Open access peer-reviewed chapter

Genetically Modified Crops and Their Impact on New Era of Agriculture

Written By

Palanisamy Dhamotharan, Sakthivel Surya Krishna, Sampathkumar Sruthi and Subburaj Monica

Submitted: 05 May 2022 Reviewed: 19 June 2022 Published: 06 September 2022

DOI: 10.5772/intechopen.105937

From the Edited Volume

Case Studies of Breeding Strategies in Major Plant Species

Edited by Haiping Wang

Chapter metrics overview

139 Chapter Downloads

View Full Metrics

Abstract

Genetically engineered crops are different from genetically modified crops. Changes in genetic make-up of crops by any conventional or any means technology fall under genetically modified crops category. In general, among different traits, herbicide and pest tolerance were more exploited in which herbicide tolerant crops occupy more than 90% of the total cultivated area of genetically engineered crops as the mode of actions of herbicides are well known and tolerant gene sources are readily available in wild weedy forms and various microorganisms. By knowing the pathway of mode of action of herbicides and pest tolerance, herbicide and pest tolerant crops were transformed by altering the structure and functions of rate limiting enzymes without affecting the normal functionalities of plants. Other than these two economically important characters, several characters were engineered in different crop plants such as disease resistant, increased yield and biomass production, male sterility and quality characters like anti-allergy factors, modified composition of fatty acid, protein, increased micronutrients and enzyme contents, reduced contents of anti-nutritional factors and toxic substances. Through genetic engineering, transformed plants are used for establishing pharmaceutical products. In terms of area coverage, soya-bean followed maize, cotton, canola, alfalfa. Apart from genetically engineered crops, genome edited crops are developed for nullifying the negative effects and up-regulating the target traits having positive effects thus ultimately assisting in achieving food and nutritional security.

Keywords

  • agriculture
  • GE/GM crops
  • RNAi
  • ZFN
  • TALEN
  • CRISPR/Cas9

1. Introduction

Plant genome engineering aims to modify crops by incorporating agronomically desirable traits, which could not be achieved through conventional plant breeding methods. Genetically modified crops vary considerably from genetically engineered crops in that the former recruits the modification through both natural and artificial means, whereas the latter is distinguished only by artificial means that would not occur naturally. Genetically modified (GM) crops have proven to be great complement to conventional crops in meeting global demands for increased yields, increased food security, decreased pesticide use, and higher nutritional quality. The modified composition of fatty acid, protein and increased micronutrients, resistance to pests and disease, male sterility, quality characters like anti-allergy factors, reduced contents of anti-nutritional factors and toxic substances and medical benefits among which herbicide and pest tolerance are more remarkably exploited. The genetically modified herbicide tolerance crops occupy more than 90% of the total cultivated area of genetically engineered crops as the mode actions of herbicides are familiar and tolerant gene sources are more obtainable in wild weedy forms and microorganisms, which paved way for transformation by altering the structure and functions of rate-limiting enzymes without affecting the normal functionalities of plants. In terms of global acreage, genetically modified soya beans are widely cultivated followed by maize, cotton, canola, and alfalfa. However, apart from genetically modified crops, genome-edited crops are generated by nullifying the negative effects of traits of interest while up-regulating the positive effects of the target traits. Genetically modified crops benefit humanity by increasing the availability and quality of food and medical care, as well as contributing to a cleaner environment and alleviating hunger and disease around the world. In this chapter, a detailed discussion addressing the global need for resistance to insect pests, disease, herbicide tolerance, stress tolerance, quality improvement, male sterility and yield improvement through genetic modification in crop plants have been made.

Advertisement

2. Insect resistance

One of the most difficult issues in plant crop cultivation is dealing with insect pests. Insect pests are primarily managed using insecticides, but the rising occurrence of insect resistance genes in many organisms could be harnessed and introduced to crop plants through the effective use of transgenic technology. The cloning of genes codes for insecticidal δ-endotoxins dates back to the early 1980s [1]. Transgenic tomato and tobacco produce modified toxin genes which provide insect resistance res the first examples of genetically modified plants [2, 3, 4]. As insect-resistance genes transferred into plants predominantly act on the digestive system of insects, researchers are currently identifying genes with distinct modes of action to combat the development of resistance in the target insects, and to enhance potency. Few noteworthy insect control proteins (ICPs) such as protease inhibitors, different enzymes, ribosomal inactivating proteins, and lectins derived from various genus and domains that have an antimetabolic or toxic effect on insect digestion are being viewed as an alternative to control insect infestation or confer resistance to plants.

2.1 Source of transgenes

2.1.1 Resistance gene from microorganisms

2.1.1.1 Cry gene from Bacillus thuringiensis

The most important and successful example of a transgene derived from the bacterium B. thuringiensis is Bt gene toxin [5]. It is a significant soil borne spore-forming bacteria that produces insecticidal crystal (Cry) proteins encoded by crygenes (cry-represents gene; Cry-represents protein) during sporulation [5, 6]. Proteinases in the insect gut proteolytically cleave the inactive protoxins to produce the active 60–65Kd toxin made up of high homology regions interspersed with (hyper)variable regions. Sequence analysis [7, 8] and X-ray crystallography were used to infer the structure and functional roles of the toxin’s three domains [9] that bind to glycoprotein receptors in the brush border membrane of the midgut epithelium of susceptible insects. The nature of the receptors that explicitly play a significant role in establishing susceptibility/resistance to a specific Bt toxin, is under intensive investigation, with a number of midgut integral membrane glycoproteins, including aminopeptidase and a cadherin-like protein, being identified [10, 11, 12, 13]. Following binding, the toxin rapidly and irreversibly inserts into the cell membrane resulting in the formation of a pore supposedly by Cry proteins leading to disruption of the electrical, K+ and pH gradients eventually causing irreversible damage to the midgut wall which gives rise to epithelial cell lysis [6, 14] paving way to gut paralysis, cessation of feeding and finally (typically after 1–3 days) death from starvation and/or septicaemia. These toxins are mainly targeted enzymes and lectins of digestive systems of Lepidopteron and coleopteran pests.

2.1.1.1.1 Codon optimization for higher expression

Specificity, efficiency and insecticidal activity of toxins are more in codon optimised proteins than native form of Bt toxin and hence the former was introduced into various plants for increased level of expression [14]. Plants harbouring modified codon with plant preference rather than bacterial preference in which G:C rich codons are preferred over A:T rich codons, and undesirable mRNA secondary structure and polyadenylation signals eliminated, produced 100X higher than plants were transformed with unmodified (native) Bt gens [15, 16]. Genes are frequently inserted with constitutive promoters such as maize ubiquitin, CaMV35S, or rice actin 1, promoted protein expression at all times and in all parts of the plant allowing broader-spectrum ICPs to be targeted at different components of the pest complex. The use of tissue specific (e.g., RsS1 promoter for phloem-specific expression or PHA-L promoter for seed-specific expression) or inducible (e.g., potato pin2 wound-induced promoter) promoters is recommended to conserve the space and time of expression of toxic proteins and thus avoid unfavourable interactions with the beneficial insect’s ecosystem [17].

The highest risk of resistance development would most likely arise from prolonged exposure to ineffective levels of the transgene, a situation that farmers would not tolerate and would almost certainly necessitate additional (different) control measures (deployment of which would in fact reduce the risk of resistance development). CaMV35S has a notoriously low/no expression in pollen. Temporal and spatial promoters are used to target the pest at its most vulnerable stage and time of infestation.

2.1.1.1.2 Cry protein classification

Cry protein classification is based on the degree of homology of Cry proteins. Primary Cry protein group: Cry proteins with less than 45 percent amino acid similarity fall into this category. Cry 1 to Cry 78, for example (in 2018). Secondary Cry protein group: Cry proteins with less than 78 percent amino acid similarity fall into this category. Cry 1A and Cry 1B are two examples. Tertiary Cry protein group: Cry proteins with less than 95% amino acid similarity fall into this category. Cry 1Aa, Cry 1Ac, etc. Cry proteins that share more than 95% of their amino acid sequences are classified as part of the Quandary Cry protein group. Cry 1Aa1, Cry 1Aa2 (Cry 1Aa25 was recently discovered in 2019). Each Cry protein of the Bt bacterium has a distinct host range cry protein (Table 1) [18].

ClassificationPercentage of homology of amino acids (%)examples
Primary group< 45Cry 1 to Cry 78
Secondary group< 78Cry 1A, Cry 1B
Tertiary group< 95Cry 1Aa, Cry 1Ac
Quandary group> 95Cry 1Aa1, Cry 1Aa2

Table 1.

Classification of cry proteins based on their amino acids classification.

Even within the Cry protein subfamily, the toxic spectrum varies depending on the host. Cry1A and Cry1C proteins, for example, are toxic to larvae of lepidopteran pests such as the codling moth (Cydia pomonella), the European corn borer (Ostrinia nubilalis) [19], and heliothine bollworms, respectively. The Cry3A protein, on the other hand, is toxic to coleopteran pests such as the Colorado potato beetle (Leptinotarsa decemlineata) [20]. So far, 26 plant species have been genetically engineered and expressed for the Bt toxin [18].

2.1.1.2 Ipt gene from Agrobacterium tumefaciens

The introduction of Isopentenyl transferase gene (ipt) isolated from Agrobacterium tumefacien encoding a key rate limiting enzyme in the cytokinin biosynthetic pathway into the tobacco with wound inducible promoter recorded reduced consumption of leaves by the tobacco hornworm (Manduca sexta) and reduced survival of the peach potato aphid (Myzus persicae) leaving negative effects on plant development, such as an underdeveloped root system and a reduced total chlorophyll content [21].

2.1.1.3 Cholesterol-oxidase gene from streptomycete fungus

A cholesterol-oxidase gene from a streptomycete fungus has also been engineered into tobacco that was highly toxic to larvae of the boll weevil (Anthonomus grandis) and retarded the growth of the tobacco budworm (Heliothis virescens) by damaging the membranes of the insect-midgut epithelium [21, 22, 23].

2.1.1.4 Chitinase gene from Serratia marcesens

A bacterial endochitinase (from Serratia marcesens) has been shown to work in tandem with Bt toxin against S. littoralis larvae [24], but not (yet) in transgenic plants.

2.1.2 Resistance gene from higher plants

Plants have co-evolved with insects for millions of years, and have developed many adaptations in terms of antifeeding and anti-shelter, among which digestive enzyme inhibitors (proteinase and amylase inhibitors) and lectins have shown significant effect on insects, which have been transferred into crop plants without major alteration, and expression has been at a similar level to codon-optimised Bt toxins [25].

2.1.2.1 Proteinase inhibitors (PI)

Serine-like proteinases (trypsin-, chymotrypsin- and elastase-like endoproteases) dominant in lepidopteran larvae [26], a wider range of dominant gut proteinases in coleopteran species [27] and thiol proteases observed in corn rootworms (Diabrotica spp.) are some of the proteinases in insect which catalyse the release of amino acids from dietary protein, thereby providing nutrients essential for normal growth and development and thus the proteinase inhibitors plays an important part of the plant’s natural defence system against herbivory by inhibiting the protein metabolism. Although the mode of action of PIs are not fully understood, it may be claimed that hypersecretion of digestive enzymes caused by the presence of the inhibitors, would result in depletion of essential amino acids [28, 29].

Serine and cysteine-proteinase inhibitors have been shown to inhibit the growth and development of a variety of insects, primarily lepidopteran and coleopteran species [29, 30]. The expression of a gene encoding a sweet potato trypsin protease inhibitor (TPI) in transgenic tobacco (at a relatively low level for a plant-derived ICP – 0.2%) results in severe growth retardation of Spodoptera litura caterpillars fed on it besides the presence of high levels of the same naturally in it [31].

The first instance of a plant-derived ICP gene being used in transgenic plants was the constitutive expression (through the CaMV35S gene promoter) of a trypsin inhibitor gene taken from cowpea (Vigna unguiculata) and in tobacco [32]. Proteinase inhibitors do not just alter gut digestive enzymes; they can also affect insect water balance, moulting, and enzyme regulation [33]. The majority of research has focused on serine-proteinase inhibitors derived from the plant families Fabaceae, Solanaceae, and Poaceae, which are mostly aimed against not only lepidopteran pests but also some coleopteran and orthopteran pests. The cowpea trypsin inhibitor (CpTI), which has been introduced into at least 10 different plant species, is the most active inhibitor discovered to date the protection provided by CpTI was less pronounced and consistent than that of tobacco containing a truncated Bt-toxin gene [34].

2.1.2.2 α-Amylase inhibitors

To block carbohydrate metabolism, a-amylase inhibitors from wheat (WAAI) and common bean (Phaseolus vulgaris) (BAAI) are utilised. When introduced into transgenic tobacco, the former increased mortality of lepidopteran larvae fed on it by 30–40% [35], while the latter, when expressed in transgenic pea seeds and driven by the pha1 gene promoter, elevated resistance to bean weevils [36, 37].

2.1.2.3 Lectins

Lectins are a diverse group of carbohydrate-binding proteins, that are toxic to insects of the orders Homoptera, Coleoptera, Lepidoptera, and Diptera. The very first demonstration of enhanced resistance of transgenic plants expressing a foreign lectin used is the gene encoding the glucose/mannose-binding lectin from pea (Pisum sativum) [38]. The mode of action of lectins against insects is unknown, but it has been shown that some bind to midgut-epithelial cells [39], and some insecticidal lectins also show significant mammalian toxicity, including lectins from P. vulgaris (phaeton haemagluttinin, PHA), and the greater insecticidal activity is shown by chitin-binding lectins from wheatgerm (WGA) expression in transgenic maize [40].

Recent interest has primarily focused on the mannose specific lectin from snowdrop (GNA), which has shown activity against peach potato aphids, potato tuber moths [41] and the rice brown planthopper (Nilaparvata lugens) [42]. GN expressed in potato and tomato significantly reduced fecundity and enhanced resistance, respectively, in laboratory experiments [42]. When overexpressed in different species of tobacco, tomato, and sweet gum, the tobacco anionic peroxidase, which is involved in crosslinking and polymerisation, inhibition of digestive enzymes, and the generation of highly reactive, toxic species, led to significant levels of resistance to several lepidopteran, coleopteran, and peach potato aphid [43]. The expression of tryptamine and tryptamine-derived alkaloids in plants may serve as anti-oviposition and antifeedant agents, or as inhibitors of larval and pupal development, and when expressed in tobacco, inhibition of reproduction of the whitefly Bemisia tabaci by observed up to 97 percent [44].

2.1.3 Resistance genes from animals

Based on in vitro testing of proteolysis inhibition by several lepidopteran larvae midgut extracts, as potential insect-resistance proteins like bovine pancreatic trypsin inhibitor (BPTI), a1-antitrypsin (a1AT), and spleen inhibitor (SI) were identified and introduced into a variety of crop. Proteinase inhibitors derived from M. sexta and expressed in cotton and tobacco were reported to impede B. tabaci reproduction [45, 46]. Despite the introduction of chitinase (from the tobacco hornworm) into tobacco, these plants only exhibited a limited level of resistance to lepidopteran larvae and peach potato aphids [47].

2.1.4 Microbial proteins

The bulk screening of microbial culture supernatants against specific pests has been one strategy to the discovery of novel insecticidal proteins. Two proteins, Vip1 and Vip2, were isolated from vegetative Bacillus cereus culture supernatants and have been shown to be acutely poisonous to maize rootworms [47]. Some B. thuringiensis vegetative culture supernatants include a protein (Vip3A) that is acutely poisonous to Agrotis and Spodoptera caterpillars [48]. These proteins activity is extremely similar to that of Bt-endotoxins, yet they are distinctly separate from them.

2.1.5 Predator toxins

Genes producing neurotoxins from predatory mites [49] and scorpions [50] have been inserted into recombinant baculoviruses, where they effectively boost the rate of killing.

Advertisement

3. Disease resistance

Since the identification of the chemicals and genes involved in disease resistance in plants, attempts have been made to develop permanent disease resistance in commercially significant crop plants. Unfortunately, many of these efforts have failed because to the complexities of disease-resistance signalling and the wide range of infection routes employed by various pathogens. Although disease-resistant transgenic plants or seeds are not currently commercially accessible, future product development looks to be feasible as our understanding of pathophysiology and plant defence deepens.

In general, plants are protected by structural defence (plant cell wall, thick and waxy epidermis, trichomes, thorns) and chemical defence (production of secondary metabolites, proteins, and digestive enzymes) [51]. Plant defence response genes are classified into susceptibility genes (S), resistance genes (R), and non-host resistance genes (NHR). Pathogen-associated molecular pattern (PAMP)-triggered immunity (PTI) is broad-spectrum resistance, effective against non-host pathogens that fail to establish virulence [52, 53], and ETI is regarded to be the basis of R-gene resistance (host resistance). NHR is regarded as a quantitative trait encompassing numerous genes and varied pathways, whereas R gene resistance is regarded as a qualitative trait that adheres to Flor’s gene-for-gene model and is dependent on a complex regulatory mechanism for pathogen detection and defence response [54, 55].

The successful transfer of R genes that express specific effector proteins in a gene-for-gene manner among plant species has resulted in long-term resistance and crop protection against a variety of pests and diseases. Polymorphisms in the coiled-coil (CC) and nucleotide-binding (NB) domains have been shown to be critical for recognition specificity, hence R genes could be edited [56]. In wheat, a two-amino acid mutation in the NB domain of the R gene (PM3F) protein was known to improve the resistance spectrum and HR response. S gene mutations can give both broad-spectrum and pathogen-specific resistance. Most S genes play functional roles in the plant, and mutations in them can cause serious pleiotropic defects. For example, CRISPR–Cas9-mediated disruption of the OsSEC3A gene in rice improved defence response against Magnaporthe oryzae, but the resistant plant also had dwarf stature and a lesion-mimic phenotype [57]. Similarly, leaf chlorosis was observed in wheat plants that had been MLO gene-edited for resistance to powdery mildew [58, 59, 60]. Rice OsSWEET14 and 11 mutations resulted in shorter height and pollen abortion [61, 62].

Understanding the processes of NHR, which has a long lifespan and a broad spectrum, is essential for breeding disease-resistant cultivars. Type I NHR is asymptomatic [63, 64], whereas Type II NHR is similar to host resistance (“gene-for-gene”) and is associated with the hypersensitive response triggered by pathogen penetration into plant tissue and activation of the resistance gene [65, 66, 67, 68]. For conferring and developing disease resistance technologies are RNAi, TALENS, CRISPR/Cas, ZFN used transfer gene in various crops (Table 2) [76].

CropDiseaseGene and its functionGE Methods usedRef.
WheatBlumeria graminisTaMLO Repress immunity against powdery mildewRNAi[69]
RiceXanthomonas oryzae pv. OryzaeEBS of OsSWEET14 and OsSWEET12, Transcriptional inductionTALEN[70, 71, 72, 73]
Wheatpowdery mildewThree homoallele TaMLOof A1, B1, D1 homoeologsTALEN[59, 60, 74]
RiceRice tungro spherical virus (RTSV)eIF4G Translation of viral RNA genomesCRISPR/ Cas9[75]
RiceMagnaphorthe oryzaeOsERF922 Negative regulator of blast resistanceRNAi CRISPR/Cas9[59, 60]

Table 2.

RNAi, CRISPR/Cas 9 and TALEN mediated gene transfer for various diseases in different crops.

Advertisement

4. Herbicide resistance

Herbicides are essential components of today’s integrated weed management strategy. To address the issue of repeated use of the same herbicide resulting in the evolution of resistant weeds, complicating their control, the development of diverse herbicide-tolerant crops is critical.

4.1 Photosystem based herbicides

Herbicides of various types disrupt various organelles and parts of plants. For example, now recognised by numerous names, including QB (Quinone), D1, encoded by the psbA gene,the 32-kDa weighted protein found in the thylakoid membrane acts as a herbicide binding protein for a variety of proteins was identified using a photoaffinity marker (azidoatrazine) [73, 75]. This is the first gene found for herbicide tolerance research. Atrazine, the most commercially important PS II herbicide, is the most extensively used herbicide in maize and sorghum as these crops can extensively degrade atrazine by glutathione conjugation. Herbicide-tolerant weeds (Senecio vulgaris) discovered in maize for atrazine, simazine, and other s-traizine category herbicides, which were previously susceptible for these herbicides, demonstrated maternal inheritance and conferred 100 times less binding affinity for herbicide in thylakoid membrane [77, 78, 79]. There have been six recorded amino acid exchanges (Table 3) [80, 81, 82, 83]. The 3D arrangement of the amino acids revealed that different herbicide classes had different binding sites on the QB protein.

Amino acid position in QB proteinWild typeMutant
264SerineAlanine
264SerineGlycine
219Valineisoleucine
251AlanineValine
255Phenyl alanineTyrosine
275LeucinePhenyl alanine

Table 3.

Amino acid exchanges in wild and mutant plant proteins.

4.1.1 Glyphosate

Glyphosate N-(phosphonomethyl) glycine is a post emergence herbicide and is a potent competitive inhibitor of the enzyme 5-enol-pyruvyl shikimic acid 3-phosphate (EPSP) synthase and it is rate limiting enzyme for aromatic amino acids synthesis pathway [84]. A gene (aroA) from gram negative bacteria Salmonella typhinutrium (now the same found in Aerobacter aerogenes [85], possessing mutation on 101th position of amino acid wherein proline is changed to serine a gene (aroA) encoding resistant EPSP enzyme against glyphosate was transformed to tobacco with octopine synthase promoter through agrobacterium transformation which led to insensitive response to herbicide [86]. The mutated gene was fused to an octopine synthetase promotor to enable expression in plants leading to regeneration of transgenic plants with glyphosate tolerance and impaired growth due to compartmentation of aromatic amino acid biosynthesis and disruption of EPSP synthase and other shikimic acid pathway enzymes located in chloroplasts. The precursor protein is directed into the chloroplast by amino-terminal regions known as transit peptides, which are eliminated during the absorption process. EPSP synthase is also a precursor protein that is enzymatically active and binds glyphosate. A fusion gene encoding the sequence of a plastidic transit peptide before the aroA sequence, which exhibits a 40-fold increase in EPSP-synthase activity in petunia plants, is likely to provide a greater level of resistance [87, 88].

4.1.2 Sulfonylurea and Imidazolinone herbicides

The herbicides imidazolinones and sulfonylureas operate by limiting acetolactate synthase (ALS), the first enzyme in the biosynthetic chain that results in the production of branched chain amino acids [89, 90, 91]. Thus, chlorsulfuron is readily detoxified by wheat, barley, or oats by phenyl ring hydroxylation and consequent conjugation with glucose. Sulfonylurea-resistant mutant strains have been isolated from species as varied as bacteria, fungi, and plant cell cultures. Using cell biology techniques, maize plants resistant to imidazolinone and tobacco plants resistant to sulfonylurea herbicides have been successfully established [90, 91].

4.1.3 Phosphinothricin

L-Phosphinothricinis (an analogue of glutamate), a component of the tripeptide “bialaphos” produced by several Streptomyces viridochromogenes, is a potent irreversible inhibitor of glutamine synthetase with herbicidal activity causing a rapid increase of ammonia concentration in plants which leads to the inhibition in photosynthesis, which derives the plants to death of the plant cell [92, 93]. Resistance to L-phosphinothricin in alfalfa was established via overproduction of a glutamine synthetase gene connected to the promoter of the 35S transcript of cauliflower mosaic virus, and the construct was incorporated into the genome of N. Tobaccum var. W38 by A. tumefaciens [94, 95]. Transgenic plants demonstrated superior resistance to high doses of commercial formulations of phosphinothricin and bialophos after integrating the resistance gene from Streptomyces hygroscopicus to tobacco and regulating it with the 35S promotor of the cauliflower mosaic virus.The treated plants showed no elevation in ammonia concentration, signifying the total shielding of the plant glutamine synthetase from the herbicide’s activity [94]. Herbicide resistance is acquired via TALENs-mediated gene mutation of OsALS in rice and ALS (SurA and SurB) in tobacco [96, 97] (Table 4).

S. NoHerbicideGeneSource of geneMutation
1.GlyphosatearoASalmonella spp.P101 → S
2.SulphonylureailvGMEscherichia coliA26 → V
ILV2Saccharomyces cerevisiaeP192 → S
SURBHraNicotiana tabacumP196 → A
W573 → L
SURBc3P196 → Q
Csr1A. thalianaP197 → S
3.PhosphinothricinGSM. sativaG245 → S,C,R
R264 → K
4.TriazinespsbAAmaranthusS264 → G
S. nigrumS264 → G
ChlamydomonasS264 → A
F255 → Y
V219 → I

Table 4.

Genes from various sources responsible for mutation in herbicide tolerance.

Advertisement

5. Stress

Abiotic stress is a highly complicated phenomena that involves biochemical and physiological changes in plant cells causing increased amounts of ROS (reactive oxygen species), that are extremely reactive, toxic and impact chlorophyll production, photosynthetic capability, and carbohydrate, protein, lipid, and antioxidant enzyme activities.

Genes that code for synthesis of osmolytes (proline, betain, etc.), water uptake and transport (ion transporters and aquaporin), transcriptional regulation and signal transduction mechanisms (MAPK) are identified from a variety of organisms and transformed into sensitive genotypes for generation of stress-protecting chemicals [98, 99]. The most efficient candidates for genetic transformation are transcription factors (TFs) that regulate the expression of a number of genes involved in conferring abiotic stress tolerance in plants [100].

5.1 Drought tolerance

Stress-related ring finger protein 1 (OsSRFP1), drought-induced SINA protein 1 (OsDIS1), and dry- and salt-tolerant protein 1 (OsDST) are negative regulators of drought tolerance, whose silencing increased antioxidant enzyme levels, reduced H2O2 concentrations, and enhanced drought tolerance in rice.

The CRISPR–Cas9 system was used to introduce novel alleles in Arabidopsis OPENSTOMATA 2 (OST2)-encoding gene—a key plasma membrane H+ ATPase causing two significant mutations at the ost2 locus that led to constitutive functioning of proton pumps, induction of necrotic lesions and exhibiting a substantially higher rate of stomatal closure coupled with a lower rate of transcriptional water loss which resulted in enhanced drought tolerance (Table 5).

CropGeneStressReference
RiceAtDREB1A from Arabidopsis thalianaResistance to drought[101]
BrCIPK1 gene from Brassica rapaEnhanced abiotic stress tolerance[102]
A. thaliana transcriptional regulator DNA polymerase II subunit B3-1 (DPB3-1) geneInduced heat tolerance[103]
Rice LSD1-like type ZFP gene OsLOL5Tolerance against salt and oxidative[104]
JERF3 transcription factor (Solanum lycopersicum)Drought tolerance[105]
OsDREB2A from Oryza sativaDrought and salinity tolerance[106]
OsERF4a from O. sativaIncreased tolerance to drought stress[107]
WheatTaPIE1 from Triticum aestivumEnhanced tolerance to Cold[108]
A. thalianaSPDS (Spermidine synthase) from Cucurbita ficifoliaMultiple environmental stress[109]
Tomato Monodehydroascorbate reductase (LeMDAR) geneIncreased tolerance to protoplast abiotic stresses[110]
Wheat WRKY2 and WRKY19 genesSalt, drought and freezing stress tolerant[111]
MaizeRab28 LEA geneOsmotic stress tolerance[112]
PotatoStrawberry D-galacturonic acid reductase (GalUR) geneIncreased stress tolerance to methyl viologen, NaCl and mannitol[113]
Rat GLOaseResistant to methyl viologen, NaCl and mannitol[113]
DHAR from A. thalianaIncreased tolerance to salt and drought[114]
SoyabeanSalinity stress tolerance and fungal resistanceOverexpression of tobacco osmotin[115]
Soybean Osmotin-like protein isolated from Solanum nigrum var. americanum (SnOLP)Increased drought tolerance[116]
Chick peatolerance by expression of VignaEnhanced salinity[117]
Increased drought toleranceoverexpression of miR408[118]
Red gramExpressing Vigna pyrroline-5-carboxylate synthetaseIncreased salt tolerance[119]
Black gramOverexpression of glyoxalase 1Salt tolerance[120]
Expression of Arabidopsis alate transporter, AtALMT1Increased aluminium tolerance[121]
ALDRXV4 geneTolerance to H2O2, drought, salt and methyl viologen induced stresses[122]
Mung beanExpression of AtNHX1Increased salt tolerance[123]
Expressing AtICE1Cold stress tolerance[124]
SugarcaneEaDREB2 from ErianthusarundinaceusDrought and salinity stress tolerance[125]
TobaccoSbpAPX from Salicornia brachiataResistance against salt, cold, drought, abscisic acid and salicylic acid stress[126]
GmERF7 transcription factor from Glycine maxIncreased tolerance to salinity[127]

Table 5.

Genes responsible for abiotic stress tolerance in various crops.

Advertisement

6. Quality

Over two billion people worldwide are malnourished as a result of nutritional stress. Genetically modified (GM) crops have the potential to fulfil the worldwide need for high-quality food through genetic engineering by doing more than merely boosting nutritional quality.

6.1 Golden rice

The carotenoid biosynthesis pathway in plants is a multistep process and is accomplished via four desaturation reactions to produce all-trans-lycopene from 15-cis-phytoene by phytoene desaturase (PDS), ζ-carotene isomerase (Z-ISO), ζ-carotene desaturase (ZDS), and carotenoid isomerase (CRTISO) whereas, a single enzyme encoded by the crtI gene accomplishes all the four reaction steps to produce lycopene in bacteria [128, 129]. In the carotenoid pathway, lycopene is an important branch point because it functions as the substrate for two competing enzymes, lycopene β-cyclase (LYCB), and lycopene ε-cyclase (LYCE). Lycopene is converted into γ-carotene, which is rapidly converted into β-carotene (has pro-vitamin A activity) by LYCB in a single pathway. The bacterial gene crtY encodes LYCB to produce β-carotene. Alternatively, with LYCE lycopene generates δ-carotene (which has no pro-vitamin A activity) [130, 131, 132].

About 60–80% by weight of total seed protein in rice is glutelin, and about 20–30% are prolamins which are the prime choice of promoter sequences responsible for expression of carotenoids in rice endosperm [133]. Six promoters of rice glutelin genes (GluA-1, GluA-2, GluA-3, GluB-3, GluB-5, and GluC) were isolated and examined in rice and listed in tables with their site of expression (Table 6) [134]. Newly reported promoters like PROLAM26 RAL2 (LOC_Os07g11330), RAL4 (LOC_Os07g11380), and CAPIP (LOC_Os06g33640) could be useful in the future [135].

S. NoPromoterExpressing region
1.GluA1, GluA2 &GluA3Peripheral region of the endosperm
2.GluB5 and GluCThe whole endosperm
3.GluB3Aleurone and subaleurone layers of rice grain

Table 6.

List of promoters and expressing region involved in Golden rice.

A combination of Daffodils phytoene synthase (psy) gene, lycopene-cyclase (lcy) gene and crtI gene from bacteria (Erwinia uredovora) PDS (phytoene desaturase) are used to generate japonica rice with ß-carotene expression (mentioned here as GR1) [136, 137]. Gtu-1 promoter was used for psy, lcy and a constitutive CaMV35S promoter with a plastid-specific transit peptide (TP) used for the expression of CrtI gene [138, 139] in GR1, Under Gtu-1 promoter, an alliance of maize phytoene synthase (Zmpsy) gene with bacterial crtI has been expressed in rice to develop an improved golden rice variety (mentioned here as GR2) [140]. The synthetic gene constructions of two carotenoid biosynthetic genes, psy from Capsicum (Capsicum annuum) and crtI from Pantoea, were also reported for golden rice development. To create the PIC (Psy-IRES-CrtI) and PAC (Psy-2A-CrtI) constructions, coupling of two genes using either the synthetic codon-optimised 2A sequence (from foot-and-mouth disease virus) or the IRES sequence (the internal ribosome entry site) were utilised (Figure 1) [141, 142].

Figure 1.

A simplified version of ß-carotene expression in golden rice.

ZFN and TALENs mediated gene mutation of IPK1 gene of the rice OsBADH2 gene respectively encoded inositol1,3,4,5,6-penta-kisphosphate 2-kinase, resulted in both herbicide tolerance and reduction of phytate in developing seeds and production of fragrant rice by increased synthesis of fragrance compound 2-acetyl-1-pyrroline (2AP) [143]. In monocot plants, the k1C gene in sorghum plant was targeted using CRISPR for gene disruption at the N-terminal ER signal peptide region, resulting in greater Lysine content and improved protein digestibility [144]. Naim et al. deployed CRISPR-Cas to target the PDS, Phytoene desaturase expressing gene in Cavendish banana (Musa acuminata) for gene knockout to target the Albinism phenotype. CRISPR-Cas9 was used to disrupt rice genes critical for defining amylose concentration, fine structure of amylopectin, and physiochemical characteristics of starch, resulting in a larger proportion of long chains in amylopectin. The GBSS gene, which encodes Granule-bound starch synthase and is responsible for amylose production in the Potato plant, was targeted using CRISPR-Cas for Gene knockouts, resulting in a product with elevated amylopectin content [145].

In Camelina sativa the FAD2 gene, which is important for fatty acid production, was targeted using CRISPR-Cas9 for gene deletion to enhance seed Oleic acid content. In tomato, CRISPR-Cas9 was used to target SlAGL6, a transcription factor that plays important roles in flower meristem and floral organ development, for gene deletion, resulting in a parthenocarpic phenotype [146]. The CAO1 and LAZY1 genes, which are responsible for synthesis of chlorophyll b from chlorophyll a and regulating shoot gravitropism, were disrupted using CRISPR-Cas to target the faulty synthesis of chlorophyll b and tiller spreading phenotypes, respectively (Table 7) [147].

CropGene transformedQuality characterReference
RiceSoybean ferritin geneIncreased iron content[148]
Phaseolus vulgaris ferritin geneEnhancement in Fe content, cysteine and phytase level[149]
Lactoferrin gene from humanIncreased iron content[150]
Ubi1-P-int (maize) & GmFAD3 cDNA(soyabean)Enhanced α-linolenic acid content[151]
amino-deoxychorismate synthase (ADCS 1) & GTP cyclohydrolase I (GTPCHI) genes100 times enhanced accumulation of vitamin B9[152]
BiP, lysine-rich binding proteinIncreased lysine content[153]
MaizeSacB gene (Bacillus amyloliquefaciens)Stable accumulation of fructan[154]
Lysine feedback-insensitive DHDPSIncreased lysine content[155]
HGGT gene from barleyTocotrienol content increase[156]
Wheat DHAR geneIncreased ascorbic acid[157]
Ferritin gene & phytase geneIncreased iron content[158]
Red gramDihydrodipicolinate synthase overexpressionimprove lysine content in seeds[159]
soyabeanDHDPS and aspartokinase from E. ColiIncreased lysine content[155]
15-kDa Maize zein proteinIncreased accumulation of C &M[160]
Chick peaRaffinose synthase 2 SilencingNutritional quality improvement[161]

Table 7.

List of different genes transformed for quality character in different crops.

Advertisement

7. Male sterility

The use of genetically engineered male sterility has a variety of applications, ranging from hybrid seed production to transgenic bioconfinement in genetically modified crops. The influence of this technique has aided in dealing with global food security concerns. The production of transgenic male sterile plants through the expression of a ribonuclease gene under the direction of an anther- or pollen-specific promoter has shown to be an efficient method of producing pollen-free elite cultivars.

7.1 Male sterility due to mutation on nuclear genes

Mariani et al. [162] achieved the first success in developing genetically engineered male sterility in crop plants by transforming tobacco and rapeseed plants with a chimeric dominant gene barnase (bacterial RNase from Bacillus amyloliquefaciense) driven by a tapetum-specific promoter (TA29) from tobacco. The coding sequences of RNase T1 from Aspergillus oryzae and barnase from B. amyloliquefaciens used to manipulate the trait were fused with the tapetum-specific TA29 promoter which is responsible for the expression of the barnase gene specifically to anther tapetal cells, causing selective destruction of the tapetal cell layer that surrounds the pollen sac by hydrolysing the tapetal cells, causing abnormal pollen formation (Figure 2). Male sterile anther carries empty exine [162]. Mariani et al. [163] demonstrated fertility restoration in TA29-barnase male sterile plants by gene encoding the barnase-specific RNase inhibitor called barstar which was isolated from same bacteria B amyloliquefaciense. When genetically engineered, male sterile plant is crossed with plant carrying TA29- barstar gene the F1 progeny shows co-expression of both genes in the anther of male fertile plants. In this system fertility restoration is due to the formation of tapetal cell-specific barnase/barstar protein complexes which completely inactivate the barnase enzyme [163]. This dominant nuclear genetic male sterility system faces same drawback as GMS system. During hybrid seed production the plants in female rows segregate in the ratio of 1:1 for male sterility and male fertility [164]. To counter the drawback of nuclear genetic male sterility system problem the barnase gene was linked to a dominant herbicide resistant gene (bar) under control of the constitutive promoter CaMV 35S which conferred resistance to broad-spectrum herbicide Basta (active ingredient is phosphinothricin or PPT). When seedlings are sprayed with Basta only the male sterile plants survive and the male fertile plants are killed as they lack bar gene.The use of bar gene allows elimination of male fertile segregants from female rows in the hybrid seed production plot thus assuring 100 per cent pure hybrid seed production [165].

Figure 2.

Simplified version of expression of transgenic male sterile system.

7.2 Male sterility due to mutation on chloroplast genes

The genetic transformation of the plastid genome has various advantages, including high level transgenic expression, expression of multigene operons, transgene maternal inheritance, and expression of bacterial genes without codon optimization [166]. Ruiz and Daniell [167] elucidated that, with chloroplast transformation, hyper-expression of β-ketothiolase encoded by the phaA gene of Acinetobacter sp. in the leaves, flower, and anther of transgenic lines gets in the way of pollen development and results in male sterility. This was restored by exposing transgenic male sterile plants to continuous illumination, which allows acetyl CoA carboxylase (ACCase) to access acetyl CoA, restoring normal fatty acid synthesis and minimising PHB production through β-ketothiola.

7.3 Male sterility due to altering metabolic process

Callose is a plant polysaccharide comprised of β-1-3 glucan that is deposited around microspore tetrads during meiosis. Tight developmental regulation and the timing of callase activity are required for optimal microspore development. The expression of modified PR-b-1-3 glucanase in transgenic tobacco plants led in the premature disintegration of the microsporocyte callose wall, resulting in mild to total male sterility [168]. Chang et al. [169] created a rice hybrid breeding method employing the rice nuclear gene Oryza sativa No Pollen 1 (OsNP1), which encodes a putative glucose–methanol–choline–oxidoreductase with involvement in tapetum degeneration and pollen exine production. The ethyl methane sulfonate-induced rice mutant, osnp1-1, was completely male sterile.

7.4 Conditional male sterility

Conditional male sterility is a situation in which plants are typically fertile, but when a specific circumstance is applied, male sterility occurs. Hawkes et al. [170] revealed the use of inactive D-glufosinate as a male sterility inducer in transgenic plants expressing a modified (F58 K, M213S) version of Rhodosporidium toruloides Damino acid oxidase (DAAO) that converts oxidised D-glufosinate to its 2-oxoderivative (2-oxo-4-methyl phosphiny to create transgenic plants, the modified DAAO encoding gene was coupled with the TAP1 promoter from Antirrhinum majus and transformed into tobacco plants. When D-glufosinate was sprayed on these transgenic plants, it caused full male sterility that lasted two or more weeks while having no effect on female fertility [170]. Guerineau et al. [171] expressed the temperature-sensitive diphtheria toxin A-chain polypeptide gene sequence under the tapetum-specific A9 promoter and generated transgenic Arabidopsis plants that were fully fertile at 26 C, but when the temperature was decreased to 18°C, male sterility was induced [171].

7.5 Male sterility through post transcriptional gene silencing

Jasmonic acid (JA), a plant hormone, is involved in several developmental signalling events in plants, including senescence, fruit ripening, anther dehiscence, and pollen maturation [172, 173]. Bae et al. [174] reported inducing male sterility by inhibiting OsAOS1 and OsAOS2 activity with the promoters of the anther-specific genes Osc4 and Osg6b, respectively. RNAi (pSK124) constructs were designed and converted into rice calli independently, concluding that the OsAOS2-RNAi vector driven by Osg6b promoter is potent enough for generating male sterility in rice.

7.6 Male sterility through modification of flavonoids

Any disruption in flavonoid production changes pigmentation and causes male sterility in plants. Fischer et al. [175] discovered the expression of a stilbene synthase (STS) gene from grape vine (VstI) driven by a 35S RNA promoter with duplicated enhancer region and a tapetum-specific promoter (Tap1) of A. majus produced male sterility strives for the substrates,4-coumaroyl CoA and malonyl CoA, which are required for sporopollenin and fatty acid biosynthesis, and hypothesised that there was a decrease in p-coumaroyl availability, resulting in impaired sporopollenin production and pollen wall formation, causing male sterility(Figure 3).

Figure 3.

Simplified version of male sterility through modification of flavonoids.

7.7 Male sterility through RNA editing

Nucleotide alterations or insertion of a nucleotide, leading to a change in the sequence of amino acid in polypeptide denotes RNA editing. Male sterility in CMS plants is connected to mitochondrial DNA rearrangement, causing the formation of novel chimeric open reading frames (ORFs), resulting in mitochondrial malfunction, such as the chimeric gene pcf-S of petunia, ORFB and ORF224 of polima in rapeseed [176]. The overexpression of unedited mitochondrial orfB gene in a transgenic strain of indica rice led to a decrease in activity of ATPase in F1F0-ATP synthase resulting in dose-dependent male sterility [177].

7.8 Heterologous male sterility

The association of CMS and new chimeric ORFs in mitochondrial DNA sequences, as well as mitochondrial dysfunction, is documented. Nizampatnam et al. [178] engineered transgenic tobacco plants to produce orfH522, a pet1-CMS-associated mitochondrial gene from sunflower that is driven by the TA29 promoter. Approximately 35% of the modified tobacco plants were completely sterile. Subsequently, by decreasing orfH522 transcripts using the RNAi approach, male fertility was restored [179].

Advertisement

8. GE for yield contributing characters

To address growing food demand as well as the challenges posed by climate change, major increases in yields of vital food crops employing transgenic technology are required. Using CRISPR-Cas9, researchers were able to increase grain number, dense erect panicles, and grain size in rice by disrupting the DEP1, Gn1a, IPA1, and GS3 genes, which are regulators of grain number, panicle architecture, grain size, and plant architecture [180]. CRISPR-Cas9 gene deletion targets the wheat genes TaGW2-B1, TaGW2-D1, and TaGW2-A1 that govern grain weight and protein content, leading to an increase in grain weight and protein content [181]. In maize, the gene ARGOS8 responded to water stress by increasing grain output [182].

Advertisement

9. Gene drive

Gene drive is a genetic phenomenon of naturally occuring skewed inheritance mechanism that implies sexual reproduction and is essentially concerned with population suppression and population modification. Synthetic gene drives deploy genome editing technologies such as CRISPR to maximise the likelihood of a certain gene being inherited from 50% to almost 100% (Figure 4).

Figure 4.

Simplified version of comparison between normal mosquito population and genetic drive mosquitos’ population. There is a huge discrepancy among the rapid adoption of GM crops for production, global markets, and consumer approval. However, the following is a list of transgenic crops that have been worldwide authorised and released for various characteristics (adopted from ISAAA database).

9.1 Working model of gene drive

When a gene drive is inserted into the genome of an animal, the progeny inherits the drive on one chromosome and a normal gene from the other parent. During early development, the CRISPR component of the drive shears the other copy. The cut is subsequently repaired using the drive as a template, resulting in two copies of the change being passed down to the progeny [183]. By suppressing the fertility gene termed doublesex on the usage of drive and thereby crashing a population of caged Anopheles gambiae mosquitos, Crisanti and his colleagues were able to prevent female mosquitos from biting or laying eggs while the drive was in place [184].

In C. capitata [185] and Aedes aegypti, a bisex RIDL system containing a tetracycline-repressible positive feedback transactivator (tTA) was successfully constructed, which does not require a specific promoter derived from the target species, but rather a minimal promoter used in conjunction with oligomerised tetO, the binding sequence. Under tight conditions, particularly in the absence of tetracycline, tTA (transactivator and lethal effector) accumulates to deadly levels in both sexes of the transgenic insect. Females must be removed from a ‘male-only’ release programme using an independent approach based on an underlying molecular mechanism that is female-specific. Functional Tra protein is only produced by females as it is encoded by a splice variant exclusively produced in females leading to functional tTA protein only produced in females rendering the system female-specific. Fu et al. [185] integrated the first intron of the sex determining gene Cctra into the DNA sequence coding for tTA to provide a sex selective component to the positive feedback system. If this Cctra fragment is spliced in the same as it is in its native context, tTA production is only allowed when the intron is spliced in the female-specific form, as the continuous coding frame of tTA is only restored in this variant. As this intron’s full splicing in its native gene is exclusive to females, tTA expression was only expected in transgenic females. The analysis of tTA transcription in transgenic C. capitata revealed a sex-specific pattern similar to the natural Cctra gene. The inserted intron spliced to produce three distinct tTA transcripts: one female-specific (F1) and two nonsex-specific (M1 and M2). The female-specific transcript was the only one that encoded a complete tTA. As a result, when grown under harsh conditions, all of the transgenic female progeny died as larvae or pupae [186]. Thus, the utilisation of gene drivers in agriculture may be primarily harnessed by implementing sterile insect techniques for successful insect pest management (Table 8).

CropsIRHRDRAbiotic stressQualityPollination controlAltered growth / yield
Maize1191286862
Cotton49441
Cowpea1
Bean1
Brinjal1
Popular2
Potato3041918
Rice331
rose2
Soyabean632291
Sugar cane33
Tomato119
Papaya4
Plum1
Petunia2
Squash2
Cucumis melo2
Sweet pepper1
Alfalfa42
apple3
Canola331020
Carnation419
Chicory33
Creeping Bentgrass1
Flax1
Brassica4
Sugar Beet3
Pine apple1
Tobacco11
Wheat1
Safflower2
Eucalyptus1

Table 8.

List of globally approved and released transgenic crops for various characters.

Advertisement

10. Discussion and conclusion

The advent of advanced targeted, ‘customizable,’ and precise new technologies for insect resistant plants, in their various embodiments and combinations, symbolises a sustainable option countering the emergence of resistant weeds, lessening agrochemical use, and curtailing adverse effects on nontarget organisms. The direct application of chimeric ODNs or siRNAs to plant cells enables for the generation of technically non-GM organisms. Inevitably, the refinement of promising techniques that are not currently destined to assign insect resistance, such as RNA manipulation with pentatricopeptide repeat proteins, the use of polygalacturonase-inhibiting proteins (PGIPs), ribozymes, and riboswitches, will likely expand the defence mechanism against pests available to researchers and farmers. However, thanks to recent advancements like RenSeq and directed molecular evolution, which enable the rapid identification of novel immune receptor genes, the pool of deployable genes for enhanced resistance to other microorganisms has grown significantly. In the near future, developments in molecular stacking and targeted gene insertion by genome editing are projected to predominate in establishing broad-spectrum resistance against both viral and nonviral diseases. Moreover, increasingly diverse, accurate, and economical genome-editing techniques like CRISPR-Cas allow for precise change of endogenous genes for disease resistance, such as susceptibility and decoy genes. On the flip side, the persistent cultivation of herbicide crops is estimated to have resulted in the resurgence of herbicide resistance in numerous weeds. Nevertheless, by benefiting from the shortcomings of the issue, the adverse consequences of herbicide-based technologies might be substantially minimised by introducing variety in weed control using alternative approaches, with an emphasis on crop rotation, herbicide rotation, and herbicide formulations. Stress-resistant plants may be generated with or without tissue culture by using simple knock-in, knock-out, replacement, fine-tuning of gene regulation, and point mutations at any gene locus. Although genome editing technology is in in its beginning phases, disruptions in specific genes can have unintended negative consequences for plant growth and development. More research is needed to fully exploit the ability of the CRISPR-Cas System in regulating abiotic stress. One rationale could be that the genes governing beneficial qualities in crops are largely quantitative trait genes, which necessitates a deeper knowledge of gene regulatory networks. As a result, expanding our knowledge of gene regulatory networks is the foundation for greater crop development. Furthermore, the approach of knocking out target genes via NHEJ has attracted a strong interest for boosting agricultural output and quality. Gene targeted insertion or substitution enable the genes to integrate more easily. The HR repair pathway is ineffectual, constraining the use of site-specific insertion and substitution. It is also expected that eliminating the HR approach to editing would lead to more precise and effective crop improvement. If the seed is the most valuable portion of the plant and the crop is mostly self-fertilised, a good fertility restoration mechanism is necessary. It is likely that by coupling inducible male sterility with apomixis, fertile plants may be obtained once the trait was fixed. It is also necessary to guarantee that apomixis is inducible, encouraging apomixis to restore to sexual reproduction and enabling the breeder the opportunity of further enhancing the hybrids by establishing appropriate combiner lines.

The world has already experienced two technical revolutions and is currently undergoing a third revolution based on biotechnology and genomics, which is predicted to yield a plethora of transgenic crops for the benefit of humanity. Genetic engineering is a radical departure from traditional breeding since it allows scientists to transfer genetic material across organisms that could not be bred earlier. The degree of public knowledge regarding the benefits and drawbacks of transgenic plants should be strengthened, laying the groundwork for the effective dissemination of research findings to real time application. The central emphasis should be on the advancement of technologies competent of bridging the gaps in modern day technology. Nonetheless, developing countries are now required to assess genetically modified (GM) crops, and they will subsequently be expected to investigate the potential use of GM trees, cattle, and fish. These advancements may provide prospects for greater output, productivity, product quality, and adaptive fitness, but they will almost likely pose challenges to developing countries’ research and regulatory capability.

References

  1. 1. Schnepf HE, Whiteley HR. Cloning and expression of bacillus thuringiensis crystal protein gene in Escherichia coli. Proceedings of the National Academy of Sciences of the United States of America. 1981;78:2893-s2897
  2. 2. Barton K, Whitely H, Yang NS. Bacillus thuringiensis d- endotoxin in transgenic Nicotiana tabacum provides resistance to lepidopteran insects. Plant Physiology. 1987;85:1103-1109
  3. 3. Fischho! DA, Bowdish KS, Perlak FJ, Marrone PG, McCormick SM, Niedermeyer JG, et al. Insect tolerant transgenic tomato plants. Bio/Technology. 1987;5:807-812
  4. 4. Vaeck M, Reynaerts A, Hofte H, Jansens S, DeBeuckleer M, Dean C, et al. Transgenic plants protected from insect attack. Nature. 1987;327:33-37
  5. 5. Peferoen M. Progress and prospects for field use of Bt genes in crops. Trends in Biotechnology. 1997;15(5):173-177
  6. 6. Crook NE, Jarrett P. Viral and bacterial pathogens of insects. Journal of Applied Bacteriology. 1991;70:91S-96S
  7. 7. Zhang Q , Hua G, Adang MJ. Effects and mechanisms of bacillus thuringiensis crystal toxins for mosquito larvae. Insect Science. 2017;24(5):714-729
  8. 8. Hodgman TC, Ellar DJ. Models for the structure and function of the bacillus thuringiensis d-endotoxins determined by compilational analysis. Journal of DNA Sequencing Mapping. 1990;1:97-106
  9. 9. Li J, Carroll J, Ellar DJ. Crystal structure of insecticidal d-endotoxin from bacillus thuringiensis at 2.5A_ resolution. Nature. 1991;353:815-821
  10. 10. Knight PJ, Crickmore N, Ellar DJ. The receptor for bacillus thuringiensis Cry1A(c) d-endotoxin in the brush border membrane of the lepidopteran Manduca sexta is aminopeptidase-N. Molecular Microbiology. 1994;11:429-436
  11. 11. Pease EA, English LH, Donovan WP. Characterization of the binding of CryIIIB2 to coleopteran brush border proteins. The FASEB Journal. 1995;9:A1313
  12. 12. Luo K, Lu Y, Adang MJ. A 106Kd form of aminopeptidase is a receptor for bacillus thuringiensis cry1C d-endotoxin in the brush border membrane of Manduca sexta. Insect Biochemistry and Molecular Biology. 1996;26:783-791
  13. 13. Yaoi K, Kadotani T, Kuwana H, Shinkawa A, Takahashi T, Iwahama H, et al. Aminopeptidase N of Bombyx mori as a candidate for the receptor of bacillus thuringiensis Cry1Aa toxin. European Journal of Biochemistry. 1997;246:652-657
  14. 14. Gill SS, Cowles EA, Pietrantonio PV. The mode of action of bacillus thuringiensis endotoxins. Annual Review of Entomology. 1992;37(1):615-634
  15. 15. Perlak FJ, Fuchs RL, Dean DA, McPherson SL, Fischhoff DA. Modification of the coding sequence enhances plant expression of insect control protein genes. Proceedings of the National Academy of Sciences of the United States of America. 1991;88:3324-3328
  16. 16. Wong EY, Hironaka CM, Fischho! DA. Arabidopsis thaliana small subunit leader and transit peptide enhance expression of bacillus thuringiensis proteins in transgenic plants. Plant Molecular Biology. 1992;20:81-93
  17. 17. Hilder VA, Boulter D. Genetic engineering of crop plants for insect resistance—A critical review. Crop Protection. 1999;18(3):177-191
  18. 18. Crickmore N, Baum J, Bravo A, Lereclus D, Narva K, Sampson K, Schnepf E, Sun M, Zeigler DR. Bacillus thuringiensis Toxin Nomenclature. 2018. Available from: http://www.btnomenclature.info/
  19. 19. Armstrong CL, Parker GB, Pershing JC, Brown SM, Sanders PR, Duncan DR, et al. Field evaluation of European cornborer control in progeny of 173 transgenic corn events expressing an insecticidal protein from bacillus thuringiensis. Crop Science. 1995;35:550-557
  20. 20. Perlak FJ, Stone TB, Muskopf YN, Petersen LJ, Parker GB, McPherson SA, et al. Genetically improved potatoes: Protection from damage by Colorado potato beetles. Plant Molecular Biology. 1993;22:313-321
  21. 21. Smigocki A et al. . In: Carozzi N, Koziel M, editors. Advances in Insect Control: The Role of Transgenic Plants. Taylor & Francis; 1997. pp. 225-236
  22. 22. Purcell JP, Greenplate JT, Jennings MG, Ryerse JS, Pershing JC, Sims SR, et al. Cholesterol oxidase: A potent insecticidal protein active against boll weevil larvae. Biochemical and Biophysical Research Communications. 1993;196(3):1406-1413
  23. 23. Cho HJ, Choi KP, Yamashita M, Morikawa H, Murooka Y. Introduction and expression of the Streptomyces cholesterol oxidase gene (ChoA), a potent insecticidal protein active against boll weevil larvae, into tobacco cells. Applied Microbiology and Biotechnology. 1995;44(1):133-138
  24. 24. Rigev A, Keller M, Strizhov M, Sneh B, Prudovsky E, Chet I, et al. Synergistic activity of a bacillus thuringinensis d-endotoxin and bacterial endochitinase against Spodoptera littoralis larvae. Applied Environmental Microbiology. 1996;62:3581-3586
  25. 25. Gatehouse AM, Davison GM, Newell CA, Merryweather A, Hamilton W, Burgess EP, et al. Transgenic potato plants with enhanced resistance to the tomato moth, Lacanobia oleracea: Growth room trials. Molecular Breeding. 1997;3(1):49-63
  26. 26. Christeller JT, Laing WA, Markwick NP, Burgess EPJ. Midgut protease activities in 12 phytophagous lepidopteran larvae: Dietary and protease inhibitor interactions. Insect Biochemistry and Molecular Biology. 1992;22(7):735-746
  27. 27. Wolfson JL, Murdock LL. Diversity in digestive proteinase activity among insects. Journal of Chemical Ecology. 1990;16(4):1089-1102
  28. 28. Gatehouse AMR, Hilder VA, Powell KS, Wang M, Davison GM, Gatehouse LN, et al. Insect-resistant transgenic plants: Choosing the gene to do the ‘job’. Biochemical Society Transactions. 1992;22(4):944-949
  29. 29. Broadway RM, Duffey SS. Plant proteinase inhibitors: Mechanism of action and effect on the growth and digestive physiology of larval Heliothis zea and Spodoptera exiqua. Journal of Insect Physiology. 1986;32(10):827-833
  30. 30. Orr GL, Strickland JA, Walsh TA. Inhibition of Diabrotica larval growth by a multicystatin from potato tubers. Journal of Insect Physiology. 1994;40(10):893-900
  31. 31. Yeh KW, Lin MI, Tuan SJ, Chen YM, Lin CJ, Kao SS. Sweet potato (Ipomoea batatas) trypsin inhibitors expressed in transgenic tobacco plants confer resistance against Spodoptera litura. Plant Cell Reports. 1997;16(10):696-699
  32. 32. Hilder VA, Gatehouse AMR, Sheerman SE, Barker RF, Boulter D. A novel mechanism of insect resistance engineered into tobacco. Nature. 1987;330:160-163
  33. 33. Boulter D. Insect pest control by copying nature using genetically engineered crops. Phytochemistry. 1993;34(6):1453-1466
  34. 34. Hoffmann MP, Zalom FG, Wilson LT, Smilanick JM, Malyj LD, Kiser J, et al. Field evaluation of transgenic tobacco containing genes encoding bacillus thuringiensis∂-endotoxin or cowpea trypsin inhibitor: Efficacy against Helicoverpa zea (Lepidoptera: Noctuidae). Journal of Economic Entomology. 1992;85(6):2516-2522
  35. 35. Carbonero P, Royo J, Diaz I, Garcia-Maroto F, GonzalezHidalgo E, Gutierez C, et al. Cereal inhibitors of insect hydrolases (a-amylases and trypsin): Genetic control, transgenic expression and insect tests. In: Workshop on engineering plants against pests and pathogens (Vol. 1). Madrid: Instituto Juan March de Estudios Investigaciones; 1993;10:71
  36. 36. Shade RE, Schroeder HE, Pueyo JJ, Tabe LM, Murdock LL, Higgins TJV, et al. Transgenic pea seeds expressing the a-amylase inhibitor of the common bean are resistant to bruchid beetles. Bio/Technology. 1994;12:793-796
  37. 37. Schroeder HE, Gollasch S, Moore A, Tabe LM, Craig S, Hardie DC, et al. Bean alpha-amylase inhibitor confers resistance to pea weevil (Bruchus pisorum) in transgenic peas (Pisum sativum L.). Plant Physiology. 1995;107:1233-1239
  38. 38. Boulter D, Edwards GA, Gatehouse AMR, Gatehouse JA, Hilder VA. Additive protective e!ects of incorporating two di!erent higher plant derived insect resistance genes in transgenic tobacco plants. Crop Protection. 1990;9:351-354
  39. 39. Gatehouse AMR, Hilder VA. Genetic manipulation of crops for insect resistance. In: Molecular Biology in Crop Protection. Dordrecht: Springer; 1994. pp. 177-201
  40. 40. Maddock SE. Expression in maize plants of wheatgerm agglutinin, a novel source of insect resistance. In: Third International Congress of Plant Molecular Biology. USA: Tucson, Arizona; 1991
  41. 41. Down RE, Gatehouse AM, Hamilton WD, Gatehouse JA. Snowdrop lectin inhibits development and decreases fecundity of the glasshouse potato aphid (Aulacorthum solani) when administered in vitro and via transgenic plants both in laboratory and glasshouse trials. Journal of Insect Physiology. 1996;42(11-12):1035-1045
  42. 42. Powell KS, Gatehouse AMR, Hilder VA, Van Damme EJM, Peumans WJ, Boonjawat J, et al. Different antimetabolic effects of related lectins towards nymphal stages of Nilaparvata lugens. Entomologia Experimentalis et Applicata. 1995;75(1):61-65
  43. 43. Dowd PF, Lagrimini LM. In: Carozzi N, Koziel M, editors. Advances in Insect Control: The Role of Transgenic Plants. Taylor & Francis; 1997. pp. 195-223
  44. 44. Thomas JC, Adams DG, Nessler CL, Brown JK, Bohnert HJ. Tryptophan decarboxylase, tryptamine, and reproduction of the whitefly. Plant Physiology. 1995;109(2):717-720
  45. 45. Thomas JC, Adams DG, Keppenne VD, Wasmann CC, Brown JK, Kanost MR, et al. Protease inhibitors of Manduca sexta expressed in transgenic cotton. Plant Cell Reports. 1995;14(12):758-762
  46. 46. Thomas JC, Adams DG, Keppenne VD, Wasmann CC, Brown JK, Kanost MR, et al. Manduca sexta encoded protease inhibitors expressed in Nicotiana tabacum provide protection against insects. Plant Physiology and Biochemistry-Paris. 1995;33:611-611
  47. 47. Estruch JJ, Carozzi NB, Desai N, Duck NB, Warren GW, Koziel MG. Transgenic plants: An emerging approach to pest control. Nature Biotechnology. 1997;15(2):137-141
  48. 48. Estruch JJ, Warren GW, Mullins MA, Nye GJ, Craig JA, Koziel MG. Vip3A, a novel bacillus thuringiensis vegetative insecticidal protein with a wide spectrum of activities against lepidopteran insects. Proceedings of the National Academy of Sciences of the United States of America. 1996;93:5389-5394
  49. 49. Tomalski MD, Miller LK. Insect paralysis by baculovirus mediated expression of a mite neurotoxin gene. Nature. 1991;352:82-85
  50. 50. Stewart LMD, Hirst M, Ferber ML, Merryweather AT, Cayley PJ, Possee RD. Construction of an improved baculovirus insecticide containing an insect specifc toxin gene. Nature. 1991;352:85-88
  51. 51. Freeman BC, Beattie GA. An Overview of Plant Defenses against Pathogens and Herbivores. The Plant Health Instructor. 2008. DOI: 10.1094. PHI-I-2008-0226-01
  52. 52. Boller T, Felix G. A renaissance of elicitors: Perception of microbe-associated molecular patterns and danger signals by pattern-recognition receptors. Annual Review of Plant Biology. 2009;60:379-407
  53. 53. Schulze-Lefert P, Panstruga R. molecular evolutionary concept connecting nonhost resistance, pathogen host range, and pathogen speciation. Trends in Plant Science. 2011;16(3):117-125
  54. 54. Flor HH. Current status of the gene-for-gene concept. Annual Review of Phytopathology. 1971
  55. 55. Jones J, Dangl JL. The plant immune system. Nature. 2006;444(7117):323-329
  56. 56. Ashikawa I. Regions outside the leucine-rich repeat domain determine the distinct resistance specificities of the rice blast resistance genes Pik and Pik-m. Molecular Breeding. 2012;30(3):1531-1535
  57. 57. Ma J, Chen J, Wang M, Ren Y, Wang S, Lei C, et al. Disruption of OsSEC3A increases the content of salicylic acid and induces plant defense responses in rice. Journal of Experimental Botany. 2018;69(5):1051-1064
  58. 58. Acevedo-Garcia J, Spencer D, Thieron H, Reinstadler A, Hammond Kosack K, Phillips AL, et al. Mlo-based powdery mildew resistance in hexaploid bread wheat generated by a nontransgenic TILLING approach. Plant Biotechnology Journal. 2017;15(3):367-378
  59. 59. Wang Q , Liu Y, He J, Zheng X, Hu J, Liu Y, et al. durable resistance to rice stripe virus. Nature Communications. 2014;5:1-8
  60. 60. Wang Y, Cheng X, Shan Q , Zhang Y, Liu J, Gao C, et al. Simultaneous editing of three homoeoalleles in hexaploid bread wheat confers heritable resistance to powdery mildew. Nature Biotechnology. 2014;32(9):947-951
  61. 61. Antony G, Zhou J, Huang S, Li T, Liu B, White F, et al. Rice xa13 recessive resistance to bacterial blight is defeated by induction of the disease susceptibility gene Os-11N3. Plant Cell. 2010;22(11):3864-3876
  62. 62. Chu Z, Yuan M, Yao J, Ge X, Yuan B, Xu C, et al. Promoter mutations of an essential gene for pollen development result in disease resistance in rice. Genes & Development. 2006;20(10):1250-1255
  63. 63. Collins NC, Thordal-Christensen H, Lipka V, Bau S, Kombrink E, Qiu JL, et al. SNARE-protein-mediated disease resistance at the plant cell wall. Nature. 2003;425(6961):973-977
  64. 64. Lu M, Tang X, Zhou JM. Arabidopsis NHO1 is required for general resistance against pseudomonas bacteria. Plant Cell. 2001;13(2):437-447
  65. 65. Mysore KS, Ryu CM. Nonhost resistance: How much do we know? Trends in Plant Science. 2004;9(2):97-104
  66. 66. Oh SK, Lee S, Chung E, Park JM, Yu SH, Ryu CM, et al. Insight into types I and II nonhost resistance using expression patterns of defense-related genes in tobacco. Planta. 2006
  67. 67. Kamoun S, Van West P, Vleeshouwers VGAA, De Groot KE, Govers F. Resistance of Nicotiana benthamiana to Phytophthora infestans is mediated by the recognition of the elicitor protein INF1. Plant Cell. 1998;10(9):1413-1425
  68. 68. Lindgren PB, Peet RC, Panopoulos NJ. Gene cluster of pseudomonas syringae pv. “Phaseolicola” controls pathogenicity of bean plants and hypersensitivity on nonhost plants. Journal of Bacteriology. 1986;168(2):512-522
  69. 69. Várallyay É, Giczey G, Burgyán J. Virus-induced gene silencing of Mlo genes induces powdery mildew resistance in Triticum aestivum. Archives of Virology. 2012;157(7):1345-1350
  70. 70. Li T, Liu B, Spalding MH, Weeks DP, Yang B. High-efficiency TALEN-based gene editing produces disease-resistant rice. Nature Biotechnology. 2012;30(5):390-392
  71. 71. Li T, Huang S, Zhou J, Yang B. Designer TAL effectors induce disease susceptibility and resistance to Xanthomonas oryzae pv. Oryzae in rice. Molecular Plant. 2013;6(3):781-789
  72. 72. Macovei A, Sevilla NR, Cantos C, Jonson GB, Slamet-Loedin I, Čermák T, et al. Novel alleles of rice eIF4G generated by CRISPR/Cas9-targeted mutagenesis confer resistance to Rice tungro spherical virus. Plant Biotechnology Journal. 2018;16(11):1918-1927
  73. 73. Trebst A, Draber W. Structure activity correlation of recent herbicides in photosynthetic reactions. In: Geissbühler H, editor. Advances in Pesticide Science. Vol. 2. New York: Pergamon Press; 1978. p. 223
  74. 74. Shan Q , Wang Y, Li J, Zhang Y, Chen K, Liang Z, et al. Targeted genome modification of crop plants using a CRISPR-Cas system. Nature Biotechnology. 2013
  75. 75. Pfister K, Steinback KE, Gardner G, Arntzen CJ. Photoaffinity labeling of an herbicide receptor protein in chloroplast membranes. Proceedings. National Academy of Sciences. United States of America. 1981;78:981
  76. 76. Li JF, Norville JE, Aach J, McCormack M, Zhang D, Bush J, et al. Multiplex and homologous recombinationmediated genome editing in Arabidopsis and Nicotiana benthamiana using guide RNA and Cas9. Nature Biotechnology. 2013;31(8):688-691
  77. 77. Ryan GF. Resistance of common groundsel to simazine and atrazine. Weed Science. 1970;18:614
  78. 78. Arntzen CJ, Ditto CL, Brewer PE. Chloroplast membrane alterations in triazine-resistant Amaranthus retroflexus biotypes. Proceedings of the National Academy of Sciences of the United States of America. 1979;76:278
  79. 79. Radesorich SR, Steinback KE, Arntzen CJ. Effect of photosystem II inhibitors on thylacoid membranes of two common groundsel (Senico vulgaris) biotypes. Weed Science. 1979;27:216
  80. 80. Hirschberg J, McIntosh L. Molecular basis of herbicide resistance in Amaranthus hybridus. Science. 1983;222:1346
  81. 81. Golden S, Haselkorn R. Mutation to herbicide resistance maps within the psbA gene of Anacystis nidulans R2. Science. 1985;229:1104
  82. 82. Erickson JM, Rahire M, Bennoun P, Delepelaire P, Diner B, Rochaix JD. Herbicide resistance in Chlamydomonas reinhardii results from a mutation in the chloroplast gene for the 32- Kilodalton protein of photosystem II. Proceedings. National Academy of Sciences. United States of America. 1984;81:3617
  83. 83. Erickson JM, Rochaix JD. In: Galan GA, editor. Abstracts 1st International Congress of Plant Molecular Biology. Vol. 54. Athens, GA: University of Georgia Center for the Education for the International Society for Plant Molecular Biology; 1985
  84. 84. Jaworski EG. Mode of action of N-phosphonomethylglycine: Inhibition of aromatic amino acid biosynthesis. Journal of Agricultural and Food Chemistry. 1972;20:1195
  85. 85. Schulz A, Sost D, Amrhein N. Intensitivity of 5-enol-pyruvylshikimic acid 3-phosphate synthase to glyphosate confers resistance to this herbicide in a strain of Aerobacter aerogenes. Archives of Microbiology. 1984;137:121
  86. 86. Steinrücken HC, Amrhein N. The herbicide glyphosate is a potent inhibitor of 5-enol-pyruvylshikimic acid 3-phosphate-synthase. Biochemical and Biophysical Research Communications. 1980;94:1207
  87. 87. Shah DM, Horsch RB, Klee HJ, Kishore GM, Winter JA, Tuner NE, et al. Engineering herbicide tolerance in transgenic plants. Science. 1986;233:478
  88. 88. Comai L, Facciotti D, Hiatt WR, Thompson G, Rose RE, Stalker PM. Expression in plants of a mutant aroA gene from salmonella typhimurium confers tolerance to glyphosate. Nature. 1985;317:741
  89. 89. Mukitch MJ, Shaner DL, Stidham MA. Imidazolinones and acetohydroxyacid synthase from high plants. Plant Physiology. 1987;83:451
  90. 90. La Rossa RA, Schloss IV. The sulfonylurea herbicide sulfometuron methyl is an extremely potent and selective inhibitor of acetolactate synthase in salmonella typhymurium. The Journal of Biological Chemistry. 1984;259:8753
  91. 91. Ray TB. Site of action of Chlorsulfuron-inhibition of valine and isoleucine biosynthesis in plants. Plant Physiology. 1981;75:827
  92. 92. Leason M, Cunliffe D, Parkin D, Lea PJ, Miflin BJ. Inhibition of piston sativum leaf glutamine synthetase EC-6.3.1.2 by methionine sulfoximine phosphinotricin and other glutamate analogues. Phytochemistry. 1982;21:855
  93. 93. Wild A, Manderscheid R. The effect of phosphinothricin on the assimilation of ammonica in plants. Zeitschrift für Naturforschung. 39c;500:1984
  94. 94. Eckes P. Isolation of L-phosphinothricin resistant plants, Abstr. NATO advanced study institute. In: Plant Molecular Biology. Copenhagen, June: Carlsberg Laboratory; 1987
  95. 95. Tischer E, Das Sarma S, Goodman HM. Nucleotide sequence of an alfalfa glutamine synthetase gene. Molecular & General Genetics. 1986;203:221
  96. 96. Zhang Y, Zhang F, Li X, Baller JA, Qi Y, Starker CG, et al. Transcription activator-like effector nucleases enable efficient plant genome engineering. Plant Physiology. 2013;161(1):20-27
  97. 97. Li T, Liu B, Chen CY, Yang B. TALEN-mediated homologous recombination produces site-directed DNA base change and herbicide-resistant rice. Journal of Genetics and Genomics. 2016;43(5):297-305
  98. 98. Bhatnagar-Mathur P, Vadez V, Sharma KK. Transgenic approaches for abiotic stress tolerance in plants: Retrospect and prospects. Plant Cell Reports. 2008;27:411-424
  99. 99. Tsai-Hung H, Jent-turn L, Yee-yung C, Ming-Tsair C. Heterology expression of the Arabidopsis C-repeat/dehydration response element binding factor 1 gene confers elevated tolerance to chilling and oxidative stresses in transgenic tomato. Plant Physiology. 2002;130:618-626
  100. 100. Pasquali G, Biricolti S, Locatelli F, Baldoni E, Mattana M. Osmyb4 expression improves adaptive responses to drought and cold stress in transgenic apples. Plant Cell Reports. 2008;27:1677-1686
  101. 101. Ravikumar G, Manimaran P, Voleti SR, Subrahmanyam D, Sundaram RM, Bansal KC, et al. Stress-inducible expression of AtDREB1A transcription factor greatly improves drought stress tolerance in transgenic indica rice. Transgenic Research. 2014;23:421-439
  102. 102. Abdula SE, Lee HJ, Ryu H, Kang KK, Nou I, Sorrells ME, et al. Enhances abiotic stress tolerance by increasing proline biosynthesis in rice. Plant Molecular Biology Reporter. 2016;34:501-511
  103. 103. Sato H, Todaka D, Kudo M, Mizoi J, Kidokoro S, Zhao Y, et al. The Arabidopsis transcriptional regulator DPB3-1 enhances heat stress tolerance without growth retardation in rice. Plant Biotechnology Journal. 2016;14(8):1756-1767
  104. 104. Guan QJ, Ma HY, Wang ZJ, Wang ZY, Bu QY, Liu SK. A rice LSD1-like-type ZFP gene OsLOL5 enhances saline-alkaline tolerance in transgenic Arabidopsis thaliana, yeast and rice. BMC Genomics. 2016;17:142
  105. 105. Zhang H, Liu W, Wan L, Li F, Dai L, Li D, et al. Functional analyses of ethylene response factor JERF3 with the aim of improving tolerance to drought and osmotic stress in transgenic rice. Transgenic Research. 2010;19:809-818
  106. 106. Mallikarjuna G, Mallikarjuna K, Reddy MK, Kaul T. Expression of OsDREB2A transcription factor confers enhanced dehydration and salt stress tolerance in rice (Oryza sativa L.). Biotechnology Letters. 2011;33:1689-1697
  107. 107. Joo J, Choi HJ, Lee YH, Kim YK, Song SI. A transcriptional repressor of the ERF family confers drought tolerance to rice and regulates genes preferentially located on chromosome 11. Planta. 2013;238:155-170
  108. 108. Zhu W, Yang L, Yang S, Gai J, Zhu Y. Overexpression of rice phosphate transporter gene OsPT2 enhances nitrogen fixation and ammonium assimilation in transgenic soybean under phosphorus deficiency. Journal of Plant Biology. 2016;59:172-181
  109. 109. Kasukabe Y, He L, Nada K, Misawa S, Ihara I, Tachibana S. Overexpression of spermidine synthase enhances tolerance to multiple environmental stresses and upregulates the expression of various stress-regulated genes in transgenic Arabidopsis thaliana. Plant and Cell Physiology. 2004;45:712-722
  110. 110. Li F, Wu QY, Sun YL, Wang LY, Yang XH, Meng QW. Overexpression of chloroplastic monodehydroascorbate reductase enhanced tolerance to temperature and methyl viologen-mediated oxidative stresses. Physiologia Plantarum. 2010;139:421-434
  111. 111. Niu CF, Wei W, Zhou QY, Tian AG, Hao YJ, Zhang WK, et al. Wheat WRKY genes TaWRKY2 and TaWRKY19 regulate abiotic stress tolerance in transgenic Arabidopsis plants. Plant, Cell and Environment. 2012;35:1156-1170
  112. 112. Amara I, Capellades M, Ludevid MD, Pages M, Goday A. Enhanced water stress tolerance of transgenic maize plants over-expressing LEA Rab28 gene. Journal of Plant Physiology. 2013;170:864-873
  113. 113. Upadhyaya CP, Akula N, Young KE, Chun SC, Kim DH, Park SW. Enhanced ascorbic acid accumulation in transgenic potato confers tolerance to various abiotic stresses. Biotechnology Letters. 2010;32:321-330
  114. 114. Eltayeb AE, Yamamoto S, Habora MEE, Yin L, Tsujimoto H, Tanaka K. Transgenic potato overexpressing Arabidopsis cytosolic AtDHAR1 showed higher tolerance to herbicide, drought and salt stresses. Breeding Science. 2011;61:3-10
  115. 115. Seo JS, Sohn HB, Noh K, Jung C, An JH, Donovan CM, et al. Expression of the Arabidopsis AtMYB44 gene confers drought/salt-stress tolerance in transgenic soybean. Molecular Breeding. 2012;29:601-660
  116. 116. Subramanyam K, Arun M, Mariashibu TS, Theboral J, Rajesh M, Singh NK, et al. Overexpression of tobacco osmotin (Tbosm) in soybean conferred resistance to salinity stress and fungal infections. Planta. 2012;236:1909-1925
  117. 117. Ghanti SKK, Sujata KG, Kumar BV, Karba NN, Janardhan Reddy K, Rao MS, et al. Heterologous expression of P5CS gene in chickpea enhances salt tolerance without afecting yield. Biologia Plantarum. 2011;55:634
  118. 118. Hajyzadeh M, Turktas M, Khawar KM, Unver T. miR408 overexpression causes increased drought tolerance in chickpea. Gene. 2015;555:186-193
  119. 119. Surekha CH, Kumari KN, Aruna LV, Suneetha G, Arundhati A, Kishor PK. Expression of the Vigna aconitifolia P5CSF129A gene in transgenic pigeonpea enhances proline accumulation and salt tolerance. Plant Cell, Tissue and Organ Culture. 2014;116:27-36
  120. 120. Bhomkar P, Upadhyay CP, Saxena M, Muthusamy A, Prakash NS, Pooggin M, et al. Salt stress alleviation in transgenic Vigna mungo L. Hepper (blackgram) by overexpression of the glyoxalase I gene using a novel cestrum yellow leaf curling virus (CmYLCV) promoter. Molecular Breeding. 2008;22:169-181
  121. 121. Saha B, Swain D, Borgohain P, Rout GR, Koyama H, Panda SK. Enhanced exudation of malate in the rhizosphere due to AtALMT1 overexpression in blackgram (Vigna mungo L.) confers increased aluminium tolerance. Plant Biology. 2020;22:701-708
  122. 122. Sidhu H, Schmidt ME, Cornelius JG, Thamilselvan S, Khan SR, Hesse A, et al. Direct correlation between hyperoxaluria/oxalate stone disease and the absence of the gastrointestinal tract-dwelling bacterium Oxalobacter formigenes: Possible prevention by gut recolonization or enzyme replacement therapy. Journal of the American Society of Nephrology. 1999;10:S334-S340
  123. 123. Sahoo DP, Kumar S, Mishra S, Kobayashi Y, Panda SK, Sahoo L. Enhanced salinity tolerance in transgenic mungbean overexpressing Arabidopsis antiporter (NHX1) gene. Molecular Breeding. 2016;36:144
  124. 124. Rout GR, Bansal A, Swain D, Jadhao KR, Shelke RG, Panda SK. Overexpression of ICE1 gene in mungbean (Vigna radiata L.) for cold tolerance. Plant Cell, Tissue and Organ Culture. 2020;143(3):593-608
  125. 125. Augustine SM, Narayan AJ, Syamaladevi DP, Appunu C, Chakravarthi M, Ravichandran V, et al. Overexpression of EaDREB2 and pyramiding of EaDREB2 with the pea DNA helicase gene (PDH45) enhance drought and salinity tolerance in sugarcane (Saccharum spp. hybrid). Plant Cell Reports. 2015;34:247-263
  126. 126. Singh N, Mishra A, Jha B. Over-expression of the peroxisomal ascorbate peroxidase (SBPAPX) gene cloned from halophyte Salicornia brachiata confers salt and drought stress tolerance in transgenic tobacco. Marine Biotechnology. 2014;16:321-332
  127. 127. Zhai Y, Wang Y, Li Y, Lei T, Yan F, Su L, et al. Isolation and molecular characterization of GmERF7, a soybean ethylene-response factor that increases salt stress tolerance in tobacco. Gene. 2013;513:174-183
  128. 128. Li F, Murillo C, Wurtzel ET. Maize Y9 encodes a product essential for 15-cis-zeta-carotene isomerization. Plant Physiology. 2007;144:1181-1189
  129. 129. Chen Y, Li F, Wurtzel ET. Isolation and characterization of the Z-ISO gene encoding a missing component of carotenoid biosynthesis in plants. Plant Physiology. 2010;153:66-79
  130. 130. Bai C, Twyman RM, Farre G, et al. A golden era-pro-vitamin A enhancement in diverse crops. In Vitro Cellular & Developmental Biology. Plant. 2011;47:205-202
  131. 131. Tian L, Magallanes-Lundback M, Musetti V, DellaPenna D. Functional analysis of beta- and epsilon-ring carotenoid hydroxylases in Arabidopsis. Plant Cell. 2003;5:1320-1332
  132. 132. Seo M, Koshiba T. Complex regulation of ABA biosynthesis in plants. Trends in Plant Science. 2002;7:41-48
  133. 133. Kawakatsu T, Yamamoto MP, Hirose S, et al. Characterization of a new rice glutelin gene GluD-1 expressed in the starchy endosperm. Journal of Experimental Botany. 2008;59:4233-4245
  134. 134. Qu LQ , Xing YP, Liu WX, et al. Expression pattern and activity of six glutelin gene promoters in transgenic rice. Journal of Experimental Botany. 2008;59(9):2417-2242
  135. 135. Xu R, Li D, Li H, et al. Isolation of four rice seed-specific promoters and evaluation of endosperm activity. Plant Cell, Tissue and Organ Culture. 2017;128:125-132
  136. 136. Burkhardt PK, Beyer P, Wunn J, et al. Transgenic rice (Oryza sativa) endosperm expressing daffodil (Narcissus pseudonarcissus) phytoene synthase accumulates phytoene, a key intermediate of provitamin a biosynthesis. The Plant Journal. 1997;11(5):1071-1078
  137. 137. Ye X, Al-Babili S, Kloti A, et al. Engineering the provitamin A (beta-carotene) biosynthetic pathway into (carotenoid-free) rice endosperm. Science. 2000;287(5451):303-305
  138. 138. Rai M, Datta K, Parkhi V, et al. Variable T-DNA linkage configuration affects inheritance of carotenogenic transgenes and carotenoid accumulation in transgenic indica rice. Plant Cell Reports. 2007;26:1221-1231
  139. 139. Parkhi V, Rai M, Tan J, et al. Molecular characterization of marker free transgenic indica rice lines that accumulate carotenoids in seed endosperm. Molecular Genetics and Genomics. 2005;274:325-336
  140. 140. Paine JA, Shipton CA, Chagger S, et al. Improving the nutritional value of Golden rice through increased pro-vitamin A content. Nature Biotechnology. 2005;23:482-487
  141. 141. Ha S-H, Liang YS, Jung H, et al. Application of two bicistronic systems involving 2A and IRES sequences to the biosynthesis of carotenoids in rice endosperm. Plant Biotechnology Journal. 2010;8:928-938
  142. 142. Ha S-H, Kim JK, Jeong YS, et al. Stepwise pathway engineering to the biosynthesis of zeaxanthin, astaxanthin and capsanthin in rice endosperm. Metabolic Engineering. 2019;52:178-189
  143. 143. Chaudhary A, Teotia S, Singh D. Tools for engineering resistance against pathogens in plants. Journal of Plant Biochemistry and Biotechnology. 2022;31:459-488
  144. 144. Li A, Jia S, Yobi A, Ge Z, Sato SJ, Zhang C, et al. Editing of an alpha-kafirin gene family increases, digestibility and protein quality in sorghum. Plant Physiology. 2018a;177:1425-1438
  145. 145. Zorrilla-Fontanesi Y, Pauwels L, Panis B, Signorelli S, Vanderschuren H, Swennen R. Strategies to revise agrosystems and breeding to control fusarium wilt of banana. Nature Food. 2020;1(10):599-604
  146. 146. Klap C, Yeshayahou E, Bolger AM, Arazi T, Gupta SK, Shabtai S, et al. Tomato facultative parthenocarpy results from SlAGAMOUSLIKE 6 loss of function. Plant Biotechnology Journal. 2017;15:634-647
  147. 147. Andersson M, Turesson H, Nicolia A, Falt AS, Samuelsson M, Hofvander P. Efficient targeted multiallelic mutagenesis in tetraploid potato (Solanum tuberosum) by transient CRISPR-Cas9 expression in protoplasts. Plant Cell Reports. 2017;36:117-128
  148. 148. Vasconcelos M, Datta K, Oliva N, Khalekuzzaman M, Torrizo L, Krishnan S, et al. Enhanced iron and zinc accumulation in transgenic rice with the ferritin gene. Plant Science. 2003;164:371-378
  149. 149. Lucca P, Hurrell R, Potrykus I. Genetic engineering approaches to improve the bioavailability and the level of iron in rice grains. Theoretical and Applied Genetics. 2001;102(2):392-397
  150. 150. Suzuki YA, Kelleher SL, Yalda D, Wu L, Huang J, Huang N, et al. Expression, characterization and biological activity of recombinant human lactoferrin in rice. Journal of Pediatric Gastroenterology and Nutrition. 2003;36:190-199
  151. 151. Anai T, Koga M, Tanaka H, Kinoshita T, Rahman SM, Takagi Y. Improvement of rice (Oryza sativa L.) seed oil quality through introduction of a soybean microsomal omega-3 fatty acid desaturase gene. Plant Cell Reports. 2003;21:988-992
  152. 152. Storozhenko S, De Brouwer V, Volckaert M, Navarrete O, Blancquaert D, Zhang GF, et al. Folate fortification of rice by metabolic engineering. Nature Biotechnology. 2007;25:1277-1279
  153. 153. Kawakatsu T, Wang S, Wakasa Y, Takaiwa F. Increased lysine content in rice grains by over-accumulation of BiP in the endosperm. Bioscience, Biotechnology, and Biochemistry. 2010;74:2529-2531
  154. 154. Caimi PG, McCole LM, Klein TM, Kerr PS. Fructan accumulation and sucrose metabolism in transgenic maize endosperm expressing a bacillus amyloliquefaciens SacB gene. Plant Physiology. 1996;110:355-363
  155. 155. Falco SC, Guida T, Locke M, Mauvais J, Sandres C, Ward RT, et al. Transgenic canola and soybean seeds with increased lysine. Biotechnology. 1995;13:577-582
  156. 156. Cahoon EB, Hall SE, Ripp KG, Ganzke TS, Hitz WD, Coughlan SJ. Metabolic redesign of vitamin E biosynthesis in plants for tocotrienol production and increased antioxidant content. Nature Biotechnology. 2003;21:1082-1087
  157. 157. Chen Z, Young TE, Ling J, Chang SC, Gallie DR. Increasing vitamin C content of plants through enhanced ascorbate recycling. Proceedings of the National Academy of Sciences of the United States of America. 2003;100:3525-3530
  158. 158. Drakakaki G, Marcel S, Glahn RP, Lund EK, Pariagh S, Fischer R, et al. Endosperm-specific co-expression of recombinant soybean ferritin and aspergillus phytase in maize results in significant increases in the levels of bioavailable iron. Plant Molecular Biology. 2005;59:869-880
  159. 159. Thu TT, Dewaele E, Claeys M, Jacobs M, Angenon G. Increasing lysine levels in pigeonpea (Cajanus cajan (L.) Millsp) seeds through genetic engineering. Plant Cell, Tissue and Organ Culture. 2007;91:135-143
  160. 160. Dinkins RD, Reddy MSS, Meurer CA, Yan B, Trick H, Thibaud-Nissen F, et al. Increased sulfur amino acids in soybean plants overexpressing the maize 15 kD azein protein. In Vitro Cellular & Developmental Biology. 2001;37:742-747
  161. 161. Valentine MF, De Tar JR, Mookkan M, Firman JD, Zhang ZJ. Silencing of soybean rafnose synthase gene reduced rafnose family oligosaccharides and increased true metabolizable energy of poultry feed. Frontiers in Plant Science. 2017;8:692
  162. 162. Mariani C, De Beuckeleer M, Truettner J, Leemans J, Goldberg RB. Induction of male sterility in plants by a chimeric ribonuclease gene. Nature. 1990;347:737-741
  163. 163. Mariani C, Gossele V, Beuckeleer MD, Block MD, Goldberg RB, Greef WD, et al. A chimaeric ribonuclease-inhibitor gene restores fertility to male sterile plants. Nature. 1992;357:384-387
  164. 164. Bedinger P. The remarkable biology of pollen. Plant Cell. 1992;4:879-887
  165. 165. Denis M, Delourme R, Gourret JP, Mariani C, Renard M. Expression of engineered nuclear male sterility in Brassica napus. Plant Physiology. 1993;101(4):1295-1304
  166. 166. De Cosa B, Moar W, Lee S-B, Miller M, Daniell H. Overexpression of the Bt cry2Aa2 operon in chloroplasts leads to formation of insecticidal crystals. Nature Biotechnology. 2001;19:71-74
  167. 167. Ruiz ON, Daniell H. Engineering cytoplasmic male sterility via the chloroplast genome by expression of b-ketothiolase. Plant Physiology. 2005;138:1232-1246
  168. 168. Worrall D, Hird DL, Hodge R, Paul W, et al. Premature dissolution of the microsporocyte callose wall causes male sterility in transgenic tobacco. The Plant Cell. 1992;4:759-771
  169. 169. Chang Z, Chen Z, Wang N, Xie G, Lu J, Yan W, et al. Construction of a male sterility system for hybrid rice breeding and seed production using a nuclear male sterility gene. Proceedings of the National Academy of Sciences. 2016;113:14145-14150
  170. 170. Hawkes T, Pline-Srnic W, Dale R, Friend E, et al. Dglufosinate as a male sterility agent for hybrid seed production. Plant Biotechnology Journal. 2011;9:301-314
  171. 171. Guerineau F, Sorensen AM, Fenby N, Scott R. Temperature sensitive diphtheria toxin confers conditional malesterility in Arabidopsis thaliana. Plant Biotechnology Journal. 2003;1:33-42
  172. 172. Ishiguro S, Kawai-Oda A, Ueda J, Nishida I, Okada K. The DEFECTIVE IN ANTHER DEHISCENCE1 gene encodes a novel phospholipase A1 catalyzing the initial step of jasmonic acid biosynthesis, which synchronizes pollen maturation, anther DEHISCENCE, and flower opening in Arabidopsis. The Plant Cell. 2001;13:2191-2209
  173. 173. Sanders PM, Lee PY, Biesgen C, Boone JD, et al. The Arabidopsis DELAYED DEHISCENCE1 gene encodes an enzyme in the jasmonic acid synthesis pathway. The Plant Cell. 2000;12:1041-1061
  174. 174. Bae HK, Kang HG, Kim GJ, Eu HJ, et al. Transgenic rice plants carrying RNA interference constructs of AOS (allene oxide synthase) genes show severe male sterility. Plant Breeding. 2010;129:647-651
  175. 175. Fischer R, Budde I, Hain R. Stilbene synthase gene expression causes changes in flower colour and male sterility in tobacco. The Plant Journal. 1997;11:489-498
  176. 176. Handa H, Gualberto JM, Grienenberger J-M. Characterization of the mitochondrial orfB gene and its derivative, orf224, a chimeric open reading frame specific to one mitochondrial genome of the “Polima” male-sterile cytoplasm in rapeseed (Brassica napus L.). Current Genetics. 1995;28:546-552
  177. 177. Chakraborty A, Mitra J, Bhattacharyya J, Pradhan S, et al. Transgenic expression of an unedited mitochondrial orfB gene product from wild abortive (WA) cytoplasm of rice (Oryza sativa L.) generates male sterility in fertile rice lines. Planta. 2015;241:1463-1479
  178. 178. Nizampatnam NR, Doodhi H, Narasimhan YK, Mulpuri S, Viswanathaswamy DK. Expression of sunflower cytoplasmic male sterility-associated open reading frame, orfH522 induces male sterility in transgenic tobacco plants. Planta. 2009;229:987-1001
  179. 179. Nizampatnam NR, Kumar VD. Intron hairpin and transitive RNAi mediated silencing of orfH522 transcripts restores male fertility in transgenic male sterile tobacco plants expressing orfH522. Plant Molecular Biology. 2011;76:557-573
  180. 180. Wang F, Wang C, Liu P, Lei C, Hao W, Gao Y, et al. Enhanced rice blast resistance by CRISPR/Cas9-targeted mutagenesis of the ERF transcription factor gene OsERF922. PLoS One. 2016;11:e0154027
  181. 181. Zhang Y, Li D, Zhang D, Zhao X, Cao X, Dong L, et al. Analysis of the functions of TaGW2 homoeologs in wheat grain weight and protein content traits. The Plant Journal. 2018;94(5):857-866
  182. 182. Shi J, Gao H, Wang H, Lafitte HR, Archibald RL, Yang M, et al. ARGOS8 variants generated by CRISPR-Cas9 improve maize grain yield under field drought stress conditions. Plant Biotechnology Journal. 2017;15:207-216. DOI: 10.1111/pbi.12603
  183. 183. Scudellari MEGAN. Hijacking evolution. Nature. 2019;571(7764):160-162
  184. 184. Kyrou K, Hammond AM, Galizi R, Kranjc N, Burt A, Beaghton AK, et al. A CRISPR–Cas9 gene drive targeting doublesex causes complete population suppression in caged Anopheles gambiae mosquitoes. Nature Biotechnology. 2018;36(11):1062-1066
  185. 185. Fu G, Condon KC, Epton MJ, Gong P, Jin L, Condon GC, et al. Female-specific insect lethality engineered using alternative splicing. Nature Biotechnology. 2007;25(3):353-357
  186. 186. Phuc HK, Andreasen MH, Burton RS, Vass C, Epton MJ, Pape G, et al. Late-acting dominant lethal genetic systems and mosquito control. BMC Biology. 2007;5(1):1-11

Written By

Palanisamy Dhamotharan, Sakthivel Surya Krishna, Sampathkumar Sruthi and Subburaj Monica

Submitted: 05 May 2022 Reviewed: 19 June 2022 Published: 06 September 2022