E, Darzynkiewicz Z (2000) Segregation of RNA and separate packaging of DNA and RNA in apoptotic bodies during apoptosis. Exp Cell Res 260: 248–256.
  25. 25. Go AT, Visser A, Mulders MA, Blankenstein MA, van Vugt JM, et al. (2007) 44 single-nucleotide polymorphisms expressed by placental RNA: assessment for use in noninvasive prenatal diagnosis of trisomy 21. Clin Chem 53: 2223–2224.
  26. 26. Kimoto Y (1998) A single human cell expresses all messenger ribonucleic acids: the arrow of time in a cell. Mol Gen Genet 258: 233–239.
  27. 27. Ko Y, Grunewald E, Totzke G, Klinz M, Fronhoffs S, et al. (2000) High percentage of false-positive results of cytokeratin 19 RT-PCR in blood: a model for the analysis of illegitimate gene expression. Oncology 59: 81–88.
  28. 28. Gala JL, Heusterspreute M, Loric S, Hanon F, Tombal B, et al. (1998) Expression of prostate-specific antigen and prostate-specific membrane antigen transcripts in blood cells: implications for the detection of hematogenous prostate cells and standardization. Clin Chem 44: 472–481.
  29. 29. Germain SJ, Sacks GP, Soorana SR, Sargent IL, Redman CW (2007) Systemic inflammatory priming in normal pregnancy and preeclampsia: the role of circulating syncytiotrophoblast microparticles. J Immunol 178: 5949–5956.
  30. 30. Sacks GP, Studena K, Sargent K, Redman CW (1998) Normal pregnancy and preeclampsia both produce inflammatory changes in peripheral blood leukocytes akin to those of sepsis. Am J Obstet Gynecol 179: 80–86.
  31. 31. Schneider JJ, Unholzer A, Schaller M, Schafer-Korting M, Korting HC (2005) Human defensins. J Mol Med 83: 587–595.
  32. 32. Zhao L, Xu S, Fjaertoft G, Pauksen K, Hakansson L, et al. (2004) An enzyme-linked immunosorbent assay for human carcinoembryonic antigen-related cell adhesion molecule 8, a biological marker of granulocyte activities in vivo. J Immunol Methods 293: 207–214.
  33. 33. Koshida S, Kobayashi D, Moriai R, Tsuji N, Watanabe N (2007) Specific overexpression of OLFM4(GW112/HGC-1) mRNA in colon, breast and lung cancer tissues detected using quantitative analysis. Cancer Sci 98: 315–320.
  34. 34. Adley BP, Smith ND, Nayar R, Yang XJ (2006) Birt-Hogg-Dube syndrome: clinicopathologic findings and genetic alterations. Arch Pathol Lab Med 130: 1865–1870.
  35. 35. Fournier T, Medjoubi NN, Porquet D (2000) Alpha-1-acid glycoprotein. Biochim Biophys Acta 1482: 157–171.
  36. 36. Thomas T (1993) Distribution of alpha 2-macroglobulin and alpha 1-acid glycoprotein mRNA shows regional specialization in rat decidua. Placenta 14: 417–428.
  37. 37. Hanemaaijer R, Sorsa T, Konttinen YT, Ding Y, Sutinen M, et al. (1997) Matrix metalloproteinase-8 is expressed in rheumatoid synovial fibroblasts and endothelial cells. Regulation by tumor necrosis factor-alpha and doxycycline. J Biol Chem 272: 31504–31509.
  38. 38. Prikk K, Maisi P, Pirila E, Sepper R, Salo T, et al. (2001) In vivo collagenase-2 (MMP-8) expression by human bronchial epithelial cells and monocytes/macrophages in bronchiectasis. J Pathol 194: 232–238.
  39. 39. Nien JK, Yoon BH, Espinoza J, Kusanovic JP, Erez O, et al. (2006) A rapid MMP-8 bedside test for the detection of intra-amniotic inflammation identifies patients at risk for imminent preterm delivery. Am J Obstet Gynecol 195: 1025–1030.
  40. 40. Arechavaleta-Velasco F, Marciano D, Diaz-Cueto L, Parry S (2004) Matrix metalloproteinase-8 is expressed in human chorion during labor. Am J Obstet Gynecol 190: 843–850.
  41. 41. Klebanoff SJ (1999) Myeloperoxidase. Proc Assoc Am Physicians 111: 383–389.
  42. 42. Kindzelskii AL, Clark AJ, Espinoza J, Maeda N, Aratani Y, et al. (2006) Myeloperoxidase accumulates at the neutrophil surface and enhances cell metabolism and oxidant release during pregnancy. Eur J Immunol 36: 1619–1628.
  43. 43. Pitkin RM, Witte DL (1979) Platelet and leukocyte counts in pregnancy. Jama 242: 2696–2698.
  44. 44. Kuhnert M, Strohmeier R, Stegmuller M, Halberstadt E (1998) Changes in lymphocyte subsets during normal pregnancy. Eur J Obstet Gynecol Reprod Biol 76: 147–151.