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

Molecular Cloning and Characterization of Two Genes Encoding Dihydroflavonol-4-Reductase from Populus trichocarpa

  • Yan Huang ,

    Contributed equally to this work with: Yan Huang, Jiqing Gou

    Affiliations Key Laboratory of Eco-Environments of Three Gorges Reservoir Region, Ministry of Education, Institute of Resources Botany, School of Life Sciences, Southwest University, Chongqing, China, Chongqing Key Laboratory of Transgenic Plant and Safety Control, Southwest University, Chongqing, China

  • Jiqing Gou ,

    Contributed equally to this work with: Yan Huang, Jiqing Gou

    Affiliation Forage Improvement Division, The Samuel Roberts Noble Foundation, Ardmore, Oklahoma, United States of America

  • Zhichun Jia,

    Affiliation Key Laboratory of Eco-Environments of Three Gorges Reservoir Region, Ministry of Education, Institute of Resources Botany, School of Life Sciences, Southwest University, Chongqing, China

  • Li Yang,

    Affiliation Key Laboratory of Eco-Environments of Three Gorges Reservoir Region, Ministry of Education, Institute of Resources Botany, School of Life Sciences, Southwest University, Chongqing, China

  • Yimin Sun,

    Affiliation Key Laboratory of Eco-Environments of Three Gorges Reservoir Region, Ministry of Education, Institute of Resources Botany, School of Life Sciences, Southwest University, Chongqing, China

  • Xunyan Xiao,

    Affiliation Key Laboratory of Eco-Environments of Three Gorges Reservoir Region, Ministry of Education, Institute of Resources Botany, School of Life Sciences, Southwest University, Chongqing, China

  • Feng Song,

    Affiliation Key Laboratory of Eco-Environments of Three Gorges Reservoir Region, Ministry of Education, Institute of Resources Botany, School of Life Sciences, Southwest University, Chongqing, China

  • Keming Luo

    kemingl@swu.edu.cn

    Affiliations Key Laboratory of Eco-Environments of Three Gorges Reservoir Region, Ministry of Education, Institute of Resources Botany, School of Life Sciences, Southwest University, Chongqing, China, Chongqing Key Laboratory of Transgenic Plant and Safety Control, Southwest University, Chongqing, China

Abstract

Dihydroflavonol 4-reductase (DFR, EC 1.1.1.219) is a rate-limited enzyme in the biosynthesis of anthocyanins and condensed tannins (proanthocyanidins) that catalyzes the reduction of dihydroflavonols to leucoanthocyanins. In this study, two full-length transcripts encoding for PtrDFR1 and PtrDFR2 were isolated from Populus trichocarpa. Sequence alignment of the two PtrDFRs with other known DFRs reveals the homology of these genes. The expression profile of PtrDFRs was investigated in various tissues of P. trichocarpa. To determine their functions, two PtrDFRs were overexpressed in tobacco (Nicotiana tabacum) via Agrobacterium-mediated transformation. The associated color change in the flowers was observed in all 35S:PtrDFR1 lines, but not in 35S:PtrDFR2 lines. Compared to the wild-type control, a significantly higher accumulation of anthocyanins was detected in transgenic plants harboring the PtrDFR1. Furthermore, overexpressing PtrDFR1 in Chinese white poplar (P. tomentosa Carr.) resulted in a higher accumulation of both anthocyanins and condensed tannins, whereas constitutively expressing PtrDFR2 only improved condensed tannin accumulation, indicating the potential regulation of condensed tannins by PtrDFR2 in the biosynthetic pathway in poplars.

Introduction

Flavonoids, very important secondary metabolites, exist widely throughout the plant kingdom. Presently, over 8,000 different compounds of flavonoids have been identified, many of which are involved in several biological processes, such as pigmentation of flowers, protection against UV-B injury, defense against pathogens and pests, pollen viability, auxin transport regulation, etc. [1], [2]. Flavonoids are divided into several structural subclasses, including flavanones, isoflavonoids, anthocyanins, flavonols, catechins, flavones and proanthocyanidins (also called condensed tannins or CTs), and are abundant in fruits, leaves and flowers [3]. The biosynthetic pathways of flavonoids are well established in plants [1], [4], [5].

In the flavonoid biosynthetic pathway, dihydroflavonol-4-reductase (DFR, EC 1.1.1.219) is one of the rate-limited enzymes that catalyzes the stereospecific reduction of three dihydroflavonols (dihydrokaempferol, dihydroquercetin and dihydromyricetin) to leucoanthocyanidins (flavan-3,4-diols) using NADPH as a cofactor [2], [6], [7]. Leucoanthocyanidins, the precursors of the anthocyanin branch, are essential for the formation of CTs [8]. Previous studies demonstrated that deactivation of the DFR gene resulted in the loss of anthocyanins and CTs in mutants of barley and Arabidopsis [9], [10]. Due to their crucial role in the flavonoid pathway, various DFR genes have been isolated from different species, such as A. thaliana, maize (Zea mays), barley (Hordeum vulgare), trembling aspen (Populus tremuloides), Medicago truncatula and Petunia hybrida [8], [11][15]. Variable DFR genes were therefore found in various genomes, i.e., a single copy DFR is present in A. thaliana, barley, tomato (Lycopersicon esculentum), grape (Vitis vinifera), snapdragon (Antirrhinum majus) and rice (Oryza sativa), while multi-copy DFRs exist in P. hybrida (Line V30), Ipomoea purpurea, I. nil and M. truncatula [7], [8], [11], [16][18].

Poplar (Populus spp.), as a model plant in trees, is extensively used in studies of tree morphology, physiology, biochemistry, ecology, genetics and molecular biology [19], [20]. A wide range of genomic and genetic resources is now available in the species, including an EST database [21], genome sequence and annotation from P. trichocarpa [22]. Previously, it was found that two duplicated PtrDFR genes are present in the P. trichocarpa genome [23]. Furthermore, a DFR mRNA was isolated from P. tremuloides, and its expression induced wounded leaves in aspen [7]. Phytochemical assays revealed that CT concentrations were significantly increased in wounded leaves. It suggested that CT synthesis is an inducible defense response in trembling aspen. However, the DFRs' functions in the biosynthetic pathway of other flavonoids, such as anthocyanidins, remain unclear.

In this study, we isolated two full-length mRNAs of PtrDFR1 and PtrDFR2, which encode DFR isoenzymes from P. trichocarpa. The expression profiles of PtrDFRs were investigated in various tissues of P. trichocarpa. Both PtrDFR1 and PtrDFR2 were overexpressed in tobacco (Nicotiana tabacum cv Xanthi) and Chinese white poplar (P. trichocarpa Carr.), and their potential biological functions in the flavonoid biosynthetic pathways were investigated.

Materials and Methods

Plant materials and bacterial strains

Poplar plants were grown in the greenhouse at 26°C under 14/10h photoperiod with supplemental light (reached 600 footcandles). Individual tissues, including leaf, stem, root and petiole, were harvested separately and frozen in liquid nitrogen until further processing. Transgenic tobacco (N. tabacum cv Xanthi) and P. tomentosa Carr. (clone 73) were grown under the same conditions.

Escherichia coli strain DH5α was used as the recipient for transformation, genetic manipulation and production of plasmid DNA for sequencing. Agrobacterium tumefaciens strain LBA4404 was used as the disarmed vector for the transformation of tobacco and poplar.

Isolation of total RNA and cloning of PtrDFR cDNAs from P. trichocarpa

Total RNA was isolated from frozen tissues of poplar (P. trichocarpa) plants using RNeasy Plant Mini Kit (Qiagen, Germany) following the manufacturer's instructions with a modification as we reported previously [24]. The leaves and petioles were excised from stems and included the fourth (young) and fifth (old) internodes from the top of the stems (1 m height). Samples from at least five stems were pooled for analysis. First-strand cDNA was synthesized from 2 µg of DNase-treated RNA with RT-AMV transcriptase (TaKaRa, Dalian, China) in a total volume of 20 µL by using oligo (dT) at 42°C for 30 min. The ORF of PtrDFRs was obtained by reverse transcription-PCR (RT-PCR) with 2 µl of cDNA from the roots, by using PtrDFR1-F: 5′-GCGAGCTCGATGGGAACAGAAGCTGAAAC-3′ (as forward primer, the SacI site is in italics) and PtrDFR1-R: 5′-GCGGATCCGTGGAACAATCAGGACGCAG-3′ (as reverse primer, the BamHI site is in italics) for PtrDFR1, and using PtrDFR2-F: 5′-GCGAGCTCCATGGGAGTAGAAGTTGAAAC-3′ (as forward primer, the SacI site is in italics) and PtrDFR2-R: 5′-GCGGATCCAAACTAAAGGGCCTCAGAATC-3′ (as reverse primer, the BamHI site is in italics) for PtrDFR2. The PCR reaction was carried out with Pfu DNA polymerase (TaKaRa, Dalian, China) in a total volume of 50 µl at 94°C for 5 min; 35 cycles of 94°C for 30 s; 56°C for 30 s; and 72°C for 2 min; followed by a final extension of 72°C for 7 min. The A tail was added to the PCR product, which was then cloned into pGEM-T easy vector according to the manufacturer's instructions (Promega), and the correct reading frame of the resulting construct was confirmed by sequencing. The PCR product was cloned into the SacI and BamHI sites of the plant binary vector pBI121. The resulting vectors, p35S:PtrDFR1 and p35S:PtrDFR2, with PtrDFR ORFs driven by the cauliflower mosaic virus 35S (CaMV 35S) promoter and the NPTII gene as a plant-selectable marker conferring kanamycin resistance were transferred into A. tumefaciens strain LBA4404 by the freeze-thaw method.

Transformation of tobacco plants

For tobacco transformation, Agrobacterium strain LBA4404 containing the binary vector was incubated in liquid YEP medium supplemented with 200 µM acetone-syringone [25], [26] at 28°C with constant shaking (200 rpm) until the culture reached an optimal density of approximately 0.6–0.8 at 600 nm. The A. tumefaciens culture was then diluted with an equal volume of liquid Murashige and Skoog medium [27]. Leaf discs from N. tabacum Xanthi were transformed as described previously [28]. Transformed plants were grown on Murashige and Skoog medium with 100 mg l−1 kanamycin under long day conditions (18/6 h photoperiod) at 25°C. Only one bud was picked from each explant to ensure independent transformants. Antibiotic-resistant plants were maintained as transgenic lines, and plantlets were transplanted to soil.

Transformation of poplar plants

Transgenic Chinese white poplar (P. tomentosa Carr. clone 73) plants were generated by Agrobacterium-mediated transformation as described previously [29]. Recombinant Agrobacterium was used to infect poplar leaf discs, and putative transgenic plants were grown and selected on woody plant medium (WPM) [30] with 100 mg l−1 kanamycin. Rooted plantlets were acclimatized in pots at 25°C in a 16/8 h photoperiod and then transferred to the greenhouse for further studies.

DNA extraction and PCR analysis

Genomic DNA was extracted from leaves (300 mg) of untransformed control and hygromycin-resistant plants using the modified CTAB extraction method as previously described [28]. To determine the presence of transgenes, putative transgenic plants were screened preliminarily by PCR analysis [31]. The following primers were designed for the NPTII gene – forward primer: 5′-AGGCTATTCGGCTATGACTGG-3′; reverse primer: 5′-TCGGGAGCGGCGATA CCGTA-3′. PCR conditions were 94°C for 3 min; 94°C for 30 s; 56°C for 30 s; and 72°C for 1 min–34 cycles in total. The PtrDFR1-F, PtrDFR1-R, PtrDFR2-F and PtrDFR2-R as described above were used for amplification of PtrDFR1 and PtrDFR2, respectively. PCR amplification was performed by using a thermocycler (Eppendorf AG, Germany). Amplified DNA was loaded on 0.8% (w/v) agarose gel and visualized after ethidium bromide staining.

RT-PCR analysis

To determine the presence of the transgenes in transgenic tobacco, total RNA was extracted from wild-type and transgenic plants using RNeasy Plant Mini Kit (Qiagen, Germany). For RT-PCR, DNase-treated RNA (2 µg) was then reverse transcribed in a total volume of 20 µl by using RT-AMV transcriptase (TaKaRa, Dalian, China) with oligo (dT) at 42°C for 30 min. Primers for Actin, which was used as an internal reference, were F: 5′-TGGACTCTGGTGATGGTGTC-3′ and R: 5′-CCTCCAATCCAAACACTGTA-3′. The number of cycles for each gene was varied to confirm that the PCR amplification was in the linear range as we described previously [32]. To detect PtrDFR expression, amplification was conducted for 26 cycles, with each cycle consisting of 94°C for 1 min; 58°C for 30 s; 72°C for 2 min; and finally 7 min at 72°C. Aliquots of individual PCR products were compared by agarose gel electrophoresis and visualized with ethidium bromide under UV light.

Quantitative real-time PCR

Total RNA extracted from leaves, roots, stems and petioles of poplar plants at different developmental stages was treated with DNase I (TaKaRa, Dalian, China) according to the manufacturer's instructions. All RNA was purified and first-strand cDNA synthesized as described above. The reverse transcribed cDNA samples were used for real-time PCR, which was performed on a BioRad IQ5 real-time PCR detection system. The forward and reverse primers for PtrDFR1 amplification were qtDFR1-F (5′-TACAATGTCCCTGCTAAGTTC-3′) and qtDFR1-R (5′-GTGGAACAATCAGGACGCAG-3′), and primers for PtrDFR2 amplification were qtDFR2-F (5′-TACAGCTTGGAGGAAATGTTC-3′) and qtDFR2-R (5′-AAACTAAAGGGCCTCAGAATC-3′). The efficiency of these primers was investigated by applying primer melting curve analysis and gel electrophoresis; both results indicated that each primer pair gave a specific and unique PCR product. A Populus Actin gene, amplified with the primers Actin-F (5′-GTGCTTCTAAGTTCCGAACAGTGC-3′) and Actin-R (5′-GACTACCAAAGTGTCTGACCACCA-3′) giving a product of 180 bp, was used as a reference for loading normalization. Quantitative real-time PCR reaction and data analysis were performed as described by Tsai et al. [22] in a 20-µl reaction volume containing 10 µl of SYBR Green Master Mix reagent (TaKaRa, Dalian, China). Each experiment was performed in duplicate and with three biological replicates along with no-template controls. Statistical differences in expression between the mean values of the control and tested samples were analyzed by one-way analysis of variance (ANOVA) and the LSD test.

Sequence comparisons and phylogenetic analysis

The nucleotide and deduced amino acid sequences were analyzed by DNAstar programs and DNAMAN software (Lynnon Corporation, USA). Amino acid sequence alignments were performed with DNAMAN software. The phylogenetic tree of DFRs was constructed by the neighbor-joining method for P. trichocarpa and other plants using DNAMAN software.

Anthocyanin measurement

Anthocyanin quantification was performed as described by Pang et al. [33]. The poplar tissues were ground in liquid nitrogen and placed into 5 mL extraction buffer (an equal volume of methanol:0.1% HCl), sonicated for 1 h and then shaken in darkness for 4 h at 120 rpm. After centrifugation at 2,500 g for 10 min, 1 mL of water was added to 1 mL of extract, followed by the addition of 1 mL of chloroform to remove chlorophyll. The absorption of the extracts was measured spectrophotometrically at 530 nm. The amount of anthocyanins was reported as (A530) g−1 fresh weight (FW). The experiment was repeated three times for each treatment.

Extraction and quantification of condensed tannins

Quantification of total condensed tannins from wild-type and transgenic poplar was performed using the vanillin-HCl method as described previously [34][36]. Leaf tissues were ground in liquid nitrogen and extracted with 5 mL of extraction solution (a 0.5% (v/v) vanillin solution in methanol containing 4% HCl). Following centrifugation at 2,500 g for 10 min, the residues were re-extracted twice, as above. Pooled supernatants were incubated for 20 min at room temperature. Samples, blanks and standards were read at 500 nm using a UV/Vis spectrophotometer, zeroing the spectrophotometer with deionized water. Blanks were subtracted from samples and condensed tannin content calculated as catechin equivalents. The concentration of condensed tannins was detected in triplicate.

Results and Discussion

Isolation and characterization of two DFR genes from Populus trichocarpa

The Populus genome was recently sequenced [21]. Based on the sequences deposited in the Populus genome database, bioinformatic analysis indicated that DFR was encoded by two genes [22] named PtrDFR1 and PtrDFR2 which consist of six exons and five introns (Figure 1A). The PtrDFR1 cDNA contains 1,375 nucleotides, including a full-length open reading frame (ORF) encoding 346 amino acids, a 71-bp 5′ untranslated region (UTR) and a 263-bp 3′ UTR (GenBank Accession No. XM_002300723). The PtrDFR2 cDNA contains a truncated ORF encoding 336 amino acids (GenBank Accession No. XM_002307631). Based on the alignment of the complete genomic sequence in the public database (JGI, http://genome.jgi-psf.org/cgi-bin/), a putative full-length ORF of PtrDFR2, encoding 376 amino acids, was obtained. Based on this sequence information, specific primers were designed to amplify the full-length mRNAs that encode for PtrDFR1 and PtrDFR2 (see Materials and Methods for more details).

thumbnail
Figure 1. Nucleotide sequence analysis of PtrDFRs and their protein sequence alignments with other DFRs.

(A) A schematic representation of the exon and intron organization of PtrDFRs. PtrDFR1 consists of six exons (black boxes) and five introns (intervening line) with a 71-bp 5′ untranslated region (UTR; white box) and a 263-bp 3′ untranslated region. PtrDFR2 also consists of six exons (black boxes) and five introns (intervening line). (B) The amino acid sequence and structure of PtrDFRs. The boxed region represents a putative NADPH binding domain at the N terminus of PtrDFR proteins. GenBank accession numbers are as follows (in parentheses): PtrDFR1 (XM_002300723), PtrDFR2 (XM_002307631), PtDFR (AY147903), AtDFR (AAA32783), MtDFR1 (AY389346), MtDFR1 (AY389347) and OsDFR (BAA36182).

https://doi.org/10.1371/journal.pone.0030364.g001

In a previous study, a cDNA clone (PtDFR) encoding dihydroflavonol-4-reductase, whose expression is induced by herbivory attack, was isolated from trembling aspen [7]. Sequence comparison revealed that PtDFR shared a much higher amino acid identity with PtrDFR1 (96.3%) than with PtrDFR2 (81.5%), indicating that the PtDFR gene may encode a DFR1 protein. DFR genes have also been cloned from other plant species, such as A. thaliana (AtDFR, GenBank Accession No. AAA32783), M. truncatula (MtDFR1, GenBank Accession No. AY389346) and Oryza sativa (OsDFR, GenBank Accession No. BAA36182). Amino acid sequence alignments showed that PtrDFR1 shared 71.4%, 71.3% and 62.0% identity with AtDFR, MtDFR1 and OsDFR, respectively (Figure 1B). The DFR enzyme catalyzes the NADPH-dependent reduction of 2R,3R-trans-dihydroflavonols to leucoanthocyanidins in the flavonoid biosynthetic pathway [37]. A putative NADP binding site (aa 10–30, VTGASGFIGSWLI/VMRLLEKGY) with very high sequence similarity with other DFRs [38] was also present in the same region (near the N-terminus) of the poplar DFR amino acid sequences (Figure 1B).

To further investigate the sequence homology of PtrDFRs to other known DFRs, a rooted phylogenetic tree was constructed by using the predicated amino acid sequences from 20 species. As shown in Figure 2, these DFR proteins were clustered into two distinct groups. Only GbDFR from gymnosperm Ginkgo biloba belonged to Group I while other DFRs from angiosperm species belonged to Group II. In angiosperm DFRs, HvDFR (Hordeum vulgare), OsDFR (O. sativa) and BfDFR (Bromheadia finlaysoniana) from monocot plant species formed a subgroup (II-a) which was distinct from the dicot subgroup (II-b). These results are consistent with a recently published phylogenetic analysis of the DFR family [22].

thumbnail
Figure 2. Phylogenetic relationships between PtrDFRs and other DFR proteins from other plant species.

Phylogenetic and evolutionary analyses were performed using the neighbor-joining method by DNAMAN software. Additional sequences include Medicago truncatula (MtDFR1, AAR27014), Populus tremuloides (PtDFR, AY147903), Arabidopsis thaliana (AtDFR, AAA32783), Fragaria x ananassa (FaDFR, AAC25960), Malus x domestica (MdDFR, AAD26204), Vitis vinifera (VvDFR, CAA53578), Dianthus caryophyllus (DcDFR, CAA91924), Gerbera hybrid (GhDFR, CAA78930), Gentiana triflora (GtDFR, AA12736), Ipomoea nil (InDFR, BAA22072), Ipomoea purpurea (IpDFR, BAA74700), Nicotiana tabacum (NtDFR, ABN80437), Solanum lycopersicum (SiDFR, CAA79154), Petunia x hybrida (PhDFR, AAF60298), Vaccinium macrocarpon (VmDFR, AAL89714), Hordeum vulgare (HvDFR, AAB20555), Bromheadia finlaysoniana (BfDFR, AAB62873) and Ginkgo biloba (GbDFR, AAU95082).

https://doi.org/10.1371/journal.pone.0030364.g002

Characterization of expression patterns of PtrDFR1 and PtrDFR2

In M. truncatula, transcript accumulation of both DFR genes was highest in young seeds and flowers, consistent with the accumulation of CTs and leucoanthocyanidins in these tissues [8]. It was established that the two DFR genes exhibited different expression patterns in Populus. PtrDFR2 transcripts were detected in both roots and leaves while PtrDFR1 expression was very weak in leaves [22]. In this study, the expression levels of the PtrDFR genes in various tissues were analyzed by quantitative real-time PCR by using gene-specific primers that can distinguish the two highly similar PtrDFR transcripts. Quantitative real-time PCR analysis showed that both PtrDFR1 and PtrDFR2 transcripts were found in each tested tissue, but were most concentrated in roots (Figure 3). PtrDFR2 transcripts were more than twice as abundant as PtrDFR1 in young petioles and were 15 times more abundant than in old petioles. In roots, the relative level of PtrDFR2 transcripts was approximately three times higher than that of PtrDFR1 transcripts. In contrast, PtrDFR2 expression was relatively lower in stems compared to PtrDFR1 expression. Neither PtrDFR gene was highly expressed in stems nor in young or mature leaves of poplar plants.

thumbnail
Figure 3. PtrDFR1 and PtrDFR2 transcript levels in different tissues of P. trichocarpa.

(A) Gel analysis of semiquantitative reverse transcriptase (RT)-PCR with transcript-specific DFR primers. (B) Expression levels were determined by qRT-PCR. Values represent averages of three biological replicates, each with two technical replicates. Actin expression of poplar was used as a control. Total RNA was isolated from poplar tissues: root (R), shoot (S), young leaf (YL), old leaf (OL), young petiole (YP) and old petiole (OP).

https://doi.org/10.1371/journal.pone.0030364.g003

Effect of ectopic expression of PtrDFR1 and PtrDFR2 on flower color in tobacco

Previous studies have shown that overexpression of the DFR genes from cranberry and M. truncatula in tobacco (N. tabacum) resulted in an increase in anthocyanin accumulation and a change in flower color [8], [39][41]. We found that the two PtrDFR genes exhibited distinct expression patterns in Populus, but it is not clear whether or not these enzymes can perform different functions in anthocyanin biosynthesis. To investigate the function of the PtrDFR proteins in vivo, transgenic tobacco plants that expressed PtrDFR genes under the control of the CaMV 35S promoter were produced. The pBI121 vector carrying the β-glucuronidase gene driven by the CaMV 35S promoter [42] was used as a control for comparison (named pBI121 control vector). PCR analysis using gene-specific primers with genomic DNA from leaf samples of putative transgenic plants identified the integration of nptII, PtrDFR1 and PtrDFR2 in the tobacco genome (Figure S1).

In general, flowers of the wild-type control tobacco plants (N. tabacum cv Xanthi) that were used for transformation exhibited white or pale pink colors under greenhouse conditions. Transgenic plants harboring the 35S:PtrDFR1 transcription cassette produced much darker pink flowers than were observed on wild-type control plants or transgenic plants carrying either the 35S:PtrDFR2 transcription cassette or the pBI121 control vector (Figure 4A). Further, anthocyanin pigments were extracted from the corollas of flowers of various plants and roughly measured spectrophotometrically. Compared to either wild-type control or pBI121 control vector transgenic plants, a significantly higher accumulation of anthocyanins was detected in the PtrDFR1 overexpression transgenic plants (Figure 4B). None of the PtrDFR2 transgenic plants showed such a significant increase. RT-PCR analyses proved overexpression of PtrDFR1 or PtrDFR2 in each correlated transgenic tobacco line (Figure 4C). Together, these results suggested that the PtrDFR1 protein may interact with the endogenous enzymes of anthocyanin biosynthetic pathways in tobacco and result in the increase of anthocyanin accumulation in vivo, whereas the PtrDFR2 protein does not seem to have such a function. Our findings are consistent with a previous report on M. truncatula [8] in which overexpression of MtDFR1 in transgenic tobacco resulted in a visible increase in anthocyanin accumulation in flowers, while MtDFR2 did not, indicating that unexpected properties and differences are found in two DFR proteins of the same orthology.

thumbnail
Figure 4. Effect of PtrDFR1 and PtrDFR2 in vivo on anthocyanin accumulation in transgenic tobacco flowers.

(A) Overexpression of PtrDFR1 resulted in a visible increase in anthocyanin accumulation in the corolla of transgenic tobacco flowers (lines 3 and 11), relative to untransformed lines (WT) and pBI121 transgenic control (pBI121) or the 35S:PtrDFR2 transgenics (5 and 9). (B) Quantitation of anthocyanin levels in transgenic tobacco flowers with a spectrophotometer. Error bars are SDs from three independent experiments. Both the lines harboring the 35S:PtrDFR1 gene had significantly higher anthocyanin levels compared with wild-type and the pBI121 transgenic control (based on a Student's t test analysis limit of P≤0.05). No lines overexpressing the 35S:PtrDFR2 gene showed significant increases in anthocyanins. (C) RT-PCR analysis of PtrDFR1 and PtrDFR2 expression in transgenic tobacco plants.

https://doi.org/10.1371/journal.pone.0030364.g004

Functional characterization of PtrDFR genes in CT biosynthesis in transgenic poplar

It has been previously demonstrated that CT concentrations are as high as 50% in cottonwood (P. angustifolia) [43], [44] and constitute up to 18% more of leaf dry weight than in aspen (P. tremuloides) [45]. CT accumulation can be induced by herbivores and pathogenic fungi in woody plants [46], [7], indicating that CTs play important roles in defense and protection against biotic stresses [47]. To date, the biosynthetic pathways for CTs and a large number of polyphenolic compounds have been well established [7], [22]. DFR encodes a key enzyme in the later steps of CT synthesis in plants, and overexpression of a DFR gene in the forage legume Lotus corniculatus resulted in an alteration of CT levels [48], [49]. To further determine the function of PtrDFRs in the CT biosynthetic pathway, their ORFs, flanked by the 35S promoter, were introduced into P. tomentosa Carr. More than 20 independent kanamycin-resistant transgenic lines harboring either p35S:PtrDFR1 or p35S:PtrDFR2 expression cassettes were produced. All transgenic plants were confirmed with PCR analysis by using gene-specific primers (Figure S2). No obvious morphological differences were observed between the transgenic and wild-type poplar plants. Three representative lines of each construction were selected for further analysis.

Spectrophotometric quantification of total CTs showed that all transgenic lines that carried the 35S:PtrDFR1 or 35S:PtrDFR2 transcription cassettes exhibited an increase in CT content in leaf tissues compared to wild-type plants and the transgenic controls of the pBI121 control vector (Figure 5B). In 35S:PtrDFR1 lines 4 and 12, a three-fold increase in the CT amount was detected in leaf tissues when compared to the wild-type control. Similarly, in 35S:PtrDFR2 lines, the concentration of total CTs in lines 5 and 9 was more than twice as high as those in the wild-type control and pBI121 transgenic control (Figure 5B). In addition, total anthocyanin levels were also measured spectrophotometrically. A dramatic increase in anthocyanin accumulation was detected in PtrDFR1 overexpression transgenic poplar plants, whereas none of the PtrDFR2 transgenic lines showed a significant increase of anthocyanins (Figure 5A). These results are consistent with the findings in transgenic tobacco as described above.

thumbnail
Figure 5. Anthocyanin and condensed tannin levels in PtrDFR1 and PtrDFR2 transgenic P. tomentosa Carr.

(A) The estimated relative anthocyanin contents at 530 nm absorbance. (B) The estimated relative condensed tannin contents at 500 nm absorbance. Error bars represent SDs from three independent experiments. Asterisks indicate a statistically significant difference between wild-type and transgenic plants (P≤0.05 by Student's t-test).

https://doi.org/10.1371/journal.pone.0030364.g005

Quantitative real-time PCR analysis confirmed that much higher levels of PtrDFR1 transcripts accumulated in 35S:PtrDFR1 transgenic lines (e.g., 4 and 12) compared to the wild-type and pBI121 transgenic control lines (Figure 6). Similarly, all 35S:PtrDFR2 lines accumulated high levels of PtrDFR2 transcripts. These results are in agreement with the previous analysis of flavonoid accumulation in the leaf tissues of 35S:PtrDFR1 lines. However, enhanced transcription levels of PtrDFR2 were not accompanied by a significant increase in the amount of total anthocyanins. Obviously, PtrDFR2 transcription levels correlate poorly with anthocyanin accumulation in Populus, but correlate well with CT accumulation. Taken together, these results suggest that the two DFR genes in P. trichocarpa may be specialized for anthocyanin synthesis or CT pathways due to gene duplication or pathway redundancy during plant evolution. Similarly, multiple DFR genes are present in L. japonicus and M. truncatula genomes, and these DFR proteins possess different catalytic activities in plants, suggesting a single duplication event followed by functional divergence or two independent duplication events followed by independent sub- or neo-functionalization events [8], [50]. To further determine their functions in the flavonoid pathway, detailed biochemical characterization of purified PtrDFR isozymes will be performed in the future.

thumbnail
Figure 6. Determination of transcript levels of PtrDFR1 and PtrDFR2 in transgenic poplar plants by quantitative real-time PCR.

The PtrDFR1 expression was significantly increased in transgenic lines 4 and 12, whereas the PtrDFR2 transcript level was higher in all three transgenic lines than in wild-type and pBI121 transgenic controls.

https://doi.org/10.1371/journal.pone.0030364.g006

Supporting Information

Figure S1.

PCR analysis of transgenic tobacco plants. (A) PCR amplification using primers specific for the production of a 741-bp NPTII fragment. (B) PCR amplification using primers specific for the production of a 1,375-bp PtrDFR1 fragment. (C) PCR amplification using primers specific for the production of a 1,128-bp PtrDFR2 fragment. M, D2000 DNA Ladder; WT, wild-type plants; pBI121, transgenic control; Plasmid, corresponding plasmid DNA (positive control); Lanes 1–16, independent transgenic lines. Numbers on the left indicate DNA marker sizes in base pairs.

https://doi.org/10.1371/journal.pone.0030364.s001

(TIF)

Figure S2.

PCR analysis of transgenic P. tomentosa Carr. plants. PCR amplification using primers designed for a 741-bp fragment of the NPTII gene using total genomic DNA as the template. (A) Of 35S:PtrDFR1 transgenic lines (2, 3, 4, 5, 6, 7, 8, 10, 11, 12 and 15). (B) Of 35S:PtrDFR2 transgenic lines (1, 2, 4, 5, 7, 8, 9, 10, 11 and 14). M, D2000 DNA Ladder; WT, wild-type plants; pBI121, transgenic control; Plasmid, corresponding plasmid DNA (positive control). Numbers on the left indicate DNA marker sizes in base pairs.

https://doi.org/10.1371/journal.pone.0030364.s002

(TIF)

Acknowledgments

We appreciate Dr. Jackie Kelley (The Samuel Roberts Noble Foundation) for reading and editing this manuscript.

Author Contributions

Conceived and designed the experiments: YH ZJ KL. Performed the experiments: YH ZJ LY XX YS FS. Analyzed the data: YH ZJ KL. Contributed reagents/materials/analysis tools: YH ZJ. Wrote the paper: YH JG KL.

References

  1. 1. Winkel-Shirley B (2001) It takes a garden. How work on diverse plant species has contributed to an understanding of flavonoid metabolism. Plant Physiol 127: 1399–1404.
  2. 2. Martens S, Knott J, Seitz CA, Janvari L, Yu SN, et al. (2003) Impact of biochemical pre-studies on specific metabolic engineering strategies of flavonoid biosynthesis in plant tissues. Biochem Eng J 14: 227–235.
  3. 3. Manach C, Scalbert A, Morand C, Remesy C, Jimenez L (2004) Polyphenols: food sources and bioavailability. Am J Clin Nutr 79: 727–747.
  4. 4. Dixon RA, Steele CL (1999) Flavonoids and isoflavonoids: An old mine for metabolic engineering. Trends in Plant Sci 4: 394–400.
  5. 5. Dixon RA, Xie DY, Sharma SB (2005) Proanthocyanidins – a final frontier in flavonoid research. New Phytol 165: 9–28.
  6. 6. Kristiansen KN, Rohde W (1991) Structure of the Hordeum vulgare gene encoding dihydroflavonol-4-reductase and molecular analysis of ant18 mutants blocked in flavonoid synthesis. Mol Gen Genet 230: 49–59.
  7. 7. Peters D, Constabel CP (2002) Molecular analysis of herbivore-induced condensed tannin synthesis: Cloning, expression of dihydroflavonol reductase from trembling aspen (Populus tremuloides). Plant J 32: 701–712.
  8. 8. Xie DY, Jackson LA, Cooper JD, Ferreira D, Paiva NL (2004) Molecular and biochemical analysis of two cDNA clones encoding dihydroflavonol-4-reductase from Medicago truncatula. Plant Physiology 134(3): 979–994.
  9. 9. Olsen O, Wang X, von Wettstein D (1993) Sodium azide mutagenesis: preferential generation of A.T→G.C transitions in the barley Ant18 gene. Proc Natl Acad Sci USA 90: 8043–8047.
  10. 10. Shirley BW, Kubasek WL, Storz G, Bruggemann E, Koornneef M, et al. (1995) Analysis of Arabidopsis mutants deficient in flavonoid biosynthesis. Plant J 8: 659–671.
  11. 11. Beld M, Martin C, Huits H, Stuitje AR, Gerats AG (1989) Flavonoid synthesis in Petunia hybrida: Partial characterization of dihydroflavonol-4-reductase genes. Plant Mol Biol 13: 491–502.
  12. 12. Sparvoli F, Martin C, Scienza A, Gavazzi G, Tonelli C (1994) Cloning and molecular analysis of structural genes involved in flavonoid and stillbene biosynthesis in grape (Vitis vinifera L.). Plant Mol Biol 24: 743–755.
  13. 13. Liew CF, Loh CS, Gho CJ, Lim SH (1998) The isolation, molecular characterization and expression of dihydroflavonol 4-reductase cDNA in the orchid, Bromheadia finlaysoniana. Plant Sci 135: 161–169.
  14. 14. Fischer TC, Halbwirth H, Meisel B, Stich K, Forkmann G (2003) Molecular cloning substrate specificity of the functionally expressed dihydroflavonol 4-reductase from Malus domestica and Pyrus communis cultivars and the consequences for flavonoid metabolism. Arch Biochem Biophys 412: 223–230.
  15. 15. Lo Piero AR, Puglisi I, Petrone G (2006) Gene characterization, analysis of expression and in vitro synthesis of dihydroflavonol 4-reductase from Citrus sinensis (L.) Osbeck. Phytochemist 67: 684–695.
  16. 16. Devic M, Guilleminot J, Debeaujon I, Bechtold N, Bensaude E, et al. (1999) The BANYULS gene encodes a DFR-like protein and is a marker of early seed coat development. Plant J 19: 387–398.
  17. 17. Inagaki Y, Johzuka-Hisatomi Y, Mori T, Takahashi S, Hayakawa Y, et al. (1999) Genomic organization of the genes encoding dihydroflavonol 4-reductase for flower pigmentation in the Japanese and common morning glories. Gene 226: 181–188.
  18. 18. Ostergaard L, Lauvergeat VV, Naested H, Mattsson O, Mundy J (2001) Two differentially regulated Arabidopsis genes define a new branch of the DFR superfamily. Plant Sci 160: 463–472.
  19. 19. Taylor G (2002) Populus: Arabidopsis for forestry. Do we need a model tree? Ann Bot 90: 681–689.
  20. 20. Jansson S, Douglas CJ (2007) Populus: a model system for plant biology. Annu Rev Plant Biol 58: 435–458.
  21. 21. Sterky F, Bhalerao RR, Unneberg P, Segerman B, Nilsson P, et al. (2004) A Populus EST resource for plant functional genomics. PNAS 101(38): 13951–6.
  22. 22. Tuskan GA, DiFazio S, Jansson S, Bohlmann J, Grigoriev I, et al. (2006) The genome of black cottonwood, Populus trichocarpa (Torr. & Gray). Science 313: 1596–1604.
  23. 23. Tsai CJ, Harding SA, Tschaplinski TJ, Lindroth RL, Yuan Y (2006) Genome-wide analysis of the structural genes regulating defense phenylpropanoid metabolism in Populus. New Phytol 172: 47–62.
  24. 24. Gou J, Strauss SH, Tsai CJ, Fang K, Chen Y, et al. (2010) Gibberellins regulate lateral root formation in Populus through interactions with auxin and other hormones. Plant Cell 22: 623–63.
  25. 25. Chen C, Dale MC, Okos MR (1990) Minimal nutritional requirements for immobilized yeast. Biotechnol Bioeng 36(10): 993–1001.
  26. 26. Jia Z, Gou J, Sun Y, Yuan L, Tang Q, et al. (2011) Enhanced resistance to fungal pathogens in transgenic Populus tomentosa Carr. by overexpression of an nsLTP-like antimicrobial protein gene from motherwort (Leonurus japonicus). Tree Physiol 30: 1599–1605.
  27. 27. Murashige T, Skoog F (1962) A revised medium for rapid growth and bioassay with tobacco tissue cultures. Physiol Plant 15: 473–497.
  28. 28. Horsch RB, Fry J, Hoffmann N, Neidermeyer J, Rogers SG, et al. (1988) Leaf disc transformation. In: Gelvin SB, Schilperoort R, editors. Plant Molecular Biology Manual. 1 p. Kluwer Academic Publishers, Dordrecht, The Netherlands.
  29. 29. Jia ZC, Sun YM, Yuan L, Tian Q, Luo KM (2010) The chitinase gene (Bbchit1) from Beauveria bassiana enhances resistance to Cytospora chrysosperma in Populus tomentosa Carr. Biotechnol Lett 32: 1325–1332.
  30. 30. Lloyd G, McCown B (1980) Commercially-feasible micropropagation of mountain laurel, Kalmia latifolia, by use of shoot-tip culture. Combined Proc Int Plant Propagators Soc 30: 421–427.
  31. 31. Luo KM, Zheng XL, Chen YQ, Xiao YH, Zhao DG, et al. (2006) The maize Knotted1 gene as a positive selectable marker gene is effective for Agrobacterium-mediated transformation in tobacco. Plant Cell Rep 25: 403–409.
  32. 32. Gou J, Ma C, Kadmiel M, Gai Y, Strauss S, et al. (2011) Tissue-specific expression of Populus C19 GA 2-oxidases differentially regulate above- and below-ground biomass growth through control of bioactive GA concentrations. New Phytol. 5 Aug 2011, DOI: 10.1111/j.1469-8137.2011.03837.x.
  33. 33. Pang YZ, Peel GJ, Wright E, Wang ZY, Dixon RA (2007) Early steps in proanthocyanidin biosynthesis in the model legume Medicago truncatula. Plant Physiol 145: 601–615.
  34. 34. Swain T, Hillis WEJ (1959) Phenolic constituents of Prunus domestica I. Quantitative analysis of phenolic constituents. J Sci Food Agric 10: 63.
  35. 35. Price ML, Van Scoyoc S, Butler LG (1978) A critical evaluation of the vanillin reaction as an assay for tannin in sorghum grain. J Agri Food Chem 26: 1214–1218.
  36. 36. Waterman PG, Mole S (1994) Analysis of Phenolic Plant Metabolites in Methods in Ecology. Oxford (UK): Blackwell. 85 p.
  37. 37. Johnson E, Yi H, Shin B, Oh BJ, Cheong H, et al. (1999) Cymbidium hybrida dihydroflavonol 4-reductase does not efficiently reduce dihydrokaempferol to produce pelargonidin-type anthocyanins. Plant J 19: 81–85.
  38. 38. Lacombe E, Hawkins S, Doorsselaere JV, Piquemal J, Goffner D, et al. (1997) Cinnamoyl CoA reductase, the first committed enzyme of the lignin branch biosynthetic pathway: cloning, expression and phylogenetic relationships. Plant J 11: 429–441.
  39. 39. Aida R, Yoshida K, Kondo T, Kishimoto S, Shibata M (2000) Copigmentation gives bluer flowers on transgenic torenia plants with the antisense dihydroflavonol-4-reductase gene. Plant Sci 160: 49–56.
  40. 40. Polashock JJ, Griesbach RJ, Sullivan RF, Vorsa N (2002) Cloning of a cDNA encoding the cranberry dihydroflavonol-4-reductase (DFR) and expression in transgenic tobacco. Plant Sci 163: 241–251.
  41. 41. Davies KM, Schwinn KE, Deroles SC, Manson DG, Lewis DH, et al. (2003) Enhancing anthocyanin production by altering competition for substrate between flavonol synthase and dihydroflavonol 4-reductase. Euphytica 131: 259–268.
  42. 42. Bevan MW (1984) Binary Agrobacterium vectors for plant transformation. Nucleic Acids Res 12: 8711–8721.
  43. 43. Driebe EM, Whitham TG (2000) Cottonwood hybridization affects tannin and nitrogen content of leaf litter and alters decomposition. Oecologia 123: 99–107.
  44. 44. Whitha TG, Young WP, Martinsen GD, Gehring CA, Schweitzer JA, et al. (2003) Community and ecosystem genetics: A consequence of the extended phenotype. Ecology 84: 559–573.
  45. 45. Lindroth RL, Hwang SY (1996) Diversity, redundancy and multiplicity in chemical defense systems of aspen. In: Romeo JT, Saunders JA, Barbosa P, editors. “Recent Advances in Phytochemistry.”. 25 p. Plenum Press, New York.
  46. 46. Schultz JC, Baldwin IT (1982) Oak leaf quality declines in response to defoliation by gypsy moth larvae. Science 217: 149–151.
  47. 47. Donaldson JR, Lindroth RL (2007) Genetics, environment, and their interaction determine efficacy of chemical defense in trembling aspen. Ecology 88: 729–739.
  48. 48. Bavage AD, Davies IG, Robbins MP, Morris P (1997) Expression of an Antirrhinum dihydroflavonol reductase gene results in changes in condensed tannin structure and accumulation in root cultures of Lotus corniculatus (birds foot trefoil). Plant Mol Biol 35: 443–458.
  49. 49. Robbins MP, Bavage AD, Strudwicke C, Morris P (1998) Genetic manipulation of condensed tannins in higher plants. II. Analysis of birdsfoot trefoil plants harboring antisense dihydroflavonol reductase constructs. Plant Physiol 116: 1133–1144.
  50. 50. Shimada N, Sasaki R, Sato S, Kaneko T, Tabata S, et al. (2005) A comprehensive analysis of six dihydroflavonol 4-reductases encoded by a gene cluster of the Lotus japonicus genome. J Exp Bot 56: 2573–2585.