Botanical Studies (2010) 51: 277-292.                                                REVIEW PAPER
An update on the progress towards the development of marker-free transgenic plants
Chandrama Prakash UPADHYAYA*, Akula NOOKARAJU, Mayank Anand GURURANI, Devanshi Chandel UPADHYAYA, Doo-Hwan KIM, Se-Chul CHUN, and Se Won PARK*
Department of Molecular Biotechnology, College of Life and Environmental Sciences, Konkuk University, 1 Hwayang-dong, Gwangjin-gu, Seoul 143 701, Republic of Korea
(Received June 19, 2009; Accepted December 16, 2009)
ABSTRACT. Selection marker genes (SMGs) have been commonly used in genetic transformation of plants for efficient selection of transformed cells, tissue, or regenerated shoots. In the majority of cases, the selec­tion is based on antibiotic or herbicide resistance. The presence of such genes within the environment or in the food supply might pose an unpredictable hazard to the ecosystem and to human health; therefore research has been initiated to develop an efficient marker-free transgenic system. Various techniques have been devel­oped in recent years to generate marker free transgenic plants and to eliminate marker genes from transgenics. These include site-specific recombination, homologous recombination, transposition, transient co-integration of the marker gene, and a co-transformation-segregation approach, but success has been limited to only a few plant species. Transgenic technology could become more reliable with the improvement of existing marker gene removal strategies and the development of novel approaches for plant genome manipulation. This review describes the contemporary strategies deployed to generate marker-free transgenic plants and marker gene removal, the merits and shortcomings of different approaches, and possible directions for future research pro­grammes.
Keywords: Bio-safety; Intra-chromosomal recombination; Marker-free; Positive selection; Selection marker gene; Site-specific recombination.
Abbreviations: SMG, Selection marker gene; GOI, Gene of interest; Lox, Locus of crossing-over; Cre, Cyclization recombination; HR, Homologous recombination; RB, Right border of T-DNA; LB, Left border of T-DNA; IGT, Identical gene terminator; TBS, Transformation booster sequence; ptDNA, Plastid DNA.
CONTENTS
INTRODUCTION.............................................................................................................................................................278
STRATEGIES TO DEVELOP MARKER-FREE TRANSGENICS.................................................................................278
Avoiding the use of selectable marker genes....................................................................................................................278
Positive selection system................................................................................................................................................279
Negative selection system...............................................................................................................................................280
Abiotic stress related genes as selection markers............................................................................................................280
STRATEGIES TO ELIMINATE MARKER GENES FROM NUCLEAR GENOME...................................................280
Co-transformation of two transgenes.............................................................................................................................280
Site specific recombination systems...............................................................................................................................281
Transposon-based marker excision methods..................................................................................................................284
STRATEGIES TO ELIMINATE MARKER GENES FROM THE PLASTID GENOME............................................285
Homologous recombination system...............................................................................................................................285
Site specific recombination system................................................................................................................................286
Transient co-integration of marker gene.........................................................................................................................286
*Corresponding authors: E-mail: prakash1@konkuk.ac.kr, Tel: +82-02-2049-6256, Fax: +82-02-447-7030 (C. P. Upadhyaya); E-mail: sewpark@konkuk.ac.kr, Tel: +82-02-450-3739, Fax: +82-02-447-7030 (S.W. Park).
278                                                                                                                                                             Botanical Studies, Vol. 51, 2010
Co-transformation-segregation.....................................................................................................................................................287
SPATIAL AND TEMPORAL EXPRESSION OF SELECTION MARKER GENE...............................................................287
CONCLUDING REMARKS......................................................................................................................................................287
LITERATURE CITED................................................................................................................................................................287
INTRODUCTION
The advent of recombinant DNA technology provided an impetus to plant genetic transformation research. Significant progress has been made using genetic engineering methods to produce transgenic plants with useful agronomic traits. Plant transformation techniques are based on the delivery, integration, and expression of defined genes in the plant cells, which can be grown to generate transgenic plants (Liu et al., 2006; Chen et al., 2008). In all plant transformation systems, to generate a substantial number of non-chimeric transformants, genes conferring resistance to selective agents such as antibiotics or herbicides are widely employed to selected transformants. With a few exceptions in which the selectable marker genes are used as the genes of interest, the marker genes generally have no agronomic value after the selection events. Moreover, once the selection marker trait has been introduced into the germplasm of a plant species, retention of this trait in the genome may be dubious. Hence, an alternative marker system must be used to incorporate subsequent transgenes. Consumer and environmental groups have expressed concern over the use of antibiotic/herbicide resistance genes from ecological and food safety perspectives. The horizontal and vertical transfer of the marker gene through the pollen and pollinators is the major environmental concern raised by the Europeans and several other communities. In order to answer these issues, one has to introduce as few foreign sequences, in addition to the gene of interest, as possible. The elimination of marker genes will not only appease some potential environmental and consumer concerns, but also remove technical barriers to plant genetic transformation (Bevan et al., 1983; Herrera-Estrella et al., 1983). Therefore, studies to avoid marker genes or to eliminate them after use have been initiated and a growing number of methods are under development (Lutz et al., 2007; Sundar et al., 2008; Rukavtsova et al., 2009). Marker-free transgenic research relies on either complete avoidance of using any marker gene or using screenable markers (genes which do not possess any potential harmful activities) to develop marker-free transgenic plants or to get rid of the selectable marker genes before transgenic plants are introduced into the fields. The elimination strategy is based on the excision of the selection marker gene out of the integrated transgene after successful selection, using several approaches like site-specific recombination, transposition, or homologous recombination (HR). However, the plastid genome (plastome or ptDNA) is an established target for plant genetic engineering for enhanced expression of foreign
proteins. Excision of the marker gene from the plastid genome is particularly important, as these are present as several thousands of copies per cell. These recombination strategies are also being used to develop marker-free transcriptomic plants. In the following discussion, these strategies are described in detail.
STRATEGIES TO DEVELOP MARKER-FREE TRANSGENICS
Avoiding the use of selectable marker genes
The simplest way to eliminate marker genes from transgenics is to avoid their use in the transformation of plants. De Vetten et al. (2003) first reported the transformation of potato cv. Kanico using Agrobacterium tumefaciens strain AGL0 and no selection marker genes. The maximum transformation efficiency obtained was 4.5% as confirmed by PCR. The only problem observed with this system was the presence of vector backbone sequences in most of the transformants, which is a character as undesirable as selection marker genes. The backbone integration study is most desirable in this method. The same strategy was successfully used by Jia et al. (2006) for producing marker-free transgenic tobacco plants at a frequency of 15%. Kim et al. (2007) reported transgenic potato plants without any selection marker gene with chloroplast targeted SOD and APx genes under oxidative stress inducible promoter with an average transformation efficiency of 2.2%. Recently, Li et al. (2009) reported marker-free transgenic tobacco with a transformation efficiency of 4%, among which 65% plants were found to be stable transformants in the T1 generation. They further described that the transgenic plants generated through no antibiotic or herbicidal selection were segregated in a Mendelian 3:1 ratio and that one-third of them were chimeric or escapes. Recently, marker-free transgenic potato plants resistant to oxidative stress (Ahmad et al., 2008) and to Potato virus X and Potato virus Y (Bai et al., 2009) were developed using this approach. In both these cases, the transgenic event was confirmed by PCR and by challenging the putative transformants to oxidative stress and to Potato virus X and Y, respectively. Similarly, Rukavtsova et al. (2009) constructed a novel plant transformation vector by removing the nptll marker gene conferring resistance to kanamycin from the vector together with its promoter and polyadenylation signal of the Agrobacterial nopaline synthase gene.
In spite of several advantages, some disadvantages are associated with this method. There is no control over
UPADHYAYA et al. — Strategies to develop marker-free transgenics                                                                                                279
selective growth of the transformants, and the researcher has to screen several putative transformants to confirm the integration of the transgene(s), which is expensive and time consuming. Uncontrolled regeneration of a large number of chimeric plants is another drawback in the case of selection without the use of antibiotics as observed in
tobacco (Li et al., 2009).
Positive selection system
Some biochemical marker genes enable the identifica­tion and selection of transformed cells without injury or death of the non-transformed cells. In this method, the selection marker genes code for enzymes that give the transformed cells a capacity to metabolize specific nutrient substances not usually metabolized by normal plants (Figure 1). This fact will provide the transformed cells an advantage over the non-transformed ones. Addition of this substance in the culture medium, as nutrient source dur­ing the regeneration process, allows normal to enhanced growth and differentiation of transformed cells while non-transformed cells will not be able to grow and regener­ate. Use of GUS, manA, and xylA genes are some of the popular positive selection systems.
The GUS gene. The ^-glucuronidase (GUS) gene isolated from E. coli (Jefferson et al., 1986) is to date the most widely used reporter gene in genetically modi­fied plants. Agronomically, the E. coli GUS gene has no value to crop plants. In this system of positive selection, a glucuronide derivative of benzyladenine (benzylad-enine N-3-glucuronide) is used as selection agent. The glucuronide supplemented in the selection medium is hy-drolyzed by the GUS enzyme produced in the transformed cells, forming active cytokinin (benzyladenine). The re­leased cytokinin acts as a stimulator for transformed cell regeneration, while non-transformed cell development is slow as the inactive form of the selective agent does not influence the non-transformed cells. This system was suc­cessfully used to effectively recover transgenic plants of tobacco (Joersbo and Okkels, 1996; Okkels et al., 1997).
The manA gene. Phosphomannose isomerase (PMI, EC 5.3.1.8) is an enzyme that converts mannose-6-phos-phate to fructose-6-phosphate, an intermediate of glycoly-sis that positively supports the growth of plant cells. This gene helps the transformed plant cells to utilize mannose as a carbon source to grow and differentiate on media containing mannose as a selection agent. In 1984, the gene coding for phosphomannose isomerase (manA or pmi) was first isolated from E. coli by Miles and Guest. How­ever, its first application in plants as a selection marker gene was reported by Joersbo et al. (1998). Though most plant species are sensitive to mannose, a few species— especially dicots like carrot, tobacco, sweet potato and legumes—have shown a considerable insensitivity to­wards this sugar. Plants like sugar beet, maize, wheat, oat, barley, tomato, potato, sunflower, rapeseed, pea, onion, and rice are extremely sensitive and have been success­fully transformed using pmi as a positive selection marker
Figure 1. T-DNA of plant transformation vector harboring positive section marker genes, GUS, P-glucuronidase; pmi, phosphomannoisomerase; xylA, Xylose isomerase A; DOGR1, 2-deoxyglucose-6-phosphate phosphatase; AtTPS1, trehalose-6-phophate synthase; used for producing marker-free transgenic plants.
(Joersbo et al., 1998, 2000; Negrotto et al., 2000; Wang et al., 2000; Bhalla-Sarin et al., 2004; Gadaleta et al., 2006;
Ku et al., 2006; Aswath et al., 2006; Penna et al., 2008).
Some plant transformation protocols that use the positive selection with pmi were found to be at least 10 times more efficient than the traditional protocols based on kanamycin selection (Wright et al., 2001). Jain et al. (2007) reported a maximum of 56% efficiency for stable transformants using mannose selection in sugarcane (Saccharum spp.).
The xylA, DOGR1 and AtTPS1 genes. The xylA gene isolated from Thermoanaerobacterium thermosul-furogenes or Streptomyces rubiginosus (Haldrup et al., 1998a) encodes for enzyme xylose isomerase, which catalyzes the isomerization of D-xylose to D-xylulose. It also catalyzes the isomerization of glucose to fructose, and is thus also termed glucose isomerase. Transgenic cells expressing the xylA gene can utilize xylose as a sole carbohydrate source and proliferate while non-transgenic cells starve. Using this system, transgenic plants of potato, tobacco, and tomato were successfully selected in xylose-containing media (Haldrup et al., 1998b; 2001). However, a problem initially encountered when xylose was used in the selection medium was the induction of callus, which resembled the auxin effect. This problem can be solved by adding auxin efflux inhibitor, triiodobenzoic acid (TIBA) (Nissen and Minocha, 1993), an antiauxin, p-chlorophe-noxyisobutyric acid (PCIB) (Kasai and Bayer, 1995), or an ethylene inhibitor, aminoethoxyvinylglycine (AVG) (Hall et al., 1985), to the selection medium.
In another system, the DOGR1 gene isolated from yeast encoding 2-deoxyglucose-6-phosphate phosphatase (2-DOG-6-P) gives resistance to 2-deoxyglucose (2-DOG). In plant cells 2-DOG is converted into 2-DOG-6-phos-phate (2-DOG-6-P), which is a competitor of glucose-6-phosphate. When over-expressed in transgenic plants, it was used as a positive selection system for tobacco and potato plants (Kunze et al., 2001). Recently, a new posi­tive selection marker gene AtTPS1, encoding trehalose-6-phophate synthase, has been developed (Leyman et al., 2008). In this system, the selection agent used is the nontoxic and common sugar glucose. Wild-type non-transformed Arabidopsis thaliana plantlets germinated on glucose had small white cotyledons and remained petite due to stoppage of the photosynthetic mechanism by ex­ternal sugar while transgenic plants expressing AtTPS1 became insensitive to glucose. The selectable marker gene
280                                                                                                                                                         Botanical Studies, Vol. 51, 2010
which encodes trehalose-6-phophate synthase catalyzes the first reaction of the two-step trehalose synthesis. With glucose (6%) as selection agent, it was possible to distin­guish green and normal-sized transgenic plants from the non-transformed ones.
Negative selection system
The negative selection method places negative se­lectable marker genes next to a positive marker gene in the same construct and is a powerful means of creating marker-free transgenic plants. In this method, transformed plants are selected for the absence of negative selection marker genes under the selection pressure of a negative se­lection marker (Figure 2). This allows for the selection of marker-free transgenic plants without performing a large number of PCRs (Jeongmoo et al., 2004). Erikson et al. (2004) characterized a novel marker gene, dao1 encoding D-amino acid oxidase. It can be used as either positive or negative marker, depending on the substrate. Therefore, it is possible to apply negative selection after a positive selection using same marker gene, dao1 , by changing sub­strates D-alanine or D-serine to D-isoleucine or D-valine.
Indoleacetic acid hydrolase (IAAH) encoded by the tms2 gene converts naphthaleneacetamide (NAM) to a potent auxin, naphthaleneacetic acid (NAA), which in­hibits seedling growth. The tms2 gene was first used as conditional selection marker gene in tobacco (Depicker et al., 1988) and in Arabidopsis (Karlin-Neumann et al., 1991). Other conditional markers proven to be effective in dicots are aux2 in cabbage (Beclin et al., 1993), the HSV-tk gene in tobacco (Czako and Marton, 1994), a bacte­rial cytochrome P450 mono-oxygenase gene (CPMO) in tobacco (O'Keefe et al., 1994) and Arabidopsis (Tissier et al., 1999), and codA in Arabidopsis (Kobayashi et al., 1995) and tobacco (Schlaman and Hooykaas, 1997). So far, the cytochrome P450 and codA are the only genes used as conditional negative selectors in monocots. Both have been proven effective in barley (Koprek et al., 1999), and only cytochrome P450 was found effective in rice (Chin et al., 1999).
Abiotic stress related genes as selection markers
This strategy is based on the fact that various genes encode proteins that protect plants from environmental stresses like drought, salt, and oxidative stress. Many such genes characterized in Arabidopsis and in several agro-nomically important crops, can be used for the develop­ment of marker-free transgenic plants. The gene for salt tolerance can be incorporated into the plant cells without the selection marker gene. After transformation, the tissue grows under the pressure of salt stress, and the explants growing well without any deformities are selected. Hence, use of the selection marker is unnecessary as the gene it­self can be used as selection marker to select the transfor-mants.
Bouchabke-Coussa et al. (2008) reported the ESKIMO1 gene involved in plant water economy, cold acclimation,
Figure 2. T-DNA of plant transformation vector harboring nega­tive section marker genes, dao1, D-amino acid oxidase; aux2, auxin synthase 2; HSV-tk, Herpes Simplex Virus thymidine kinase; CPMO, cytochrome P450 mono-oxygenase and coda, Cytosine deaminase used for producing marker-free transgenic plants.
and salt tolerance as a potential target for plant transfor­mation in an effort to increase abiotic stress resistance. At­tempts were made to introduce yeast ENAl gene encoding Na+-ATPase under the control of the CaMV35S promoter into BY2 cells, which were able to grow in modified LS medium containing 120 mM of LiCl (Yoshida and Shinmyo, 2000). Recently, an Arabidopsis AtHOL1 gene with high S-adenosyl-L-methionine-dependent methyl-transferase activity towards the thiocyanate ion (NCS-) was used as positive selection gene for an efficient select­able marker for screening the transformed seedlings of Arabidopsis (Midorikawa et al., 2009). Transgenic plants could show normal growth and development on a medium containing 3.0-5.0 mM potassium thiocyanate, which is toxic to non-transgenic plants. Veena et al. (1999) and Sanan-Mishra et al. (2005) explored the potential role of glyI gene (glyoxalase I) and PDH45 gene (pea DNA helicase45), respectively, in overcoming salinity stress. The glyI or PDH45 mRNA induced in response to high salt was over-expressed in transgenic tobacco plants that conferred salinity tolerance. These genes can be used as potential agents for the selection of transgenic plants on the medium supplemented with salts.
STRATEGIES TO ELIMINATE MARKER GENES FROM THE NUCLEAR GENOME
Co-transformation of two transgenes
Co-transformation is a simple and highly effective method to eliminate marker genes from the nuclear ge­nome of transgenic plants. Co-transformation involves the transformation of plant cells with two plasmids that target insertion at two different loci in the plant genome. One plasmid carries a selection marker gene while the other carries a gene of interest. Agrobacterium mediated co-transformation is achieved in three ways (Figure 3). In the first method, two different vectors are carried in two separate Agrobacterium strains (McKnight et al., 1987; Deblock and Debrouwer, 1991; De Neve et al., 1997); the second method involves two different vectors placed in the same Agrobacterium cell (De Framond et al., 1986; Daley et al., 1998; Sripriya and Veluthambi, 2008); while the third method uses two T-DNAs that can be borne by a single binary vector (two T-DNA system) (Komari et al., 1996; Xing et al., 2000; Matthew et al., 2001; McCormac et al., 2001; Miller et al., 2002). The selection marker gene
UPADHYAYA et al. — Strategies to develop marker-free transgenics                                                                                                281
Figure 3. Schematic representation of the strategy for marker gene removal via Agrobac-terium-mediated co-transformation with T-DNAs carrying marker and gene of interest (GOI). A, Marker and GOI are in separate vectors in separate Agrobacterium strains; B, Marker and GOI in separate vectors in single Agrobacterium strain; C, Marker and GOI borne by two T-DNAs in the same vector.
can be eliminated from the nuclear genome of the trans-genic plants at the time of segregation and recombination that occurs during sexual reproduction by selecting the transgenic lines carrying the gene of interest. Xue et al. (2003) generated transgenic barley plants producing high levels of cellulase (1,4-P-glucanase using a binary plasmid containing hph and cel-hyb1 in separate T-DNAs), where the selection marker gene (hph) was subsequently elimi­nated from transgenic lines through segregation of hph from synthetic thermo-tolerant hybrid cellulose (cel-hyb1) in T1 progeny. Marker-free double haploid transgenic rice plants were generated via Agrobacterium-mediated trans­formation with a "double-T-DNA" binary vector followed by anther culture, with an efficiency of 12.2% (Zhu et al.,
2007). Recently, Shiva Prakash et al. (2009) generated
marker-free transgenic plants of maize by co-bombard­ment of immature embryos with TDNAs containing a gene of interest, GUS, and an antibiotic selection marker, nptII, separately. Marker-free transgenic GUS positive lines were selected in the T1 generation after excision of the selection marker gene by recombination.
Though simple and effective, this method carries sev­eral unavoidable limitations. It is time consuming and compatible only for sexually propagated fertile plants. Secondly, the tight linkage between co-integrated DNAs may limit the efficiency of co-transformation. Indeed, the integration of marker gene and transgene is an indiscrimi­nate event. Both may integrate in the same loci and only a proportion of plants carrying the selectable marker will also carry the desired gene at an unlinked site (Scutt et al., 2002). Also, this method may not be suitable for species with very low transformation efficiency.
Site specific recombination systems
Recombination is a well known phenomenon in bio­logical systems that takes place between the two homolo-
gous DNA molecules. In site specific recombination, DNA strand exchange takes place between segments possessing only a limited degree of sequence homology (Coates et al., 2005). Three site specific recombination systems are well known and described for the elimination of selection marker genes. These are the Cre/lox site specific recombi­nation system (Dale and Ow, 1991), the FLP/FRT recom­bination system from Saccharomyces cerevisiae (Landy, 1989; Kilby et al., 1995; Lyznik et al., 1996), and the R/ RS recombination system from Zygosaccharomyces rouxii (Onouchi et al., 1995; Sugita et al., 2000).
The recombination sites are typically between 30 to 200 nucleotides in length and consist of two motifs with a partial inverted repeat symmetry. The recombinase binds to these motifs, which flank a central crossover sequence at which the recombination takes place (Figure 4). In a pioneering work, Dale and Ow (1991) and Qin et al. (1994) used the Cre recombinase of the E. coli bacteriophage P1 to remove a selection marker gene flanked by lox target sites from a transgenic locus in transformed tobacco. The unique ability of Cre to catalyze a crossover between di­rectly repeated lox sites flanking any fragment of DNA has been exploited to remove selectable marker genes from a number of transgenic plants. Basically, the strategy in­volves the generation of plants that express the cre gene and crossing them with plants in which the selection mark­er gene is flanked by lox sites. The marker gene is excised in the F1 generation, and the cre gene is segregated away in the subsequent generations. In this system, the selection marker gene can be eliminated either by re-transformation (Odell et al., 1990; Dale and Ow, 1991; Russell et al., 1992) or by crossing over (Bayley et al., 1992; Russell et al., 1992; Gleave et al., 1999; Hoff et al., 2001; Aru-mugam et al., 2007; Chakraborti et al., 2008) (Figure 5A). This system has been very well studied for a generation of marker-free transgenic plants of Arabiodopsis, maize, to-
282                                                                                                                                                         Botanical Studies, Vol. 51, 2010
Figure 4. Schematic representation of the mechanism of action of microbial recombinase systems, Cre-lox, Flp/Frt and RS/R. The gene construct contains the Lox/Flp/R recombinase sequences flanked by the selection marker gene while the gene of interest lies out­side the recombinase sequences (a). When the transgenics containing the Cre/FRT/RS is crossed with the other transgenic containing the Cre, FLP or R recombinase gene sequences (b), the recombination process starts (c) and the marker genes are removed (d) from chromosome. Remaining Cre recombinase sequence in chromosome is segregated out during the advancement of the generation. SMG; Selection marker gene, GOI; Gene of interest. Symbol cycle in the figure represent the recombinase (Cre, FLP or R) proteins.
Figure 5. Strategy of Cre-lox site-specific recombination system for excision of selection marker gene from transgenic plants. A, The lox gene construct and the Cre gene construct to be crossed with lox transgenics; B, The inducible system where the Cre gene is under the control of inducible promoter (IP).
UPADHYAYA et al. — Strategies to develop marker-free transgenics                                                                                              283
bacco, and rice (Zuo et al., 2001; Zhang et al., 2003; Yuan et al., 2004; Sreekala et al., 2005; Song et al., 2008). Jia et al. (2006) reported a self operating Cre-lox recombination system which uses a movement function-improved To­bacco Mosaic Virus (TMV) vector, m30B, to express Cre recombinase for elimination of the selection marker gene nptII from transgenic tobacco plants. Recently, Zhang et al. (2009) generated marker-free transgenic tomato plants with improved abiotic stress tolerance, employing a Cre/ loxP DNA recombination system in which AtIpk2b, an inositol polyphosphate 6-/3-kinase gene from Arabidopsis thaliana was constitutively overexpressed. The expres­sion of AtIpk2b conferred improved resistance to drought, cold, and oxidative stress in both sets of transgenic tomato plants.
Inducible system for Cre-lox recombination. Mark­er gene removal from transgenic plants using the Cre-lox recombination system of bacteriophage P1 requires re-transformation and out-crossing approaches which are laborious and time consuming (Dale and Ow, 1991). In order to initiate the Cre-lox recombination for removal of the marker gene, other novel inducible site-specific recombination systems have been applied (Figure 5B). However, several approaches were developed to overcome these shortcomings, including the use of some chemical inducers (Schaart et al., 2004; Yuan et al., 2004; Zhang et al., 2006) and heat shock (Wang et al., 2005; Cuellar et al., 2006). Marker-free transgenic tomato plants express­ing cry1Ac were obtained, using a chemically regulated Cre-lox -mediated site specific recombination system. The marker gene nptII was eliminated by two directly oriented loxP sites located between the CaMV35S promoter and a cry1Ac without promoter. Upon induction by P-estradiol (2 uM), the selection marker gene fused with Cre recom­binase and flanked by two lox sites was auto excised from the plant genome, thus generating marker-free transgenic plants (Zuo et al., 2001; Zhang et al., 2006). Recently, Lin et al. (2008) developed marker-free transgenic rice plants using a chemically-induced site specific recombi-nase system. They reported a chemical induction method for creating selectively terminable transgenic rice using benzothiadiazole (Bentazon), a herbicide used for weed control in major crops like rice, maize, wheat, cotton, and soybean. Similarly, Ma et al. (2009) reported on a marker-free transgenic tomato using a salicyclic acid-inducible Cre/loxP recombination system.
A Cre/loxP recombination system was used for marker elimination by Kuroda et al. (2003). Under it, a caseinolytic protease P1 (clpP1) was deleted, resulting in removal of the tobacco plant shoot system. Shan et al. (2006) used the heat-inducible system in a FLP/frt site-specific recombinase system. Under this, the expression of FLP was tightly under the control of the heat shock protein, hsp. Two different constructs were used, the frt-containing vector (pCAMBIA1300-betA-frt-als-frt) and the FLP expression vector (pCAMBIA1300-hsp-FLP-
hpt). Through the process of re-transformation, the FLP recombinase gene was introduced into transgenic (betA-
frt-als-frt) tobacco. In the re-transgenic plants after heat shock treatment, the marker gene als, flanked by two frt sites, could be excised by the inducible expression of FLP recombinase under the control of hsp promoter. A heat inducible strategy for the elimination of selection marker genes was also reported in vegetatively propagated plants like potato (Cuellar et al., 2006). Recently, Deb Roy et al. (2008) reported a heat inducible Cre/loxP site specific re­combination system to remove nptII gene from A. thaliana transgenic plants transformed with glyI gene. The cre gene was driven by the heat-inducible promoter (hsp), and the nptII gene is flanked by lox sequences. These inducible site-specific recombination systems can also be applied in vegetatively propagated crop plants for marker gene excision.
In spite of several advantages, these methods carry some limitations. Due to the prolonged presence of bacte­rial recombinases in the plants, there may be unwanted changes in the plant genome at the sites of transgene exci­sion.
Auto Excision strategy to eliminate marker gene. The earlier methods of excision like heat- and chemical-inducible systems are time consuming, and the marker gene is eliminated in the generation following segrega­tion. A novel and ideal method to eliminate the selection marker gene in a single generation has been developed. This method is referred to as the "auto excision strategy" in which the marker gene is easily eliminated in the T1 seeds of the transgenic plants. and the next generation of the transgenic plants becomes marker-free. An auto exci­sion system is controlled by pollen and/or seed specific promoters. It relies on floral specific promoters to regulate the expression of Cre recombinase to generate marker-free transgenic plants. Functionally characterized promoters are used in the strategy, and the system was successfully dem­onstrated in rice (Bai et al., 2008). This approach was me­diated by the Cre/loxP recombination system, and the cre gene was under the control of the floral specific promoter OsMADS45. Using this system, the marker gene nptII was completely removed from the T1 progeny of the rice with 37.5% efficiency (Bai et al., 2008). Verweire et al. (2007) developed homologous marker-free transgenic plants of Arabidopsis thaliana introducing a germline-specific auto-excision vector containing a cre gene under the control of a germline-specific promoter derived from APETALA1 and SOLO DANCERS genes from Arabidopsis thaliana. Using this method, the frequency of regeneration of marker-free transgenic lines obtained in Arabidopsis was 83-100%. Highly efficient auto-excision of selective markers has been successfully achieved in tobacco (Mlynarova et al., 2006; Luo et al., 2007; Kopertekh et al., 2009). Kopertekh et al. (2009) reported a developmentally regulated Cre-lox site-specific recombination system for excision of a bar marker gene by using seed-specific napin promoter.
In spite of all these advantages, an auto excision strat­egy has its limitations. It is, for example, successful only in flowering plants, and it is not useful for vegetatively
284                                                                                                                                                         Botanical Studies, Vol. 51, 2010
propagated plants like grapes, potato, or banana.
Intrachromosomal homologous recombination (ICR). This is based on the use of homologous recombina­tion to eliminate selectable marker genes after insertion. A 352 bp attP region of bacteriophage X is a target for three specific proteins that mediate integration and excision of the phage genome within the E. coli (Figure 6). The pro­cess of bacteriophage integration involves a phage-encod-ed X integrase and a bacterially encoded integration host factor (IHF). This strategy was successfully used in trans­genic tobacco to remove selection marker genes like gfp, nptll, and tms2 flanked by the attP region (Zubco et al., 2000). The marker gene excision was achieved by select­ing transformed tobacco calli initially on kanamycin-con-taining media and subsequently culturing on kanamycin-free media to allow for the loss of the nptll gene by ICR. The detection of ICR events was based on the acquisition of sensitivity to kanamycin and confirmed by the loss of a negative selection tms2 marker gene. ICR events in plants were found to be very rare, with only ten such events de­tectable in all of the cells of a 6-week-old tobacco plant (Puchta et al., 1995).
The ICR method of marker gene removal is relatively simple as it does not require the expression of a heterolo-gous recombinase. In addition, this technique does not require any sexual reproduction steps and could therefore be applied for vegetatively propagated plants too. This method requires transfer of transformed tissues to a selec­tion medium followed by transfer to a non-selection me­dium to allow the loss of the selection marker gene. Such lengthy propagation may increase the risk of somaclonal mutations (Ow, 2001). Further, this method has to be test­ed on a large number of plant species before its large scale implementation.
Multi-auto transformation (MAT) vector system.
The MAT vector system represents a sophisticated ap-
proach for the removal of marker genes from the nuclear genome of the transgenic plants (Ebinuma et al., 1997). This system involves placing a chosen transgene adjacent to a multigenic element flanked by reciprocal recombina­tion (RS) sites. A copy of the selectable isopentyl trans-ferase (ipt) gene from A. tumefaciens is inserted between these RS recombinase sites together with the yeast R recombinase gene, and this entire assembly is placed in T-DNA region. The MAT vector system allows the removal of the R recombinase gene along with the ipt gene. The system does not require sexual crosses for the removal of marker or recombinase genes and can also be used in vegetatively propagated plants. The only limitation is that the prolonged expression of recombinase in plant tissues may cause unwanted recombinase effects. This limitation can be overcome by the use of more recent version of the MAT vector system that allows a delay in the excision of the ipt and R recombinase genes by using the chemically-inducible glutathione S-transferase promoter from maize to drive R recombinase gene expression (Sugita et al., 2000). The one-step MAT vector system for removal of the selection marker gene has been successfully demonstrated in tobacco and hybrid aspen (Zuo et al., 2001), rice (Endo et al., 2002), and apricot (Noguera et al., 2006).
Transposon-based marker excision methods
Transposons or the "jumping genes" have been used as a tool to excise the marker sequence from the gene of interest. The strategy makes use of the Ac/Ds transposition system and is primarily based on the fact that the DNA sequences located in the Ds repeats can be translocated to excise along with the Ds element (Figure 7). This method involves Agrobacterium-mediated transformation fol­lowed by intragenomic relocation of the transgene of in­terest, and its subsequent segregation from the selectable marker in the progeny (Goldsbrough et al., 1993) or direct excision of the marker gene from the genome (Ebinuma et al., 1997). Both strategies were developed using the maize Ac/Ds transposable element, and the system could be adapted successfully to use similar autonomous trans-posable elements. The Ac transposase activity requires the expression of plant promoters (Honma et al., 1993). Ebinuma et al. (1997) reported the feasibility of this strat­egy in tobacco and eliminated the ipt marker gene from transgenic tobacco plants. Transgenic plants constitutively expressing the ipt gene elevated cytokinin to auxin ratios resulting in a loss of apical dominance and suppression of root formation, referred to as shooty phenotype. Trans­formed tobacco leaf discs with a T-DNA containing nptII and gus gene and a chimeric Ac element, which included a 35S-ipt gene, showed an extremely shooty phenotype. Upon sub-culturing these phenotypic distinct shoots, nor­mal shoots were developed, which indicated the deletion of ipt gene expression. Jin et al. (2003) reported on an Ac/ Ds transposon system for removal of the hpt selection marker gene to obtain marker-free transgenic plants in rice (Oryza sativa L.). The Ds element containing the gene of interest, bar, was constructed next to the selection marker
Figure 6. The intrachromosomal recombination system for marker gene removal from transgenic plants involving a phage-encoded X integrase and a bacterially encoded Integration Host Factor (IHF). TBS; Transformation booster sequence, SMG, Selection marker gene, GOI; Gene of interest, attP; attachment P region of bacteriophage.
UPADHYAYA et al. — Strategies to develop marker-free transgenics                                                                                             285
the drawbacks of this method.
STRATEGIES TO ELIMINATE MARKER GENES FROM THE PLASTID GENOME
The plastid genome is present in multiple copies in each cell, and this multiplies up to large numbers like 104 copies per cell, particularly in leaf tissues. This gives a considerably high level of transgene expression, which may be useful for applications requiring high concentra­tions of proteins like the engineering of drought resistance or the production of pharmaceuticals in plants (Daniell et al., 1998, 2002). Modification of the plastid genome rep­resents an attractive alternative to engineering of the plant nuclear genome for some applications. As the chloroplast genome is maternally inherited, it reduces—but does not eliminate—the possibility of transgene escape via pollina­tion into its wild relatives. Removal of marker genes from the engineered chloroplast genome is particularly impor­tant as their very high copy number could lead to high levels of unwanted marker gene products. Some strategies for removing selection marker genes from the chloroplast genome of transplastomic plants have been suggested.
Homologous recombination system
In the case of chloroplast transformation, transgene in­tegration occurs through homologous recombination. Re­moval of the marker gene is possible through homologous recombination between identical gene promoters or termi­nators flanking the gene of interest and selection marker gene (Figure 8). This technique was successfully demon­strated for the first time in the unicellular green alga Chla-mydomonas reinhardtii (Fischer et al., 1996) and in higher plants (Iamtham and Day, 2000) using recombinative ex­cision between the similar sets of promoter or terminator sequences in the transplastomic plants. After the removal of antibiotic selection, these excision events accumulated to a high frequency, leading to a homoplastic, marker-free state in approximately 25% of the transgenic lines in the subsequent generation (Iamtham and Day, 2000).
Dufourmantel et al. (2007) applied a variant of the homology-based marker excision strategy to produce a marker-free soybean and tobacco plants with strong herbicide tolerance. A plastid transformation vector carrying an aadA that disrupts an herbicide resistance gene was used. Initial selection was done using spectinomycin,
Figure 7. An Ac/Ds Transposon-based method for the removal of marker gene from nuclear genome of transgenic plants. The method is based on intragenomic relocation of GOI and its subsequent segregation from the selection marker gene in the progeny. GOI: Gene of interest.
gene hpt to get Ds-T-DNA. Rice plants were transformed by A. tumefaciens containing Ac-T-DNA and Ds-T-DNA, respectively. These two transformants crossed with each other to produce the F1 plant containing both Ac and Ds elements. F1 plant was then selfed to produce F2 progeny in which the T-DNA insert and transposed Ds element be­came segregated. Recently, a maize transposon Ac-based system was utilized for excision of marker gene from rice transformants. The transposon system was induced by salicylic acid to excise transposon along with marker gene (Charng et al., 2008). A T-DNA bearing the cry1B endotoxin gene flanked by Ac-Ds transposon termini and cloned in the 5' untranslated region of a gfp gene cassette along with the gene encoding AcTpase was successfully transferred to a japonica cultivar of rice, generating 68 in­dependent transformation events (Cotsaftis et al., 2002).
The major advantage of this strategy is that, the marker-free transgenic plants can easily be screened at the T0 generation, avoiding the need for sexual reproduction and indicating the applicability of the strategy to the vegeta-tively propagated crops also. In spite of all the advantages, a few limitations are inevitable, like the very low regen­eration frequency of marker-free transgenic plants and the genomic instability of transgenic plants because of the continued presence of heterologous transposons (Scutt et al., 2002). The requirement of genetic crossing or segrega­tion for separating the transgene and the marker gene is a time consuming process and can thus be counted as one of
Figure 8. Schematic representation showing the mechanism of marker gene removal from chloroplast by homologous recom­bination. Selection marker gene is excised through homologous recombination between identical gene promoters or terminators flanking GOI and marker gene GOI; Gene of interest, IGT; Identical gene terminator, SMG; selection marker gene.
286                                                                                                                                                         Botanical Studies, Vol. 51, 2010
and the aadA marker gene was excised by homologous recombination, yielding a transplastomic plant carrying a complete herbicide resistance gene, 4-hydroxyphenylpyru-vate dioxygenase (HPPD).
Site specific recombination system
This system utilizes a two-step protocol. In the first step, construction of transplastomic plants carrying a marker gene flanked by two directly oriented recombinase target sites will be developed. In the second step, the marker will be removed by introducing a plastid-targeted recombinase in to the plant genome (Lutz and Maliga, 2007). So far, two recombinases (Cre and Int) were reported for excision of the marker gene from the plastid genome. The first method, the Cre-lox system, works similarly to nuclear transgenics. This system was successfully demonstrated in tobacco (Corneille et al., 2001; Hajdukiewicz et al., 2001). The system eliminates the selection marker gene flanked by two directly oriented lox sites. The nuclear cre gene can subsequently be removed by segregation in the seed progeny. A modified Cre-lox system is a highly efficient tool for obtaining marker-free transplastomic plants. Plastid transgene constructions contain a selection marker gene flanked by lox recombination sites. This method was found to be dependant on whether the cre gene was introduced into the nuclear genome of a transplastomic line by direct transformation or by sexual crossing (Corneille et al., 2001). In cases where cre was introduced by re-transformation, excision homologous recombination events between similar genetic elements present in the transgene construction were observed at efficiencies comparable to those of the Cre-mediated excision events observed by Iamtham and Day (2000). However, excision events by homologous recombination were not observed when the cre gene was introduced through sexual crosses.
The only disadvantage of a Cre recombinase-mediated marker gene removal from the chloroplast genome is that it requires crossing the transformed plants to remove the nuclear-encoded recombinase gene and its associated genetic marker. This system is thus applicable only to sexually propagated plants.
In the second method, the phiC31 phage integrase enzyme Int is used to insert the gene of interest into the ptDNA (Lutz et al., 2004). The marker gene will be flanked with directly oriented nonidentical phage attP (215 bp) and bacterial attB (54 bp) attachment sites, which are recognized by Int recombinase. The excision of the marker gene will be achieved after transformation of the nucleus with an int gene encoding plastid-targeted Int (Kittiwong-wattana et al., 2007).
Transient co-integration of the marker gene
This method involves incorporation of the marker gene and gene of interest in the tDNA together by two homolo­gous recombination events via the targeting sequences. Placing the marker gene outside the targeting region en­ables selection of a cointegrated structure that forms by re­combination via only one of the targeting sequences. When selection for antibiotic resistance is stopped, the second recombination event can take place, and the marker gene is lost (Figure 9). This strategy was first demonstrated by Klaus et al. (2004) by linking gene of interest to a missing photosynthetic genes (petA, rpoA). This method is more convenient if repeated transformation is done with variants of the same plastid gene, and this approach requires an ini­tial investment, obtaining a knockout mutant to use in the visually-aided complementation assay (Klaus et al., 2003).
Figure 9. Transient co-integration based marker removal from chloroplast genome of transplastomic plants. In this system the marker gene is placed outside the targeting region that enables selection for a cointe-grate structure and is later lost upon re­moval of antibiotic selection through second recombination. M/nptII: Marker, neomycin phosphotransferase-II, LF and RF: Left and right homologous fragments, GOI: Gene of interest, GUS: Reporter gene encoding P-glucuronidase, aadA: Gene for Spectino-mycin resistance.
UPADHYAYA et al. — Strategies to develop marker-free transgenics                                                                                             287
Co-transformation-segregation
This method of co-transformation of two transgenes works similarly to the way nuclear transgenics does. This strategy involves transformation with two plasmids that carry target insertions at two different ptDNA locations. Selection for the marker yields transplastomic clones that also carry an insertion of the non-selected gene. This ap­proach relies on the heteroplastomic ptDNA population obtained by cotransformation with two independently targeted genes. The feasibility of the approach was first shown in a unicellular alga, C. reinhardtii, with a single chloroplast (Kindle et al., 1991; Newman et al., 1991; Rof-fey et al., 1991). Later this approach was tried in tobacco to obtain marker-free plants that lack the antibiotic resis­tance gene and are resistant to the glyphosate herbicide (Carrer and Maliga, 1995; Ye et al., 2003).
SPATIAL AND TEMPORAL EXPRESSION
OF SELECTION MARKER GENE
This strategy involves the regulation of selection marker gene expression through an inducible promoter, which can be preferentially expressed, either temporarily or spatially at the time/site of transformation. This system allows the expression of a selection marker gene for only a limited period of time such as when transformants are selected without the expression of the markers in mature plant. This system was first successfully employed in tobacco for suppression of nptll gene driven by the wound-inducible promoter AoPRi (Ozcan et al., 1993). This promoter was expressed at the wound site of tobacco leaf discs, to al­low efficient selection of transformants. Mature leaf tissue showed very little expression of nptll and in some trans-formants it was virtually negligible.
CONCLUDING REMARKS
The ever-increasing demand for producing marker-free transgenic crops arose from various public concerns regarding the bio-safety of genetically modified (GM) crops. Many GM crops contain selection marker genes which provide resistance against commonly used antibiot­ics or herbicides, and a common public perception about such crops is that, these genes could be passed from food to bacteria in the guts of humans and animals. Several legislative attempts to address these issues have already been made by the governing bodies of some European and American nations. These issues led to an extensive inter­est among researchers for the development of systems to produce marker-free transgenic plants, for their wider consumer acceptance. The range of options used is wide, ranging from attempts to transform without using selec­tion markers to highly sophisticated transposon-based gene removal systems. Although, co-transformation of transgenes with selection marker genes is a widely popular technique to eliminate the undesired marker genes, it also has a few constraints. Among them is its restricted appli-
cation, as it can not be used for vegetatively propagated plants. Another novel technique to eliminate marker genes from the transgenic plants is site-specific recombination under the control of inducible promoters, which seems quite promising. Integration of foreign genes into the plas-tid genome prevents the spread of transgenes, as plastids are maternally inherited. However, because of the high copy number and the prokaryotic expression signals of the selection markers, it is advisable to remove the markers after the generation of transplastomic plants (Iamtham and Day, 2000). The thrust of future research will be develop­ing the most efficient systems for developing marker-free transgenic plants to address the concerns of ecologists and environmentalists regarding the bio-safety of genetically engineered plants. This research is in its infancy, and it is sure to expedite the refined study of plant systems, which will consequently increase the purview of biotechnology-based research in the near future.
Acknowledgements. The authors wish to thank the anonymous reviewers for their critical comments on the manuscript. Work in our laboratory focusing on the topic of this review is funded by a research grant from Konkuk University.
LITERATURE CITED
Ahmad, R., Y.H. Kim, M.D. Kim, M.N. Phung, W.I. Chung, H.S. Lee, S.S. Kwak, and S.Y. Kwon. 2008. Develop­ment of selection marker-free transgenic potato plants with enhanced tolerance to oxidative stress. J. Plant Biol. 51(6): 401-407.
Arumugam, N., V. Gupta, A. Jagannath, A. Mukhopadhyay, A.K. Pradhan, P.K. Burma, and D. Pental. 2007. A passage through in vitro culture leads to efficient production of marker-free transgenic plants in Brassica juncea using the Cre-lox system. Transgenic Res. 16: 703-712.
Aswath, C.R., S.Y. Mo, D.H. Kim, and S.W. Park. 2006. Agro-bacterium and biolistic transformation of onion using non-antibiotic selection marker phosphomannose isomerase. Plant Cell Rep. 25: 92-99.
Bai, X., Q. Wang, and C. Chu. 2008. Excision of a selective marker in transgenic rice using a novel Cre/loxP system controlled by a floral specific promoter. Transgenic Res. 17: 1035-1043.
Bai, Y., Z. Guo, X. Wang, D. Bai, and W. Zhang. 2009. Genera­tion of double-virus-resistant marker-free transgenic potato plants. Prog. Nat. Sci. 19: 543-548.
Bayley, C.C., M. Morgan, E.C. Dale, and D.W. Ow. 1992. Ex­change of gene activity in transgenic plants catalyzed by the Cre-lox site-specific recombination system. Plant Mol. Biol. 18: 353-361.
Beclin, C., F. Charlot, E. Botton, L. Jouanin, and C. Dore. 1993. Potential use of aux2 gene from Agrobacterium rhizogenes as a conditional negative marker in transgenic cabbage. Transgenic Res. 2: 4855.
288                                                                                                                                                         Botanical Studies, Vol. 51, 2010
Bevan, M.W., R.B. Flavell, and M.D. Chilton. 1983. A chimeric antibiotic resistance gene as a selectable marker for plant cell transformation. Nature 304: 184-187.
Bhalla-Sarin, N., P. Bhomkar, S. Debroy, N. Sharma, M. Sax-ena, C.P. Upadhyaya, A. Muthusamy, M. Pooggin, and T. Hohn. 2004. Transformation of Vigna mungo (blackgram) for abiotic stress tolerance using marker-free approach. New Directions for a Diverse Planet: Proceedings of the 4th International Crop Congress Brisbane, Australia, Oct., pp.
454.
Bouchabke-Coussa, O., M.L. Quashie, J.S. Redondo, M.N. For-tabat, C. Gery, A. Yu, D. Linderme, J. Trouverie, F. Granier, E. Teoule, and M.D. Tardif. 2008. ESKMO1 is a key gene involved in water economy as well as cold acclimation and salt tolerance. BMC Plant Biol. 8: 125.
Carrer, H. and P. Maliga. 1995. Targeted insertion of foreign genes into the tobacco plastid genome without physical linkage to the selectable marker gene. Biotechnology 13:
791-794.
Chakraborti, D., A. Sarkar, A. Hossain, H.A. Mondal, D. Schuer-mann, B. Hohn, B.K. Sarmah, and S. Das. 2008. Cre/lox system to develop selectable marker-free transgenic tobacco plants conferring resistance against sap sucking homopteran insect. Plant Cell Rep. 27: 1623-1633.
Charng, Y.C., K.T. Li, H.K. Tai, N.S. Lin, and J. Tu. 2008. An inducible transposon system to terminate the function of a selectable marker in transgenic plants. Mol. Breed. 21: 359­368.
Chen, H.J., I-C. Wen, G.J. Huang, W.C. Hou, and Y.H. Lin. 2008. Expression of sweet potato asparaginyl endopeptidase caused altered phenotypic characteristics in transgenic Ara-bidopsis. Bot. Stud. 49: 109-117.
Chin, H.G., M.S. Choe, S.H. Lee, S.H. Park, J.C. Koo, N.Y. Kim, J.J. Lee, B.G. Oh, G.H. Yi, S.C. Kim, H.C. Choi, M.J. Cho, and C.D. Han. 1999. Molecular analysis of rice plants harboring an Ac/Ds transposable element-mediated gene trapping system. Plant J. 19: 615-623.
Coates, B.S., R.L. Hellmich, and L.C. Lewis. 2005. Two differ­entially expressed ommochrome-binding protein-like genes (obp1 and obp2) in larval fat body of the European corn borer Ostrinia nubilalis. J. Insect Sci. 5: 19.
Corneille, S., K. Lutz, Z. Svab, and P. Maliga. 2001. Efficient elimination of selectable marker genes from the plastid ge­nome by the CRE-lox site-specific recombination system. Plant J. 27: 171-178.
Cotsaftis, O., C. Sallaud, J.C. Breitler, D. Meynard, R. Greco, A. Pereira, and E. Guiderdoni. 2002. Transposon-mediated generation of T-DNA- and marker-free rice plants express­ing a Bt endotoxin gene. Mol. Breed. 10: 165-180.
Cuellar, W., A. Gaudin, D. Solorzano, A. Casas, L. Nopo, P. Chudalayandi, G. Medrano, J. Kreuze, and M. Ghislain. 2006. Self-excision of the antibiotic resistance gene nptII using a heat inducible Cre-loxP system from transgenic po­tato. Plant Mol. Biol. 62: 71-82.
Czako, M. and L. Marton. 1994. The herpes simplex virus
thymidine kinase gene as a conditional negative-selection marker gene in Arabidopsis thaliana. Plant Physiol. 104: 1067-1071.
Dale, E.C. and D.E. Ow. 1991. Gene transfer with subsequent removal of the selection gene from the host genome. Proc. Natl. Acad. Sci. USA 88: 10558-10562.
Daley, M., V.C. Knauf, K.R. Summerfelt, and J.C. Turner. 1998. Co-transformation with one Agrobacterium tumefaciens strain containing two binary plasmids as a method for pro­ducing marker-free transgenic plants. Plant Cell Rep. 19: 489-496.
Daniell, H., M.S. Khan, and L. Allison. 2002. Milestones in
chloroplast genetic engineering: an environmentally friend­ly era in biotechnology. Trends Plant Sci. 7: 84-91.
Daniell, H., R. Datta, S. Varma, S. Gray, and S.B. Lee. 1998. Containment of herbicide resistance through genetic en­gineering of the chloroplast genome. Nature Biotech. 16:
345-348.
De Block, M. and D. Debrouwer. 1991. Two T-DNA's co-trans­formed into Brassica napus by a double Agrobacterium tu-mefaciens infection are mainly integrated at the same locus. Theor. Appl. Genet. 82: 257-263.
De Framond, A.J., E.W. Back, W.S. Chilton, L. Kayes, and M.D. Chilton. 1986. Two unlinked T-DNAs can transform the same tobacco plant cell and segregate in the F1 generation. Mol. Gen. Genet. 202: 125-131.
De Neve, S. de Buck, A. Jacobs, V.M. Montagu, and A. De-picker. 1997. T-DNA integration patterns in co-transformed plant cells suggest that T-DNA repeats originate from co-integration of separate T-DNAs. Plant J. 11: 15-29.
De Vetten, N., A.M. Wolters, K. Raemakers, I. van der Meer, R. ter Stege, E. Heeres, P. Heeres, and R. Visser. 2003. A transformation method for obtaining marker-free plants of a cross-pollinating and vegetatively propagated crop. Nat.
Biotechnol. 21: 439-442.
Deb Roy, S, M. Saxena, P.S. Bhomkar, M. Pooggin, T. Hohn, and N.B. Sarin. 2008. Generation of marker free salt toler­ant transgenic plants of Arabidopsis thaliana using the glyI gene and cre gene under inducible promoters. Plant Cell, Tissue & Org. Cult. 95: 1-11.
Depicker, A.G., A.M. Jacobs, and M.C. Van Montagu. 1988. A negative selection scheme for tobacco protoplast-derived cells expressing the T-DNA gene. Plant Cell Rep. 7: 63-66.
Dufourmantel, N., M. Dubald, M. Matringe, H. Canard, F. Gar-con, C. Job, E. Kay, Jean-Pierre Wisniewski, Jean-Marc Ferullo, B. Pelissier, A. Sailland, and G. Tissot. 2007. Gen­eration and characterization of soybean and marker-free tobacco plastid transformants over-expressing a bacterial 4-hydroxyphenylpyruvate dioxygenase which provides strong herbicide tolerance. Plant Biotechnol. J. 5: 118-133.
Ebinuma, H., K. Sugita, E. Matsunaga,and M. Yamakado. 1997. Selection of marker-free transgenic plants using the iso-pentenyl transferase gene. Proc. Natl. Acad. Sci. USA 94:
2117-2121.
Endo, S., K. Sugita, M. Sakai, H. Tanaka, and H. Ebinuma.
UPADHYAYA et al. — Strategies to develop marker-free transgenics                                                                                                    289
2002. Single-step transformation for generating marker-free transgenic rice using the ipt-type MAT vector system. Plant
Erikson, O., M. Hertzberg, and T. Nasholm. 2004. A conditional marker gene allowing both positive and negative selection in plants. Nat. Biotechnol. 22: 455-458.
Fischer, N., O. Stampacchia, K. Redding, and J.D. Rochaix. 1996. Selectable marker recycling in the chloroplast. Mol.
Gen. Genet. 251: 373-380.
Gadaleta, A., A. Giancaspro, A. Blechl, and A.B. Flanco. 2006. Phosphomannose-isomerase, PMI, as a selectable marker gene for durum wheat transformation. J. Cereal Sci. 43: 31­37.
Gleave, A.P., D.S. Mitra, S.R. Mudge, and B.A.M. Morris. 1999. Selectable marker-free transgenic plants without sexual crossing: transient expression of cre recombinase and use of a conditional lethal dominant gene. Plant Mol. Biol. 40: 223-235.
Goldsbrough, A.P., C.N. Lastrella, and J.I. Yoder. 1993. Transpo­sition mediated re-positioning and subsequent elimination of marker genes from transgenic tomato. Biotechnology 11: 1286-1292.
Hajdukiewicz, P.T., L. Gilbertson, and J.M. Staub. 2001. Mul­tiple pathways for Cre/lox-mediated recombination in plas-tids. Plant J. 27: 161-70.
Haldrup, A., M. Noerremark, and F.T. Okkels. 2001. Plant selec­tion principle based on xylose isomerase. In vitro Cell Dev. Biol.-Plant 37: 114-119.
Haldrup, A., S.G. Petersen, and F.T. Okkels. 1998a. The xylose
isomerase gene from Thermoanaerobacterium thermosulfu-rogenes allows effective selection of transgenic plant cells using D-xylose as the selection agent. Plant Mol. Biol. 37:
287-296.
Haldrup, A., S.G. Petersen, and F.T. Okkels. 1998b. Positive selection: a plant selection principle based on xylose isom-erase, an enzyme used in the food industry. Plant Cell Rep.
18: 76-81.
Hall, J.C., P.K. Kassi, M.S. Spencer, and W.H.V. Born. 1985. An evaluation of the role of ethylene in herbicidal injury in­duced by picloram or clopyralid in rapeseed and sunflower
Herrera-Estrella, L., M. De Block, E. Messens, J.P. Hernalsteens, V.M. Montagu, and J. Schell. 1983. Chimeric genes as dominant selectable markers in plant cells. EMBO J. 2(6): 987-995.
Hoff, T., K.M. Schnorr, and J. Mundy. 2001. A recombinase-me-diated transcriptional induction system in transgenic plants. Plant Mol. Biol. 45: 41-49.
Honma, M.A., B.J. Baker, and C.S. Waddell. 1993. High-frequency germinal transposition of DsALS in Arabidopsis.
Proc. Natl. Acad. Sci. USA 90: 6242-6246.
Iamtham, S. and A. Day. 2000. Removal of antibiotic resistance genes from transgenic tobacco plastids. Nat. Biotechnol. 18: 1172-1176.
Jain, M., K. Chengalrayan, A. Abouzid, and M. Gallo. 2007. Prospecting the utility of a PMI/mannose selection system for the brecovery of transgenic sugarcane (Saccharum spp. hybrid) plants. Plant Cell Rep. 26: 581-590.
Jefferson, R.A., S.M. Burgess, and D. Hirsh. 1986. beta-Glucuronidase from Escherichia coli as a gene-fusion
marker. Proc. Natl. Acad. Sci. USA 83(22): 8447-8451.
Jeongmoo, P., L. Young, K. Bong, and C. Won. 2004. Co-trans­formation using a negative selectable marker gene for the production of selectable marker gene-free transgenic plants. Theor. Appl. Genet. 109: 1562-1567.
Jia, H., Y. Pang, X. Chen, and R. Fang. 2006. Removal of the se­lectable marker gene from transgenic tobacco plants by ex­pression of Cre recombinase from a Tobacco Mosaic Virus vector through agroinfection. Transgenic Res. 15: 375-384.
Jin, W.Z., R.J. Duan, F. Zhang, S.Y. Chen, Y.R. Wu, and P. Wu. 2003. Application of Ac/Ds transposon system to genetate marker gene free transgenic plants in rice. Sheng Wu Gong
Cheng Xue Bao 19(6): 668-673.
Joersbo, M. and F.T. Okkels. 1996. A novel principle for selec­tion of transgenic plant cells: positive selection. Plant Cell Rep. 16: 219-221.
Joersbo, M., I. Donaldson, J. Kreiberg, S.G. Petersen, J. Brunst-edt, and F.T. Okkels. 1998. Analysis of mannose selec­tion used for transformation of sugar beet. Mol. Breed. 4: 111-117.
Joersbo, M., J.D. Mikkelsen, and J. Brunstedt. 2000. Relation­ship between promoter strength and transformation fre­quencies using mannose selection for the production of transgenic sugar beet. Mol. Breed. 6: 207-213.
Karlin-Neumann, G.A., J.A. Brusslan, and E.M. Tobin. 1991. Phytochrome control of the tms2 gene in transgenic Arabi-dopsis: a strategy for selecting mutants in the signal trans-
duction pathway. Plant Cell 3: 573-582.
Kasai, F. and D.E. Bayer. 1995. Effect of 2,4-Dichlorophen-oxyacetic acid, antiauxins, and metabolic perturbations on cytoplasmic and vascular pH of corn root tips measured by
in vivo 32 P-NMR. Pestic. Biochem. Physiol. 51: 160-161.
Kilby, N.J., J.D. Davies, M.R. Snaith, and J.A.H. Murray. 1995. FLP recombinase in transgenic plants: constitutive activity in stably transformed tobacco and generation of marked cell clones in Arabidopsis. Plant J. 8: 637-652.
Kim, M.S., H.S. Kim, H.N. Kim, Y.S. Kim, K.H. Baek, Y.I. Park, H. Joung, and J.H. Jeon. 2007. Growth and tuberiza-tion of transgenic potato plants expressing sense and anti-sense sequences of Cu/Zn superoxide dismutase from lily
chloroplasts. J. Plant Biol. 50: 490-495.
Kindle, K.L., K.L. Richards, and D.B. Stern. 1991. Engineer­ing the chloroplast genome: techniques and capabilities for chloroplast transformation in Chlamydomonas reinhardtii. Proc. Natl. Acad. Sci. USA 88: 1721-1725.
Kittiwongwattana, C., K.A. Lutz, M. Clark, and M. Maliga. 2007. Plastid marker gene excision by the phiC31 phage site-specific recombinase. Plant Mol. Biol. 64: 137-143.
290                                                                                                                                                         Botanical Studies, Vol. 51, 2010
Klaus, S.M.J., F.C. Huang, C. Eibl, H.-U. Koop, and T.J. Golds. 2003. Rapid and proven production of transplastomic tobac­co plants by restoration of pigmentation and photosynthesis. Plant J. 35: 811-821.
Klaus, S.M.J., F.C. Huang, T.J. Golds, and H.-U. Koop. 2004.
Generation of marker-free plastid transformants using a transiently cointegrated selection gene. Nat. Biotechnol. 22:
225-229.
Kobayashi, T.H.S., J. Stougaard, and H. Ichikawa. 1995. A con­ditional negative selection for Arabidopsis expressing a bacterial cytosine deaminase gene. Jpn. J. Genet. 70: 409­422.
Komari, T., Y. Hiei, Y. Saito, N. Murai, and T. Kumashiro. 1996. Vectors carrying two separate T-DNAs for co-transforma­tion of higher plants mediated by Agrobacterium tumefa-ciens and segregation of transformants free from selection markers. Plant J. 10: 165-174.
Kopertekh, L., I. Broer, and J. Schiemann. 2009. Developmen-tally regulated site-specific marker gene excision in trans­genic B. napus plants. Plant Cell Rep. 28: 1075-1083.
Koprek, T., D. McElroy, J. Louwerse, R. William-Carrier, and P.G. Lemaux. 1999. Negative selection systems for trans-genic barley (Hordeum vulgare L.): comparison of bacterial codA- and cytochrome P450 gene-mediated selection. Plant J. 19: 719-726.
Ku, J.J., Y.W. Park, and Y.D. Park. 2006. A non-antibiotic selec­tion system uses the phosphomannose-isomerase (PMI) gene for Agrobacterium-mediated transformation of Chi­nese cabbage. J. Plant Biol. 49: 115-122.
Kunze, I., M. Ebneth, U. Heim, M. Geiger, U. Sonnewald, and K. Herbers. 2001. 2-Deoxyglucose resistance: a novel selection marker for plant transformation. Mol. Breed. 7: 221-227.
Kuroda, H. and P. Maliga. 2003. The plastid clpP1 protease gene is essential for plant development. Nature 425: 86-89.
Landy, A. 1989. Dynamic, structural, and regulatory aspects of X site-specific recombination. Ann. Rev. Biochem. 58: 913-949.
Leyman, B., N. Avonce, M. Ramon, P. Van Dijck, J.M. Thev-elein, and G. Iturriaga. 2008. New selection marker for plant transformation. In: Methods in Molecular Biology. In P. Balbas and A. Lorence (eds.), Recombinant gene expression: Reviews and protocols. Humana Press Inc., To-
towa, NJ. Vol. 267, pp. 385-396.
Li, B., C. Xie, and H. Qiu. 2009. Production of selectable marker-free transgenic tobacco plants using a non-selection approach: chimerism or escape, transgene inheritance, and efficiency. Plant Cell Rep. 28: 373-386.
Lin, C., F. Jun, X. Xu, T. Zhao, J. Cheng, J. Tu, G. Ye, and Z.
Shen. 2008. A built-in strategy for containment of transgen-ic plants: Creation of selectively terminable transgenic rice.
PLoS ONE 3: 1818.
Liu, N.Y., W.J. Hsieh, H.C. Liu, and Y.Y. Charng. 2006. Hsa32, a phosphosulfolactate synthase-related heatshock protein, is not involved in sulfolipid biosynthesis in Arabidopsis. Bot. Stud. 47: 389-395.
Luo, K., H. Duan, D. Zhao, X. Zheng, W. Deng, Y. Chen, C.N.
Stewart, R. McAvoy, X. Jiang, Y. Wu, A. He, Y. Pei, and Y.
Li. 2007. 'GM-gene-deleter': fused loxP-FRT recognition sequences dramatically improve the efficiency of FLP or CRE recombinase on transgene excision from pollen and seed of tobacco plants. Plant Biotechnol. J. 5: 263-274.
Lutz, K, S. Corneille, A.K. Azhagiri, Z. Svab, and P. Maliga. 2004. A novel approach to plastid transformation utilizes the phiC31 phage integrase. Plant J. 37: 906-913.
Lutz, K.A. and P. Maliga. 2007. Construction of marker-free transplastomic plants. Curr. Opinion Biotechnol. 18: 107­114.
Lyznik, L.A., K.V. Rao, and T.K. Hodges. 1996. FLP-mediated recombination of FRT sites in maize genome. Nucleic Acids
Res. 24: 3784-3789.
Ma, B.G., X.Y. Duan, J.X. Niu, C. Ma, Q.N. Hao, L.X. Zhang, and H.P. Zhang. 2009. Expression of stilbene synthase gene in transgenic tomato using salicylic acid-inducible Cre/ loxP recombination system with self-excision of selectable marker. Biotechnol. Lett. 31: 163-169.
Matthew, P.R., M.B. Wang, P.M. Waterhouse, S. Thornton, S.J. Fieg, F. Gubler, and J.V. Jacobsen. 2001. Marker gene elimination from transgenic barley, using co-transformation with adjacent "twin T-DNAs" on a standard Agrobacterium transformation vector. Mol. Breed. 7: 195-202.
McCormac, A.C., M.R. Fowler, D.F. Chen, and M.C. Elliott. 2001. Efficient co-transformation of Nicotiana tabacum by two independent T-DNAs, the effect of T-DNA size and im­plications for genetic separation. Transgenic Res. 10: 143­155.
McKnight, T.D., M.T. Lillis, and R.B. Simpson. 1987. Segrega­tion of genes transferred to one plant cell from two separate Agrobacterium strains. Plant Mol. Biol. 8: 439-445.
Midorikawa, K., Y. Nagatoshi, and T. Nakamura. 2009. A selec­tion system for transgenic Arabidopsis thaliana using potas­sium thiocyanate as the selective agent and AtHOL1 as the selective marker. Plant Biotechnol. 26: 341-344.
Miles, J.S. and J.R. Guest. 1984. Nucleotide sequence and tran-scriptional start point of the phosphomannose isomerase gene (manA) of Escherichia coli. Gene 32: 41-48.
Miller, M., L. Tagliani, N. Wang, B. Berka, and D. Bidney. 2002. High efficiency transgene segregation in co-transformed maize plants using an Agrobacterium tumefaciens 2 T-DNA binary system. Transgenic Res. 11: 381-396.
Mlynarova, L., A.J. Conner, and J.P.H. Nap. 2006. Directed microspore-specific recombination of transgenic alleles to prevent pollen-mediated transmission of transgenes. Plant
Biotechnol. J. 4: 445-452.
Negrotto, D., M. Jolley, S. Beer, A.R. Wenck, and G. Hansen. 2000. The use of phosphomannose-isomerase as a select­able marker to recover transgenic maize plants (Zea mays L.) via Agrobacterium transformation. Plant Cell Rep. 19: 798-803.
Newman, S.M., N.W. Gillham E.H. Harris, A.M. Johnson, and J.E. Boynton. 1991. Targeted disruption of chloroplast
UPADHYAYA et al. — Strategies to develop marker-free transgenics                                                                                              291
genes in Chlamydomonas reinhardtii. Mol. Gen. Genet.
230: 65-74.
Nissen, P. and S.C. Minocha. 1993. Inhibition of 2,4-D by so­matic embryogenesis in carrot as explored by its reversal by difluoromethylornithine. Physiol. Plant 89: 673-680.
Noguera, S.L., C. Petri, and Burgos. 2006. Production of mark­er-free transgenic plants after transformation of Apricot cultivars. In F. Romojaro, F. Dicenta, P. Martinez-Gomez (eds.), XIII International Symposium on Apricot Breeding and Culture, Murcia, Spain.
O'Keefe, D.P., J.M. Tepperman, C. Dean, K.J. Leto, D.L. Erbes
and J.T. Odell. 1994. Plant expression of a bacterial cyto-chrome P450 that catalyzes activation of a sulfonylurea pro-
herbicide. Plant Physiol. 105: 473-482.
Odell, J., P. Caimi, B. Saucer, and S. Russel. 1990. Site directed -recombination in the genome of transgenic tobacco. Mol. Gen. Genet. 223:369-378.
Okkels, F.T., J. Ward, and M. Joersbo. 1997. Synthesis of cytoki-nin glucuronides for the selection of transgenic plant cells. Phytochemistry 46: 801-804.
Onouchi, H., R. Nishihama, M. Kudo, Y. Machida, and C. Ma-chida. 1995. Visualization of site-specific recombination catalyzed by a recombinase from Zygosaccharomyces rouxii in Arabidopsis thaliana. Mol. Gen. Genet. 247: 653-660.
Ow, D.W. 2001. The right chemistry for marker gene removal.
Nat. Biotechnol. 19: 115-116.
Ozcan, S., K. Freidel, A. Leuker, and M. Ciriacy. 1993. Glucose uptake and catabolite repression in dominant HTR1 mutants of Saccharomyces cerevisiae. J. Bacteriol. 175: 5520-5528.
Penna, S., M.B. Ramaswamy, and V.A. Bapat. 2008. Mannose based selection with Phosphomannose isomerase (PMI) gene as a positive selectable marker for rice genetic trans­formation. J. Crop Sci. Biotechnol. 11(4): 233-236.
Puchta, H., P. Swoboda, S. Gal, M. Blot, and B. Hohn. 1995.
Somatic intrachromosomal homologous recombination events in populations of plant siblings. Plant Mol. Biol. 28:
281-292.
Qin, M., C. Bayley, T. Stockton, and D.W. Ow. 1994. Cre recom-binase-mediated site-specific recombination between plant chromosomes. Proc. Natl. Acad. Sci. USA 91: 1706-1710.
Roffey, R.A., J.H. Golbeck, C.R. Hille, and R.T. Sayre. 1991. Photosynthetic electron transport in genetically altered photosystem II reaction centers of chloroplasts. Proc. Natl. Acad. Sci. USA 88: 9122-91266.
Rukavtsova, E.B., N.S. Zakharchenko, S.V. Pigoleva, A.A. Yukhmanova, E.N. Chebotareva, and Ya. I. Bur'yanov. 2009. Obtaining marker-free transgenic plants. Biochem. Bioph. 426: 143-146
Russell, S., J.L. Hoopes, and J.T. Odell. 1992. Directed excision of a transgene from the plant genome. Mol. Gen. Genet.
234: 49-59.
Sanan-Mishra, N., X.H. Phan, S.K. Sopory, and N. Tuteja. 2005. Pea DNA helicase 45 overexpression in tobacco confers high salinity tolerance without affecting yield. Proc. Natl.
Acad. Sci. USA 102: 509-514.
Schaart, J.G., F.A. Krens, K.T.B. Pelgrom, O. Mendes, and G.J.A. Rouwendal. 2004. Effective production of marker-free transgenic strawberry plants using inducible site-specif­ic recombination and a bifunctional selectable marker gene. Plant Biotechnol. J. 2: 233-240.
Schlaman, H.R.M. and P.J.J. Hooykaas. 1997. Effectiveness of the bacterial gene codA encoding cytosine deaminase as a negative selectable marker in Agrobacterium-mediated plant transformation. Plant J. 11: 1377-1385.
Scutt, C.P., E. Zubko, and P. Meyer. 2002. Techniques for re­moval of marker genes from transgenic plants. Biochimie 84: 1119-1126.
Shan, X.Y., B.L. Shan, and J.R. Zhang. 2006. Production of marker-free transgenic tobacco plants by Flp/frt recombina­tion system. Chinese J. Biotechnol. 22: 744-749.
Shiva Prakash, N., R. Bhojaraja, S.K. Shivbachan, G..G. Hari Priya, T.K. Nagraj, V. Prasad, V. Srikanth Babu, T.L. Jay-aprakash, S. Dasgupta, T.M. Spencer, and R.S. Boddupalli. 2009. Marker-free transgenic corn plant production through co-bombardment. Plant Cell Rep. 28: 1655-1668.
Song, H.Y., X.S. Ren, S.I. Jun, C.Q. Li, and M. Song. 2008. Construction of marker-free GFP transgenic tobacco by cre/ lox site-specific recombination system. Agri. Sci. China 7:
1061-1070.
Sreekala, C., L. Wu, K. Gu, D. Wang, D. Tian, and Z. Yin. 2005. Excision of a selectable marker in transgenic rice (Oryza sativa L.) using a chemically regulated Cre/loxP system.
Plant Cell Rep. 24: 86-94.
Sripriya, R. and R.K. Veluthambi. 2008. Generation of selectable marker-free sheath blight resistant transgenic rice plants by efficient co-transformation of a cointegrate vector T-DNA and a binary vector T-DNA in one Agrobacterium tumefa-
ciens strain. Plant Cell Rep. 27: 1635-1644.
Sugita, K., T. Kasahara, E. Matsunaga, and H. Ebinuma. 2000. A transformation vector for the production of marker-free transgenic plants containing a single copy transgene at high frequency. Plant J. 5: 461-469.
Sundar, I.K. and N. Sakhtivel. 2008. Advances in selectable marker genes for plant transformation. J. Plant Physiol. 165: 1698-1716.
Tissier, A.F., S. Marillonnet, V. Klimyuk, K. Patel, M.A. Torres, G. Murphy, and J.D.G. Jones. 1999. Multiple independent defective Suppressor-mutator transposon insertions in Arabidopsis: A tool for functional genomics. Plant Cell 11: 1841-1852.
Veena, V., S. Reddy, and S.K. Sopory. 1999. Glyoxalase I from Brassica juncea: molecular cloning, regulation, and its overexpression confer tolerance in transgenic tobacco under
stress. Plant J. 17: 385-395.
Verweire, D., K. Verleyen, S. De Buck, M. Claeys, and G. Angenon. 2007. Marker-free transgenic plants through genetically programmed auto-excision. Plant Physiol. 145: 1220-1231.
292                                                                                                                                                         Botanical Studies, Vol. 51, 2010
Wang, A.S., R.A. Evans, P.R. Altendorf, J.A. Hanten, M.C.
Doyle, and J.L. Rosichan. 2000. A mannose selection sys­tem for production of fertile transgenic maize. Plant Cell
Rep. 19: 654-660. Wang, Y., B.J. Wang, Y.L. Chen, J.F. Hu, J. Li, and Z.P. Lin.
2005. Inducible excision of selectable marker gene from transgenic plants by the Cre/lox site-specific recombination system. Transgenic Res. 14: 605-614.
Wright, M., J. Dawson, E. Dunder, J. Suttie, J. Reed, C. Kramer, Y. Chang, R. Novitzky, H. Wang, and L. Artim-Moore. 2001. Efficient biolistic transformation of maize (Zea mays L.) and wheat (Triticum aestivum L.) using the phospho-mannose isomerase gene, pmi, as the selectable marker. Plant Cell Rep. 20: 429-436.
Xing, A., Z. Zhang, S. Sato, P. Staswick, and T. Clement. 2000.
The use of two T-DNA binary system to derive marker-free transgenic soybeans. In vitro Cell. Dev. Biol.- Plant 36:
456-463.
Xue, G.P., M. Patel, J.S. Johnson, D.J.Smyth, and C.E. Vickers. 2003. Selectable marker-free transgenic barley producing a high level of cellulase (1,4-P-glucanase) in developing grains. Plant Cell Rep. 21: 1088-1094.
Ye, G.N., S. Colburn, C.W. Xu, P.T.J. Hajdukiewicz, and J.M. Staub. 2003. Persistance of unselected transgenic DNA dur­ing a plastid transformation and segregation approach to herbicide resistance. Plant Physiol. 133: 402-410.
Yoshida, K. and A. Shinmyo. 2000. Transgene expression sys­tems in plants, a natural bioreactor. J. Biosci. Bioeng. 90:
353-362.
Yuan, Y., Y.J. Liu, and T. Wang. 2004. A new cre/lox system for deletion of selectable marker gene. Acta Bot. Sin. 46: 862­866.
Zhang, W., S. Subbarao, P. Addae, A. Shen, C. Armstrong, V. Peschke, and L. Gilbertson. 2003. Cre/lox-mediated marker gene excision in transgenic maize (Zea mays L.) plants. Theor. Appl. Genet. 107: 1157-1168.
Zhang, Y., H. Li, B. Ouyang, Y. Lu, and Z. Ye. 2006. Chemical-induced autoexcision of selectable markers in elite tomato plants transformed with a gene conferring resistance to lepi-dopteran insects. Biotechnol. Lett. 28: 1247-1253.
Zhang, Y., H. Liu, B. Li, J.T. Zhang, Y. Li, and H. Zhang. 2009. Generation of selectable marker-free transgenic tomato resistant to drought, cold and oxidative stress using the Cre/ loxP DNA excision system. Transgenic Res. 18: 607-619.
Zhu, L., Y. Fu, W. Liu, G. Hu, H. Si, K. Tang, and Z. Sun. 2007. Rapid generation of selectable marker-free transgenic rice with three target genes by co-transformation and anther
culture. Rice Sci. 14: 239-246.
Zubco, E., C. Scutt, and P. Meyer. 2000. Intrachromosomal recombination between attP regions as a tool to remove selectable marker genes from tobacco transgenes. Nat. Bio-
technol. 18: 442-445.
Zuo, J., Q.W. Niu, S.G. Moller, and N.H. Chua. 2001. Chemical-regulated, site-specific DNA excision in transgenic plants. Nat. Biotechnol. 19: 157-161.