Next Article in Journal
Thyroid Hormones and Antioxidant Systems: Focus on Oxidative Stress in Cardiovascular and Pulmonary Diseases
Previous Article in Journal
Carbon Nanotube-Induced Pulmonary Granulomatous Disease: Twist1 and Alveolar Macrophage M1 Activation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Characterization of a Maize Wip1 Promoter in Transgenic Plants

1
Institute of Crop Sciences, Chinese Academy of Agricultural Sciences, Zhongguancun South Street 12, Beijing 100081, China
2
Institute of Industrial Crops, Henan Academy of Agricultural Sciences, Zhengzhou 450002, Henan, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2013, 14(12), 23872-23892; https://doi.org/10.3390/ijms141223872
Submission received: 2 October 2013 / Revised: 22 November 2013 / Accepted: 25 November 2013 / Published: 6 December 2013
(This article belongs to the Section Biochemistry)

Abstract

:
The Maize Wip1 gene encodes a wound-induced Bowman-Birk inhibitor (BBI) protein which is a type of serine protease inhibitor, and its expression is induced by wounding or infection, conferring resistance against pathogens and pests. In this study, the maize Wip1 promoter was isolated and its function was analyzed. Different truncated Wip1 promoters were fused upstream of the GUS reporter gene and transformed into Arabidopsis, tobacco and rice plants. We found that (1) several truncated maize Wip1 promoters led to strong GUS activities in both transgenic Arabidopsis and tobacco leaves, whereas low GUS activity was detected in transgenic rice leaves; (2) the Wip1 promoter was not wound-induced in transgenic tobacco leaves, but was induced by wounding in transgenic rice leaves; (3) the truncated Wip1 promoter had different activity in different organs of transgenic tobacco plants; (4) the transgenic plant leaves containing different truncated Wip1 promoters had low GUS transcripts, even though high GUS protein level and GUS activities were observed; (5) there was one transcription start site of Wip1 gene in maize and two transcription start sites of GUS in Wip1::GUS transgenic lines; (6) the adjacent 35S promoter which is present in the transformation vectors enhanced the activity of the truncated Wip1 promoters in transgenic tobacco leaves, but did not influence the disability of truncated Wip1231 promoter to respond to wounding signals. We speculate that an ACAAAA hexamer, several CAA trimers and several elements similar to ACAATTAC octamer in the 5′-untranslated region might contribute to the strong GUS activity in Wip1231 transgenic lines, meanwhile, compared to the 5′-untranslated region from Wip1231 transgenic lines, the additional upstream open reading frames (uORFs) in the 5′-untranslated region from Wip1737 transgenic lines might contribute to the lower level of GUS transcript and GUS activity.

1. Introduction

It is necessary to use different promoters when multiple genes are cloned in tandem in transgenic plants because multiple copies of a single promoter could lead to the silencing of the transgenes. Promoters may be classified as constitutive, tissue-specific, or inducible based on their expression patterns. The most widely used constitutive promoters include the CaMV 35S promoter [1], the maize ubiquitin promoter [2], and the rice actin promoter [3]. Several other strong promoters have also been isolated and identified from plants or viruses [46]. Although many promoters have been isolated and tested, only a few have been successfully used in agricultural biotechnology. The constitutive expression of certain genes might be harmful to their host plants, affecting plant growth and development; inducible or tissue-specific promoters would be more useful than constitutive promoters in these cases.
The function of a promoter is normally determined by the combinatorial action of multiple regulatory elements, i.e., enhancers and cis-elements, and by the interactions between regulatory elements and nuclear protein factors [7]. For the CaMV 35S promoter, the region from −46 to +1 upstream of the gene is the basal promoter and the region from −343 to −46 acts as an enhancer [1]. Although the CaMV 35S promoter is presumed to be constitutive, some reports have shown that it is developmentally regulated [8] and can be affected by abiotic stress [9]. Negative cis-regulatory regions determine the tissue-specific activity of a number of promoters [10]. Besides the cis elements in DNA sequence affect transcription, various cis elements in mRNA have the functions to control the translation efficiency. Many reports have shown that the 5′-untranslated region (5′-UTR) and 3′-untranslated region (3′-UTR) can increase the translation efficiency [11]. Upstream open reading frames (uORFs) that are located in the 5′-UTR have been reported to generally downregulate the expression of the major open reading frame (mORF) through ribosomal stalling or reducing initiation efficiency [12]. However, a few uORFs can upregulate the expression of mORF [13]. Additionally, more and more reports showed that the 35S promoter interferes with the tissue specificity, strength [14] or inducibility (our data, unpublished) of its adjacent promoter. The onset and extent of influence depends on the responsive nature of the adjacent promoter and the distance between the 35S promoter and adjacent promoter [15].
The Bowman-Birk inhibitor (BBI) is a type of serine protease inhibitor, and its expression is induced by wounding or infection, conferring resistance against pathogens and pests [16]. It was reported that the BBI confers heavy metal and multiple drug tolerance in yeast [17]. There are seven BBI genes in rice and overexpression of the rice BBI2-3 conferred fungal pathogen resistance in transgenic rice plants [18]. The maize Wip1 gene, encoding a wound-induced BBI protein, has been cloned, and the promoter of the gene might respond to wound signals [19,20]. In contrast to most BBIs, which inhibit trypsin and chymotrypsin proteases, both of the inhibitory domains of maize Wip1 inhibit chymotrypsin [20]; however, these two domains have different evolutionary histories and ecological functions [21].
Here, we isolated the maize Wip1 promoter and characterized its function in transgenic plants, showing that Wip1 promoters showed different properties in dicot and monocot plants, and some truncated promoters under the influence of adjacent 35S promoter can drive high GUS activities even with a low GUS transcriptional level.

2. Results

2.1. Isolation and Analysis of the Maize Wip1 Promoter

To isolate the maize Wip1 promoter, we performed a BLAST analysis with the mRNA sequence of maize Wip1 (X71396) in www.maizegdb.org [22]. The maize Wip1 sequence was homologous to the genome sequence from 12,274,189 to 12,273,525 on chromosome 8, and one intron was found in the Wip1 coding region. Based on the maize genomic sequence, a 1737 bp fragment (chr8: 12,275,876–12,274,140) upstream of the translation start site of Wip1 was amplified by PCR. The cis-elements in the promoter sequence were analyzed using the PlantCARE [23] and PLACE [24] databases, and the transcription start sites were predicated on the website: http://www.fruitfly.org/seq_tools/promoter.html [25]. Two putative TATA boxes and two corresponding potential transcription start sites were present in the promoter sequence. Six W-boxes and three GCC-boxes were found in the Wip1 promoter. In addition, several CAAT-boxes, which are common cis-acting elements in promoter and enhancer regions, were also found (Figure 1).

2.2. Truncated Wip1 Promoter Has Strong Activity in Transgenic Arabidopsis and Tobacco Leaves

To analyze the function of the Wip1 promoter, several expression vectors (pWip1737, pWip1500, pWip1231, pWip1191, pWip791, pWip491) were constructed (Figure 2A, Tables 1 and 2). The constructed plasmids were transformed into tobacco, and at least ten independent transgenic tobacco lines for each construct were obtained. The transformations were confirmed by PCR analysis of genomic DNA using primers specific for GUS (data not shown). We performed quantitative analysis of GUS activity using leaf samples from 40-day-old T3 progeny of plants. The results showed that the Wip1231 promoter had strong activity, while no or weak activity was detected for the other truncated promoters (Figure 2B). To confirm the results of the GUS activity analysis, the seedlings were subjected to histochemical staining. Strong GUS staining was observed for the seedlings that were transgenic for the Wip1231 promoter and 35S promoter, whereas no or light staining was observed for the seedlings that were transgenic for the full-length and other truncated promoters (Figure 3), consistent with the GUS activity analysis results. To investigate whether the truncated Wip1231 promoter also had strong activity in other plant species, the constructs were transformed into Arabidopsis. As expected, the Wip1231 promoter also drove high GUS activity in transgenic Arabidopsis plant leaves (Figure 2C), further confirming the high activity of the truncated Wip1231 promoter.
To find out the potential negative cis-acting elements, detailed 5′ and 3′ deletions of Wip1737, based on Wip1231, were performed, involving the generation of an additional 15 deletion constructs (pWip1231A, pWip1231C, pWip931, pWip1231a, pWip1231b, pWip1231c, pWip1231Aa, pWip1231Ab, pWip1231Ac, pWip1231Ba, pWip1231Bb, pWip1231Bc, pWip1231Ca, pWip1231Cb, pWip1231Cc; Figure 4A, Tables 1 and 2). These constructs were transformed into Arabidopsis, and the GUS activity in the transgenic plant leaves was analyzed. The results showed that Wip1231A, Wip1231C, and Wip931, which were of the same 3′ end as Wip1231, had high activity similar to that of Wip1231, while the other Wip1 promoter fragments presented very low levels of activity (Figure 4C). To further confirm the functions of these three fragments, three constructs (pWip1231A, pWip1231C and pWip931) were also transformed into tobacco, respectively. High GUS activities were observed in the transgenic tobacco leaves transformed with each of these three promoters (Figure 4B).

2.3. Wip1 Promoters Are not Induced by Wounding and Have Different Activity in Different Organs in Transgenic Tobacco

It has been shown that the expression of Wip1 gene is wound-induced in maize [19,20]. We investigated whether the Wip1 promoter is wound-induced in transgenic tobacco plant leaves that were wounded with three different kinds of wounding methods, respectively. Under three different treatment conditions, to our surprise, wounding did not induce the increase of the activity of Wip1 promoters (Figure 5), indicating that Wip1 promoters are not wound-induced in transgenic tobacco.
The GUS activities in different organs of the transgenic tobacco lines transformed with different truncated Wip1 promoters or the control 35S promoter were analyzed. The results showed that the Wip1231 promoter has high activity in the leaf and stem but low activity in the root and seed (Figure 6).

2.4. Wip1231 Promoter Drives Low GUS Transcriptional Level

Generally, the level of protein is consistent with the level of mRNA, but sometimes they are discordant. To confirm Wip1231 is a strong promoter, we investigated the transcriptional levels of GUS in the transgenic tobacco plants. To our surprise, lower transcriptional levels of GUS were detected in transgenic plants containing Wip1 promoters than in transgenic plants containing CaMV 35S promoter. There were higher GUS transcriptional level in Wip1231::GUS transgenic lines than in Wip1737::GUS and Wip1500::GUS lines (Figure 7). The results of Western blot analysis showed that there were high GUS protein levels in both Wip1231::GUS and 35S::GUS transgenic lines, whereas GUS protein was undetectable in Wip1737::GUS, Wip1500::GUS and WT plants (Figure S1). These results indicate that the strong GUS activity in Wip1231::GUS transgenic lines can be attributed to high efficient translation of GUS protein. We speculate that there might be some translation enhancer in the 5′-UTR of GUS transcripts in Wip1231::GUS transgenic plants.

2.5. Wip1 Promoters Have Multiple GUS Transcription Start Sites in Transgenic Tobacco

To find out the exact 5′-UTRs, we further determined the accurate transcription start sites in maize and transgenic tobacco plants by the 5′-RACE method. The native transcription start site of Wip1 in maize was identified as the nucleotide G which is located at 61 bp upstream of the Wip1 translation initiation site and 3 bp downstream of the predicated one (Figures 1 and 8). For Wip1737::GUS transgenic tobacco line, a strong band and one weak band were amplified, and sequencing results revealed that two transcription start sites are present (Figures 1 and 8). The latter is the same as the transcription start site of Wip1 gene in maize. The former was identified as nucleotide C which is located at 532 bp upstream of the ATG translation initiation site and identical to the predicated one, and this transcription start site was also identified in Wip1231::GUS transgenic tobacco line. So there is a 272 bp-length untranslated region in GUS transcripts in Wip1231::GUS lines. In the 272 bp 5′-UTR there is an ACAAAA element, several CAA trimers and several elements similar to ACAATTAC octamer (Figure S2). The elements might contribute to the high translation efficiency of GUS transcripts in Wip1231 transgenic lines. The sequences of 5′-UTR from Wip1231 and Wip1737 transgenic lines were comparatively analyzed to find that there were more uORFs in the latter (Figures S3–S5). Maybe the additional uORFs lead to the lower level of GUS transcript and GUS activity in Wip1737 transgenic lines.

2.6. Wip1231 Promoters Were Influenced by the Adjacent 35S Promoter Sequence

In this study, the transformation vectors were constructed from a derivative of the pCAMBIA1300 binary vector that includes a CaMV 35S promoter controlling a hptII selectable marker gene. It has been reported that the CaMV 35S enhancer can influence the expression of a transgene within the same transformation construct [14,30,31]. To investigate whether the 35S enhancer influences the truncated Wip1231, we constructed transformation vectors p1300-1231-NOS and p1300-35S-NOS, where the 35S promoter was replaced with a NOS promoter to control the hptII selectable marker gene, and the reporter gene GUS was controlled by the Wip1231 promoter and 35S promoter, respectively. Then they were transformed into tobacco, and transgenic tobacco plants were obtained. The GUS activity in transgenic plants transformed with p1300-1231-NOS was very low, compared to transgenic lines transformed with pWip1231. However, transgenic plants transformed with p1300-35S-NOS still had high GUS activity, which were similar to the transgenic lines transformed with p1300-221 (Figure 9). The Wip1231 promoter was still not induced by wounding in transgenic tobacco plants which was transformed with p1300-1231-NOS (Figure 9). The above results showed that adjacent 35S promoter did enhance the expression of GUS reporter gene that was controlled by truncated Wip1231, and the truncated Wip1231 does not respond to wound signals in transgenic tobacco plants.

2.7. Wip1 Promoter Is Wound Inducible in Transgenic Rice Plants

Expression of the Wip1 gene is induced through wounding in maize. However, the GUS gene, driven by Wip1 promoters, was not induced by wounding in transgenic tobacco which is a dicot. This encouraged us to analyze whether Wip1 promoters have the same function in monocot species. Constructs pWip1737, pWip1500 and pWip1231 were transformed into rice, respectively, and the transgenic rice events were confirmed by PCR analysis of the genomic DNA (data not shown). Low GUS activities were detected in transgenic rice plant leaves containing different Wip1 promoters, compared with that in transgenic rice transformed with the CaMV 35S promoter. When transgenic rice leaves were wounded, significant increase of GUS activity were observed in all the transgenic lines containing different Wip1 promoters, whereas the GUS activity in transgenic plants containing the 35S promoter did not change (Figure 10). This indicates that the Wip1 promoter is induced by wounding in rice.

3. Discussion

Arabidopsis and tobacco have long been used as simple and high-efficiency transformation systems in which to analyze gene function. In addition to protein-coding genes, the promoters from rice or maize have been studied in one of these two systems [3234]. In this study, we used both systems to study the function of maize Wip1 promoter. To our surprise, expression of GUS driven by truncated maize Wip1 fragments was not induced by wound in transgenic tobacco plants (Figure 5). This may be due to the fact that the maize Wip1 promoter comes from a monocot, and its function may be different in monocot and dicot species. Although it has been reported that the wound-inducible PinII promoter from potato was able to drive gene expression in response to wound signals in rice [28], many plant promoters exhibit different functionality in dicots and monocots. For example, the widely used maize ubiquitin promoter conferred lower expression of a reporter gene in the tobacco (dicot) protoplasts than in maize (monocot) protoplasts [2]. Several dicot promoters also have been shown to have lower activities than monocot promoters when transformed into monocot species [35].
Few strong promoters have been identified to date, and it remains important for researchers in plant biotechnology to identify promoters that are as strong as or stronger than the CaMV 35S promoter. In transgenic plants, high GUS activities were observed in transgenic lines containing truncated Wip1 promoters even though low GUS transcriptional levels were detected. The above results indicate that truncated Wip1 promoters are not strong, and the reason for the high GUS activity driven by the truncated Wip1 promoters might be due to the increased translation of the GUS gene. It has been shown that the 35S-GUS and 35S-ABI4-GUS lines had at least 50 fold differences in transcript level, however more than 300 fold difference in GUS activity was observed [36], indicating that transcriptional level is not correlated with the protein level in some cases. Many factors, i.e., promoter activity, mRNA stability, protein translation efficiency, can affect the expression level of foreign genes in transgenic plants. The high GUS activity driven by truncated Wip1 promoters indicate that Wip1 promoters are also useful in plant biotechnology, even though these promoters led to low transcriptional level.
Normally, a promoter shows different properties in monocot and dicot plants. It has been shown that the CaMV 35S promoter can drive high levels of transgene expression in dicot plants, whereas its activity is relatively lower in monocot plants [37]. In transgenic rice, lower activity of the Wip1231 promoter compared to the 35S promoter was observed, and wound treatment significantly increased GUS activities in transgenic rice containing different Wip1 promoters. These results are contrary to the results in transgenic tobacco, indicating that the Wip1 promoter also showed different properties in monocot and dicot plants; this will limit the usage of the Wip1 promoter as a general (cross-species) promoter in transgenic crops.
5′-RACE results revealed that only one native transcription start site was present in the maize Wip1 gene, and a new one was identified in both Wip1737::GUS and Wip1231::GUS transgenic tobacco lines. We analyzed the 272 bp 5′-UTR of GUS mRNA from Wip1231::GUS transgenic tobacco lines to find that there is an ACAAAA hexamer, several separate CAA trimers and several elements similar to ACAATTAC octamer (Figure S2). The ACAAAA hexamer exists in many 5′-UTRs of plant genes and has been speculated to enhance translation [38]. The poly(CAA) region and ACAATTAC octamer have been confirmed to enhance translation [3941]. The features of the sequence may contribute to the property of the 5′-UTR for high efficient translation. It has been reported that uORF plays an important role in the regulation of gene expression by different mechanisms, for example, by ribosomal stalling, reducing initiation efficiency [12]. uORF can downregulate gene expression by accelerating mRNA degradation or upregulating gene expression by reforming the 5′-UTR through the product of the uORF [42]. Six upstream ATGs (uATGs) were found in the 272 bp UTR. Among the six uATGs, the first three are in two very short uORFs, the fourth and fifth are in the reading frame of the GUS gene and will produce GUS fusion protein, and the sixth uATG should not have the opportunity to initiate translation (Figure S3). We suggest that the two uORFs may have little possibility to influence the mORF translation. Strong GUS activity in transgenic plants transformed with pWip1231 construct and related pWip1231A, pWip1231C and pWip931 might be due to that they all have the same 3′ end and produce the same 5′-UTR containing two uATGs which produce GUS fusion proteins, and strong translation may initiate from the two uATGs.
14 uATGs were found in the 532 bp UTR of GUS mRNA from Wip1737::GUS transgenic tobacco lines. These uATGs form more uORFs than the six uATGs in 272 bp UTR of GUS mRNA from Wip1231::GUS transgenic lines. The fourth and fifth uATGs in the 272 bp UTR which can produce GUS fusion protein, are in frame with a termination code in the 532 bp UTR and form a uORF. Interestingly, the uORF also exists in the 361 bp UTR of GUS mRNA from Wip1231a::GUS transgenic tobacco lines if the lines have the same transcription start site as Wip1231::GUS lines. We suggest that the uORF may lead to the low GUS activity of transgenic plants transformed by pWip1737 and pWip1231a. We also suggest that the uORF might accelerate the degradation of mRNA from Wip1731::GUS and Wip1500::GUS transgenic tobacco lines, because the levels of mRNA in these lines were lower than that in Wip1231::GUS transgenic tobacco lines. We cannot exclude other factors that contribute to the attained results.
A few reports have shown that the CaMV 35S enhancer has the ability to alter the expression of nearby genes by affecting nearby promoters [14,30,31]. In the T-DNA region of transformation plasmids used for the promoter analysis, the selectable marker gene hptII was controlled with double enhanced CaMV 35S promoter. We confirmed that the 35S promoter influenced the GUS expression in Wip1231::GUS transgenic plants, because low GUS activity was observed when the 35S promoter was replaced with the NOS promoter. A problem may appear if the 35S promoter exists nearby an inducible or tissue-specific promoter that drives target transgene expression, but the problem might be negligible when the adjacent promoter is a strong constitutive one. The disadvantage of the 35S promoter can be overcome by using enhancer-blocking insulators to disturb the interaction between the 35S promoter and adjacent promoters [43]. Every coin has two sides; maybe the property of the 35S promoter can be used to develop a “strong constitutive” promoter controlling transgene expression in plants, such as the Wip1231 and several other truncated Wip1 promoters studied in the paper, although these truncated Wip1 promoters drive low transcriptional level of transgene.

4. Experimental Section

4.1. Isolation and Analysis of the Maize Wip1 Promoter

Maize genomic DNA was isolated from leaves according to a previously described method [44]. A 1737 bp fragment designated as the full-length promoter sequence was amplified using the primers F1737 (5′-AACTGCAGGGCTCCGTTCTACTTGACT-3′) and R1737 (5′-CGGGATCCGGTCTCGGACGAGCTGTTCTT-3′). The underlined letters indicate PstI and BamHI restriction sites. Regulatory motifs were identified using PlantCARE [23] and PLACE [24], the transcription start sites were predicated on the website: http://www.fruitfly.org/seq_tools/promoter.html [25].

4.2. Plasmid Construction

The plasmid pBI221 was digested with HindIII and EcoRI. The CaMV 35S promoter-GUS-NOS fragment was purified and ligated into plasmid pCAMBIA1300 that had been digested with the same enzymes to construct the plasmid pCAMBIA1300-221. The amplified 1737 bp full-length promoter sequence was digested with restriction enzymes PstI and BamHI and inserted upstream of the GUS gene in plasmid pCAMBIA1300-221, replacing the CaMV 35S promoter to construct a new plasmid, pWip1737 (Figure S6). The full-length Wip1 promoter was replaced with other truncated Wip1 promoters to construct the other vectors. The details for the primers used to amplify the truncated promoters are shown in Figure 1 and Tables 1 and 2.
To construct vectors where 35S promoter was replaced with NOS promoter to control hptII selectable gene, NOS promoter was amplified using Gateway binary vector pGWB454 (AB294466.1) as template with primer F-NOS (5′-CCAACATGGTGGCATCATGAGCGGAGAATTAAG-3′) and primer R-NOS (5′-CTCGAGAGATCCGGTGCAGATTATTTG-3′). The underlined letters indicate BstXI and XhoI restriction sites, respectively. The vector pWip1231 was digested with BstXI and XhoI to produce three fragments, a 10 kb fragments containing vector bone, a 1.1 kb fragment containing hptII gene and a 0.8 kb fragment containing 35S promoter. The 10 and 1.1 kb fragments were gel purified separately and the 0.8 kb one was discarded. The 10 kb purified fragment ligated with the amplified NOS promoter fragment was digested by the same endonucleases to form the medium vector that was digested by XhoI. Then the purified product ligated with the 1.1 kb purified fragment in correct orientation to form the vector p1300-1231-NOS. The new control vector p1300-35S-NOS derived from pCAMBIA1300-221 was constructed similarly. The restriction endonuclease sites were in marked in detail in supplementary Figure S1.

4.3. Plant Transformation

For the tobacco plant transformations, the plant expression plasmids were transferred into competent Agrobacterium tumefaciens (strain LBA4404) cells by freeze-thaw treatment. The transformed Agrobacterium were selected on YEB-agar plates containing 100 mg/L kanamycin and 100 mg/L streptomycin. Recombinant Agrobacterium were infiltrated into the young tobacco (Nicotiana tobacum) leaves according to the described method [45]. For the Arabidopsis transformations, vector constructs were transformed into Agrobacterium tumefaciens strain GV3101. Arabidopsis ecotype Columbia Col-0 was transformed using the floral dip method [46]. The obtained tobacco and Arabidopsis T0 seeds were selected on MS medium plates containing 30 mg/L Hygromycin B to eliminate the non-transgenic plants.
For the rice transformation, vector constructs were transformed into Agrobacterium tumefaciens strain EHA105. Three-week-old calli derived from mature seed (Rryza sativa L. cv. Kita-ake) were used for the transformation. The transformation was performed according to the described method [47].

4.4. Wounding of Transgenic Plant Leaves

The second leaves from apex of 40-day-old T3 tobacco plants grown in greenhouse were wounded using three different kinds of methods. The first wound treatment was performed according to Walker-Simmons, et al. [26] wherein the leaves were crushed across the midvein by a hemostat, with a second wounding 20 h later, and the leave samples were harvested 24 h after the first wounding. The second wound treatment was performed according to An, et al. [27] wherein the leaves were cut into about 1 cm2 sections and the leaf slices were put in Murashige and Skoog (MS) liquid medium containing 3% sucrose (pH 5.8). The leaf slices were wounded by making several small holes in each section with forceps, and then were placed in tissue culture incubator at 28 °C for 24 h under light (3000 Lux), then the samples were harvested for the experiment. The third wound treatment was performed according to the method described by Xu, et al. [28] wherein the leaves were cut three times perpendicularly to the midvein by a scissor along both edges of the leaf blade without damaging the midvein. The first leaf under the wounding one was cut off before wounding as unwounded sample, frozen in liquid nitrogen and stored at −80 °C before extracting crude protein.
For the wounding of rice leaves, the second leaves from apex of 40-day-old T2 rice plants grown in greenhouse were wound according to the method described by Xu, et al. [28] wherein the leaves were cut at intervals of 1 cm perpendicularly to veins by a scissor along both edges of the leaf blade without damaging the midvein, and the leave samples were harvested 24 h after wounding. The first leaf under the wounding one was cut off before wounding as unwounded sample, frozen in liquid nitrogen and stored at −80 °C before extracting crude protein.

4.5. Analysis of RNA Level by qRT-PCR

The second leaf from the apex of 40-day-old T3 transgenic tobacco was sampled. Total RNA was extracted using EasyPure™ Plant RNA Kit (TransGen, Beijing, China). The RNA was reverse transcribed to cDNA using TransScript® II First-Strand cDNA Synthesis SuperMix Kit (TransGen, Beijing, China), and qRT-PCR was performed using the TransStart® Green qPCR SuperMix Kit from Beijing TransGen Biotech Co. Ltd. (Beijing, China). The primers used for qRT-PCR were designed according to the sequences of the GUS gene from vector pBI221 (AF502128.1) and tobacco Actin gene (U60495.1). The sense primer for GUS gene is S-GUS: 5′-CCAACTCCTACCGTACCTC-3′, and the antisense primer is A-GUS: 5′-TCGAAACCAATGCCTAAA-3′. The sense primer for Actin gene is S-ACTIN: 5′-AAGGGATGCGAGGATGGA-3′, and the antisense primer is A-ACTIN: 5′-CAAGGAAATCACCGCTTTGG-3′. The instrument of ABI 7300 was used. The reaction volume is 20 μL. The amplification program was: 95 °C 2 min for pre-denature, then 40 cycles of 95 °C 10 s for denature, 55 °C 30 s for annealing, 72 °C 31 s for elongation were followed. The fluorescence data was collected at the end of elongation and analyzed by the 2−ΔΔCt method [29].

4.6. Analysis of the Transcription Start Sites by 5′ Rapid Amplification of cDNA Ends (5′-RACE)

According to the sequences of maize Wip1 gene (X71396.1) and GUS gene from vector pBI221 (AF502128.1), two specific primers, wip1-R: 5′-CAAAAAGGACTGCGACCCCGTCTG-3′ and GUS-R: 5′-GTTCTGCGACGCTCACACCGATACC-3′, were designed. wip1-R was used to amplify the 5′ end of Wip1 mRNA in maize. GUS-R was used to amplify the 5′ end of GUS mRNA in transgenic tobacco. Total RNA was extracted from leaves of 10-day-old maize inbred line Z31 and 30-day-old transgenic tobacco plants using EasyPure™ Plant RNA Kit (TransGen, Beijing, China). The generation of RACE-ready cDNA and rapid amplification of cDNA end were performed strictly according to the user manual of SMARTer™ RACE cDNA Amplification Kit (Clontech, Mountain View, CA, USA).

4.7. GUS Activity Assay and Histochemical Staining

T2 transgenic tobacco seeds were germinated on MS medium plates with 30 mg/L Hygromycin B to eliminate the non-transgenic plants. The surviving 14-day-old seedlings were immersed in GUS staining solution [48] and incubated overnight at 37 °C. Then, the samples were de-pigmented with 70% ethanol at 37 °C until the chlorophyll had completely disappeared.
For quantitative analysis, young leaves of 40-day-old T3 tobacco seedlings and 20-day-old T3Arabidopsis seedlings were homogenized in 500 μL extraction buffer (50 mM sodium phosphate buffer, pH 7.0, 10 mM EDTA, 0.1% Triton X-100, 10 mM β-mercaptoethanol) and clarified by centrifugation at 13,000 rpm for 10 min at 4 °C. 10 μL supernatant was used to assay the GUS activity in 100 μL extraction buffer containing 2 mM 4-methylumbelliferyl-β-D-glucuronide. Fluorescence was measured using Skanlt 2.4.3 RE for Varioskan Flash (Thermo, Waltham, MA, USA). Protein concentration was determined using the Model 680 spectrophotometer (Bio-Rad, Philadelphia, PA, USA) using the Coomassie Brilliant Blue method [49].

5. Conclusions

We used standard methods to study the function of the maize Wip1 promoter and found that several truncated Wip1 promoters led to high GUS activity in transgenic plants, even though the GUS transcriptional level was relatively low. The high GUS activity may be ascribed to high translation efficiency. The activity of the truncated Wip1 promoter fragments were influenced by the adjacent 35S promoter, however they may also be used as strong constitutive promoters to drive transgenes in plant biotechnology when a 35S promoter exists nearby. This may be a way to develop new “strong constitutive” promoters.

Supplementary Information

ijms-14-23872-s001.pdf

Acknowledgments

We thank Jianmin Wan and Xin Zhang from Institute of Crop Sciences, Chinese Academy of Agricultural Sciences (Beijing, China) for the help on the rice transformation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fang, R.X.; Nagy, F.; Sivasubramaniam, S.; Chua, N.H. Multiple cis regulatory elements for maximal expression of the cauliflower mosaic virus 35S promoter in transgenic plants. Plant Cell 1989, 1, 141–150. [Google Scholar]
  2. Christensen, A.H.; Sharrock, R.A.; Quail, P.H. Maize polyubiquitin genes: Structure, thermal perturbation of expression and transcript splicing, and promoter activity following transfer to protoplasts by electroporation. Plant Mol. Biol 1992, 18, 675–689. [Google Scholar]
  3. Zhang, W.; McElroy, D.; Wu, R. Analysis of rice Act1 5′ region activity in transgenic rice plants. Plant Cell 1991, 3, 1155–1165. [Google Scholar]
  4. Cazzonelli, C.I.; McCallum, E.J.; Lee, R.; Botella, J.R. Characterization of a strong, constitutive mung bean (Vigna radiata L.) promoter with a complex mode of regulation in planta. Transgenic Res 2005, 14, 941–967. [Google Scholar]
  5. Stavolone, L.; Kononova, M.; Pauli, S.; Ragozzino, A.; de Haan, P.; Milligan, S.; Lawton, K.; Hohn, T. Cestrum yellow leaf curling virus (CmYLCV) promoter: A new strong constitutive promoter for heterologous gene expression in a wide variety of crops. Plant Mol. Biol 2003, 53, 663–673. [Google Scholar]
  6. Xie, Y.; Liu, Y.; Meng, M.; Chen, L.; Zhu, Z. Isolation and identification of a super strong plant promoter from cotton leaf curl Multan virus. Plant Mol. Biol 2003, 53, 1–14. [Google Scholar]
  7. Dynan, W.S. Modularity in promoters and enhancers. Cell 1989, 58, 1–4. [Google Scholar]
  8. Sunilkumar, G.; Mohr, L.; Lopata-Finch, E.; Emani, C.; Rathore, K.S. Developmental and tissue-specific expression of CaMV 35S promoter in cotton as revealed by GFP. Plant Mol. Biol 2002, 50, 463–474. [Google Scholar]
  9. Boyko, A.; Molinier, J.; Chatter, W.; Laroche, A.; Kovalchuk, I. Acute but not chronic exposure to abiotic stress results in transient reduction of expression levels of the transgene driven by the 35S promoter. New Biotechnol 2010, 27, 70–77. [Google Scholar]
  10. Yin, T.; Wu, H.; Zhang, S.; Lu, H.; Zhang, L.; Xu, Y.; Chen, D.; Liu, J. Two negative cis-regulatory regions involved in fruit-specific promoter activity from watermelon (Citrullus vulgaris S.). J. Exp. Bot 2009, 60, 169–185. [Google Scholar]
  11. Wilkie, G.S.; Dickson, K.S.; Gray, N.K. Regulation of mRNA translation by 5′- and 3′-UTR-binding factors. Trends Biochem. Sci 2003, 28, 182–188. [Google Scholar]
  12. Tran, M.K.; Schultz, C.J.; Baumann, U. Conserved upstream open reading frames in higher plants. BMC Genomics 2008, 9, 361. [Google Scholar]
  13. Wu, J.; Miller, B.L. Aspergillus asexual reproduction and sexual reproduction are differentially affected by transcriptional and translational mechanisms regulating stunted gene expression. Mol. Cell Biol 1997, 17, 6191–6201. [Google Scholar]
  14. Zheng, X.; Deng, W.; Luo, K.; Duan, H.; Chen, Y.; McAvoy, R.; Song, S.; Pei, Y.; Li, Y. The cauliflower mosaic virus (CaMV) 35S promoter sequence alters the level and patterns of activity of adjacent tissue-and organ-specific gene promoters. Plant Cell Rep 2007, 26, 1195–1203. [Google Scholar]
  15. Yang, Y.; Singer, S.D.; Liu, Z. Evaluation and comparison of the insulation efficiency of three enhancer-blocking insulators in plants. Plant Cell Tissue Organ Cult 2011, 105, 405–414. [Google Scholar]
  16. Qi, R.F.; Song, Z.W.; Chi, C.W. Structural features and molecular evolution of Bowman-Birk protease inhibitors and their potential application. Acta Biochim. Biophys. Sin 2005, 37, 283–292. [Google Scholar]
  17. Shitan, N.; Horiuchi, K.; Sato, F.; Yazaki, K. Bowman-birk proteinase inhibitor confers heavy metal and multiple drug tolerance in yeast. Plant Cell Physiol 2007, 48, 193–197. [Google Scholar]
  18. Qu, L.J.; Chen, J.; Liu, M.; Pan, N.; Okamoto, H.; Lin, Z.; Li, C.; Li, D.; Wang, J.; Zhu, G.; et al. Molecular cloning and functional analysis of a novel type of Bowman-Birk inhibitor gene family in rice. Plant Physiol 2003, 133, 560–570. [Google Scholar]
  19. Eckelkamp, C.; Ehmann, B.; Schopfer, P. Wound-induced systemic accumulation of a transcript coding for a Bowman-Birk trypsin inhibitor-related protein in maize (Zea mays L.) seedlings. FEBS Lett 1993, 323, 73–76. [Google Scholar]
  20. Rohrmeier, T.; Lehle, L. WIP1, a wound-inducible gene from maize with homology to Bowman-Birk proteinase inhibitors. Plant Mol. Biol 1993, 22, 783–792. [Google Scholar]
  21. Tiffin, P.; Gaut, B.S. Molecular evolution of the wound-induced serine protease inhibitor wip1 in Zea and related genera. Mol. Biol. Evol 2001, 18, 2092–2101. [Google Scholar]
  22. Maize Genetics and Genomics Database. Available online: http://www.maizegdb.org/ (accessed on 20 May 2010).
  23. Lescot, M.; Déhais, P.; Thijs, G.; Marchal, K.; Moreau, Y.; van de Peer, Y.; Rouzé, P.; Rombauts, S. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res 2002, 30, 325–327. [Google Scholar]
  24. Higo, K.; Ugawa, Y.; Iwamoto, M.; Korenaga, T. Plant cis-acting regulatory DNA elements (PLACE) database: 1999. Nucleic Acids Res 1999, 27, 297–300. [Google Scholar]
  25. Reese, M.G. Neural Network Promoter Prediction. Available online: http://www.fruitfly.org/seq_tools/promoter.html (accessed on 27 October 2010).
  26. Walker-Simmons, M.; Holländer-Czytko, H.; Andersen, J.K.; Ryan, C.A. Wound signals in plants: A systemic plant wound signal alters plasma membrane integrity. Proc. Natl. Acad. Sci. USA 1984, 81, 3737–3741. [Google Scholar]
  27. An, G.; Mitra, A.; Choi, H.K.; Costa, M.A.; An, K.; Thornburg, R.W.; Ryan, C.A. Functional analysis of the 3′control region of the potato wound-inducible proteinase inhibitor II gene. Plant Cell 1989, 1, 115–122. [Google Scholar]
  28. Xu, D.; McElroy, D.; Thornburg, R.W.; Wu, R. Systemic induction of a potato pin2 promoter by wounding, methyl iasmonate, and abscisic acid in transgenic rice plants. Plant Mol. Biol 1993, 22, 573–588. [Google Scholar]
  29. Livak, K.J.; Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2−ΔΔCt method. Methods 2001, 25, 402–408. [Google Scholar]
  30. Yoo, S.Y.; Bomblies, K.; Yoo, S.K.; Yang, J.W.; Choi, M.S.; Lee, J.S.; Weigel, D.; Ahn, J.H. The 35S promoter used in a selectable marker gene of a plant transformation vector affects the expression of the transgene. Planta 2005, 221, 523–530. [Google Scholar]
  31. Singer, S.D.; Cox, K.D.; Liu, Z. Both the constitutive cauliflower mosaic virus 35S and tissue-specific AGAMOUS enhancers activate transcription autonomously in Arabidopsis thaliana. Plant Mol. Biol 2010, 74, 293–305. [Google Scholar]
  32. Chen, X.; Wang, Z.; Gu, R.; Fu, J.; Wang, J.; Zhang, Y.; Wang, M.; Zhang, J.; Jia, J.; Wang, G. Isolation of the maize Zpu1 gene promoter and its functional analysis in transgenic tobacco plants. Plant Cell Rep 2007, 26, 1555–1565. [Google Scholar]
  33. Lü, S.; Gu, H.; Yuan, X.; Wang, X.; Wu, A.M.; Qu, L.; Liu, J.Y. The GUS reporter-aided analysis of the promoter activities of a rice metallothionein gene reveals different regulatory regions responsible for tissue-specific and inducible expression in transgenic Arabidopsis. Transgenic Res 2007, 16, 177–191. [Google Scholar]
  34. Ren, Y.; Zhao, J. Functional analysis of the rice metallothionein gene OsMT2b promoter in transgenic Arabidopsis plants and rice germinated embryos. Plant Sci 2009, 176, 528–538. [Google Scholar]
  35. Wilmink, A.; van de Ven, B.C.; Dons, J.J. Activity of constitutive promoters in various species from the Liliaceae. Plant Mol. Biol 1995, 28, 949–955. [Google Scholar]
  36. Finkelstein, R.; Lynch, T.; Reeves, W.; Petitfils, M.; Mostachetti, M. Accumulation of the transcription factor ABA-insensitive (ABI)4 is tightly regulated post-transcriptionally. J. Exp. Bot 2011, 62, 3971–3979. [Google Scholar]
  37. Park, S.H.; Yi, N.; Kim, Y.S.; Jeong, M.H.; Bang, S.W.; Choi, Y.D.; Kim, J.K. Analysis of five novel putative constitutive gene promoters in transgenic rice plants. J. Exp. Bot 2010, 61, 2459–2467. [Google Scholar]
  38. Caspar, T.; Quail, P.H. Promoter and leader regions involved in the expression of the Arabidopsis ferredoxin A gene. Plant J 1993, 3, 161–174. [Google Scholar]
  39. Gallie, D.R.; Sleat, D.E.; Watts, J.W.; Turner, P.C.; Wilson, T.M.A. Mutational analysis of the tobacco mosaic virus 5′-leader for altered ability to enhance translation. Nucleic Acids Res 1988, 16, 883–893. [Google Scholar]
  40. Gallie, D.R.; Walbot, V. Identification of the motifs within the tobacco mosaic virus 5′-leader responsible for enhancing translation. Nucleic Acids Res 1992, 20, 4631–4638. [Google Scholar]
  41. De Amicis, F.; Patti, T.; Marchetti, S. Improvement of the pBI121 plant expression vector by leader replacement with a sequence combining a poly (CAA) and a CT motif. Transgenic Res 2007, 16, 731–738. [Google Scholar]
  42. Morris, D.R.; Geballe, A.P. Upstream open reading frames as regulators of mRNA translation. Mol. Cell. Biol 2000, 20, 8635–8642. [Google Scholar]
  43. Singer, S.D.; Cox, K.D.; Liu, Z. Enhancer-promoter interference and its prevention in transgenic plants. Plant Cell Rep 2011, 30, 723–731. [Google Scholar]
  44. Aljanabi, S.M.; Martinez, I. Universal and rapid salt-extraction of high quality genomic DNA for PCR-based techniques. Nucleic Acids Res 1997, 25, 4692–4693. [Google Scholar]
  45. Horsch, R.; Fry, J.; Hoffman, N.; Eichholz, D.; Rogers, S.; Fraley, R. A simple and general method for transferring genes into plants. Science 1985, 227, 1229–1231. [Google Scholar]
  46. Clough, S.J.; Bent, A.F. Floral dip: A simplified method for Agrobacterium-mediated transformation of Arabidopsis thaliana. Plant J 1998, 16, 735–743. [Google Scholar]
  47. Goto, F.; Yoshihara, T.; Shigemoto, N.; Toki, S.; Takaiwa, F. Iron fortification of rice seed by the soybean ferritin gene. Nat. Biotechnol 1999, 17, 282–286. [Google Scholar]
  48. Jefferson, R.A.; Kavanagh, T.A.; Bevan, M.W. GUS fusions: Beta-glucuronidase as a sensitive and versatile gene fusion marker in higher plants. EMBO J 1987, 6, 3901–3907. [Google Scholar]
  49. Bradford, M.M. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem 1976, 72, 248–254. [Google Scholar]
Figure 1. The maize Wip1 promoter sequence. The start codon (ATG) and putative transcription start sites are indicated in larger letters. The putative TATA boxes are shadowed and italicized. The W-box and GC-box are shadowed only. The primers used for PCR amplification are underlined and given corresponding names. Triangle (▲) indicates the transcription start site of GUS in Wip1231::GUS line; rhombuses (♦) indicate the transcription start site of GUS in Wip1737::GUS line; asterisk (★) means the transcription start site of Wip1 in maize.
Figure 1. The maize Wip1 promoter sequence. The start codon (ATG) and putative transcription start sites are indicated in larger letters. The putative TATA boxes are shadowed and italicized. The W-box and GC-box are shadowed only. The primers used for PCR amplification are underlined and given corresponding names. Triangle (▲) indicates the transcription start site of GUS in Wip1231::GUS line; rhombuses (♦) indicate the transcription start site of GUS in Wip1737::GUS line; asterisk (★) means the transcription start site of Wip1 in maize.
Ijms 14 23872f1
Figure 2. The Wip1 promoters and their activities in transgenic tobacco and Arabidopsis plant leaves. (A) Schematics of the full-length and truncated Wip1 promoters. The numbers and/or letters in front of each schematic represent the names of the corresponding constructs; (B) The GUS activities of 40-day-old transgenic tobacco plants containing different constructs. Error bars represent the S.E. of n independent transgenic lines, n = 9 for 1737 and 1500; n = 16 for 1231 and 35S; n = 3 for 1191, 791, 491 and WT; and (C) The GUS activity of 20-day-old transgenic Arabidopsis plants. Error bars represent the S.E. of n independent transgenic lines, n = 6 for 1737 and 1500; n = 20 for 1231 and 35S; n = 3 for WT. ND indicates not determined.
Figure 2. The Wip1 promoters and their activities in transgenic tobacco and Arabidopsis plant leaves. (A) Schematics of the full-length and truncated Wip1 promoters. The numbers and/or letters in front of each schematic represent the names of the corresponding constructs; (B) The GUS activities of 40-day-old transgenic tobacco plants containing different constructs. Error bars represent the S.E. of n independent transgenic lines, n = 9 for 1737 and 1500; n = 16 for 1231 and 35S; n = 3 for 1191, 791, 491 and WT; and (C) The GUS activity of 20-day-old transgenic Arabidopsis plants. Error bars represent the S.E. of n independent transgenic lines, n = 6 for 1737 and 1500; n = 20 for 1231 and 35S; n = 3 for WT. ND indicates not determined.
Ijms 14 23872f2
Figure 3. Histochemical GUS staining of 14-day-old transgenic tobacco plants. T2 transgenic tobacco seeds were germinated on MS medium plates with 30 mg/L Hygromycin B to eliminate the non-transgenic plants. The surviving 14-day-old seedlings were immersed in GUS staining solution and incubated overnight at 37 °C. Then, the samples were de-pigmented with 70% ethanol at 37 °C until the chlorophyll had completely disappeared.
Figure 3. Histochemical GUS staining of 14-day-old transgenic tobacco plants. T2 transgenic tobacco seeds were germinated on MS medium plates with 30 mg/L Hygromycin B to eliminate the non-transgenic plants. The surviving 14-day-old seedlings were immersed in GUS staining solution and incubated overnight at 37 °C. Then, the samples were de-pigmented with 70% ethanol at 37 °C until the chlorophyll had completely disappeared.
Ijms 14 23872f3
Figure 4. The finely truncated Wip1 promoters and their activity in transgenic Arabidopsis and tobacco plant leaves. (A) Schematics of the finely truncated Wip1 promoters. The numbers or/and letters in front of each schematic indicate the names of the corresponding constructs; (B) The GUS activities of transgenic tobacco plants. Error bars represent the S.E of n independent transgenic lines, n = 8 for 1737, 1231 and 35S; n = 18 for 1231A, 1231C and 931; n = 3 for WT. ND indicates not determined; and (C) The GUS activities of transgenic Arabidopsis plants containing different constructs. Error bars represent the S.E. of n independent transgenic lines, n = 3 for WT, and n = 6 for the others.
Figure 4. The finely truncated Wip1 promoters and their activity in transgenic Arabidopsis and tobacco plant leaves. (A) Schematics of the finely truncated Wip1 promoters. The numbers or/and letters in front of each schematic indicate the names of the corresponding constructs; (B) The GUS activities of transgenic tobacco plants. Error bars represent the S.E of n independent transgenic lines, n = 8 for 1737, 1231 and 35S; n = 18 for 1231A, 1231C and 931; n = 3 for WT. ND indicates not determined; and (C) The GUS activities of transgenic Arabidopsis plants containing different constructs. Error bars represent the S.E. of n independent transgenic lines, n = 3 for WT, and n = 6 for the others.
Ijms 14 23872f4
Figure 5. GUS activity in transgenic tobacco leaves treated with wounding. The second leaves from apex of 40-day-old T3 tobacco plants grown in greenhouse were wounded using three different kinds of methods. GUS activities were assessed 24 h after wounding. A-NW and A-W mean non-wounded or wounded samples using the method described in Walker-Simmons, et al. [26]. B-NW and B-W mean non-wounded or wounded samples using the method described in An, et al. [27]. C-NW and C-W mean non-wounded or wounded samples using the method described in Xu, et al. [28]. Data was shown as average ± S.E. of 6 independent transgenic lines.
Figure 5. GUS activity in transgenic tobacco leaves treated with wounding. The second leaves from apex of 40-day-old T3 tobacco plants grown in greenhouse were wounded using three different kinds of methods. GUS activities were assessed 24 h after wounding. A-NW and A-W mean non-wounded or wounded samples using the method described in Walker-Simmons, et al. [26]. B-NW and B-W mean non-wounded or wounded samples using the method described in An, et al. [27]. C-NW and C-W mean non-wounded or wounded samples using the method described in Xu, et al. [28]. Data was shown as average ± S.E. of 6 independent transgenic lines.
Ijms 14 23872f5
Figure 6. GUS activity of different tissues of transgenic tobacco plants. Data was shown as average ± S.E. of n independent transgenic lines, n = 3 for 1737, 1500 and WT; n = 6 for 1231 and 35S.
Figure 6. GUS activity of different tissues of transgenic tobacco plants. Data was shown as average ± S.E. of n independent transgenic lines, n = 3 for 1737, 1500 and WT; n = 6 for 1231 and 35S.
Ijms 14 23872f6
Figure 7. Transcriptional level of GUS in transgenic tobacco leaves. Data was shown as average ± S.E. of 6 independent transgenic lines. Letters above the columns indicate statistically significant differences at p < 0.05 level. Relative transcript levels were calculated using the 2−ΔΔCt method [29] with Actin as a housekeeping gene.
Figure 7. Transcriptional level of GUS in transgenic tobacco leaves. Data was shown as average ± S.E. of 6 independent transgenic lines. Letters above the columns indicate statistically significant differences at p < 0.05 level. Relative transcript levels were calculated using the 2−ΔΔCt method [29] with Actin as a housekeeping gene.
Ijms 14 23872f7
Figure 8. 5′-RACE analysis of the transcription start sites. M, DNA molecular marker; (1) Wip1 in maize; (2) Wip1231::GUS line; and (3) Wip1737::GUS line.
Figure 8. 5′-RACE analysis of the transcription start sites. M, DNA molecular marker; (1) Wip1 in maize; (2) Wip1231::GUS line; and (3) Wip1737::GUS line.
Ijms 14 23872f8
Figure 9. GUS activity in tobacco plant leaves transformed with constructs p1300-35S-NOS or p1300-1231-NOS. The second leaves from apex of 40-day-old T3 tobacco plants grown in greenhouse were used. NW and W mean non-wounded or wounded samples using the method described in Walker-Simmons, et al. [26]. Data was shown as average ± S.E of n independent transgenic lines, n = 15 for p1300-1231-NOS; n = 13 for p1300-35S-NOS.
Figure 9. GUS activity in tobacco plant leaves transformed with constructs p1300-35S-NOS or p1300-1231-NOS. The second leaves from apex of 40-day-old T3 tobacco plants grown in greenhouse were used. NW and W mean non-wounded or wounded samples using the method described in Walker-Simmons, et al. [26]. Data was shown as average ± S.E of n independent transgenic lines, n = 15 for p1300-1231-NOS; n = 13 for p1300-35S-NOS.
Ijms 14 23872f9
Figure 10. GUS activity in transgenic rice leaves treated with wounding. The second leaves from apex of 40-day-old rice plants grown in a greenhouse were wounded using the method described in Xu, et al. [28]. NW and W mean non-wounded or wounded samples. Data was shown as average ± S.E. of 6 independent transgenic lines. Asterisks (**) means the significant difference at p < 0.01 level.
Figure 10. GUS activity in transgenic rice leaves treated with wounding. The second leaves from apex of 40-day-old rice plants grown in a greenhouse were wounded using the method described in Xu, et al. [28]. NW and W mean non-wounded or wounded samples. Data was shown as average ± S.E. of 6 independent transgenic lines. Asterisks (**) means the significant difference at p < 0.01 level.
Ijms 14 23872f10
Table 1. The sequence of the primers used in the paper.
Table 1. The sequence of the primers used in the paper.
PrimersSequence
F17375′-AACTGCAGGGCTCCGTTCTACTTGACT-3′
R17375′-CGGGATCCGGTCTCGGACGAGCTGTTCTT-3′
F12315′-AACTGCAGTTTGTTTTGGATTATAAT-3′
F11915′-AACTGCAGCGCGTCGCCGTCTCTTACTGT-3′
F7915′-AACTGCAGTTCTTTTTAGTTGTCGCTGGA-3′
F4915′-AACTGCAGGGGAAGATTATTACTATACAC-3′
R12315′-CGGGATCCTTCCTCGTGATCATTTCCG-3′
F1231A5′-AACTGCAGCCTTACATAGTGGTTGGT-3′
F1231B5′-AACTGCAGGCTTCTATTTTATCACAC-3′
F1231C5′-AACTGCAGGGCGTCCATTTGTATCTC-3′
R1231a5′-CGGGATCCATGGATGGATAGACGCTT-3′
R1231b5′-CGGGATCCGCCGAGATCTGAACATGC-3′
R1231c5′-CGGGATCCCCTCTCGTAGAAGCAAGC-3′
Table 2. The primers used to amplify different truncated promoters.
Table 2. The primers used to amplify different truncated promoters.
The amplified fragmentForward PrimerReverse Primer
Wip1500F1231R1737
Wip1231F1231R1231
Wip1191F1191R1737
Wip791F791R1737
Wip491F491R1737
Wip1231AF1231AR1231
Wip931F1191R1231
Wip1231CF1231CR1231
Wip1231aF1231R1231a
Wip1231bF1231R1231b
Wip1231cF1231R1231c
Wip1231AaF1231AR1231a
Wip1231AbF1231AR1231b
Wip1231AcF1231AR1231c
Wip1231BaF1231BR1231a
Wip1231BbF1231BR1231b
Wip1231BcF1231BR1231c
Wip1231CaF1231CR1231a
Wip1231CbF1231CR1231b
Wip1231CcF1231CR1231c

Share and Cite

MDPI and ACS Style

Zhang, S.; Lian, Y.; Liu, Y.; Wang, X.; Liu, Y.; Wang, G. Characterization of a Maize Wip1 Promoter in Transgenic Plants. Int. J. Mol. Sci. 2013, 14, 23872-23892. https://doi.org/10.3390/ijms141223872

AMA Style

Zhang S, Lian Y, Liu Y, Wang X, Liu Y, Wang G. Characterization of a Maize Wip1 Promoter in Transgenic Plants. International Journal of Molecular Sciences. 2013; 14(12):23872-23892. https://doi.org/10.3390/ijms141223872

Chicago/Turabian Style

Zhang, Shengxue, Yun Lian, Yan Liu, Xiaoqing Wang, Yunjun Liu, and Guoying Wang. 2013. "Characterization of a Maize Wip1 Promoter in Transgenic Plants" International Journal of Molecular Sciences 14, no. 12: 23872-23892. https://doi.org/10.3390/ijms141223872

Article Metrics

Back to TopTop