Open access

Growing Uses of 2A in Plant Biotechnology

Written By

Garry A. Luke, Claire Roulston, Jens Tilsner and Martin D. Ryan

Published: 15 April 2015

DOI: 10.5772/59878

From the Edited Volume

Biotechnology

Edited by Deniz Ekinci

Chapter metrics overview

2,945 Chapter Downloads

View Full Metrics

1. Introduction

The combination, ‘pyramiding’ or ‘stacking’ of multiple genes in plants is a fundamental aspect of modern plant research and biotechnology. The most widely adopted stacked traits (herbicide tolerance and insect protection) provide growers with benefits of increased crop yield, simplified management of weed control and reduced insecticide use. The global acreage of stacked traits or more precisely genetically modified organisms bearing stacked traits is expected to increase rapidly in the near future, with the introduction of nutritional and/or industrial traits to satisfy the needs of consumers and producers [1]. Several approaches have been used to stack multiple genes into plant genomes and then to coordinate expression [2-4]. Stacking approaches include sexual crossing between plants carrying distinct transgenes [5,6], sequential re-transformation [7], and single-plasmid [8] or multiple-plasmid co-transformation [9]. These strategies, however, suffer from the inherent weakness that co-expression of the heterologous proteins is unreliable.

Due to limited genomic coding space, many viruses encode more than one protein from a single mRNA transcript. Internal ribosome entry site (IRES) sequences serve as a launching pad for internal initiation of translation, allowing expression of two or more genes from a single transcript [reviewed in 10]. A number of IRES motifs from plant [11] and animal [12] viruses have been used to direct the expression of multiple recombinant proteins in plants and plant cells [13,14]. However, widespread use of IRES motifs in plant biotechnology is limited: they are not small (~600 base pairs), adding to the size of the transgene; translation efficiency of a gene placed after the IRES is much lower than that of a gene located before the IRES [14]. One promising gene/protein strategy adopted by some viruses to ensure a balance of proteins in vivo is to express polyprotein precursors with cleavable linkers between the proteins of interest [15]. Several groups demonstrated the potential of this approach by co-expressing two proteins separated by the tobacco etch virus (TEV) NIa protease recognition sequence (heptapeptide cleavage recognition sequence ENLYFQS) together with the NIa proteinase [16-18]. The utility of the NIa protease is limited due to the presence of a nuclear-localizing signal (NLS) within the protease and the amount of energy necessary to express the 49 kDa protease. It is also possible to use linker sequences that are putative substrates of known endogenous plant proteases [19].

To bypass the need for an endogenous or recombinant accessory protease acting on the translated polypeptide product a different approach involves the use of self-processing viral 2A peptide bridges [reviewed in 20, 21]. The designation “2A” derives from the systematic nomenclature of protein domains within the polyproteins of picornaviruses. In foot-and-mouth disease virus (FMDV) and some other picornaviruses the oligopeptide 2A region of the polyprotein manipulates the ribosome to “skip” the synthesis of the glycyl-prolyl peptide bond at its own carboxyl terminus leading to the release of the nascent protein and translation of the downstream sequence [22]. Under the monikers of “Skipping”, “Stop-Carry On” and “StopGo” translation, it allows the stoichiometric production of multiple, discrete, protein products from a single transgene [23,24]. Several recent review articles have amply covered the role of 2A biotechnology in animal systems [20, 25]. This summary-review will provide an up-to-date overview of 2A and cover the wider application of 2A-polyproteins to the expression of multiple proteins in plants.

Advertisement

2. The 2A story — The end of the beginning

2.1. The co-translational model of 2A-mediated “cleavage”

FMDV, like other members of the family Picornaviridae, is a non-enveloped RNA virus which contains a single-stranded, positive-sense RNA molecule of approximately 8500 nt that functions as an mRNA [26]. This (+) RNA encodes a high molecular mass polyprotein that undergoes co-translational processing to yield the structural proteins (1A, 1B, 1C and 1D, commonly known as VP4, VP2, VP3 and VP1 respectively) which comprise the viral capsid, and the non-structural proteins (2A, 2B, 2C, 3A, 3B, 3Cpro, and 3Dpol) that control the viral life cycle within host cells [27]. The 2A oligopeptide is only 18 amino acids (aa) long (-LLNFDLLKLAGDVESNPG-) defined by the co-translational “cleavage” at its C-terminus and a post-translational cleavage at its N-terminus, mediated by the virus-encoded proteinase 3Cpro [28]. Analysis of recombinant FMDV polyproteins [29] and artificial polyprotein systems in which 2A was inserted between two reporter proteins [22] showed that just 2A, plus the N-terminal proline of the downstream protein 2B was sufficient for highly efficient co-translational “cleavage” (Figure 1, Panel A). Quantification of products using in vitro cell-free translation systems showed the product upstream of 2A accumulated in a molar excess over that downstream – at variance with a proteolytic model of 2A which predicts a 1:1 stoichiometry of the cleavage products [23, 30, 31]. We and others have shown that 2A is not a proteinase, nor a substrate for a host-cell proteinase, but an autonomous element mediating a co-translational “recoding” event [27, 29]. From these observations we proposed a model of the 2A reaction based on hydrolysis of the nascent chain from ribosome-associated tRNA at the peptidyl-transferase centre [23, 24, 30]. For in-depth reviews of the model see [24, 32, 33].

Figure 1.

Schematic overview of 2A function and gene fusion constructs. Panel A: Two individual polypeptides can be generated from one transcript using F2A to link the individual genes. Panel B: F2A cleavage efficiency. All constructs shared a common core consisting of CFP-F2A-RABD2a. Panel C: Sequences encoding HA-tagged CAH1 were fused in-frame to wild type and mutant versions of genes encoding the GTPases RABD2a, SAR1, and ARF1 linked by F2A-the pre-protein has the endomembrane targeting sequence. These synthetic polyproteins were efficiently cleaved when transiently expressed in protoplasts and in planta. CFP, enhanced cyan fluorescent protein; GUS, β-glucuronidase; SP, ER signal peptide of CAH1 protein; GS, Golgi targeting signal of N-acetylglucosaminyl transferase 1; RABD2a, Arabidopsis RABD2a GTPase; SAR1, Nicotiana tabacum SAR1p; ARF1, Arabidopsis ADP-ribosylation factor 1; HA, hemagglutinin epitope tag; CAH1, Arabidopsis α-CAH1 (adapted from [60]).

2A comprises two parts, an N-terminal region (without sequence conservation) predicted to form an alpha helix, and a C-terminal motif,-DxExNPG, followed by a proline required for the reaction. Recently it was shown that the synonymous codon usage of this conserved motif is biased [34]. The amino acids E,S,N,P,G,P tend to use GAG, TCC, AAC, CCT, GGG and CCC respectively. The results also indicate that the synonymous codon usage of the 2A peptide has no effect on 2A activity. In summary, our results indicate the conserved –DxExNPG motif within the peptidyl transferase centre (PTC) of the ribosome is restricted and it forms a tight turn, shifting the ester bond between the C-terminal glycine and tRNAGly (in the P site of the ribosome) into a conformation which rules out nucleophilic attack by prolyl-tRNAPro (in the A site)-no peptide bond is formed. Although no stop codon is involved, eukaryotic translation release (termination) factors 1 and 3 (eRF1/eRF3) release the nascent protein from the ribosome [35-37]. Due to its mode of action, the 2A peptide has been described as a “cis-acting hydrolase element” (CHYSEL) [32]. Our model of this translational recoding event predicts two outcomes, either ribosomes terminate translation, or, translation of the downstream sequences resumes. Skipping induced by 2A sequences gives approximately equal expression of the proteins upstream and downstream of the 2A site as measured by: i) CAT and GUS enzyme activity [38]; ii) cell free translation in vitro and Western blot [22, 23, 31, 39, 40]; iii) GFP/FACS with antibiotic resistance [41]; iv) co-fluorescence reporting [42, 43]; v) fluorescence resonance energy transfer (FRET) analysis [44] and vi) protein segregation in transgenic animals [45, 46]. Since these sequences act co-translationally, artificial polyprotein systems may include signal sequences to localize different protein translation products to discrete sub-cellular sites.

2.2. 2A and 2A-like sequences

Probing databases for the presence of the “signature” motif (-DxExNPGP-) showed that “2A-like” sequences were present in several genera of the Picornaviridae (aphtho-, cardio-, erbo-, tescho and certain parechoviruses), single-stranded RNA insect viruses (iflaviruses, dicistroviruses, tetraviruses), double-stranded RNA viruses of the Reoviridae (type C and non-ABC rotaviruses, cypoviruses) and penaeid shrimp viruses [47, reviewed in 20,48]. Previously we demonstrated the activity of “2A-like” sequences within non-long terminal repeat retrotransposons (non-LTRs) of Trypanosoma brucei, T.cruzi, T.vivax, and T.congolense [31, 49] and more recently within the non-LTRs of a wide range of multicellular organisms: Xenopus tropicalis (African claw-toed frog, vertebrate), Branchiostoma floridae (Amphioxus, Florida lancelet, cephalochordate), Strongylocentrotus purpuratus (purple sea urchin, echinoderm), Aplysia californica (California sea slug, mollusc), Crassostrea gigas (Pacific oyster, mollusc), Lottia gigantean (Owl limpet, mollusc) and Nematostella vectensis (sea anemone, cnidarian) [50]. Presently, in silico searches have identified the 2A motif in a range of putative retrotransposon domains (Table 1).

Cellular 2A Name 2A sequence
Non-LTR retrotransposons
Trypanosoma spp
T.brucei Ingi -RSLGTCKRAISSIIRTKMBVSGDVEENPGP-
T.cruzi L1Tc -QRYTYRLRAVCDAQRQKLLLSGDIEQNPGP-
Strongylocentrotus purpuratus STR-32_SP -NSSCVLNIRSTSHLAILLLLSGQVEPNPGP-
(purple sea urchin) STR-51_SP -SRPILYYSNTTASFQLSTLLSGDIEPNPGP-
STR-61_SP -GARIRYYNNSSATFQTILMTCGDVDPNPGP-
STR-69_SP -CRRIAYYSNSDCTFRLELLKSGDIQSNPGP-
STR-197_SP -KHPILYYTNGESSFQIELLSCGDINPNPGP-
Crassostrea gigas (Pacific oyster) CR1-1_CGi -SRHIVVYNFYLQFFMFLLLLCGDIEVNPGP-
Lottia gigantean (owl limpet) CR1_1_LG -TLLNDTFSSILYYCFILIIRSGDIELNPGP-
Aplysia californica (California sea slug) Ingi-1_AC -PGFFLGGQHNPAWLARLLILAGDVEQNPGP-
Retrotransposons (unpublished) Accession number
Rhipicephalus pulchellus (Zebra tick) JAA55454.1 -SFVFTSLYADIVHCLCSLLLSGDVELNPGP-
JAA55744.1 -HCCLSIVVQCCDVIRSLLLLAGDIETNPGP-
Biomphalaria glabrata (freshwater snail) AC233255.1 -KWKFSVRDSRIKYLSLLILIAGDVESNPGP-
Caenorhabditis elegans (roundworm) Z49911.1 -LCETPSLPHTTFLKRKLLVRSGDVESNPGP-
AL132860.1 -LCETPSLPHTTFLKRKLLVRSGDVESNPGP-
F0081226.1 -LCETPSLPHTTFLKRKLLVRSGDVESNPGP-
Chlorella variabilis (Green alga) EFNS2199.1 -LRLPCSCSTTALIKRMKLLLSGDVEENPGP-

Table 1.

Active 2A cellular sequences.

The –DxExNPGP- motif conserved among 2A/2A-like sequence is shown in red.


Chimeric polyproteins incorporating 2A have been widely tested in eukaryotic systems, including mammalian [22], plant [38], insect [51], yeast [39] and fungal cells [52].The 2A system does not work in prokaryotic cells-the reported proteolysis activity of 1D-2A in Escherichia coli cells [53] was not detected in equivalent constructions in our laboratory showing “cleavage” specificity for eukaryotic systems alone [54]. The unique activity of 2A peptides has led to their use as tools for co-expression of two (or more) proteins in biomedicine and biotechnology [reviewed in 20, 21, 55]. The most widely used 2A sequence is derived from the FMDV (hereafter referred to as “F2A”) [42]. Other 2A peptides used successfully include “T2A” from Thosea asigna virus (TaV), “E2A” from equine rhinitis virus (ERAV) and “P2A” from porcine teschovirus-1 (PTV-1) [Table 2]. Comparing the in vitro activity of different 2As inserted between GFP and GUS, we have shown that T2A20 has the highest cleavage efficiency followed by E2A20, P2A20, and F2A20 [31]. In 2A peptide-linked TCR:CD3 constructs, Szymczak and colleagues demonstrated that F2A22 and T2A18 have higher efficiency than E2A20 [44]. In human cell lines, zebrafish and mice, cleavage and targeting of NLS-EGFP and mCherry-CAAX to the nucleus and plasma membrane, respectively, was the most efficient in P2A19-linked constructs followed by T2A18, E2A20 and F2A22 [56]. To allay public fears and opposition to plants carrying a transgenic viral sequence, efficient 2A-like cellular sequences could be used (Table 1) [Unpublished Data].

Virus Abbreviation 2A/2A-like sequence References
Picornaviridae
Foot-and-mouth disease virus FMDV -PVKQLLNFDLLKLAGDVESNPGP- [38,44,83,99,106,150]
Equine rhinitis A virus ERAV -QCTNYALLKLAGDVESNPGP- [44]
Porcine teschovirus -1 PTV-1 -ATNFSLLKQAGDVEENPGP- [45]
Tetraviridae
Thosea asigna virus TaV -EGRGSLLTCGDVESNPGP- [44,46]

Table 2.

Examples of 2A/2A-like sequences used in biomedicine and biotechnology

The –DxExNPGP-motif conserved among 2A/2A-like sequence is shown in red.


2.3. Intracellular protein targeting Of 2A constructs

For effective technologies, some synthesized proteins must be transported across membranes and directed towards other sites in order to function. Protein targeting occurs either co-translationally (targeting to endoplasmic reticulum [ER], Golgi, vacuole, plasma membrane) or post-translationally (targeting to nucleus, mitochondria, chloroplast, etc) and is orchestrated by distinct signal sequences encoded within the polypeptide [42]. In plants, the original FMDV-2A sequence was tested in various artificial polyproteins using reporter genes chloramphenicol acetyltransferase (CAT), β-glucuronidase (GUS) and green fluorescent protein (GFP) expressed in transgenic tobacco plants. This preliminary series of studies suggested that 2A cleaves proteins properly in plant cells [38, 57] and directs protein targeting to different cellular compartments via either co-or post-translational mechanisms [58]. Subsequently, Samalova and co-workers questioned its use in plant systems, suggesting that the 2A sequence was dispensable for efficient cleavage of polyproteins carrying a single internal signal peptide – it appears signal peptide cleavage by signal peptidase was responsible for processing the polyprotein. The use of a self-cleaving 2A was required when both halves of the fusion were translocated across the ER membrane, however, the upstream product was mis-sorted to the vacuole. Furthermore, it was shown that the FMDV 2A peptide resulted in low rates of polypeptide separation in plant cells when placed downstream of common fluorescent proteins (GFP and RFP derivatives) [43].

The Arabidopsis carbonic anhydrase (CAH1) is one of the few plant proteins known to be targeted to the chloroplast via the secretory pathway – the pre-protein has the endomembrane targeting sequence. The need for post-translational modifications, such as N-glycosylation, for proper folding, and to enhance stability and/or function of these proteins probably explains the use of this alternative trafficking pathway [59]. Recently, the FMDV 2A co-expression system was re-assessed to study the effects of three Ras-like small GTPase proteins, RAB2a, ARF1, and SAR1 on CAH1 protein trafficking in plant cells [60]. Members of this superfamily share several common structural features and act as molecular switches that regulate many aspects of plant vesicular transport [61,62]. Rabs regulate virtually all steps of membrane traffic from the specification of membrane identity to the accuracy of vesicle targeting [63]. ARF1 has been shown to play a critical role in COPI-mediated retrograde trafficking, while SAR1 is involved in COPII-mediated ER-to-Golgi protein transport [reviewed in 64]. In this study, targeting information and the sequence N-terminal of 2A proved to be important for efficient cleavage when translated by membrane-bound and cytosolic ribosomes respectively (Figure1, Panel B). In addition, expected subcellular localization of the fluorescent marker protein suggested no significant mis-targeting of the 2A-tagged markers. After optimization of 2A cleavage efficiency, mutant forms of the three small GTPases (HACAH1-2A/RABD2a/SAR1/ARF1, Figure1, Panel C) were successfully used to study trafficking of CAH1 through the endomembrane system demonstrating the versatility of 2A in plant systems.

Exchange factors for ARF GTPases (ARF-GEFs) regulate vesicle trafficking in a variety of organisms. In animals and fungi, there are eight ARF-GEF families, but only the apparently ancestral GBF and BIG families are present in plants, suggesting that plant ARF-GEFs have acquired multiple roles in different trafficking pathways [65, 66]. In Arabidopsis the ARF-GEFs GNOM-like 1 (GNL1) and its close homologue GNOM jointly regulate the retrograde COPI-mediated traffic from the Golgi to the ER, which is the ancient eukaryotic function of the GBF1 class [67]. Another line of research by Teh and Moore (2007) revealed secretory traffic is resistant to the trafficking inhibitor brefeldin A (BFA), whereas endosomal recycling involves GNOM – GNL1 is a BFA-resistant GBF protein that functions with the BFA-sensitive ARF GEF GNOM [68]. The 20aa 2A peptide from FMDV was used in this study to construct polyproteins that expressed trafficked fluorescent protein markers in fixed stoichiometry in different cellular compartments: N-ST-RFP-2A-GFP-HDEL produces a Golgi-localized RFP (red) and an ER-localized GFP (green), N-secRFP-2A-GFP-HDEL produces an ER-localized GFP and an RFP that is targeted to the vacuole via the ER and Golgi.

2.4. The use of 2A multigene expression strategies in plant science — Caveats and proposals

The take-home message from F2A mutagenesis experiments is that the sequence is largely intolerant to amino acid substitution over its entire length [31, 37]. While mutations of conserved amino acids have, in general, more pronounced effects than changes to non-conserved ones [31], variations at most positions within the peptide reduce activity – 2A peptides are optimized to function as a whole [37]. Sequences immediately upstream of 2A are known to be either critical or very important for activity [57, 69-72]. Longer versions of F2A with extra sequences derived from the capsid protein (“1D”) – upstream of 2A in the FMDV polyprotein – produce higher levels of cleavage [23, 29, 47]. Specifically, N-terminal extension of 2A by 5aa of 1D improved “cleavage”, but extension by 14aa of 1D or longer (21 and 39aa) produced complete “cleavage” and an equal stoichiometry of the up-and downstream translation products [23]. After “fine-tuning” of the F2A sequence we suggest that researchers opt for F2A30 (+11aa 1D). This 2A proved to be the most favourable in terms of both length and cleavage efficiency and was unaffected by the sequence of the upstream gene [73,74]. In the case of shorter 2As, cleavage efficiency has been improved by insertion of various spacer sequences such as Gly-Ser-Gly or Ser-Gly-Ser-Gly [41, 44, 45, 75-77], the V5 epitope tag (-GKPUPNPLLGLDST-) [78], or a 3xFlag epitope tag [79] ahead of the 2A sequence. If opting for a shorter sequence, users should be aware activity can be affected by the short amino acid tract linking the protein upstream with 2A introduced by the cloning strategy. For example, the F2A20 encoded by pGFP-F2A20-GUS was highly active [29, 31], whereas the pGFP-F2A20-CherryFP was noticeably lower [73]. The only difference was the short “linker” between GFP and F2A created by the cloning strategy:-SGSRGAC-(pGFPF-2A20-GUS; linker derived from Xba1 and Sph1 restriction sites) and-RAKRSLE-(pGFPF-2A20-CherryFP; linker derived from furin and Xho1 restriction site) [73]. Taken together, these observations are consistent with our translational model in which 2A activity is a product of its interaction with the exit tunnel of the ribosome which is thought to accommodate a nascent peptide of 30-40 amino acids [80].

When using the 2A system, it should be noted that the 2A oligopeptide remains as a C-terminal extension of the upstream fusion partner and the downstream protein must have an N-terminal proline residue. Although an N-terminal proline confers a long half-life upon a protein [81], it does prevent many N-terminal post-translational modifications that may be essential for activity. If this is the case, proteins that require authentic termini can be introduced as the first polyprotein domain. The need to target proteins to different subcellular locations within plant cells by C-terminal localization signals may be compromised if they contain a 2A-extension. In the case of proteins translocated into the ER, a strategy was adopted to include a furin proteinase cleavage site between the upstream protein and 2A [82,83]. Furin is a subtilisin-like serine endoprotease that cleaves precursors on the C-terminal side of the consensus sequence –Arg-X-Lys/Arg-Arg (-RX(K/R)R-) in the trans-Golgi network (TGN) [84,85]. The furin cleavage sequences -RKRR-,- RRRR-, and – RRKR-consisting of only basic amino acids, which can be efficiently cut by carboxypeptidases ( ), was used to remove 2A peptide-derived amino acids from the upstream antibody heavy chain during protein secretion [83]. Proteins expressed in plants could have their 2A extensions removed by endogenous proteinases acting on similar hybrid linker peptides. In 2004, François and colleagues connected the first nine amino acids (SN AADEVAT) of the LP4 peptide to the 20aa F2A to generate a hybrid linker peptide, LP4-2A [86]. LP4 is the fourth linker peptide of the naturally occurring polyprotein precursor originating from seed of Impatiens balsamina [87]. Cleavage of the polyprotein with plant defensin DmAMP1 from Dahlia merckii at its amino-terminus and plant defensin RsAFP2 from Raphanus sativus at its carboxy-terminus resulted in the release and targeting of (DmAMP1-SN) and RsAFP2 to different cellular compartments [86]. Recently, 2A and LP4-2A were used to connect the Bacillus thuringiensis (Bt) cry1Ah gene, which encodes a protein exhibiting strong insecticidal activity, and the mG2-epsps gene, which encodes a protein tolerant to glyphosate, the world’s most important and widely used herbicide [72]. The expression level of the two genes linked by LP4-2A was higher than those linked by 2A, regardless of the order of the genes within the vector. Furthermore, tobacco plants transformed with the LP4-2A fusion vectors showed better pest resistance and glyphosate tolerance compared to plants transformed with the 2A fusion constructs.

Advertisement

3. 2A in plant biotechnology

3.1. A strategy to improve transgene expression from the Chlamydomonas nuclear genome

Micro-algae have the potential to be low-cost bioreactors for recombinant protein (RP) production due to their relatively rapid growth rates, favourable transformation time, ease of containment and rapid scalability [88-90]. The availability of a complete genome sequence [91] coupled with the ability to manipulate all three genomes (chloroplast, nuclear and mitochondrial) makes Chlamydomonas reinhardtii an attractive species for biotechnologists [90,92]. While transgene expression from the alga’s nuclear genome offers several advantages over chloroplast expression, such as post-translational modifications and protein targeting and/or secretion [93], yields of target RPs are often inadequate for industrial purposes. In an attempt to overcome this limitation, Rasala and colleagues constructed a C.reinhardtii nuclear expression vector using F2A to fuse GFP or xylanase 1 (xyn1) from Trichoderma reesi to the bleomycin/zeocin antibiotic resistance gene sh-ble [94]. High-value xylanases have numerous applications in the textile, paper, pulp, food and feed industries [reviewed in 95]. Efficient cleavage of GFP from Ble-2A was observed in the algal cytoplasm, leading to high level GFP expression and built-in resistance to zeocin. Co-expression of Xyn1 with Ble-2A led to the selection of transformants with higher xylanase activity (~100-fold) compared to unfused Xyn1. Subsequently, the ble-2A expression system was used to secrete enzymatically active Xyn1 by insertion of a secretion signal peptide (SP) between F2A and Xyn1 (ble2A-SP-xyn1). In a follow-on study, the same strategy was used to express and compare six recombinant fluorescent proteins (FPs: blue mTagBFP, cyan mCerulean, green CrGFP, yellow Venus, orange tdTomato and red mCherry) in C.reinhardtii [96]. All FPs were easily detectable in live cells, and the ble-2A-FP polyproteins were efficiently processed to yield unfused FPs. CrGFP was shown to be the least fluorescent due to its low signal-to-noise ratio, while the FPs with longer emission wavelengths (Venus, TdTomato and mCherry) had the highest signal-to-noise ratios. In this study, the ble-2A vector was used to tag an endogenous gene (α-tubulin) with a fluorescent protein tag (mCerulean) that was readily detectable in Chlamydomonas using standard live-cell imaging techniques. Taken together, these results suggest this new expression system will become an important tool for both algal biotechnology and basic research.

3.2. Engineering plant metabolomes

Manipulating plant metabolomes ranges from modifications, building extensions or branches onto existing biochemical pathways to very extensive changes-such as the rice C4 project. Here, the aspiration is to covert rice from a C3 plant (3-carbon molecule present in the first product of carbon fixation) to a more efficient C4 plant – eliminating photorespiration. As a major global crop species, rice has been the subject of intensive research. The ‘Golden Rice’ project was undertaken to address the problem of vitamin A deficiency (VAD), and is discussed in some detail here since it provides an interesting ‘vignette’ of the progress in transgenesis and plant metabolome engineering. It should be noted, however, that another ‘biofortified’ crop also designed to reduce VAD is the “Super-banana”, originally developed for the Ugandan population, which recently gained approval to begin human trials in the United States. In this case, however, knowledge gained from a cultivar identified in Papua New Guinea, enabled development of super-bananas by genome editing, rather than transgenesis, of a commercial cultivar.

A shortage of dietary vitamin A leads to VAD resulting in impairment in sight, increases in the severity of a range of infectious diseases and is estimated to lead to the premature deaths of 650-700,000 children under the age of 5 each year: a particular problem in parts of S.E. Asia and Africa. Whilst no rice cultivars produce provitamin A within the endosperm, the precursor geranylgeranyl diphosphate (GGDP) is produced. This precursor could be converted into β-carotene, which functions as provitamin A in humans. To convert GGDP to β-carotene was originally thought to require the activities of phytoene synthase, phytoene desaturase, ζ-carotene desaturase and lycopene β-cyclase, although recent analyses have shown endogenous rice enzymes can substitute for lycopene β-cyclase in the conversion of lycopene to β-carotene. The agrobacterial vector construct pB19hpc encodes the daffodil phytoene desaturase gene (psy) driven by the endosperm-specific glutelin (Gt1) promoter, and the phytoene desaturase (crtI) gene from the bacterium Erwinia uredovora, driven by the cauliflower mosaic virus (CaMV) 35S promoter, arranged in tandem (Figure 2, Panel A) [97]. Whilst the psy gene already encoded an N-terminal transit peptide sequence, the crtI sequence was modified to encode the N-terminal transit peptide sequence from the pea Rubisco small subunit – ensuring both proteins were imported into plastids, the site of GGDP biosynthesis. Selectable marker genes (aphIV and nptII) were also included. To complete the β-carotene biosynthetic pathway, plants were co-transformed with vectors pZPsC and pZLcyH. Vector pZPsC carries psy and crtI (as pB19hpc), but lacked the aphIV marker expression cassette. Vector pZLcyH encodes lycopene β-cyclase (lcy) from Narcissus driven by the rice glutelin promoter and the marker aphIV gene controlled by the CaMV 35S promoter, arranged in tandem. Again, lycopene β-cyclase carried a functional transit peptide to direct plastid import. Whilst this proved to be a successful strategy – it is rare one’s experimental results are given a Papal blessing, the use of multiple promoters in such a strategy of transgenesis often leads to ‘interference’: known as ‘transcriptional interference’ or ‘promoter suppression’-probably caused by competitive binding of transcription factors and/or modification of DNA structure at one site that affects the other site. Such problems may be overcome by the inclusion of ‘insulators’ which serve to segregate an enhancer and an adjacent promoter into independent domains-‘enhancer blocking’.

Figure 2.

Strategies of plant transgenesis. In all panels, genes / transcription termination / poly(A) addition sequences are shown as boxed areas and promoters as arrows. Panel A: The transcription of the daffodil phytoene desaturase gene (psy), Erwinia phytoene desaturase (crtI) and the Narcissus lycopene β-cyclase (lcy) genes are driven either by the endosperm-specific glutelin (Gt1) or CaMV 35S promoters, as indicated. In all cases shown in this figure, sequences encoding the N-terminus of CRT1 were preceded by those encoding functional transit peptides-to ensure import into plastids. The aphIV and nptII selectable markers used are shown, together with transcription terminators / poly(A) addition signals (light grey boxes) and Agrobacterium vector left and right borders (LB, RB). Panel B: In the second generation Golden Rice, Erwinia phytoene desaturase (crtI) gene was co-expressed with the maize phytoene desaturase gene (psy) gene under the control of tandem rice glutelin promoters (Glu1). The Agrobacterial vector also encoded a selectable marker cassette comprising the maize polyubiquitin (Ubi1) promoter with intron, hygromycin resistance (hptII) and nos terminator are shown as grey boxes. Panel C: In the third generation of Golden Rice, Capsicum phytoene synthase (psy) and Pantoea carotene desaturase (crtI) were co-expressed by linkage into a single ORF via a synthetic 2A sequence that was optimized for rice codons (pPAC construct), or, the genes were linked by the insertion of an IRES into the intergenic region (pPIC construct). In both cases, transcription was driven by the rice globulin promoter (GIb). The Agrobacterial constructs also comprised a selectable marker (bar) driven by the 35S promoter and were flanked by a 5’-matrix attachment region (Mar) from the chicken lysozyme gene are not shown. Panel D: Sequences encoding the maize proteinase inhibitor (mpi) and the potato carboxypeptidase inhibitor (pci) were fused into a single ORF. The two proteinase inhibitors were linked using either (i) the processing site of the Bacillus thuringiensis Cry1B precursor protein (C) or (ii) the 2A sequence from foot-and-mouth disease virus (F2A). In both cases the wound-inducible mpi promoter (arrow) was used to drive the expression of the mpi-pci fusion genes. Panel E: Genes encoding the 3,3’-β-hydroxylase (crtZ) and 4,4’-β-oxygenase (crtW) from marine bacteria (Paracoccus spp.) were linked via F2A to form a single ORF. Again, both gene sequences were preceded by those encoding functional transit peptides to ensure import into plastids. Selectable markers (e.g. lacZ and hptII) are shown as grey boxes.

The first-generation Golden rice encoded the daffodil phytoene synthase gene, but this daffodil enzyme proved to be a rate-limiting step in β-carotene biosynthesis, since substitution with psy from maize produced a 23-fold increase in total carotenoids (with a preferential accumulation of β-carotene) compared to the original Golden Rice (Figure 2, Panel B) [98]. Again, tandem (rice glutelin) promoters were used. To avoid the problems of promoter interference, a third generation of Golden Rice was developed using two systems both of which produce multiple proteins from a single transcript mRNA. In this strategy, the crtI and psy genes were linked by either a 2A sequence (creating a single ORF; pPAC construct), or, the two genes were separated by an IRES (pPIC construct; Figure 2, Panel C) [99]. In this strategy phytoene synthase from Capsicum was co-expressed with phytoene desaturase from Pantoea from a single transcription unit in both cases. The rice globulin promoter drove the expression of [Psy-2A-CrtI], or, [Psy-IRES-CrtI]. The endosperm of transgenic PAC rice had a much more intense golden colour than did the PIC rice transformants, demonstrating that 2A was more efficient than an IRES in co-expression of PST and CRTI and hence the synthesis of β-carotene. Indeed, immunoblot analyses of CRTI (the downstream protein in both cases) showed that 2A was nine-fold more effective than an IRES. It is well-known that in the case of a bicistronic mRNAs using an IRES, the gene downstream of the IRES is only translated ~1/10th as efficiently as the gene upstream of the IRES. In comparison to IRESes, the use of 2A and 2A-like sequences has the advantages of producing stoichiometric levels of the translational products and can be used to express multiple (>2) proteins at similar levels.

In light of the developments in the synthesis of β-carotene outlined above, it is interesting to note that whilst higher plants synthesize carotenoids, they do not possess the ability to form ketocarotenoids – potent antioxidants with numerous reported health benefits. Organisms capable of synthesizing ketocarotenoids are rare, although an early report showed that co-expression of the 3,3’-β-hydroxylase (crtZ) and 4,4’-β-oxygenase (crtW) from marine bacteria (Paracoccus spp.), linked via 2A (Figure 2, Panel E), lead to the formation of ketolated carotenoids (astaxanthin, canthaxanthin and 4-ketozeaxanthin) from β-carotene and its hydroxylated intermediates by the construction of an astaxanthin pathway [100].

3.3. Improving resistance to abiotic/biotic stresses

Abiotic stresses such as drought, excessive salinity, high and low temperature are critical factors limiting the productivity of agricultural crops. The development of genetically engineered plants with enhanced tolerance presents an important challenge in plant gene technology. A common response of plants to these environmental stresses is the accumulation of sugars and other compatible solutes. Trehalose is a nonreducing disaccharide that functions as a stress protection metabolite in many organisms [reviewed in 101]. In yeast, trehalose-6-phosphate synthase (TPS1) and trehalose-6-phosphate phosphatase (TPS2) enzymes catalyse the conversion of glucose-6-phosphate and uridine diphosphate (UDP)-glucose to trehalose in a two-step pathway [102]. Several efforts have been undertaken to engineer the drought-and salt-tolerance of economically important plants using TPS1 and/or TPS2 genes from yeast and bacteria [103-105]. Both TPS1 and TPS2 genes of Zygosaccharomyces rouxii were co-introduced into potato plants as a ZrTPS2-F2A-ZrTPS1 polyprotein in an attempt to develop stress-tolerant transgenic plants [106]. The resulting plants showed increased tolerance to drought and no visible phenotypic alterations. Glycinebetaine (N,N,N-trimethyl glycine; betaine) is regarded as an extremely effective compatible solute, which is able to restore and maintain the osmotic balance of living cells, in response to high salinity, cold and drought [107]. In plants, betaine is produced by the two-step oxidation of choline via the two enzymes choline monooxygenase (CMO) and betaine aldehyde dehydrogenase (BADH). Transferring betaine synthesis genes CMO and BADH from the halophyte Suaeda salsa to Pichia pastoris produced “CMO-F2A-BADH” recombinant yeasts with higher tolerance to salt, methanol, and high temperature stress [108]. The result indicates that this strategy could be used to improve the tolerance to stress of commercially important crops such as potato, rice, tomato, and tobacco, which do not accumulate betaine.

Plants are substrates for a wide range of pests and pathogens, including fungi, bacteria, viruses, nematodes, insects, and parasitic plants [109]. To defend themselves against pathogen attack, plants produce a battery of antimicrobial peptides (AMPs), secondary metabolites and reactive oxygen species [reviewed in 110-112]. AMPs (such as defensins) are attractive candidates for transgenic applications for several reasons: their diverse antimicrobial activity, low toxicity for non-target cells and low cost in terms of energy and biomass involved in their expression [113]. To achieve resistance against a broader range of pathogens in plants, co-expression of transgenes encoding AMPs with different biochemical targets is an attractive approach. In the case of plant defensins DmAMP1 and RsAFP2 (see above), biological activity of the hybrid protein was higher compared to the individual parental proteins [86]. The potato AMPs snakin-1(SN1) and defensin-1 (PTH1) were fused to improve plant protection against phytopathogens [114]. SN1 is active against both bacterial and fungal species, whereas PTH1 shows primarily antifungal activity [115, 116]. Antimicrobial activity of SN1 and PTH1 (linked by the F2A sequence) as a single-fusion protein in E.coli systems was better against the majority of tested microorganisms compared with the activity of individual proteins. In a sense, this (finding) is a surprise since F2A does not display any cleavage activity in E.coli [54]. Nevertheless, increased antibacterial and antifungal activity was reported in tobacco and potato plants expressing the snakin-defensin hybrid.

The expression of plant proteinase inhibitors is one strategy for increasing resistance against insects. The maize serine proteinase inhibitor (MPI) and the potato carboxypeptidease inhibitor (PCI) were co-expressed in rice using two strategies (Figure 2, Panel D) [117]. The first was to link the two gene sequences into a single ORF via a sequence encoding the proteinase processing site of the B. thuringiensis Cry1B precursor protein. The rationale here is that the translation (fusion protein) product would be post-translationally cleaved into MPI and PCI by an endogenous rice proteinase. The second strategy was to link the mpi and pci genes via 2A: here, the translation products would be translated as discrete products ([MPI-2A]+PCI). Both co-expression strategies were successful and both types of rice transgene showed increased resistance to the striped stem borer (Chilo suppressalis). Whilst both strategies benefit from co-expression using a single transgene, the additional merit of using 2A for similar approaches is that post-translational processing by an endogenous proteinase limits the sub-cellular targeting of proteins downstream of the Cry1B linker to the cytoplasm or other sub-cellular sites using post-translational import – not the exocytic pathway which requires the co-translational recognition of signal sequences by signal recognition particle. Since 2A works within the ribosome, proteins downstream of 2A can be modified to comprise N-terminal signal sequences and direct these translation product(s) to the exocytic pathway.

Glucosinolates (GLSs) present in cruciferous plants (e.g. cabbage, broccoli, and oilseed rape) play a defensive role against generalist insects [118] and pathogens [119]. However, crucifer-specialist insect herbivores like the economically important pest diamondback moth (DBM; Plutella xylostella) frequently use GLSs to stimulate oviposition [120]. An increase in the global area of brassica crops between 1993 and 2009 coupled with the high fecundity of DBM especially in tropical regions has resulted in development of resistance to many broad-spectrum insecticides used in the field [121]. Recently, genetic engineering has been used to produce non-host GLS-containing plants as a first step towards the creation of “dead-end trap crops”. The transfer of the six-step benzylglucosinolate (BGLS) pathway into tobacco plants using only two ORFs, consisting of the first three genes (GGP1-F2A-CYP83B1-F2A-CYP79A2) and last three genes (SOT16-F2A-UGT74B1-F2A-SUR1), give rise to BGLS-producing plants [122,123]. Importantly, these non-host plants were more attractive for DBM oviposition than wild-type tobacco plants. As larvae are unable to survive, the strategy of engineering oviposition cues into non-host plants offers an alternative trap crop approach to crop protection. The combination of abiotic and biotic stresses presents an added degree of complexity, as responses at a molecular level are largely controlled by different signalling pathways that can act antagonistically [124]. The pyramiding of several defence genes may therefore provide further opportunities for creating broad-spectrum stress tolerance in agronomically important crop plants [125].

3.4. Cost-effective production of cellulose degrading enzymes for biomass-to-fuel conversion

Dwindling fossil resources and increasing energy demands are driving the development of alternative feedstocks for producing fuels and chemicals. Cellulosic feedstocks such as crop residues, wood products and dedicated crops (e.g. switchgrass, salix) are among leading alternatives because they are sufficiently abundant, low cost and do not compete with food sources. The bioconversion of lignocellulose biomass into fuels involves three major transformations: the production of saccharolytic enzymes (cellulases and hemicellulases), the hydrolysis of carbohydrate components present in pretreated biomass to sugars, and fermentation of sugars to produce fuels such as ethanol and butanol [reviewed in 126, 127]. Unfortunately, the high cost of enzymes is a major barrier in the biomass-to-fuel industry [128]. Observed results indicate in planta enzyme expression offers a potential method for low cost large-scale enzyme production [129, 130]. Current indications are subcellular targeting [131, 132] and simultaneous expression of recombinant cellulolytic enzymes [133] are key factors in optimizing their accumulation in transgenic plants. Transgenic expression of 2A-linked cellulase enzymes (β-glucosidase, BglB; xylanase, Xyl11; exoglucanase, E3; endoglucanase, Cel5A) in chloroplast-targeted tobacco plants induced synergistic effects that led to more efficient hydrolysis of lignocellulose materials for bioethanol production [134]. Chloroplast transit peptides, small subunit of Rubisco complex (Rs) and Rubisco activase (Ra) [135], were fused to the N-terminal of the enzyme genes. This study found a synergistic effect between BglB and Cel5A in the (RsBglB-F2A-RaCel5A) lines and between E3 and Cel5A expressed in the (RsE3-F2A-RaCel5A) lines. The enzymes had higher activities which led to enhanced carboxymethyl cellulose (CMC) hdydrolysis into glucose and cellobiose. A similar observation was made by Jung et al., [2010] with chloroplast-derived BglB and Cel5A [136]. Furthermore, supplementing the protein extracts of transgenic (RsBglB-F2A-RaCel5A) with CBH11 exoglucanase increased hydrolysis activity. While the cost or yield may not be the same for all chloroplast-derived enzymes, these are important steps in cellulose bioconversion.

3.5. Recombinant plant viruses as (Co-)expression systems

Plant viruses are both important crop pathogens and very efficient vectors for transient protein over-expression. The majority have comparatively small RNA genomes, encoding a limited number of genes. Modifying such viruses for research or biotechnology purposes is often impeded by limitations imposed by the genome organization. Absolute genome size may be restricted by virion capsid size, open reading frames can overlap each other as well as regulatory sequence elements, and duplication of regulatory sequences to drive expression of foreign genes can lead to genomic instability. 2A peptides can be used to overcome some of these limitations.

One major use for plant virus-based over-expression vectors is the production of immunogenic epitopes in plants. Using plants as hosts has the benefits of a eukaryotic expression system with fast growth rates that can be produced on large scales and shares no pathogens with humans or animals. Viral vectors naturally achieve extremely high over-expression levels and can systemically infect whole plants from small, inexpensive inocula and avoiding the requirement to produce transgenic plants. To achieve this, the modified virus has to retain its infectivity and ability to move through the plant. A common approach to overexpress foreign proteins from a plant virus genome is to fuse them to the capsid protein (CP). CPs are often the most highly expressed viral proteins and this ensures efficient overexpression without the need to re-engineer any regulatory sequences in the viral genome. If the epitopes are displayed on the virus particle surface, they are also easy to purify. However, CP fusions are not always tolerated as they can interfere with viral encapsidation and spread. 2A peptides can be used to rescue encapsidation and infectivity of CP-fusion viruses by providing a pool of unfused CP.

The first such use in a modified plant virus was the Potato virus X (PVX) ‘overcoat’ virus, in which a green fluorescent protein is fused to the N-terminus of the CP via an FMDV 2A linker [137]. The 2A-mediated ‘cleavage’ is incomplete so that both free CP and the GFP-CP fusion are produced, enabling formation of fully infectious virus particles which also incorporate the GFP-CP fusion. GFP is exposed at the virion surface, permitting tracking of infection and imaging of virions [137]. A similar ‘overcoat’ principle was also used to image infections of the related Plantago asiatica mosaic virus (PlAMV) [138]. Obviously this overcoat approach also enables the fusion of other peptides or proteins to the PVX CP. This was first demonstrated by expression of a single chain antibody against the herbicide diuron [139]. Approximately 100-250 μg protein/g leaf fresh weight of the antibody was produced, and ‘overcoat’ virions were easily purified by diuron-based immune-capture.

Subsequently, a number of antigenic epitopes have been expressed using PVX ‘overcoat’ vectors, including Rotavirus inner capsid protein [140], Classical swine fever virus glycoprotein [141], tuberculosis antigen ESAT-6 [142], and a consensus epitope from a Hepatitis C virus envelope protein [143]. Between 1-125 μg protein or virus/g leaf fresh weight, or up to 0.5-1% of total soluble protein, were obtained in these studies, and ‘overcoat’ virions could be purified by centrifugation [141, 143]. In one case, greater amounts of protein were achieved with an alternative expression strategy using a duplicated subgenomic promoter [140], but such constructs tend to be genetically very unstable compared to 2A-based ‘overcoat’ vectors [144]. PVX virions carrying the Classical swine fever virus glycoprotein epitope produced an immunoprotective response in rabbits [141], and antibodies from a mouse immunized with PVX particles displaying the hepatitis C virus R9 epitope reacted with sera from infected patients [143]. For smaller epitopes, direct fusion to the PVX CP was sometimes possible, but for all larger foreign peptides, the 2A-mediated partial separation was required to enable encapsidation of systemically infective virus [143]. Filamentous viruses like PVX are particularly suited for surface presentation of antigens as they have a large number of CP units (~1270/PVX virion), and ‘overcoat’ overexpression vectors using 2A have also been developed for the related viruses Pepino mosaic virus (PepMV) [144] and PlAMV [145]. So far, plant virus-expressed epitopes have been purified mainly as virion-attached surface peptides, but with 2A linkers, it is also possible to produce them mainly as free proteins. For efficient surface display, an optimal ratio of free and fused CP has to be found that maximizes virus-displayed epitopes whilst still enabling efficient encapsidation. The range of 2A-like sequences with different ‘cleavage’ activities [31, 37] will be useful in the development of further ‘overcoat’ vectors.

Expression vectors using 2A peptides have also been developed based on Cowpea mosaic virus (CPMV) [146], Bean pod mottle virus (BPMV) [147], and Wheat streak mosaic virus (WSMV) [148]. In the case of the CPMV and WSMV vectors, 2A linkers were used to overcome the problem that these viruses encode large polyproteins that are processed into functional subunits by viral proteases. Foreign proteins inserted into the polyprotein open reading frame need to be released and 2A sequences provide an alternative to viral protease cleavage sites. In CPMV, using 2A instead of viral cleavage motifs reduced the number of additional amino acids attached to the over-expressed protein, and as in PVX ‘overcoat’ vectors, some of the foreign protein is displayed on virus particles, which are very stable and easy to purify. In WSMV, use of FMDV 2A or (and also FMDV 1D/2A) sequences resulted in more efficient release of the foreign protein (GFP) than with viral proteinase sites [148]. WSMV vectors enable protein expression in cereal hosts. In the soybean-infecting BPMV vectors, 2A linkers were used both to enable insertion of foreign genes into the viral polyprotein open reading frame, and to facilitate simultaneous co-expression of two different foreign proteins.

Advertisement

4. Food for thought

The first demonstration the 2A was active in plant cells used an artificial polyprotein which comprised two reporter proteins flanking 2A [38]. This co-expression system was soon adopted by plant virologists for use in both rod-shaped and icosahedral virus particles either as high-level expression systems, or, to produce particles ‘decorated’ with fluorescent proteins, immunogens, single-chain antibodies etc. [137-147]. Here, plants are used simply as ‘bioreactors’ for production of recombinant proteins / virus particles – the plants are not transgenic. In the case of transgenic plants the first reports of the use of 2A to co-express multiple proteins were as a ‘proof-of-principle’ or research tools [38, 57], but within a few years plants were being genetically engineered to demonstrate how nutritional properties could be improved [105, 149]. Whilst the use of 2A rapidly expanded in the arenas of animal biotechnology and biomedicine (e.g. monoclonal antibody production, cancer gene therapies, production of pluripotent stem cells: reviewed in [25]), progress in transgenic plants was slower-due to a number of reasons, including the ‘trickle-down’ effects on plant biotechnology from the EU policies concerning genetically-modified plants. Over the past few years, however, the 2A co-expression system has been used in the development of methods to engineer plant genomes [149], the expression of high-value proteins, the improvement of plant tolerance to biotic and abiotic stresses, the improvement of nutritional properties through metabolome engineering [vide supra] and the expression of plant storage proteins with amino acid content more suited to human nutrition [150]. The drive to improve agricultural productivity through the development of ‘dual-use’ crops necessitates complex strategies of plant engineering and it seems clear that the use of 2A will continue to expand.

Advertisement

Acknowledgments

The authors gratefully acknowledge the long-term support of the UK Biotechnology and Biological Sciences Research Council (BBSRC). The University of St Andrews is a charity registered in Scotland no. SCO13532.

References

  1. 1. James C. Global Status Of Commercialized Biotech/GM Crops: 2011. ISSSA Brief No.43. ISSSA, Ithaca.
  2. 2. François IEJA, Broekaert WF, Cammue BPA. Different Approaches For Multi-Transgene-Stacking In Plants. Plant Science 2002; 281-295.
  3. 3. Halpin C. Gene Stacking In Transgenic Plants – The Challenge For 21st Century Plant Biotechnology. Plant Biotechnology Journal 2005;3 141-155.
  4. 4. Dafny-Yelin M. and Tzfira T. Delivery of Multiple Transgenes to Plant Cells. Plant Physiology 2007; 145 1118-1128.
  5. 5. Ma JK., Hiatt A, Hein M, Vine ND, Wang F, Stabila P, van Dolleweerd C, Mostov K, Lehner T. Generation And Assembly Of Secretory Antibodies In Plants. Science 1995;268 716-719.
  6. 6. Bizily SP, Rugh CL, Meagher RB. Phytodetoxification Of Hazardous Organomercurials By Genetically Engineered Plants. Nature Biotechnology 2000;18 213-217.
  7. 7. Lapierre C, Pollet B, Petit-Conil M, Toval G, Romero J, Pilate G, Leplé JC, Boerjan W, Ferret VV, de Nadai V, Jouanin L. Structural Alterations Of Lignins In Transgenic Poplars With Depressed Cinnamyl Alcohol Dehydrogenase Or Caffeic Acid O-Methyltransferase Activity Have An Opposite Impact On The Efficiency Of Industrial Kraft Pulping. Plant Physiology 1999;119 153-163.
  8. 8. Halpin C, Barakate A, Askari BM, Abbott JC, Ryan MD. Enabling Technologies For Manipulating Multiple Genes On Complex Pathways. Plant Molecular Biology 2001;47 295-310.
  9. 9. Chen L, Marmey P, Taylor NJ, Brizard J-P, Espinoza C, D’Cruz P, Huet H, Zhang S, de Kochko A, Beachy RN, Fauquet CM. Expression And Inheritance Of Multiple Transgenes In Rice Plants. Nature Biotechnology 1998; 16 1060-1064.
  10. 10. Komar AA, Hatzoglou M. Internal Ribosome Entry Sites In Cellular mRNAs: Mystery Of Their Existence. Journal of Biological Chemistry June 2005;280 23425-23428.
  11. 11. Toth RL, Chapman S, Carr F, Santa Cruz S. A Novel Strategy For The Expression Of Foreign Genes From Plant Virus Vectors. FEBS Letters 2001;489 215-219.
  12. 12. Groppelli E, Belsham GJ, Roberts LO. Identification Of Minimal Sequences Of The Rhopalosiphum padi Virus 5’ Untranslated Region Required For Internal Initiation Of Protein Synthesis In Mammalian, Plant And Insect Translation Systems. Journal of General Virology 2007;88 1583-1588.
  13. 13. Urwin PE, Yi L, Martin H, Atkinson HJ, Gilmartin PM. Functional Characterization Of The EMCV IRES In Plants. Plant Journal 2000;24 583-589.
  14. 14. Urwin PE, Zubko EI, Atkinson HJ. The Biotechnological Application And Limitation Of IRES To Deliver Multiple Defence Genes To Plant Pathogens. Physiological and Molecular Plant Pathology 2002;61(2) 103-108.
  15. 15. Spall VE, Shanks M, Lomonossoff GP. Polyprotein Processing as a Strategy for Gene Expression in RNA Viruses. Seminars in Virology 1997;8 15-23.
  16. 16. Marcos JF, Beachy RN. In-Vitro Characterization Of A Cassette To Accumulate Multiple Proteins Through Synthesis Of A Self-Processing Polypeptide. Plant Molecular Biology 1994;24 495-503.
  17. 17. von Bodman K, Domier LL, Farrand SK. Expression Of Multiple Eukaryotic Genes From A Single Promoter In Nicotiana Bio/Technology 1995;13 587-591.
  18. 18. Dasgupta S, Collins GB, Hunt AG. Co-Ordinated Expression Of Multiple Enzymes In Different Subcellular Compartments In Plants. Plant Journal 1998;16 107-116.
  19. 19. Zhang B, Rapolu M, Huang L, Su WW. Coordinate Expression Of Multiple Proteins In Plant Cells By Exploiting Endogenous Kex2p-Like Protease Activity. Plant Biotechnology Journal 2011;9 970-981.
  20. 20. Luke GA, Escuin H, de Felipe P, Ryan MD. 2A To The Fore, Research, Technology And Applications. In: Harding S. (ed) Biotechnology and Genetic Engineering Reviews. Loughborough: Nottingham University Press; 2009. Vol.26, 223-260.
  21. 21. Luke GA. Translating 2A Research into Practice. In: Agbo EC. (ed.) Innovations in Biotechnology. Rijeka: InTech; 2012. 165-186.
  22. 22. Ryan MD, Drew J. Foot-And-Mouth Disease Virus 2A Oligopeptide Mediated Cleavage Of An Artificial Polyprotein. The EMBO Journal 1994;13 928-933.
  23. 23. Donnelly MLL, Luke GA, Mehrotra A, Li X, Hughes LE, Gani D, Ryan MD. Analysis Of The Aphthovirus 2A/2B Polyprotein “Cleavage” Mechanism Indicates Not A Proteolytic Reaction, But A Novel Translational Effect: A Putative Ribosomal “Skip”. Journal of General Virology 2001;82 1013-1025.
  24. 24. Brown JD, Ryan MD. Ribosome “Skipping”: “Stop-Carry On” or “StopGo” Translation. In: Atkins JF and Gesteland RF (eds.) Recoding: Expansion of Decoding Rules Enriches Gene Expression. New York: Springer; 2010. 101-122.
  25. 25. Luke GA, Ryan MD. The Protein Coexpression Problem In Biotechnology And Biomedicine: Virus 2A And 2A-Like Sequences Provide A Solution. Future Virology 2013;8(10) 1-13.
  26. 26. Forss S, Strebel K, Beck E, Schaller H. Nucleotide Sequence And Genome Organization Of Foot-And-Mouth Disease Virus. Nucleic Acids Research 1984;12 6587-6601.
  27. 27. Ryan MD, Belsham GJ, King AM. Specificity Of Enzyme-Substrate Interactions In Foot-And-Mouth Disease Virus Polyprotein Processing. Virology 1989;173(1) 35-45.
  28. 28. Belsham GJ. Distinctive Features Of The Foot-And-Mouth Disease Virus, A Member Of The Picornavirus Family; Aspects Of Virus Protein Synthesis, Protein Processing And Structure. Progress in Biophysics and Molecular Biology 1993;60 241-260.
  29. 29. Ryan MD, King AM, Thomas GP. Cleavage Of Foot-And-Mouth Disease Virus Polyprotein Is Mediated By Residues Located Within A 19 Amino Acid Sequence. Journal of General Virology 1991;72 2727-2732.
  30. 30. Ryan MD, Donnelly MLL, Lewis A, Mehrotra AP, Wilkie J, Gani D. A Model For Non-Stoichiometric, Co-Translational Protein Scission In Eukaryotic Ribosomes. Bioorganic Chemistry 1999;27 55-79.
  31. 31. Donnelly MLL, Hughes LE, Luke GA, Li X, Mendoza H, ten Dam E, Gani D, Ryan MD. The “Cleavage” Activities Of FMDV 2A Site-Directed Mutants And Naturally Occurring “2A-Like” Sequences. Journal of General Virology 2001;82 1027-1041.
  32. 32. Ryan MD, Luke GA, Hughes LE, Cowton VM, ten Dam E, Li X, Donnelly MLL, Mehrotra A, Gani D. The Aphtho-and Cardiovirus “Primary” 2A/2B Polyprotein “Cleavage”. In: B.L Semler and E.Wimmer (eds.) Molecular Biology of Picornaviruses. Washington DC: ASM Press; 2002. 213-223.
  33. 33. Martĩnez-Salas E, Ryan MD. Translation and Protein Processing, In: The Picronaviruses, eds E. Ehrenfeld, E.Domingo. & R.P. RoosWashington DC:ASM Press; 2010. 141-161.
  34. 34. Gao Z-L, Zhou J-H, Zhang J, Ding Y-Z, Liu Y-S. The Silent Point Mutations At The Cleavage Site Of 2A/2B Have No Effect On The Self-Cleavage Activity Of 2A Of Foot-And-Mouth Disease Virus. Infection, Genetics and Evolution. 2014;28C 101-106.
  35. 35. Doronina VA, de Felipe P, Wu C, Sharma P, Sachs MS, Ryan MD, Brown JD. Dissection Of A Co-Translational Nascent Chain Separation Event. Biochemical Society Transactions 2008;36(4) 712-716.
  36. 36. Doronina VA, Wu C, de Felipe P, Sachs MS, Ryan MD, Brown JD. Site-Specific Release of Nascent Chains from Ribosomes at a Sense Codon. Molecular and Cellular Biology 2008;28(13) 4227-4239.
  37. 37. Sharma P, Yan F, Doronina VA, Escuin-Ordinas H, Ryan MD, Brown JD. 2A Peptides Provide Distinct Solutions To Driving Stop-Carry On Translational Recoding. Nucleic Acids Research 2012;40(7) 3143-3151.
  38. 38. Halpin C, Cooke SE, Barakate A, El Amrani A, Ryan MD. Self-processing 2A-polyproteins – a system for co-ordinate expression of multiple proteins in transgenic plants. The Plant Journal 1999;17(4) 453-459.
  39. 39. de Felipe P, Hughes LE, Ryan MD, Brown JD. Co-translational, Intraribosomal Cleavage of Polypeptides by the Foot-and-mouth Disease Virus 2A Peptide. The Journal of Biological Chemistry 2003;13 11441-11448,
  40. 40. Torres V, Barra L, Garcés F, Ordenes K, Leal-Ortiz S, Garner CC, Fernandez F, Zamorano P. A Bicistronic Lentiviral Vector Based On The 1D/2A Sequence Of Foot-And-Mouth Disease Virus Expresses Proteins Stoichiometrically. Journal of Biotechnology 2010;146 138-142.
  41. 41. Lorens JB, Pearsall DM, Swift SE, Peelle B, Armstrong R, Demo SD, Ferrick DA, Hitoshi Y, Payan DG, Anderson D. Stable, Stoichiometric Delivery Of Diverse Protein Functions. Journal of Biochemical and Biophysical Methods 2004;58 101-110.
  42. 42. de Felipe P, Ryan MD. Targeting of Proteins Derived from Self-Processing Polyproteins Containing Multiple Signal Sequences. Traffic 2004;5 616-626.
  43. 43. Samalova M, Fricker M, Moore I. Ratiometric Fluorescence-Imaging Assays of Plant Membrane Traffic Using Polyproteins. Traffic 2006;7 1701-1723.
  44. 44. Szymczak AL, Workman CJ, Wang Y, Vignali KM, Dilioglou S, Vanin EF, Vignali DA. Correction Of Multi-Gene Deficiency In Vivo Using A Single “Selfcleaving” 2A Peptide-Based Retroviral Vector. Nature Biotechnology 2004;22(5) 589-594.
  45. 45. Provost E, Rhee J, Leach SD. Viral 2A Peptides Allow Expression Of Multiple Proteins From A Single ORF In Transgenic Zebrafish Embryos. Genesis 2007;45(10) 625-629.
  46. 46. Trichas G, Begbie J, Srinivas S. Use Of The Viral 2A Peptide For Bicistronic Expression In Transgenic Mice. BMC Biology 2008;6 40.
  47. 47. Luke GA, de Felipe P, Lukashev A, Kallioinen SE, Bruno EA, Ryan MD. The Occurrence, Function, and Evolutionary Origins of “2A-like” Sequences in Virus Genomes. Journal of General Virology 2008;89 1036-1042.
  48. 48. Luke GA, Pathania US, Roulston C, de Felipe P, Ryan MD. DxExNPGP – Motives For The Motif. Recent Research Developments in Virology. 2014;9 25-42. ISBN:978-81-308-0540-5.
  49. 49. Heras SR, Thomas MC, García-Canadas M, de Felipe P, García-Perez JL, Ryan MD, Lopez MC. L1Tc Non-LTR Retrotransposons From Trypanosoma Cruzi Contain A Functional Viral-Like Self-Cleaving 2A Sequence In Frame With The Active Proteins They Encode. Cellular and Molecular Life Sciences 2006;63 1449-1460.
  50. 50. Odon V, Luke GA, Roulston C, de Felipe P, Ruan L, Escuin-Ordinas H, Brown JD, Ryan MD, Sukhodub A. APE-Type Non-LTR Retrotransposons of Multicellular Organisms Encode Virus-Like 2A Oligopeptide Sequences, Which Mediate Translational Recoding during Protein Synthesis. Molecular Biology and Evolution 2013;30(8) 1955-1965.
  51. 51. Blazar BR, Osborn MJ, Panoskaltsis-Mortari A, Bell SK, Vignali DAA, Ryan MD, Wilber AC, McIvor RS, Tolar J. A Picornaviral 2A-Like Sequence-Based Tricistronic Vector Allowing For High-Level Therapeutic Gene Expression Coupled To A Dual-Reporter System. Molecular Therapy 2005;12(3) 569-574.
  52. 52. Suzuki N, Geletka LM, Nuss DL. Essential And Dispensable Virus-Encoded Replication Elements Revealed By Efforts To Develop Hypoviruses As Gene Expression Vectors. Journal of Virology 2000;74(16) 7568-7577.
  53. 53. Dechamma HJ, Ashok Kumar C, Nagarajan G, Suryanarayana VVS. Processing Of Multimer FMD Virus VP1-2A Protein Expressed In E.Coli Into Monomers. Indian Journal of Experimental Biology 2008;46 760-763.
  54. 54. Donnelly MLL, Gani D, Flint M, Monaghan S, Ryan MD. The Cleavage Activities Of Aphthovirus And Cardiovirus 2A Proteins. Journal of General Virology 1997;78 13-21.
  55. 55. de Felipe P, Luke GA, Hughes LE, Gani D, Halpin C, Ryan MD. E unum pluribus: Multiple Proteins From A Self-Processing Polyprotein. Trends in Biotechnology 2006;24(2) 68-75.
  56. 56. Kim JH, Lee S-R, Li L-H, Park H-J, Park J-H, Lee KY, Kim M-K, Shin BH, Choi S-Y. High Cleavage Efficiency of a 2A Peptide Derived from Porcine Teschovirus-1 in Human Cell Lines. Zebrafish and Mice. PLos ONE 2011;6(4) Article ID e18556.2011.
  57. 57. Ma C, Mitra A. Expressing Multiple Genes In A Single Open Reading Frame With The 2A Region Of Foot-And-Mouth Disease Virus As A Linker. Molecular Breeding 2002;9 191-199.
  58. 58. El-Amrani A, Barakate A, Askari BM, Li X, Roberts AG, Ryan MD, Halpin C. Coordinate Expression and Independent Subcellular Targeting of Multiple Proteins from a Single Transgene. Plant Physiology 2004;135 16-24.
  59. 59. Burén S, Ortega-Villasante C, Blanco-Rivero A, Martínez-Bernardini A, Shutova T, Shevela D, Messinger J, Bako L, Villarejo A, Samuelsson G. Importance of Post-Translational Modifications for Functionality of a Chloroplast-Localized Carbonic Anhydrase (CAH1) in Arabidopsis thaliana. PloS ONE 2011;6(6): e21021.doi:10.1371/journal.pone.0021021.
  60. 60. Burén S, Ortega-Villasante C, Ötvös K, Samuelsson G, Bakó L, Villarejo S. Use of the Foot-and-Mouth Disease Virus 2A Peptide Co-Expression System to Study Intracellular Protein Trafficing in Arabidopsis PLoS ONE 2012;7(12):e51973.doi:10.1371/journal.pone.0051973
  61. 61. Yang Z. Small GTPases: Versatile Signaling Switches in Plants. The Plant Cell 2002;Supplement S375-S388.
  62. 62. Jürgens G. Membrane Trafficking in Plants. Annual Review of Cell and Developmental Biology. 2004;20 481-504.
  63. 63. Hutagalung AH, Novick PJ. Role of Rab GTPases in Membrane Traffic and Cell Physiology. Physiological Reviews 2011;91 119-149.
  64. 64. Cevher-Keskin B. ARF1 and SAR1 GTPases in Endomembrane Trafficking in Plants. International Journal of Molecular Sciences 2013;14 18181-18199.
  65. 65. Cox R, Mason-Gamer RJ, Jackson CL, Segev N. Phylogenetic Analysis Of Sec7-Domain-Containing Arf Nucleotide Exchangers. Molecular Biology of the Cell 2004;15 1487-1505.
  66. 66. Mouratou B, Biou V, Joubert A, Cohen J, Shields DJ, Geldner N, Jürgens G, Melançon P, Cherfils J. The Domain Architecture Of Large Guanine Nucleotide Exchange Factors For The Small GTP-Binding Protein Arf. BMC Genomics 2005;6 20.
  67. 67. Richter S, Geldner N, Schrader J, Wolters H, Stierhof Y-D, Rios G, Koncz C, Robinson DG, Jürgens G. Functional Diversification Of Closely Related ARF-Gefs In Protein Secretion And Recycling. Nature 2007;448 488-492.
  68. 68. Teh O-K, Moore I. An ARF-GEF Acting At The Golgi And In Selective Endocytosis In Polarized Plant Cells. Nature 2007;448 493-496.
  69. 69. Lengler J, Holzmuller H, Salmons B, Gunzburg WH, Renner M. FMDV-2A Sequence And Protein Arrangement Contribute To Functionality Of CYP2B1-Reporter Fusion Protein. Analytical Biochemistry 2005;343 116-124.
  70. 70. Chinnasamy D, Milsom MD, Shaffer J, Neuenfeldt J, Shaaban AF, Margison GP, Fairbairn LJ, Chinnasamy N. Multicistronic Lentiviral Vectors Containing The FMDV 2A Cleavage Factor Demonstrate Robust Expression Of Encoded Genes At Limiting MOI. Virology Journal 2006;3 14.
  71. 71. Rothwell DG, Crossley R, Bridgeman JS, Sheard V, Zhang Y, Sharp TV, Hawkins RE, Gilham DE, McKay TR. Functional Expression Of Secreted Proteins From A Bicistronic Retroviral Cassette Based On FMDV 2A Can Be Position-Dependent. Human Gene Therapy 2010;21(11) 1631-1637.
  72. 72. Sun H, Lang Z, Zhu L, Huang D. Acquiring Transgenic Tobacco Plants With Insect Resistance And Glyphosate Tolerance By Fusion Gene Transformation. Plant Cell Reports 2012;31 1877-1887.
  73. 73. Minskaia E, Ryan MD. Protein Coexpression Using FMDV 2A: Effect of “Linker” Residues. BioMed Research International Volume 2013, Article ID 291730.
  74. 74. Minskaia E, Nicholson J, Ryan MD. Optimisation Of The Foot-And-Mouth Disease Virus 2A Co-Expression System For Biomedical Applications. BMC Biotechnology 2013;13 67.
  75. 75. Holst J, Szymczak-Workman AL, Vignali KM, Burton AR, Workman CJ, Vignali DAA. Generation Of T-Cell Receptor Retrogenic Mice. Nature Protocols 2006;1(1) 406-417.
  76. 76. Holst J, Vignali KM, Burton AR, Vignali DAA. Rapid Analysis Of T-Cell Selection In Vivo Using T Cell-Receptor Retrogenic Mice. Nature Methods 2006;3(3) 191-197.
  77. 77. Wargo JA, Robbins PF, Li Y, Zhao Y, El-Gamil M, Caragacianu D, Zheng Z, Hong JA, Downey S, Schrump DS, Rosenberg SA, Morgan RA. Recognition Of NY-ESO-1+Tumor Cells By Engineered Lymphocytes Is Enhanced By Improved Vector Design And Epigenetic Modulation Of Tumor Antigen Expression. Cancer Immunology, Immunotherapy 2009;58(3) 383-394.
  78. 78. Yang S, Cohen CJ, Peng PD, Zhao Y, Cassard L, Yu Z, Zheng Z, Jones S, Restifo NP, Rosenberg SA, Morgan RA. Development Of Optimal Bicistronic Lentiviral Vectors Facilitates High-Level TCR Gene Expression And Robust Tumor Cell Recognition. Gene Therapy 2008;15(21) 1411-1423.
  79. 79. Tan Y, Liang H, Chen A, Guo X. Coexpression Of Double Or Triple Copies Of The Rabies Virus Glycoprotein Gene Using A “Self-Cleaving” 2A Peptide-Based Replication-Defective Human Adenovirus Serotype 5 Vector. Biologicals 2010;38 586-593.
  80. 80. Hardesty B, Kramer G. Folding Of Nascent Peptide On The Ribosome. Progress in Nucleic Acid Research & Molecular Biology 2001;66 41-66.
  81. 81. Varshavsky A. The N-End Rule. Cell 1992;69(5) 725-735.
  82. 82. Fang J, Qian JJ, Yi S, Harding TC, Tu GH, VanRoey M, Jooss K. Stable Antibody Expression At Therapeutic Levels Using The 2A Peptide. Nature Biotechnology 2005;23(5) 584-590.
  83. 83. Fang JM, Yi SL, Simmons A, Tu GH, Nguyen M, Harding TC, VanRoey M, Jooss K. An Antibody Delivery System For Regulated Expression Of Therapeutic Levels Of Monoclonal Antibodies in vivo. Molecular Therapy 2007;15(6) 1153-1159.
  84. 84. Molloy SS, Bresnahan PA, Leppla SH, Klimpel KR, Thomas G. Human Furin Is a Calcium-dependent Serine Endoprotease That Recognizes the Sequence Arg-X-X-Arg and Efficiently Cleaves Anthrax Toxin Protective Antigen. The Journal of Biological Chemistry 1992;267 16396-16402.
  85. 85. Liu G, Thomas L, Warren RA, Enns CA, Cunningham CC, Hartwig JH, Thomas G. Cytoskeletal Protein ABP-280 Directs The Intracellular Trafficking Of Furin And Modulates Proprotein Processing In The Endocytic Pathway. The Journal of Cell Biology 1997;139(7) 1719-1733.
  86. 86. François IEJA, Hemelrijck WV, Aerts AM. Processing In Arabidopsis thaliana Of A Heterologous Polyprotein Resulting In Differential Targeting Of The Individual Plant Defensins. Plant Science 2004;166(1) 113-121.
  87. 87. Tailor RH, Acland DP, Attenborough S, Cammue BP, Evans IJ, Osborn RW, Ray JA, Rees SB, Broekaert WF. A Novel Family Of Small Cysteine-Rich Antimicrobial Peptides From Seed Of Impatiens balsamina Is Derived From A Single Precursor Protein. Journal of Biological Chemistry 1997;272(39) 24480-24487.
  88. 88. Franklin SE, Mayfield SP. Prospects For Molecular Farming In The Green Alga Chlamydomonas reinhardtii. Current Opinion in Plant Biology 2004;7 159-165.
  89. 89. Mayfield SP, Manuell AL, Chen S, Wu J, Tran M, Siefker D, Muto M, Marin-Navarro J. Chlamydomonas reinhardtii Chloroplasts As Protein Factories. Current Opinions in Biotechnology 2007;18 126-133.
  90. 90. Specht E, Miyake-Stoner S, Mayfield S. Micro-Algae Come Of Age As A Platform For Recombinant Protein Production. Biotechnology Letters 2010;32 1373-1383.
  91. 91. Merchant SS, Prochnik SE, Vallon O, Harris EH, Karpowicz SJ, et al. The Chlamydomonas Genome Reveals The Evolution Of Key Animal And Plant Functions. Science 2007;318 245-251.
  92. 92. Kindle KL. High-Frequency Nuclear Transformation Of Chlamydomonas reinhardtii. Proceedings of the National Academy of Sciences USA 1990;87 1228-1232.
  93. 93. Leon-Banares R, Gonzalez-Ballester D, Galvan A, Fernandez E. Transgenic Microalgae As Green Cell-Factories. Trends in Biotechnology 2004;22 45-52.
  94. 94. Rasala BA, Lee PA, Shen Z, Briggs SP, Mendez M, Mayfield SP. Robust Expression and Secretion of Xylanase 1 in Chlamydomonas reinhardtii by Fusion to a Selection Gene and Processing with the FMDV 2A Peptide. PLoS ONE 2012 7(8): e43349.doi:10.1371/journal.pone.0043349.
  95. 95. Dhiman SS, Sharma J, Battan B. Industrial Applications And Future Prospects Of Microbial Xylanases: A Review. BioResources 2008;3(4) 1377-1402.
  96. 96. Rasala BA, Barrera DJ, Ng J, Plucinak TM, Rosenberg JN, Weeks DP, Oyler GA, Peterson TC, Haerizadeh F, Mayfield SP. Expanding The Spectral Palette Of Fluorescent Proteins For The Green Microalga Chlamydomonas reinhardtii. The Plant Journal 2013;74(4) 545-556.
  97. 97. Ye X, Al-Babili S, Klöti A, Zhang J, Lucca P, Beyer P, Potrykus I. Engineering The Provitamin A (Beta-Carotene) Biosynthetic Pathway Into (Carotenoid-Free) Rice Endosperm. Science 2000;287(5451):303-305.
  98. 98. Paine JA, Shipton CA, Chaggar S; Howells RM, Kennedy MJ, Vernon G, Wright SY, Hinchliffe E, Adams JL, Silverstone Al, Drake R. Improving The Nutritional Value Of Golden Rice Through Increased Pro-Vitamin A Content. Nature Biotechnology 2005;23(4) 482–487.
  99. 99. Ha SH, Liang YS, Jung H, Ahn M-J, Suh S-C, Kweon S-J, Kim D-H, Kim Y-M, Kim J-K. Application Of Two Bicistronic Systems Involving 2A And IRES Sequences To The Biosynthesis Of Carotenoids In Rice Endosperm. Plant Biotechnology Journal 2010;8(8) 928-938.
  100. 100. Ralley L, Enfissi EMA, Misawa N, Schuch W, Bramley PM, Fraser PD. Metabolic Engineering Of Ketocarotenoid Formation In Higher Plants. Plant Journal 2004;39 477-486.
  101. 101. Iordachescu M, Imai R. Trehalose and Abiotic Stress in Biological Systems. In: Abiotic Stress in Plants-Mechanisms and Adaptations, Prof. Arun Shanker (Ed.), pp215-234. ISBN: 978-953-307-394-1, InTech, DOI: 10.5772/22208. Available from: http://www.intechopen.com/books/abiotic-stress-in-plants-mechanisms-and-adaptations/trehalose-and-abiotic-stress-in-biological-systems
  102. 102. Goddijn OJM, Van Dun K. Trehalose Metabolism In Plants. Trends Plant Science 1999;4 315-319.
  103. 103. Holmström KO, Mantyla E, Welin B, Mandal A, Palva ET, Tunnela OE, Londesborough J. Drought Tolerance In Tobacco. Nature 1996;379 683-684.
  104. 104. Goddijn OJM, Verwoerd TC, Voogd E, Krutwagen RWHH, de Graaf PTHM, van Kun K, Poels J, van Dun K, Ponstein AS, Damm B, Pen J. Inhibition Of Trehalose Activity Enhances Trehalose Accumulation In Transgenic Plants. Plant Physiology 1997;113(1) 181-190.
  105. 105. Jang IC, Oh SJ, Seo JS, Choi WB, Song SI, Kim CH, Kim YS, Seo HS, Choi YD, Nahm BH, Kim JK. Expression Of A Bifunctional Fusion Of The Escherichia Coli Genes For Trehalose-6-Phosphate Synthase And Trehalose-6-Phosphate Phosphatase In Transgenic Rice Plants Increases Trehalose Accumulation And Abiotic Stress Tolerance Without Stunting Growth. Plant Physiology 2003;131(2) 516-524.
  106. 106. Kwon SJ, Hwang EW, Kwon HB. Genetic Engineering Of Drought Resistant Potato Plants By Co-Introduction Of Genes Encoding Trehalose-6-Phosphate Phosphatase Of Zygosaccharomyces rouxii. Korean Journal Genetics 2004;26(2) 199-206.
  107. 107. Le Rudulier D, Strom AR, Dandekar AM, Smith LT, Valentine RC. Molecular Biology Of Osmoregulation. Science 1984;224 1064-1068.
  108. 108. Wang S, Yao Q, Tao J, Qiao Y, Zhang Z. Co-Ordinate Expression Of Glycine Betaine Synthesis Genes Linked By The FMDV 2A Region In A Single Open Reading Frame In Pichia pastoris. Applied Genetics and Molecular Biotechnology 2007;77 891-899.
  109. 109. Hammond-Kosack KE, Jones JDG. Response To Plant Pathogens. In: Buchannan B, Gruissern W, Jones R. eds Biochemistry and molecular biology of plants. Rockville, MD: American Society of Plant Physiologists. 2002 1102-1157. ISBN:978-0-943088-39-6.
  110. 110. Lay FT, Anderson MA. Defensins, Components Of The Innate Immune System In Plants. Current Protein and Peptide Science 2005;6 85-101.
  111. 111. Manners JM. Hidden Weapons Of Microbial Destruction In Plant Genomes. Genome Biology 2007;8 225. Doi:10.1186/9b-2007-8-9-225.
  112. 112. Mittler R, Vanderrauwera S, Gollery M, van Breusegem F. Reactive Oxygen Network Of Plants. Trends in Plant Science 2004;9(10) 490-498.
  113. 113. Boman HG. Antibacterial Peptides: Key Components Needed In Immunity. Cell 1991;65(2) 205-207.
  114. 114. Kovalskaya N, Zhao Y, Hammond RW. Antibacterial and Antifungal Activity of a Snakin-Defensin Hybrid Protein Expressed in Tobacco and Potato Plants. The Open Science Journal 2011;5 29-42.
  115. 115. Segura A, Moreno M, Madueno F, Molina A, Garcia-Olmedo F. Snakin-1, A Peptide From Potato That Is Active Against Plant Pathogens. Molecular Plant-Microbe Interactions 1999;12 16-23.
  116. 116. Broekaert WF, Terras FRG, Cammue BPA, Osborn RW. Plant defensins: Novel Antimicrobial Peptides As Components Of The Host Defense System. Plant Physiology 1995;108 1353-1358.
  117. 117. Quilis J, López-García B, Meynard D, Guiderdoni E, San Segundo B. Inducible Expression Of A Fusion Gene Encoding Two Proteinase Inhibitors Leads To Insect And Pathogen Resistance In Transgenic Rice. Plant Biotechnology Journal 2014;12 367-377.
  118. 118. Wittstock U, Kliebensteiin D, Lambrix VM, Reichelt M, Gershenzon J. Glucosinolate Hydrolysis And Its Impact On Generalist And Specialist Insect Herbivores. In: Romeo JT (ed) Integrative phyto-chemistry: From ethnobotany to molecular ecology. 2003 vol 37. Elsevier, Amsterdam, pp 101-125.
  119. 119. Brader G, Mikkelsen MD, Halkier BA, Tapio Palva E. Altering Glucosinolate Profiles Modulates Disease Resistance In Plants. Plant Journal 2006;46 758-767.
  120. 120. Sun JY, Sonderby IE, Halkier BA, Jander G, de Vos M. Non-Volatile Intact Indole Glucosinolates Are Host Recognition Cues For Ovipositing Plutella xylostella. Journal of Chemical Ecology 2009;35 1427-1436.
  121. 121. Furlong MJ, Wright DJ, Dosdall LM. Diamondback Moth Ecology And Management: Problems, Progress, And Prospects. Annual Reviews in Entomology 2013;58 517-541.
  122. 122. Geu-Flores F, Olsen CE, Halkier BA. Towards Engineering Glucosinolates Into Non-Cruciferous Plants. Planta 2009;229 261-270.
  123. 123. Møldrup ME, Geu-Flores F, de Vos M, Olsen CE, Sun J, Jander G, Halkier BA. Engineering Of Benzylglucosinolate In Tobacco Provides Proof-Of-Concept For Dead-End Trap Crops Genetically Modified To Attract Plutella Xylostella (Diamondback Moth). Plant Biotechnology Journal 2012;10 435-442.
  124. 124. Atkinson NJ, Urwin PE. The Interaction Of Plant Biotic And Abiotic Stresses: From Genes To Field. Journal of Experimental Botany 2012;63(10) 3523-3544.
  125. 125. Witcombe JR, Hollington PA, Howarth CJ, Reader S, Steele KA. Breeding For Abiotic Stresses For Sustainable Agriculture. Philosophical Transactions of the Royal Society B: Biological Sciences 2008;363 703-716.
  126. 126. Kumar R, Singh S, Singh OV. Bioconversion Of Lignocellulosic Biomass: Biochemical And Molecular Perspectives. Journal of Industrial Microbiology and Biotechnology 2008;35 377-391.
  127. 127. Kumar P, Barrett DM, Delwiche MJ, Stroeve P. Methods for Pretreatment of Lignocellulosic Biomass for Efficient Hydrolysis and Biofuel Production. Industrial and Engineering Chemistry Research 2009;48 3713-3729.
  128. 128. Klein-Marcuschamer D, Oleskowicz-Popiel P, Simmons BA, Blanch HW. The Challenge Of Enzyme Cost In The Production Of Lignocellulosic Biofuels. Biotechnology and Bioengineering 2012;109 1083-1087.
  129. 129. Tremblay R, Diao H, Huner N, Jevnikar AM, Ma S. The Development, Characterization, And Demonstration Of A Novel Strategy For Purification Of Recombinant Proteins Expressed In Plants. Transgenic Research 2011;20 1357-1366.
  130. 130. Jung S, Lee D-S, Kim Y-O, Joshi CP, Bae H-J. Improved Recombinant Cellulose Expression In Chloroplast Of Tobacco Through Promoter Engineering And 5’ Amplification Promoting Sequence. Plant Molecular Biology 2013;83 317-328.
  131. 131. Dai Z, Hooker BS, Quesenberry RD, Thomas SR. Optimization Of Acidothermus Cellulolyticus Endoglucanase (E1) Production In Transgenic Tobacco Plants By Transcriptional, Post-Transcription And Post-Translational Modification. Transgenic Research 2005;14 627-643.
  132. 132. Harrison MD, Geijskes J, Coleman HD, Shand K, Kinkema M, Palupe A, Hassall R, Sainz M, Lloyd R, Miles S, Dale JL. Accumulation Of Recombinant Cellobiohydrolase And Endoglucanase In Leaves Of Mature Transgenic Sugar Cane. Plant Biotechnology Journal 2011;9 884-896.
  133. 133. Verma D, Kanagaraj A, Jin S, Singh ND, Kolattukudy PE, Daniell H. Chloroplast-Derived Enzyme Cocktails Hydrolyse Lignocellulosic Biomass And Release Fermentable Sugars. Plant Biotechnology Journal 2010;8 332-350.
  134. 134. Lee D-S, Lee K-H, Jung S, Jo E-J, Han K-H, Bae H-J. Synergistic Effects Of 2A-Mediated Polyproteins On The Production Of Lignocellulose Degradation Enzymes In Tobacco Plants. Journal of Experimental Botany 2012;63(13) 4797-4810.
  135. 135. Kim S, Lee DS, Choi IS, Ahn SJ, Kim YH, Bae HJ. Arabidopsis Thaliana Rubisco Small Subunit Transit Peptide Increases The Accumulation Of Thermotoga Maritima Endoglucanase Cel5A In Chloroplasts Of Transgenic Tobacco Plants. Transgenic Research 2010;19 489-497.
  136. 136. Jung S, Kim S, Bae H, Lim H-S, Bae H-J. Expression Of Thermostable Bacterial β-Glucosidase (Bglb) In Transgenic Tobacco Plants. Bioresource Technology 2010;101 7144-7150.
  137. 137. Santa Cruz S, Chapman S, Roberts AG, Roberts IM, Prior DAM, Oparka KJ. Assembly And Movement Of A Plant Virus Carrying A Green Fluorescent Protein Overcoat. Proceedings Of The National Academy Of Sciences USA 1996; 93 6286-6290.
  138. 138. Yamaji Y, Maejima K, Komatsu K, Shiraishi T, Okano Y, Himeno M, Sugawara K, Neriya Y, Minato N, Miura C, Hashimoto M, Namba S. Lectin-Mediated Resistance Impairs Plant Virus Infection At The Cellular Level. Plant Cell 2012; 24 778-793.
  139. 139. Smolenska L, Roberts IM, Learmonth D, Porter AJ, Harris WJ, Wilson TMA, Santa Cruz S. Production Of A Functional Single Chain Antibody Attached To The Surface Of A Plant Virus. FEBS Letters. 1998; 441 379-382.
  140. 140. O’Brien GJ, Bryant CJ, Voogd C, Greenberg HB, Gardner RC, Bellamy AR. Rotavirus VP6 Expressed By PVX Vectors In Nicotiana Benthamiana Coats PVX Rods And Also Assembles Into Virus Like Particles. Virology 2000; 270 444-453.
  141. 141. Marconi G, Albertini E, Barone P, De Marchis F, Lico C, Marusic C, Rutili D, Veronesi F, Porceddu A. In planta production of two peptides of the classical swine fever virus (CSFV) E2 glycoprotein fused to the coat protein of potato virus X. BMC Biotechnology. 2006; 6 29.
  142. 142. Zelada AM, Calamante G, de la Paz Santangelo M, Bigi F, Verna F, Mentaberry A, Cataldi A. Expression Of Tuberculosis Antigen ESAT-6 In Nicotiana Tabacum Using A Potato Virus X-Based Vector. Tuberculosis 2006; 86 263-267.
  143. 143. Uhde-Holzem K, Schlösser V, Viazov S, Fischer R, Commandeur A. Immunogenic properties of chimeric potato virus X particles displaying the hepatitis C virus hypervariable region I peptide R9. Journal Of Virological Methods 2010; 166 12-20.
  144. 144. Sempere RN, Gómez P, Truniger V, Aranda MA. Development Of Expression Vectors Based On Pepino Mosaic Virus. Plant Methods 2011; 7 6.
  145. 145. Minato N, Komatsu K, Okano Y, Maejima K, Ozeki J, Senshu H, Takahashi S, Yamaji Y, Namba S. Efficient Foreign Gene Expression In Planta Using A Plantago Asiatica Mosaic Virus-Based Vector Achieved By The Strong RNA-Silencing Suppressor Activity Of Tgbp1. Archives Of Virology. 2014; 150 885-896.
  146. 146. Gopinath K, Wellink J, Porta C, Taylor KM, Lomonossoff GP, van Kammen A. Engineering Cowpea Mosaic Virus RNA-2 Into A Vector To Express Heterologous Proteins In Plants. Virology 2000; 267 159-173.
  147. 147. Zhang C, Bradshaw JD, Whitham SA, Hill JH. The Development Of An Efficient Multipurpose Bean Pod Mottle Virus Viral Vector Set For Foreign Gene Expression And RNA Silencing. Plant Physiology 2010; 153 52-65.
  148. 148. Tatineni S, McMechan AJ, Hein GL, French R. Efficient And Stable Expression Of GFP Through Wheat Streak Mosaic Virus-Based Vectors In Cereal Hosts Using A Range Of Cleavage Sites: Formation Of Dense Fluorescent Aggregates For Sensitive Virus Tracking. Virology 2011; 410 268-281.
  149. 149. Christian M, Qi Y, Zhang Y, Voytas DF. Targeted mutagenesis of Arabidopsis thaliana using engineered TAL effector nucleases. G3 (Bethesda). 2013; 3 1697-1705.
  150. 150. Randall J, Sutton D, Ghoshroy S, Bagga S, Kemp JD. Co-Ordinate Expression Of Beta-And Delta-Zeins In Transgenic Tobacco. Plant Science 2004; 167 367-372.

Written By

Garry A. Luke, Claire Roulston, Jens Tilsner and Martin D. Ryan

Published: 15 April 2015