Open access peer-reviewed chapter - ONLINE FIRST

Chloroplasts: The Future of Large-Scale Protein Production

Written By

Brenda Julian Chávez, Stephanie Solano Ornelas, Quintín Rascón Cruz, Carmen Daniela González Barriga, Sigifredo Arévalo Gallegos, Blanca Flor Iglesias Figueroa, Luis Ignacio Siañez Estrada, Tania Siqueiros Cendón, Sugey Ramona Sinagawa García and Edward Alexander Espinoza Sánchez

Submitted: 28 April 2023 Reviewed: 11 May 2023 Published: 30 May 2023

DOI: 10.5772/intechopen.111829

Chloroplast Structure and Function IntechOpen
Chloroplast Structure and Function Edited by Muhammad Sarwar Khan

From the Edited Volume

Chloroplast Structure and Function [Working Title]

Prof. Muhammad Sarwar Khan

Chapter metrics overview

129 Chapter Downloads

View Full Metrics

Abstract

Chloroplast engineering has matured considerably in recent years. It is emerging as a promising tool to address the challenges related to food security, drug production, and sustainable energy posed by an ever-growing world population. Chloroplasts have proven their potential by efficiently expressing transgenes, encapsulating recombinant proteins, and protecting them from cellular machinery, making it possible to obtain highly functional proteins. This quality has also been exploited by interfering RNA technology. In addition to the practical attributes offered by chloroplast transformation, such as the elimination of position effects, polycistronic expression, and massive protein production, the technique represents an advance in biosafety terms; however, even if its great biotechnological potential, crops that have efficiently transformed are still a proof of concept. Despite efforts, other essential crops have remained recalcitrant to chloroplast transformation, which has limited their expansion. In this chapter, we address the most recent advances in this area and the challenges that must be solved to extend the transformation to other crops and become the de facto tool in plant biotechnology.

Keywords

  • chloroplast
  • engineering
  • agronomical traits
  • biofuels
  • biopharmaceuticals
  • mass protein production

1. Introduction

It has recently been projected that the world population will reach 9.8 billion people in the next three decades, considering that plants provide about 80% of the food that is consumed and that traditional agriculture is inefficient [1, 2, 3], meeting their food and public health needs constitutes a challenge. Therefore, seeking to meet the requirements of this growth, biotechnology has used plants as platforms for the production of proteins, and this has been exploited by different industries such as energetic, pharmaceutical, and agricultural, which have already obtained fortified foods, vaccines, antibodies, plants that exhibit resistance to pests and diseases as well as plants with low content of cell wall components [4].

Modified plants have usually been developed by inserting genes directly into the nuclear genome because of the facilities this offers, such as the possibility of using different transformation methods, high transformation frequencies, and the fact that different plant species can be transformed using a single vector type. However, the use of chloroplast transformation seems to be increasing because, although it has lower levels of transgene integration and there are difficulties in obtaining homoplasmic plants [5], it allows the containment of transgenes, which can even be expressed in a polycistronic way, it avoids position effects, and above all, increase the production levels of recombinant proteins. In addition, an advantage that can be exploited is that chloroplasts can function as biocapsules, accumulating proteins of therapeutic interest, and these can pass through the digestive tract without disturbance [6, 7].

Although chloroplast transformation has been implemented in a broader set of plant species, it has not yet been satisfactorily implemented in important crops, such as cereals, considering that these provide about 50% of the proteins in the human diet [8], limits the expansion of this technology. However, although chloroplast transformation still has challenges, it offers an opportunity to lower the production costs of recombinant proteins without biosafety concerns. In this chapter, we explore the advances in the genetic engineering of chloroplasts to produce proteins of industrial interest and the challenges that must be overcome to establish this technology as the preferred tool in the expression of recombinant proteins.

Advertisement

2. Genetics of the chloroplast

Chloroplasts are small organelles primarily associated with their role in the photosynthetic process, which is possible because they can obtain electrons from water [910]. However, other major metabolic processes also occurred within them, such as the biosynthesis of phytohormones, vitamins, pigments, starch, or phenols [11, 12]. Also, the biochemical processes that occur in the chloroplast largely determine the plant’s adaptation to its environment [13, 14].

Even though the chloroplasts are semi-autonomous, the control of chloroplast biogenesis and the chloroplast metabolic processes is mainly under nuclear control because, from the ~3000 proteins that there are in the chloroplast, 97% of them are encoded by the nuclear genome and imported via the Toc/Tic machinery [15, 16, 17]. Nevertheless, some critical proteins are produced in the chloroplast. There are involved in the transcription, replication, and translation, as well as the proteins that make up the complexes of NADPH plastoquinone oxidoreductase, cytochrome b6f, and the ATP-synthase subunit, as well as the photosystems I and II, envelope membrane proteins, ß-subunit of acetyl-CoA-carboxylase, and cytochrome C biogenesis [16, 18].

The proteins produced in the chloroplast are encoded by ~120 genes, which are arranged in a circular double-strand genome of quadripartite structure [16, 19, 20] of which, depending on the species and the tissue age, there are multiple copies [21], e.g., 22 copies in potato leaves and 900 copies in wheat leaves. However, the reason for this is not yet known [22, 23].

It has been considered that most chloroplast genomes are highly conserved in terms of their organization and content. However, the length genome varies from ~150 to 220 kb and always appears to be associated with the contraction or expansion of the Inverted Repeat regions [11, 16, 24], which are essential to gene conservation, replication initiation, stabilization, and the evolution of the chloroplast genome [21, 25, 26]. Nevertheless, this is still under discussion because there are species of angiosperms (pea and alfalfa) and gymnosperms (pine) that lack these regions and species such as Euglena gracilis that have three sequences but, like repeated sequences clustered in tandem [27, 28, 29].

Chloroplast genomes are organized into nucleoids whose number, size, composition, and structural organization are varied, even between adjacent nucleoids [16]. These nucleoids are attached to the thylakoid membrane through different proteins such as MFP1, TCP34, and pTAC16, and seem to be involved in the regulation of gene expression through their association with different SWIB-domain proteins, CND41, sulfite reductase, and WHIRLY proteins [30, 31, 32, 33]. In addition, there is strong evidence that nucleoids serve as platforms for forming ribosomes [34, 35]. Therefore, it is currently accepted that nucleoids participate in replication, transcription, translation, post-translational regulation of gene expression, and repair of chloroplast genomes.

Although the number of nucleoids and genomes per nucleoid is variable, it is considered that a nucleoid comprises 10 to 20 copies of plastid DNA. The large number of chloroplast genomes with the possibility of polycistronic expression has made chloroplast genetic engineering rapidly growing [36, 37].

Advertisement

3. Incorporation of new traits in plants through chloroplast transformation

The growing demand for agricultural products has influenced the search for alternatives that maximize their production, including developing modified plants [38]. These crops have allowed an increase in agricultural production, which has also translated into higher profits for farmers [39, 40], to such a degree that by 2018, the commercialization/planting of modified plants had already been adopted by 71 countries [41], which is a rapid growth rate given that it was introduced to the market in 1995, just 23 years prior. The incorporation of new characteristics to plants has maintained an interest in modified crops, so much so that to date, 88% of cotton and 82% of corn grown in the United States is genetically modified [42], and this trend continues, in March 2023 Brazil approved the wheat-HB4 (Triticum aestivum) crop that exhibits the Hahb-4 transcription factor from Helianthus annuus that confers drought stress tolerance, as reported by the International Service for the Acquisition of Agri-biotech Applications (event: IND-ØØ412–7) [43].

The development of the first modified plants was achieved by nuclear transformation with the intention of conferring resistance to biotic and abiotic stresses [44, 45, 46], and although this method of improvement has been maintained the leadership of plant genetic engineering, to date, we are still grappling with problems such as gene silencing, random transgene integration, inappropriate gene expression regulation, genomic instability, interference with other genes, and selection issues. These problems can affect the expression and function of the transgene, as well as its stability and heritability. Also, there is an ecological risk of the crop-to-crop gene flow [47, 48, 49].

To reduce the problems associated with nuclear transformation, biotechnology has gone deeper into the genetic engineering of chloroplasts, which considerably improves the expression levels of recombinant proteins and decreases the unwanted effects associated with modified crops.

The genetic engineering of chloroplasts seems to be growing rapidly, which is possible thanks to the standardization of DNA delivery protocols, the understanding of the mechanisms that govern transformation efficiency, as well as the increase in sequenced chloroplast genomes, which have gone from 6768 in 2021 to 10,712 reported in the RefSeq database from the National Center for Biotechnology Information (NCBI) today, and of which 1072 of them have been reported only so far in 2023.

Although establishing chloroplast transformation in different crops has been challenging, it has already been successfully established in different species (Table 1), and currently, been developed plants that express enzymes of industrial value, antigenic proteins for vaccine production, proteins of pharmaceutical interest, and proteins for the production of biofuels and biomaterials [36, 94, 95, 96, 97], as well as proteins that confer resistance to pests and diseases [77, 98, 99].

FamilySpeciesRef.
PlantsDicotsSolanaceaeNicotiana sylvestris, Nicotiana tabacum var. Petit Havana, and Nicotiana benthamiana[50, 51, 52, 53]
Petunia hybrida var. Pink Wave[54]
Solanum melongena[55]
Solanum lycopersicum[56]
Capsicum annuum var. G4[57]
Solanum tuberosum cv. Desirée and line 1607[58, 59]
BrassicaceaeBrassica napus cv. FY-4[60]
Brassica oleracea var. capitata[61]
Brassica oleracea var. botrytis[62]
Lesquerella fendleri[63]
Brassica napus[64]
Arabidopsis thaliana[65]
FabaceaeMedicago sativa cv. Longmu 803[66]
Glycine max[67]
ApiaceaeDaucus carota cv. Half long[68]
MalvaceaeGossypium hirsutum cv. Coker310FR[69]
AsteraceaeArtemisia annua[70]
Lactuca sativa cv. Verônica, cv Flora and cv. Cisco[71, 72, 73]
PlantaginaceaeScoparia dulcis[74]
CucurbitaceaeMomordica charantia[75]
AmaranthaceaeBeta vulgaris[76]
SalicaceaePopulus alba[77, 78]
MonocotsPoaceaeSaccharum officinarum[79]
Oryza sativa var. Japonica line 19 and Hwa-Chung[80, 81]
LiverwortsMarchantiaceaeMarchantia polymorpha[82]
MossFunariaceaePhyscomitrella patens[83]
AlgaeMicroalgaeIsochrysidaceaeTisochrysis lutea[84]
MonodopsidaceaeNannochloropsis oceanica[85]
PhaeodactylaceaePhaeodactylum tricornutum[86]
Red algaeCyanidiaceaeCyanidioschizon merolae[87]
BangiaceaePyropia yezoensis[88]
PorphyridiophyceaePorphyridium sp. UTEX 637[89]
Green algaeDunaliellaceaeDunaliella tertiolecta[90]
HaematococcaceaeHaematococcus pluvialis[91]
EuglenaceaeEuglena gracilis[92]
ChlamydomonadaceaeChlamydomonas reindhartii[93]

Table 1.

Species transformed by stable chloroplast transformation.

3.1 Chloroplast transformation to confer resistance to pests

Annually, about 40% of agricultural products are lost before harvest due to attacks by insects, weeds, and plant pathogens, increasing by 20% after harvest [100]. Therefore, using pesticides has been one of the alternatives that have allowed controlling these losses, increasing their use in cropland from 1.55 kg Ha−1 in 1990 to 2.69 kg Ha−1 in 2019 [101].

The use of pesticides has contributed significantly to agricultural production. However, the emergence of pests resistant to these chemicals (>17,000 cases of resistance amongst 612 species globally by 2020) [102] has resulted in a greater reliance on higher doses and the introduction of new pesticides, which is a worrying trend since it can lead to environmental contamination, and more significant risks for human health; by the way, each year, millions of people around the world experience unintentional acute poisoning by pesticides (~385 million reported cases), of which 11,000 cases end in death [103]. Although this figure is already worrying in itself, it must be considered that 50% of pesticides are organophosphates. These are of particular concern since they pass from the roots to the leaves and can be consumed by the public, which increases the risk of poisoning [104, 105, 106].

Pursuing sustainable and effective alternatives to protect crops without the negative consequences of excessive pesticide use resulted in the introduction of the first insecticidal gene, Cry1Ab, to plants through nuclear transformation [44]. Since then, numerous studies have demonstrated the efficacy of Cry proteins against various pests, including Plutella xylostella Sesamia inferens, Chilo suppressalis, Herpitogramma licarisalis, Mycalesis gotama, Cnaphalocrocis medinalis, Scirpophaga incertulas, Naranga anescens, Parnara guttata, and Elasmopalpus lignosellus [107, 108, 109, 110, 111]. Cry proteins have also been tested to control nematodes, Phthiraptera, Orthoptera, mites, Coleoptera, Lepidoptera, Diptera, Hymenoptera, Hemiptera, and protozoa [112, 113]. The usefulness of these proteins has led to their continued study and exploration, with over 700 insecticidal proteins already reported (http://www.lifesci.sussex.ac.uk/home/Neil_Crickmore/Bt/). However, insects’ resistance to Bt crops has been reported (from 3 cases in 2005 to 16 in 2016), affecting crops containing Cry1Ab, Cry1Ac, Cry1A.105, Cry1Fa, Cry2Ab, Cry3Bb, mCry3A, eCry3.1Ab, and Cry34/35Ab [114, 115]. On the other hand, there is a direct observation of resistance by Diabrotica virgifera virgifera to the eCry3.1Ab protein in the field before the plants were commercialized [116, 117]; such resistance may increase through cross-resistance.

Although to reduce the risks of insect resistance to Bt crops have been used strategies such as multiple toxins, protein engineering, ultra-high doses, and refugees [118], the existence of other risks of the nuclear transformation, such as ecological risks and collateral effects in the expression of the genes, have motivated the use of chloroplast engineering for the expression of proteins with insecticidal potential, which began with the expression of the Cry1Ac protein in tobacco chloroplast obtaining protection against lepidopteran Heliothis virescens, Helicoverpa zea and Spodoptera exigua [45].

Other Cry proteins such as Cry9Aa2, Cry1Ia5, Cry2Aa2, Cry1C, and Cry1Ab have also been expressed in chloroplast from soybean, cabbage, and poplar, including protecting the plants against Phthorimaea operculella, Hyphantria cunea, P. xylostella, Lymantria dispar, Anticarsia gemmatalis and Helicoverpa armigera [77, 119, 120, 121, 122, 123], and until today, the insects did not show resistance to Cry proteins compartmentalized in the chloroplast; therefore, the accumulation proteins in the chloroplast, either through their direct expression within it or by redirecting nuclear-expressed proteins to the chloroplast for storage, seems to be a viable option to avoid the development of resistance. Although this last option it as already been analyzed with the expression of Tvip3A*, Cry1Ac, Cry1Ah, and Cry2A proteins [124, 125, 126, 127], and is very attractive because it benefits from the facilities offered by nuclear transformation, the compartmentalization of proteins does not eliminate the risks of environmental contamination, nor any other undesired effect caused by gene insertion into the nuclear genome. Therefore it requires a more detailed analysis.

Other proteins with insecticidal activity have been expressed using chloroplasts, such as a chloroperoxidase from Pseudomonas pyrrocinia and the pta gene from Pinellia ternata agglutinin being effective in the control of Alternaria alternata, Aspergillus flavus, Escherichia coli, Fusarium moniliforme, Pseudomonas syringae, Verticillium dahliae, Colletotrichum destructivum, and Fusarium verticillioides [128, 129, 130].

Although the Cry proteins with which Bt crops have been nuclearly armed have shown excellent performance in controlling different pests, the expression of lectins and other proteins with insecticidal capacity from Bacillus toyonensis and Lysinibacillus sphaericus has also been shown to be effective in controlling pests such as Alphitobius diaperinus, Spodoptera exigua, Cydia pomonella, Anthonomus grandis, Aedes aegypti, and Myzus persicae [131, 132]. Furthermore, while the expression of these proteins in chloroplasts has not been thoroughly investigated, it is important to note them because despite the efficacy of Bt crops in controlling Lepidoptera and Coleoptera, control of Hemiptera has not been entirely successful, as many Hemipteran species have now become significant pests of Bt crops [133, 134].

3.1.1 Expression of RNAi for pest control

In recent years, it has been proposed that the expression of interference RNA (RNAi) for pest control in the chloroplast is a promising alternative to protein expression, and although the RNAi has been achieved previously by nuclear transformation using both double-stranded RNA (dsRNA) and long hairpin RNA (hpRNA) [135, 136, 137], the obtained level protection in plants has been insufficient for practical application [138].

One problem in achieving high levels of protection has been the impossibility of accumulating large amounts of unprocessed RNAi in the host because the nuclear-expressed RNAi are processed by the cellular machinery leaving little raw RNAi available for ingestion by the host [139, 140]. Furthermore, when RNAi is ingested, the host’s digestive system degrades another part of the ingested RNAi, resulting in a reduced amount of RNAi available for effective interference [141]. Therefore, the chloroplast has been visualized as a viable solution because it offers three crucial advantages 1) there is no RNAi processing machinery in these organelles, 2) they can produce and accumulate large amounts of RNAi [142, 143, 144], and 3) the chloroplast acts as a natural bioencapsulation method to protect the RNAi when consumed by insects [145, 146].

There is limited information on RNAi expression in the chloroplast, but the published reports show an area of high potential. For example, recently [138] expressed dsRNA targeting the MpDhc64C gene, a newly identified target gene whose silencing causes the lethality of the green peach aphid Myzus persicae. Results revealed that transplastomic plants exhibited significant resistance to aphids, which in turn showed reduced survival, decreased fecundity, and decreased weight of survivors, a similar effect to the obtained by Ren, Cao [147] with the use of dsRNA β-Actin to control of the beetle Henosepilachna vigintioctopunctata in transplastomic potato plants.

Other RNAi studies in chloroplasts have shown results that go the same way. In 2015, Jin et al. [141] silenced the V-ATPase, chitin synthase (Chi gene), and cytochrome P450 monooxygenase (P450 gene) from H. armigera using dsRNA, reducing the weight and growth of larvae. Also, a hpRNA targeting acetylcholinesterase (ACE gene) of H. armigera and a dsRNA targeting β-actin and Vps32 (ACT and SHR genes, respectively) from Colorado potato beetle provided substantial protection against H. armigera and reduced growth of larvae of Colorado potato beetle in transplastomic tobacco and potato plants [140, 143].

Although the RNAi expression in chloroplast has been successfully established with promising results of accumulation and stability, future studies should be focused on elucidating the ideal length of the RNAi since there are reports of dsRNA with a length of 200 nt more protective than a dsRNA of 60 nt or > 200 nt; also, it has been reported that the cellular machinery can degrade the long RNAi producing siRNA that have a less insecticidal effect [136, 140]. Therefore, the adequate length of the RNAi that must be expressed is still not entirely clear, and this is important because the length and type of RNAi, whether it is hpRNA or dsRNA, dramatically influences the accumulation [138, 142, 144].

Other aspects that should also be considered are the suitable target genes and the implementation of efficient RNAi delivery methods, even though topical RNAi has recently been reported and seems a cost-effective alternative [148]. Currently, microinjection and overall ingestion are the most used. Moreover, this last, in practice, should be carefully considered because there are pests that do not have direct contact with chloroplasts and considerably impair RNAi efficacy [149].

In 2017, was approved the first variety of transgenic maize Smartstax PRO®, which nuclearly express a dsRNA of the Snf7 (Sucrose non-fermenting 7) gene from D. virgifera virgifera [150, 151, 152], whose commercial launch was in 2022, which supports the idea that RNAi technology has great potential, and although there is no product developed from chloroplast transformation on the market yet, RNAi technology may help chloroplast engineering in this process.

3.2 Chloroplast transformation to confer abiotic stress tolerance

Reactive oxygen species (ROS) are forms of oxygen partially reduced and produced during biotic and abiotic stress. Although ROS are commonly associated with stress, they are also produced under cellular respiration and photosynthetic processes [153]. Nevertheless, although ROS are produced in normal metabolic processes, the uncontrolled production of ROS causes an alteration of the correct function of the cells. Therefore, decreasing ROS in plants is still an essential target in biotechnology.

Nuclearly have been expressed genes to promote the tolerance to different abiotic stresses such as cold stress (CsWRKY46 gene) [154], oxidative stress (ScVTC2 gene) [155], drought and salt stress tolerance (ZmSNAC13 and ZmWRKY86 gene, respectively) [156, 157], metal tolerance (OsMYB-R1 and SbMT-2 gene) [158, 159, 160], ROS decrease (CfAPX gene) [160] and tolerance to waterlogging (HaOXR2 gene) [161] and have been obtained satisfactory results; nevertheless, trying to improve the results obtained by nuclear expression, genes have been expressed in chloroplasts aimed at increasing the antioxidant pathway, intervening in the glycine betaine (GB) pathway such as codA gene from Arthrobacter globiformis [98, 162, 163].

Other proteins have also been expressed in the chloroplast, such as flavodoxin (fld gene) [164], arabitol dehydrogenase (ArDH) [165], otsB-A operon (trehalose phosphate synthase/phosphatase) [166], γ-tocopherol methyltransferase (TMT gene) [167], betaine aldehyde dehydrogenase (badh gene) [68], dehydroascorbate reductase (DHAR gene), superoxide dismutase (MnSOD gene), glutathione reductase (gor gene) [168], glutathione-S-transferase (GST gene) [162, 169], homogentisate phytyltransferase (HPT gene), conferring tolerance to salt, cold, UV-B radiation, heavy metal, and osmotolerance.

Despite the potential and benefits of plastid transformation in protecting plants, it is challenging to confer abiotic stress resistance due to the involvement of various metabolic pathways. Therefore, providing a robust resistance would require multiple gene expressions. These must be carefully selected since the expression of certain types of genes can cause pleiotropic effects because the encoded proteins could interfere with the structure and function of the thylakoids or decrease the levels of ATP production [170, 171]. Nevertheless, despite the challenges of achieving stress resistance, the population’s constant growth requires finding strategies to address this need.

3.3 Expression of hydrolytic enzymes in chloroplasts

Lignocellulose residues are the most abundant raw material and a highly renewable carbon source on earth [172], and due to particularities as its abundance, availability, and sustainability is believed to be a solution to solve fossil fuel shortage [173, 174, 175], to such a degree that global ethanol and biodiesel production is projected to rise to 132 billion liters and 50 billion liters, respectively, by 2030 [176]. However, although lignocellulosic compounds are abundant [177, 178], a large part of this renewable energy is beyond our reach; that is, despite being produced each year more than 40 million tons of non-edible plant material, lignocellulose is not the most important feedstock for biofuel production because it is not efficiently processed [176, 179].

The processing problem is caused by the presence of lignin that imbibes both the cellulose as well as hemicellulose and limits its degradation [180, 181, 182]. Therefore, physical and chemical methods have been used to increase biomass degradation. However, apart from the fact that these methods can represent a potential ecological risk, they increase production costs.

In the search for alternatives to improve biomass degradation, the use of multiple enzymes has been recurring, and in this sense, different organisms have been the source of them [183]; nevertheless, the hydrophobic nature of the substrate, the enzymes cost, the concentration of required enzymes, as well as the release of phenolic compounds during the enzymatic reaction such as xylan that can inhibit enzyme activity, have limited their use [184, 185, 186], forcing the search for new and more efficient enzymes as well as an efficient method for their accumulation; therefore, the protein expression in the chloroplast it has presented as a promissory strategy to aboard the challenges of the degradation of lignocellulosic biomass.

The expression of enzymes capable of degrading cell wall polymers into chloroplast has already been tested successfully. However, although different genes of pectinases, manganese peroxidase, cutinase, and laccase enzymes from Fusarium solani, Streptomyces thermocarboxydus, Phanerochaete chrysosporium, Trichoderma reesei, and Pleurotus ostreatus has been tested with promissory results [187, 188, 189, 190], cellulases have been the subject of the most intensive search because they represent 40% of plant biomass [177, 178].

Different cellulase enzymes have already been expressed in the chloroplastic compartment with satisfactory results, e.g., xylanases (xynA, xyl, xyn2, Xyl10B, xynA, xyn10A, and xyn11B genes) from Bacillus subtilis, Trichoderma reesei, Clostridium cellulovorans, Thermotoga maritima, Clostridium cellulovorans, and Alicyclobacillus acidocaldarius [191, 192, 193, 194], as well as endo-, exo-glucanases, and β-glucosidases (cel6A, cel9A, cel6B, Cel6, Cel7, EndoV, CelK1, Cel3, TF6A, Bgl-1, bgl1C, EGPh, celA, and celB genes) from Trichoderma reesei, Thermomonospora fusca, Pyrococcus horikoshii, Chaetomium globosum, Paenibacillus sp., Phanerochaete chrysosporium, Aspergillus niger, and Thermotoga neapolitana [36, 195, 196, 197, 198, 199, 200]. Despite the expressed proteins being functional, it should be considered that the microorganisms that possess efficient mechanisms for cellulose degradation have redundant genes often [201]. Therefore, high degradation depends not on a particular protein but on its enzymatic cocktail, which should be considered to improve enzymatic processes.

Regardless of the intense search for cell wall hydrolytic enzymes that have been carried out, there is still a need to identify efficient cellulose enzymes [184] because they are critical factors in paper recycling, cotton processing, juice extraction, detergent production, food industry, animal feed additives and largely determine the price of biofuels [202, 203, 204]. For this reason, it is necessary to use strategies that reduce cellulase enzyme production costs as much as possible to make them commercially viable on a larger scale. Further studies will continue to be carried out to express cellulases in chloroplasts.

3.4 Chloroplast engineering for the biopharmaceutical industry

Population growth implies a constant demand for medicines, representing a lucrative opportunity for the pharmaceutical industry, which only in 2022 billed ~1.48 trillion U.S. dollars, an increase of 4.23% concerning 2021. However, most of the world’s population cannot access medicines due to unaffordable prices [205, 206, 207]. Therefore, products with nutritional and pharmaceutical value are gaining importance to such an extent that there are currently 1775 products with different formulations, biosimilars, and biobetters approved by the Foods and Drugs Administration (FDA) and European Medicines Agency (EMA) for use in humans [208], which can reduce the costs of medical treatments.

Chloroplasts have already begun to contribute to the production of plasma proteins, vaccines, antibodies, and enzymes [97, 209, 210], which is especially relevant since two of the current challenges in treatments with proteins for oral consumption are obtaining high protein concentrations and the possible degradation that they may experience when passing through the digestive system [206]. These challenges could be overcome by the accumulation and bioencapsulation of proteins in chloroplasts [211].

Two decades have passed since the first candidate antigen against a human disease was expressed [212], and since then, other proteins have been expressed, e.g., A27L immunogenic protein (A27L gene) [213], ESAT6, Mtb72F [214, 215], Angiotensin-converting enzyme 2 (ACE2 gene), Angiotensin (1–7) (Ang-[1–7] gene) [7, 211], Coagulation factor VIII (F8 gene) [216], E7 Human papillomavirus antigen (E7 gene) [217], Coagulation Factor IX (F9 gene) [96], SAG1 Surface antigen (SAG1 gene) [95], EDIII-1, EDIII-3, EDIII-4 (ediii-1, ediii-3, and ediii-4 genes) [218], KETc1, KETc7, KETc12, GK1, TSOL18/HP6-Tsol [219], Griffithsin protein (grft gene) [220], S1D [221], and Human epidermal growth factor (hEGF gene) [94] in plants; and recently, was cloned the gene IL29 in alga to produce human interleukin 29. The proteins expressed in the chloroplastic compartment have been obtained in high concentrations retaining their functionality, showing the potential value of expressing biopharmaceutical proteins in chloroplasts [222].

One of the challenges currently faced by therapeutic proteins is to stimulate their passage through the epithelial mucosa [223]; however, this challenge has been efficiently addressed by expressing the proteins in the chloroplast fused with the cholera toxin B subunit (CTB), and this has already been tested with the expression of exendin-4 (CTB-EX4) and acid alpha-glucosidase (CTB-GAA) [6, 224]. Fusing the proteins to CTB is an important approach as it has been reported that it could help promote oral tolerance; however, other non-toxic fusion proteins have been reported, such as human transferrin, although there are no reports with this fused protein in the chloroplast.

There is a need for biobetters/biosimilars on the market, and expressing therapeutic proteins in the chloroplast is not only feasible but could solve delivery and efficacy issues. To date, no one plant/chloroplast-based vaccine against human diseases is available to the public. On the other hand, although transplastomic plants have been cultivated for just over a decade with the consent of the United States Department of Agriculture Animal and Plant Health Inspection Service USDA-APHIS, they have not been scaled to the commercial level even though these plants do not fit USDA-APHIS regulation 7 CFR part 340 [206].

Advertisement

4. Conclusions

Given that the world population has been projected to increase by just over 20% in 30 years, it becomes evident that it will be necessary to adopt new technologies to face this situation in areas such as agriculture, energy, and pharmaceuticals, which are key sectors to guarantee well-being and sustainability in the future. For almost three decades, chloroplast transformation has shown important qualities that can help us address the challenges, offering greater efficiency and performance in the expression of modified genes and more precision and control in genetic modification accompanied by greater environmental safety and biosafety. While some challenges need to be addressed, such as the optimization of regulatory regions, limited expression of proteins in non-photosynthetic tissues, the understanding of the mechanisms that govern pleiotropic effects, and the current inability to transform cereals efficiently, it is a fact that it can play a fundamental role in the production of products.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. FAO. Plant health 2020 [Internet]. 2020. Available from: https://www.fao.org/plant-health-2020/en/. [Accesed: March 12, 2023]
  2. 2. Liu Y-x, Li F, Gao L, Tu Z-l, Zhou F, Lin Y-j. Advancing approach and toolbox in optimization of chloroplast genetic transformation technology. Journal of Integrative Agriculture. 2023; In Press. DOI: 10.1016/j.jia.2023.02.031
  3. 3. UN. Department of Economic and Social Affairs, Population Division, World population prospects 2022 [Internet]. 2022. Available from: https://population.un.org/wpp/. [Accesed: April 11, 2023]
  4. 4. Xu J, Towler M, Weathers PJ. Platforms for plant-based protein production. In: Pavlov A, Bley T, editors. Bioprocessing of Plant In Vitro Systems. Cham: Springer International Publishing; 2018. pp. 509-548. DOI: 10.1007/978-3-319-54600-1_14
  5. 5. Rozov SM, Sidorchuk YV, Deineko EV. Transplastomic plants: Problems of production and their solution. Russian Journal of Plant Physiology. 2022;69(2):20. DOI: 10.1134/S1021443722020157
  6. 6. Kwon KC, Nityanandam R, New JS, Daniell H. Oral delivery of bioencapsulated exendin-4 expressed in chloroplasts lowers blood glucose level in mice and stimulates insulin secretion in beta-TC 6 cells. Plant Biotechnology Journal. 2012;11(1):77-86. DOI: 10.1111/pbi.12008
  7. 7. Shenoy V, Kwon K-C, Rathinasabapathy A, Lin S, Jin G, Song C, et al. Oral delivery of angiotensin-converting enzyme 2 and angiotensin-(1-7) bioencapsulated in plant cells attenuates pulmonary hypertension. Hypertension. 2014;64(6):1248-1259. DOI: 10.1161/HYPERTENSIONAHA.114.03871
  8. 8. Labuschagne M. A review of cereal grain proteomics and its potential for sorghum improvement. Journal of Cereal Science. 2018;84:151-158. DOI: 10.1016/j.jcs.2018.10.010
  9. 9. Farquhar J, Zerkle AL, Bekker A. Geological constraints on the origin of oxygenic photosynthesis. Photosynthesis Research. 2011;107(1):11-36. DOI: 10.1007/s11120-010-9594-0
  10. 10. Taylor NL, Stroher E, Millar AH. Arabidopsis organelle isolation and characterization. Methods in Molecular Biology. 2014;1062:551-572. DOI: 10.1007/978-1-62703-580-4_29
  11. 11. Daniell H, Lin CS, Yu M, Chang WJ. Chloroplast genomes: Diversity, evolution, and applications in genetic engineering. Genome Biology. 2016;17(1):134. DOI: 10.1186/s13059-016-1004-2
  12. 12. Singhal R, Pal R, Dutta S. Chloroplast engineering: Fundamental insights and its application in amelioration of environmental stress. Applied Biochemistry and Biotechnology. 2023;195(4):2463-2482. DOI: 10.1007/s12010-022-03930-8
  13. 13. Zhang Y, Zhang A, Li X, Lu C. The role of chloroplast gene expression in plant responses to environmental stress. International Journal of Molecular Sciences. 2020;21(17):6082. DOI: 10.3390/ijms21176082
  14. 14. Mamaeva A, Taliansky M, Filippova A, Love AJ, Golub N, Fesenko I. The role of chloroplast protein remodeling in stress responses and shaping of the plant peptidome. New Phytologist. 2020;227(5):1326-1334. DOI: 10.1111/nph.16620
  15. 15. Grabsztunowicz M, Koskela MM, Mulo P. Post-translational modifications in regulation of chloroplast function: Recent advances. Frontiers in Plant Science. 2017;8:240. DOI: 10.3389/fpls.2017.00240
  16. 16. Kusnetsov VV. Chloroplasts: Structure and expression of the plastid genome. Russian Journal of Plant Physiology. 2018;65(4):465-476. DOI: 10.1134/S1021443718030044
  17. 17. Leister D. Chloroplast research in the genomic age. TRENDS in Genetics. 2003;19(1):47-56. DOI: 10.1016/S0168-9525(02)00003-3
  18. 18. Green BR. Chloroplast genomes of photosynthetic eukaryotes. The Plant Journal. 2011;66(1):34-44. DOI: 10.1111/j.1365-313X.2011.04541.x
  19. 19. Gruenstaeudl M, Nauheimer L, Borsch T. Plastid genome structure and phylogenomics of Nymphaeales: Conserved gene order and new insights into relationships. Plant Systematics and Evolution. 2017;303(9):1251-1270. DOI: 10.1007/s00606-017-1436-5
  20. 20. Asaf S, Khan AL, Khan AR, Waqas M, Kang S-M, Khan MA, et al. Complete chloroplast genome of Nicotiana otophora and its comparison with related species. Frontiers in Plant Science. 2016;7:843. DOI: 10.3389/fpls.2016.00843
  21. 21. Heinhorst S, Cannon GC. DNA replication in chloroplasts. Journal of Cell Science. 1993;104(1):1-9. DOI: 10.1242/jcs.104.1.1
  22. 22. Scott NS, Tymms MJ, Possingham JV. Plastid-DNA levels in the different tissues of potato. Planta. 1984;161(1):12-19. DOI: 10.1007/BF00951454
  23. 23. Boffey SA, Leech RM. Chloroplast DNA levels and the control of chloroplast division in light-grown wheat leaves. Plant Physiology. 1982;69(6):1387-1391. DOI: 10.1104/pp.69.6.1387
  24. 24. Choi IS, Jansen R, Ruhlman T. Lost and found: Return of the inverted repeat in the legume clade defined by its absence. Genome Biology and Evolution. 2019;11(4):1321-1333. DOI: 10.1093/gbe/evz076
  25. 25. Palmer JD, Thompson WF. Chloroplast DNA rearrangements are more frequent when a large inverted repeat sequence is lost. Cell. 1982;29(2):537-550. DOI: 10.1016/0092-8674(82)90170-2
  26. 26. Wolfe KH, Li WH, Sharp PM. Rates of nucleotide substitution vary greatly among plant mitochondrial, chloroplast, and nuclear DNAs. Proceedings of the National Academy of Sciences. 1987;84(24):9054-9058. DOI: 10.1073/pnas.84.24.9054
  27. 27. Rawson JRY, Kushner SR, Vapnek D, Kirby Alton N, Boerma CL. Chloroplast ribosomal RNA genes in Euglena gracilis exist as three clustered tandem repeats. Gene. 1978;3(3):191-209. DOI: 10.1016/0378-1119(78)90032-X
  28. 28. Kim JS, Kim HT, Kim J-H. The largest plastid genome of monocots: A novel genome type containing at residue repeats in the slipper orchid Cypripedium japonicum. Plant Molecular Biology Reporter. 2015;33(5):1210-1220. DOI: 10.1007/s11105-014-0833-y
  29. 29. Kim JI, Shin H, Škaloud P, Jung J, Yoon HS, Archibald JM, et al. Comparative plastid genomics of Synurophyceae: Inverted repeat dynamics and gene content variation. BMC Evolutionary Biology. 2019;19(1):20. DOI: 10.1186/s12862-018-1316-9
  30. 30. Melonek J, Matros A, Trösch M, Mock HP, Krupinska K. The core of chloroplast nucleoids contains architectural SWIB domain proteins. The Plant Cell. 2012;24(7):3060-3073. DOI: 10.1105/tpc.112.099721
  31. 31. Sato N, Nakayama M, Hase T. The 70-kDa major DNA-compacting protein of the chloroplast nucleoid is sulfite reductase. FEBS Letters. 2001;487(3):347-350. DOI: 10.1016/S0014-5793(00)02342-5
  32. 32. Krupinska K, Desel C, Frank S, Hensel G. WHIRLIES are multifunctional DNA-binding proteins with impact on plant development and stress resistance. Frontiers in Plant Science. 2022;13:880423. DOI: 10.3389/fpls.2022.880423
  33. 33. Murakami S, Kondo Y, Nakano T, Sato F. Protease activity of CND41, a chloroplast nucleoid DNA-binding protein, isolated from cultured tobacco cells. FEBS Letters. 2000;468(1):15-18. DOI: 10.1016/S0014-5793(00)01186-8
  34. 34. Bohne AV. The nucleoid as a site of rRNA processing and ribosome assembly. Frontiers in Plant Science. 2014;5:257. DOI: 10.3389/fpls.2014.00257
  35. 35. Melonek J, Oetke S, Krupinska K. Multifunctionality of plastid nucleoids as revealed by proteome analyses. Biochimica et Biophysica Acta (BBA) - Proteins and Proteomics. 2016;1864(8):1016-1038. DOI: 10.1016/j.bbapap.2016.03.009
  36. 36. Espinoza-Sánchez EA, Torres-Castillo JA, Rascón-Cruz Q , Zavala-García F, Sinagawa-García SR. Production and characterization of fungal β-glucosidase and bacterial cellulases by tobacco chloroplast transformation. Plant Biotechnology Reports. 2016;10(2):61-73. DOI: 10.1007/s11816-016-0386-7
  37. 37. Rascón-Cruz Q , González- Barriga CD, Iglesias-Figueroa BF, Trejo-Muñoz JC, Siqueiros-Cendón T, Sinagawa-García SR, et al. Plastid transformation: Advances and challenges for its implementation in agricultural crops. Electronic Journal of Biotechnology. 2021;51:95-109. DOI: 10.1016/j.ejbt.2021.03.005
  38. 38. Bu Y, Sun B, Zhou A, Zhang X, Takano T, Liu S. Overexpression of AtOxR gene improves abiotic stresses tolerance and vitamin C content in Arabidopsis thaliana. Journal BMC Biotechnology. 2016;16(1):69. DOI: 10.1186/s12896-016-0299-0
  39. 39. Brookes G, Barfoot P. Farm income and production impacts of using GM crop technology 1996-2016. GM Crops & Food. 2018;9(2):59-89. DOI: 10.1080/21645698.2018.1464866
  40. 40. Tschofen M, Knopp D, Hood E, Stöger E. Plant molecular farming: Much more than medicines. Journal of Annual Review of Analytical Chemistry. 2016;9:271-294. DOI: 10.1146/annurev-anchem-071015-041706
  41. 41. ISAAA. Global Status of Commercialized Biotech/GM Crops in 2019. In: BRIEF 55. International Service for the Acquisition of Agri-Biotech Applications. 2019
  42. 42. USDA. Economic Research Service U.S. Department of Agriculture, Recent trends in GE adoption [Internet]. 2020. Available from: https://www.usda.gov/. [Accesed: April 13, 2023]
  43. 43. ISAAA. GM Approval Database/event: IND-ØØ412-7 [Internet]. 2023. Available from: https://www.isaaa.org/gmapprovaldatabase/event/default.asp?EventID=574. [Accesed: April 13, 2023]
  44. 44. Vaeck M, Reynaerts A, Höfte H, Jansens S, De Beuckeleer M, Dean C, et al. Transgenic plants protected from insect attack. Nature. 1987;328(6125):33-37. DOI: 10.1038/328033a0
  45. 45. McBride KE, Svab Z, Schaaf DJ, Hogan PS, Stalker DM, Maliga P. Amplification of a chimeric bacillus gene in chloroplasts leads to an extraordinary level of an insecticidal protein in tobacco. Bio/Technology. 1995;13(4):362-365. DOI: 10.1038/nbt0495-362
  46. 46. Daniell H, Datta R, Varma S, Gray S, Lee S-B. Containment of herbicide resistance through genetic engineering of the chloroplast genome. Nature Biotechnology. 1998;16(4):345. DOI: 10.1038/nbt0498-345
  47. 47. Songstad D, Petolino J, Voytas DF, Reichert N. Genome editing of plants. Journal of Critical Reviews in Plant Sciences. 2017;36(1):1-23. DOI: 10.1080/07352689.2017.1281663
  48. 48. Rajeevkumar S, Anunanthini P, Ramalingam S. Epigenetic silencing in transgenic plants. Journal of Frontiers in Plant Science. 2015;6:693. DOI: 10.3389/fpls.2015.00693
  49. 49. Dively GP, Huang F, Oyediran I, Burd T, Morsello S. Evaluation of gene flow in structured and seed blend refuge systems of non-Bt and Bt corn. Journal of Pest Science. 2020;93(1):439-447. DOI: 10.1007/s10681-016-1637-y
  50. 50. Svab Z, Hajdukiewicz P, Maliga P. Stable transformation of plastids in higher plants. Proceedings of the National Academy of Sciences. 1990;87(21):8526-8530. DOI: 10.1073/pnas.87.21.8526
  51. 51. Svab Z, Maliga P. High-frequency plastid transformation in tobacco by selection for a chimeric aadA gene. Proceedings of the National Academy of Sciences. 1993;90(3):913-917. DOI: 10.1073/pnas.90.3.913
  52. 52. Davarpanah SJ, Jung SH, Kim YJ, Park Y-I, Min SR, Liu JR, et al. Stable plastid transformation in Nicotiana benthamiana. Journal of Plant Biology. 2009;52(3):244-250. DOI: 10.1007/s12374-009-9023-0
  53. 53. Maliga P, Svab Z. Engineering the plastid genome of Nicotiana sylvestris, a diploid model species for plastid genetics. In: Birchler J, editor. Plant Chromosome Engineering: Methods and Protocols. Totowa, NJ: Humana Press; 2010. DOI: 10.1007/978-1-61737-957-4_2
  54. 54. Zubko MK, Zubko EI, Kv Z, Meyer P, Day A. Stable transformation of petunia plastids. Transgenic Research. 2004;13(6):523-530. DOI: 10.1007/s11248-004-2374-x
  55. 55. Singh AK, Verma SS, Bansal KC. Plastid transformation in eggplant (Solanum melongena L.). Transgenic Research. 2010;19(1):113-119. DOI: 10.1007/s11248-009-9290-z
  56. 56. Ruf S, Hermann M, Berger IJ, Carrer H, Bock R. Stable genetic transformation of tomato plastids and expression of a foreign protein in fruit. Nature Biotechnology. 2001;19(9):870. DOI: 10.1038/nbt0901-870
  57. 57. Kota S, Lakkam R, Kasula K, Narra M, Qiang H, Rao Allini V, et al. Construction of a species-specific vector for improved plastid transformation efficiency in Capsicum annuum L. 3 Biotech. 2019;9(6):226. DOI: 10.1007/s13205-019-1747-z
  58. 58. Sidorov VA, Kasten D, Pang SZ, Hajdukiewicz PT, Staub JM, Nehra NS. Stable chloroplast transformation in potato: Use of green fluorescent protein as a plastid marker. The Plant Journal. 1999;19(2):209-216. DOI: 10.1046/j.1365-313X.1999.00508.x
  59. 59. Valkov VT, Gargano D, Manna C, Formisano G, Dix PJ, Gray JC, et al. High efficiency plastid transformation in potato and regulation of transgene expression in leaves and tubers by alternative 5′ and 3′ regulatory sequences. Transgenic Research. 2011;20(1):137-151. DOI: 10.1007/s11248-010-9402-9
  60. 60. Cheng L, Li H-P, Qu B, Huang T, Tu J-X, Fu T-D, et al. Chloroplast transformation of rapeseed (Brassica napus) by particle bombardment of cotyledons. Journal of Plant Cell Reports. 2010;29(4):371-381. DOI: 10.1007/s00299-010-0828-6
  61. 61. Liu CW, Lin CC, Chen JJ, Tseng MJ. Stable chloroplast transformation in cabbage (Brassica oleracea L. var. capitata L.) by particle bombardment. Plant Cell Reports. 2007;26(10):1733-1744. DOI: 10.1007/s00299-007-0374-z
  62. 62. Nugent GD, Coyne S, Nguyen TT, Kavanagh TA, Dix PJ. Nuclear and plastid transformation of Brassica oleracea var. botrytis (cauliflower) using PEG-mediated uptake of DNA into protoplasts. Plant Science. 2006;170(1):135-142
  63. 63. Skarjinskaia M, Svab Z, Maliga P. Plastid transformation in Lesquerella fendleri, an oilseed Brassicacea. Transgenic Research. 2003;12(1):115-122. DOI: 10.1023/A:1022110402302
  64. 64. Hou B-K, Zhou Y-H, Wan L-H, Zhang Z-L, Shen G-F, Chen Z-H, et al. Chloroplast transformation in oilseed rape. Transgenic Research. 2003;12(1):111-114. DOI: 10.1023/A:1022180315462
  65. 65. Sikdar S, Serino G, Chaudhuri S, Maliga P. Plastid transformation in Arabidopsis thaliana. Plant Cell Reports. 1998;18(1-2):20-24. DOI: 10.1007/s002990050525
  66. 66. Wei Z, Liu Y, Lin C, Wang Y, Qa C, Dong Y, et al. Transformation of alfalfa chloroplasts and expression of green fluorescent protein in a forage crop. Biotechnology Letters. 2011;33(12):2487-2494. DOI: 10.1007/s10529-011-0709-2
  67. 67. Dufourmantel N, Pelissier B, Garcon F, Peltier G, Ferullo JM, Tissot G. Generation of fertile transplastomic soybean. Plant Molecular Biology. 2004;55(4):479-489. DOI: 10.1007/s11103-004-0192-4
  68. 68. Kumar S, Dhingra A, Daniell H. Plastid-expressed betaine aldehyde dehydrogenase gene in carrot cultured cells, roots, and leaves confers enhanced salt tolerance. Plant Physiology. 2004;136(1):2843-2854. DOI: 10.1104/pp.104.045187
  69. 69. Kumar S, Dhingra A, Daniell H. Stable transformation of the cotton plastid genome and maternal inheritance of transgenes. Plant Molecular Biology. 2004;56(2):203-216. DOI: 10.1007/s11103-004-2907-y
  70. 70. Kaushal C, Abdin MZ, Kumar S. Chloroplast genome transformation of medicinal plant Artemisia annua. Plant Biotechnology Journal. 2020;18(11):2155-2157. DOI: 10.1111/pbi.13379
  71. 71. Queiroz LN, Maldaner FR, Mendes ÉA, Sousa AR, D’Allastta RC, Mendonça G, et al. Evaluation of lettuce chloroplast and soybean cotyledon as platforms for production of functional bone morphogenetic protein 2. Journal of Transgenic Research. 2019;28(2):213-224. DOI: 10.1007/s11248-019-00116-7
  72. 72. Lelivelt CL, McCabe MS, Newell CA, Bastiaan deSnoo C, Van Dun KM, Birch-Machin I, et al. Stable plastid transformation in lettuce (Lactuca sativa L.). Plant Molecular Biology. 2005;58(6):763-774. DOI: 10.1007/s11103-005-7704-8
  73. 73. Kanamoto H, Yamashita A, Asao H, Okumura S, Takase H, Hattori M, et al. Efficient and stable transformation of Lactuca sativa L. cv. Cisco (lettuce) plastids. Transgenic Research. 2006;15(2):205-217. DOI: 10.1007/s11248-005-3997-2
  74. 74. Muralikrishna N, Srinivas K, Kumar KB, Sadanandam A. Stable plastid transformation in Scoparia dulcis L. Physiology and Molecular Biology of Plants. 2016;22(4):575-581. DOI: 10.1007/s12298-016-0386-7
  75. 75. Narra M, Kota S, Velivela Y, Ellendula R, Allini VR, Abbagani S. Construction of chloroplast transformation vector and its functional evaluation in Momordica charantia L. 3 Biotech. 2018;8(3):140. DOI: 10.1007/s13205-018-1160-z
  76. 76. De Marchis F, Wang Y, Stevanato P, Arcioni S, Bellucci M. Genetic transformation of the sugar beet plastome. Transgenic Research. 2008;18(1):17. DOI: 10.1007/s11248-008-9193-4
  77. 77. Wu Y, Xu L, Chang L, Ma M, You L, Jiang C, et al. Bacillus thuringiensis cry1C expression from the plastid genome of poplar leads to high mortality of leaf-eating caterpillars. Tree Physiology. 2019;39(9):1525-1532. DOI: 10.1093/treephys/tpz073
  78. 78. Okumura S, Sawada M, Park YW, Hayashi T, Shimamura M, Takase H, et al. Transformation of poplar (Populus alba) plastids and expression of foreign proteins in tree chloroplasts. Transgenic Research. 2006;15(5):637-646. DOI: 10.1007/s11248-006-9009-3
  79. 79. Mustafa G, Khan MS. Transmission of engineered plastids in sugarcane, a C4 monocotyledonous plant, reveals that sorting of preprogrammed progenitor cells produce heteroplasmy. Plants. 2020;10(1):26. DOI: 10.3390/plants10010026
  80. 80. Wang Y, Wei Z, Xing S. Stable plastid transformation of rice, a monocot cereal crop. Biochemical and biophysical research communications. 2018;503(4):2376-2379
  81. 81. Lee SM, Kang K, Chung H, Yoo SH, Xu XM, Lee SB, et al. Plastid transformation in the monocotyledonous cereal crop, rice (Oryza sativa) and transmission of transgenes to their progeny. Molecules and Cells. 2006;21(3):401-410
  82. 82. Chiyoda S, Linley PJ, Yamato KT, Fukuzawa H, Yokota A, Kohchi T. Simple and efficient plastid transformation system for the liverwort Marchantia polymorpha L. suspension-culture cells. Transgenic Research. 2007;16(1):41-49. DOI: 10.1007/s11248-006-9027-1
  83. 83. Sugiura C, Sugita M. Plastid transformation reveals that moss tRNAArg-CCG is not essential for plastid function. The Plant Journal. 2004;40(2):314-321. DOI: 10.1111/j.1365-313X.2004.02202.x
  84. 84. Bo Y, Wang K, Wu Y, Cao H, Cui Y, Wang L. Establishment of a chloroplast transformation system in Tisochrysis lutea. Journal of Applied Phycology. 2020;32(5):2959-2965. DOI: 10.1007/s10811-020-02159-4
  85. 85. Gan Q , Jiang J, Han X, Wang S, Lu Y. Engineering the chloroplast genome of oleaginous marine microalga Nannochloropsis oceanica. Frontiers in plant science. 2018;9:439. DOI: 10.3389/fpls.2018.00439
  86. 86. Xie W-H, Zhu C-C, Zhang N-S, Li D-W, Yang W-D, Liu J-S, et al. Construction of novel chloroplast expression vector and development of an efficient transformation system for the diatom Phaeodactylum tricornutum. Marine Biotechnology. 2014;16(5):538-546. DOI: 10.1007/s10126-014-9570-3
  87. 87. Zienkiewicz M, Krupnik T, Drożak A, Golke A, Romanowska E. Transformation of the Cyanidioschyzon merolae chloroplast genome: Prospects for understanding chloroplast function in extreme environments. Plant Molecular Biology. 2017;93(1):171-183. DOI: 10.1007/s11103-016-0554-8
  88. 88. Kong F, Zhao H, Liu W, Li N, Mao Y. Construction of plastid expression vector and development of genetic transformation system for the seaweed Pyropia yezoensis. Marine Biotechnology. 2017;19(2):147-156. DOI: 10.1007/s10126-017-9736-x
  89. 89. Lapidot M, Raveh D, Sivan A, Arad SM, Shapira M. Stable chloroplast transformation of the unicellular red alga Porphyridium species. Plant Physiology. 2002;129(1):7-12. DOI: 10.1104/pp.011023
  90. 90. Georgianna DR, Hannon MJ, Marcuschi M, Wu S, Botsch K, Lewis AJ, et al. Production of recombinant enzymes in the marine alga Dunaliella tertiolecta. Algal Research. 2013;2(1):2-9. DOI: 10.1016/j.algal.2012.10.004
  91. 91. Gutierrez CL, Gimpel J, Escobar C, Marshall SH, Henríquez V. Chloroplast genetic tool for the green microalgae Haematococcus pluvialis (Chlorophyceae, Volvocales). Journal of Phycology. 2012;48(4):976-983. DOI: 10.1111/j.1529-8817.2012.01178.x
  92. 92. Doetsch NA, Favreau MR, Kuscuoglu N, Thompson MD, Hallick RB. Chloroplast transformation in Euglena gracilis: Splicing of a group III twintron transcribed from a transgenic psbK operon. Current Genetics. 2001;39(1):49-60. DOI: 10.1007/s002940000174
  93. 93. Goldschmidt-Clermont M. Transgenic expression of aminoglycoside adenine transferase in the chloroplast: A selectable marker for site-directed transformation of Chlamydomonas. Journal of Nucleic Acids Research. 1991;19(15):4083-4089. DOI: 10.1093/nar/19.15.4083
  94. 94. Morgenfeld MM, Vater CF, Alfano EF, Boccardo NA, Bravo-Almonacid FF. Translocation from the chloroplast stroma into the thylakoid lumen allows expression of recombinant epidermal growth factor in transplastomic tobacco plants. Transgenic Research. 2020;29:295-305. DOI: 10.1007/s11248-020-00199-7
  95. 95. Albarracín RM, Becher ML, Farran I, Sander VA, Corigliano MG, Yácono ML, et al. The fusion of toxoplasma gondii SAG1 vaccine candidate to Leishmania infantum heat shock protein 83-kDa improves expression levels in tobacco chloroplasts. Biotechnology Journal. 2015;10(5):748-759. DOI: 10.1002/biot.201400742
  96. 96. Su J, Zhu L, Sherman A, Wang X, Lin S, Kamesh A, et al. Low cost industrial production of coagulation factor IX bioencapsulated in lettuce cells for oral tolerance induction in hemophilia B. Biomaterials. 2015;70:84-93. DOI: 10.1016/j.biomaterials.2015.08.004
  97. 97. Molina A, Hervas-Stubbs S, Daniell H, Mingo-Castel AM, Veramendi J. High-yield expression of a viral peptide animal vaccine in transgenic tobacco chloroplasts. Plant Biotechnology Journal. 2004;2(2):141-153. DOI: 10.1046/j.1467-7652.2004.00057.x
  98. 98. You L, Song Q , Wu Y, Li S, Jiang C, Chang L, et al. Accumulation of glycine betaine in transplastomic potato plants expressing choline oxidase confers improved drought tolerance. Planta. 2019;249(6):1963-1975. DOI: 10.1007/s00425-019-03132-3
  99. 99. Roudsari FM, Salmanian AH, Mousavi A, Hashemi Sohi H, Jafari M. Regeneration of glyphosate-tolerant Nicotiana tabacum after plastid transformation with a mutated variant of bacterial aroA gene. Iranian Journal of Biotechnology. 2009;7(4):247-253
  100. 100. Paoletti MG, Pimentel D. Environmental risks of pesticides versus genetic engineering for agricultural Pest control. Journal of Agricultural and Environmental Ethics. 2000;12(3):279-303. DOI: 10.1023/A:1009571131089
  101. 101. FAO. Pesticides Indicators [Internet]. 2019. Available from: https://www.fao.org/faostat/en/#data/EP. [Accesed: March 11, 2023]
  102. 102. APRD. Arthropod Pesticide Resistance Database (APRD) [Internet]. 2022. Available from: https://www.pesticideresistance.org/. [Accesed: March 12, 2023]
  103. 103. Boedeker W, Watts M, Clausing P, Marquez E. The global distribution of acute unintentional pesticide poisoning: Estimations based on a systematic review. BMC Public Health. 2020;20(1):1875. DOI: 10.1186/s12889-020-09939-0
  104. 104. Fauzi NIM, Fen YW, Omar NAS, Hashim HS. Recent advances on detection of insecticides using optical sensors. Sensors. 2021;21(11):3856. DOI: 10.3390/s21113856
  105. 105. Olisah C, Rubidge G, Human LRD, Adams JB. A translocation analysis of organophosphate pesticides between surface water, sediments and tissues of common reed Phragmites australis. Chemosphere. 2021;284:131380. DOI: 10.1016/j.chemosphere.2021.131380
  106. 106. Sapbamrer R, Hongsibsong S. Organophosphorus pesticide residues in vegetables from farms, markets, and a supermarket around Kwan Phayao Lake of northern Thailand. Archives of Environmental Contamination and Toxicology. 2014;67(1):60-67. DOI: 10.1007/s00244-014-0014-x
  107. 107. Singsit C, Adang MJ, Lynch RE, Anderson WF, Wang A, Cardineau G, et al. Expression of a bacillus thuringiensis cryIA (c) gene in transgenic peanut plants and its efficacy against lesser cornstalk borer. Transgenic Research. 1997;6(2):169-176. DOI: 10.1023/a:1018481805928
  108. 108. Shu Q , Ye G, Cui H, Cheng X, Xiang Y, Wu D, et al. Transgenic rice plants with a synthetic cry1Ab gene from bacillus thuringiensis were highly resistant to eight lepidopteran rice pest species. Molecular Breeding. 2000;6(4):433-439. DOI: 10.1023/A:1009658024114
  109. 109. Cao J, Zhao JZ, Tang J, Shelton A, Earle E. Broccoli plants with pyramided cry1Ac and cry1C Bt genes control diamondback moths resistant to Cry1A and Cry1C proteins. Theoretical and Applied Genetics. 2002;105(2):258-264. DOI: 10.1007/s00122-002-0942-0
  110. 110. Cao J, Shelton AM, Earle ED. Development of transgenic collards (Brassica oleracea L., var. acephala) expressing a cry1Ac or cry1C Bt gene for control of the diamondback moth. Crop Protection. 2005;24(9):804-813. DOI: 10.1016/j.cropro.2004.12.014
  111. 111. Cao J, Shelton AM, Earle ED. Sequential transformation to pyramid two Bt genes in vegetable Indian mustard (Brassica juncea L.) and its potential for control of diamondback moth larvae. Plant Cell Reports. 2007;27(3):479. DOI: 10.1007/s00299-007-0473-x
  112. 112. Ye W, Zhu L, Liu Y, Crickmore N, Peng D, Ruan L, et al. Mining new crystal protein genes from bacillus thuringiensis on the basis of mixed plasmid-enriched genome sequencing and a computational pipeline. Applied Environmental Microbiology. 2012;78(14):4795-4801. DOI: 10.1128/AEM.00340-12
  113. 113. Guo Y, Weng M, Sun Y, Carballar-Lejarazú R, Wu S, Lian C. Bacillus thuringiensis toxins with nematocidal activity against the pinewood nematode Bursaphelenchus xylophilus. Journal of Invertebrate Pathology. 2022;189:107726. DOI: 10.1016/j.jip.2022.107726
  114. 114. Welch KL, Unnithan GC, Degain BA, Wei J, Zhang J, Li X, et al. Cross-resistance to toxins used in pyramided Bt crops and resistance to Bt sprays in Helicoverpa zea. Journal of Invertebrate Pathology. 2015;132:149-156. DOI: 10.1016/j.jip.2015.10.003
  115. 115. Tabashnik BE, Carrière Y. Surge in insect resistance to transgenic crops and prospects for sustainability. Nature Biotechnology. 2017;35(10):926-935. DOI: 10.1038/nbt.3974
  116. 116. Zukoff SN, Ostlie KR, Potter B, Meihls LN, Zukoff AL, French L, et al. Multiple assays indicate varying levels of cross resistance in Cry3Bb1-selected field populations of the Western corn rootworm to mCry3A, eCry3.1Ab, and Cry34/35Ab1. Journal of Economic Entomology. 2016;109(3):1387-1398. DOI: 10.1093/jee/tow073
  117. 117. Bernardi D, Salmeron E, Horikoshi RJ, Bernardi O, Dourado PM, Carvalho RA, et al. Cross-resistance between Cry1 proteins in fall armyworm (Spodoptera frugiperda) may affect the durability of current pyramided bt maize hybrids in Brazil. PLoS One. 2015;10(10):e0140130-e0140130. DOI: 10.1371/journal.pone.0140130
  118. 118. Huang F, Buschman LL, Higgins RA, McGaughey WH. Inheritance of resistance to bacillus thuringiensis toxin (Dipel ES) in the european corn borer. Science. 1999;284(5416):965-967. DOI: 10.1126/science.284.5416.965
  119. 119. Liu C-W, Lin C-C, Yiu J-C, Chen JJW, Tseng M-J. Expression of a bacillus thuringiensis toxin (cry1Ab) gene in cabbage (Brassica oleracea L. var. capitata L.) chloroplasts confers high insecticidal efficacy against Plutella xylostella. Theoretical and Applied Genetics. 2008;117(1):75-88. DOI: 10.1007/s00122-008-0754-y
  120. 120. De Cosa B, Moar W, Lee SB, Miller M, Daniell H. Overexpression of the Bt cry2Aa2 operon in chloroplasts leads to formation of insecticidal crystals. Nature Biotechnology. 2001;19(1):71-74. DOI: 10.1038/83559
  121. 121. Chakrabarti SK, Lutz KA, Lertwiriyawong B, Svab Z, Maliga P. Expression of the cry9Aa2 B.t. gene in tobacco chloroplasts confers resistance to potato tuber moth. Transgenic Research. 2006;15(4):481-488. DOI: 10.1007/s11248-006-0018-z
  122. 122. Dufourmantel N, Tissot G, Goutorbe F, Garçon F, Muhr C, Jansens S, et al. Generation and analysis of soybean plastid transformants expressing bacillus thuringiensis Cry1Ab protoxin. Plant Molecular Biology. 2005;58(5):659-668. DOI: 10.1007/s11103-005-7405-3
  123. 123. Reddy VS, Leelavathi S, Selvapandiyan A, Raman R, Giovanni F, Shukla V, et al. Analysis of chloroplast transformed tobacco plants with cry1Ia5 under rice psbA transcriptional elements reveal high level expression of Bt toxin without imposing yield penalty and stable inheritance of transplastome. Molecular Breeding. 2002;9(4):259-269. DOI: 10.1023/A:1020357729437
  124. 124. Muzaffar A, Kiani S, Khan MAU, Rao AQ , Ali A, Awan MF, et al. Chloroplast localization of Cry1Ac and Cry2A protein-an alternative way of insect control in cotton. Biological Research. 2015;48(1):1-11. DOI: 10.1186/s40659-015-0005-z
  125. 125. Wu J, Luo X, Zhang X, Shi Y, Tian Y. Development of insect-resistant transgenic cotton with chimeric TVip3A* accumulating in chloroplasts. Transgenic Research. 2011;20(5):963-973. DOI: 10.1007/s11248-011-9483-0
  126. 126. Kiani S, Mohamed BB, Shehzad K, Jamal A, Shahid MN, Shahid AA, et al. Chloroplast-targeted expression of recombinant crystal-protein gene in cotton: An unconventional combat with resistant pests. Journal of Biotechnology. 2013;166(3):88-96. DOI: 10.1016/j.jbiotec.2013.04.011
  127. 127. Li X, Li S, Lang Z, Zhang J, Zhu L, Huang D. Chloroplast-targeted expression of the codon-optimized truncated cry1Ah gene in transgenic tobacco confers a high level of protection against insects. Plant Cell Reports. 2013;32(8):1299-1308. DOI: 10.1007/s00299-013-1444-z
  128. 128. Wang Y-P, Wei Z-Y, Zhang Y-Y, Lin C-J, Zhong X-F, Wang Y-L, et al. Chloroplast-expressed MSI-99 in tobacco improves disease resistance and displays inhibitory effect against rice blast fungus. International Journal of Molecular Sciences. 2015;16(3):4628-4641. DOI: 10.3390/ijms16034628
  129. 129. Jin S, Zhang X, Daniell H. Pinellia ternata agglutinin expression in chloroplasts confers broad spectrum resistance against aphid, whitefly, lepidopteran insects, bacterial and viral pathogens. Plant Biotechnology Journal. 2011;10(3):313-327. DOI: 10.1111/j.1467-7652.2011.00663.x
  130. 130. Ruhlman TA, Rajasekaran K, Cary JW. Expression of chloroperoxidase from pseudomonas pyrrocinia in tobacco plastids for fungal resistance. Plant Science. 2014;228:98-106. DOI: 10.1016/j.plantsci.2014.02.008
  131. 131. Sauka DH, Peralta C, Pérez MP, Onco MI, Fiodor A, Caballero J, et al. Bacillus toyonensis biovar Thuringiensis: A novel entomopathogen with insecticidal activity against lepidopteran and coleopteran pests. Biological Control. 2022;167:104838. DOI: 10.1016/j.biocontrol.2022.104838
  132. 132. Aasen SS, Hågvar EB. Effect of potato plants expressing snowdrop lectin (GNA) on the performance and colonization behaviour of the peach-potato aphid Myzus persicae. Journal of Acta Agriculturae Scandinavica, Section B-Soil Plant Science. 2012;62(4):352-361. DOI: 10.1080/09064710.2011.619996
  133. 133. Wu K-M, Lu Y-H, Feng H-Q , Jiang Y-Y, Zhao J-Z. Suppression of cotton bollworm in multiple crops in China in areas with bt toxin-containing cotton. Science. 2008;321(5896):1676-1678. DOI: 10.1126/science.1160550
  134. 134. Lu Y, Wu K, Jiang Y, Xia B, Li P, Feng H, et al. Mirid bug outbreaks in multiple crops correlated with wide-scale adoption of bt cotton in China. Science. 2010;328(5982):1151-1154. DOI: 10.1126/science.1187881
  135. 135. Gq S, Sink KC, Walworth AE, Cook MA, Allison RF, Lang GA. Engineering cherry rootstocks with resistance to Prunus necrotic ring spot virus through RNA i-mediated silencing. Plant Biotechnology Journal. 2013;11(6):702-708. DOI: 10.1111/pbi.12060
  136. 136. Bolognesi R, Ramaseshadri P, Anderson J, Bachman P, Clinton W, Flannagan R, et al. Characterizing the mechanism of action of double-stranded RNA activity against western corn rootworm (Diabrotica virgifera virgifera LeConte). PLoS One. 2012;7(10):e47534. DOI: 10.1371/journal.pone.0047534
  137. 137. Mao Y-B, Tao X-Y, Xue X-Y, Wang L-J, Chen X-Y. Cotton plants expressing CYP6AE14 double-stranded RNA show enhanced resistance to bollworms. Transgenic Research. 2011;20(3):665-673. DOI: 10.1007/s11248-010-9450-1
  138. 138. Wu M, Dong Y, Zhang Q , Li S, Chang L, Loiacono FV, et al. Efficient control of western flower thrips by plastid-mediated RNA interference. Proceedings of the National Academy of Sciences. 2022;119(15):e2120081119. DOI: 10.1073/pnas.21200811
  139. 139. Gordon KHJ, Waterhouse PM. RNAi for insect-proof plants. Nature Biotechnology. 2007;25(11):1231-1232. DOI: 10.1038/nbt1107-1231
  140. 140. Zhang J, Khan SA, Hasse C, Ruf S, Heckel DG, Bock R. Full crop protection from an insect pest by expression of long double-stranded RNAs in plastids. Science. 2015;347(6225):991-994. DOI: 10.1126/science.1261680
  141. 141. Jin S, Singh ND, Li L, Zhang X, Daniell H. Engineered chloroplast dsRNA silences cytochrome p450 monooxygenase, V-ATPase and chitin synthase genes in the insect gut and disrupts Helicoverpa armigera larval development and pupation. Plant Biotechnology Journal. 2015;13(3):435-446. DOI: 10.1111/pbi.12355
  142. 142. Burke W, Kaplanoglu E, Kolotilin I, Menassa R, Donly C. RNA interference in the tobacco hornworm, Manduca sexta, using plastid-encoded long double-stranded RNA. Frontiers in Plant Science. 2019;10:313. DOI: 10.3389/fpls.2019.00313
  143. 143. Bally J, McIntyre GJ, Doran RL, Lee K, Perez A, Jung H, et al. In-plant protection against Helicoverpa armigera by production of long hpRNA in chloroplasts. Frontiers in Plant Science. 2016;7:1453. DOI: 10.3389/fpls.2016.01453
  144. 144. He W, Xu W, Xu L, Fu K, Guo W, Bock R, et al. Length-dependent accumulation of double-stranded RNAs in plastids affects RNA interference efficiency in the Colorado potato beetle. Journal of Experimental Botany. 2020;71(9):2670-2677. DOI: 10.1093/jxb/eraa001
  145. 145. Turner C, Davy M, MacDiarmid R, Plummer K, Birch N, Newcomb R. RNA interference in the light brown apple moth, Epiphyas postvittana (Walker) induced by double-stranded RNA feeding. Insect Molecular Biology. 2006;15(3):383-391. DOI: 10.1111/j.1365-2583.2006.00656.x
  146. 146. Baum JA, Bogaert T, Clinton W, Heck GR, Feldmann P, Ilagan O, et al. Control of coleopteran insect pests through RNA interference. Nature Biotechnology. 2007;25(11):1322-1326. DOI: 10.1038/nbt1359
  147. 147. Ren B, Cao J, He Y, Yang S, Zhang J. Assessment on effects of transplastomic potato plants expressing Colorado potato beetle β-actin double-stranded RNAs for three non-target pests. Pesticide Biochemistry and Physiology. 2021;178:104909. DOI: 10.1016/j.pestbp.2021.104909
  148. 148. Niu J, Yang W-J, Tian Y, Fan J-Y, Ye C, Shang F, et al. Topical dsRNA delivery induces gene silencing and mortality in the pea aphid. Pest Management Science. 2019;75(11):2873-2881. DOI: 10.1002/ps.5457
  149. 149. Dong Y, Yang Y, Wang Z, Wu M, Fu J, Guo J, et al. Inaccessibility to double-stranded RNAs in plastids restricts RNA interference in Bemisia tabaci (whitefly). Pest Management Science. 2020;76(9):3168-3176. DOI: 10.1002/ps.5871
  150. 150. ISAAA. GM Approval Database [Internet]. 2023. Available from: https://www.isaaa.org/gmapprovaldatabase/. [Accesed: March 12, 2023]
  151. 151. EPA. U.S. Environmental Protection Agency, Registers Innovative Tool to Control Corn Rootworm [Internet]. 2017. Available from: https://www.epa.gov/. [Accesed: March 17, 2023]
  152. 152. Head GP, Carroll MW, Evans SP, Rule DM, Willse AR, Clark TL, et al. Evaluation of SmartStax and SmartStax PRO maize against western corn rootworm and northern corn rootworm: Efficacy and resistance management. Pest Management Science. 2017;73(9):1883-1899. DOI: 10.1002/ps.4554
  153. 153. Mittler R. Oxidative stress, antioxidants and stress tolerance. Trends in Plant Science. 2002;7(9):405-410. DOI: 10.1016/s1360-1385(02)02312-9
  154. 154. Zhang Y, Yu H, Yang X, Li Q , Ling J, Wang H, et al. CsWRKY46, a WRKY transcription factor from cucumber, confers cold resistance in transgenic-plant by regulating a set of cold-stress responsive genes in an ABA-dependent manner. Plant Physiology and Biochemistry. 2016;108:478-487. DOI: 10.1016/j.plaphy.2016.08.013
  155. 155. Miao R, Zang W, Yuan Y, Zhang Y, Zhang A, Pang Q. The halophyte gene ScVTC2 confers resistance to oxidative stress via AsA-mediated photosynthetic enhancement. Plant Physiology and Biochemistry. 2021;169:138-148. DOI: 10.1016/j.plaphy.2021.11.020
  156. 156. Fang X, Li W, Yuan H, Chen H, Bo C, Ma Q , et al. Mutation of ZmWRKY86 confers enhanced salt stress tolerance in maize. Plant Physiology and Biochemistry. 2021;167:840-850. DOI: 10.1016/j.plaphy.2021.09.010
  157. 157. Luo P, Chen Y, Rong K, Lu Y, Wang N, Xu Z, et al. ZmSNAC13, a maize NAC transcription factor conferring enhanced resistance to multiple abiotic stresses in transgenic Arabidopsis. Plant Physiology and Biochemistry. 2022;170:160-170. DOI: 10.1016/j.plaphy.2021.11.032
  158. 158. Patel MK, Pandey S, Patel J, Mishra A. A type 2 metallothionein (SbMT-2) gene cloned from Salicornia brachiata confers enhanced Zn stress-tolerance in transgenic tobacco by transporting Zn2+ and maintaining photosynthesis efficacy. Environmental and Experimental Botany. 2021;191:104626. DOI: 10.1016/j.envexpbot.2021.104626
  159. 159. Tiwari P, Indoliya Y, Chauhan AS, Pande V, Chakrabarty D. Over-expression of rice R1-type MYB transcription factor confers different abiotic stress tolerance in transgenic Arabidopsis. Ecotoxicology and Environmental Safety. 2020;206:111361. DOI: 10.1016/j.ecoenv.2020.111361
  160. 160. Zhang Y, Yang L, Zhang M, Yang J, Cui J, Hu H, et al. CfAPX, a cytosolic ascorbate peroxidase gene from Cryptomeria fortunei, confers tolerance to abiotic stress in transgenic Arabidopsis. Plant Physiology and Biochemistry. 2022;172:167-179. DOI: 10.1016/j.plaphy.2022.01.011
  161. 161. Torti P, Raineri J, Mencia R, Campi M, Gonzalez DH, Welchen E. The sunflower TLDc-containing protein HaOXR2 confers tolerance to oxidative stress and waterlogging when expressed in maize plants. Plant Science. 2020;300:110626. DOI: 10.1016/j.plantsci.2020.110626
  162. 162. Le Martret B, Poage M, Shiel K, Nugent GD, Dix PJ. Tobacco chloroplast transformants expressing genes encoding dehydroascorbate reductase, glutathione reductase, and glutathione-S-transferase, exhibit altered anti-oxidant metabolism and improved abiotic stress tolerance. Plant Biotechnology Journal. 2011;9(6):661-673. DOI: 10.1111/j.1467-7652.2011.00611.x
  163. 163. Zhang J, Tan W, Yang X-H, Zhang H-X. Plastid-expressed choline monooxygenase gene improves salt and drought tolerance through accumulation of glycine betaine in tobacco. Plant Cell Reports. 2008;27(6):1113. DOI: 10.1007/s00299-008-0549-2
  164. 164. Ceccoli RD, Blanco NE, Segretin ME, Melzer M, Hanke GT, Scheibe R, et al. Flavodoxin displays dose-dependent effects on photosynthesis and stress tolerance when expressed in transgenic tobacco plants. Planta. 2012;236(5):1447-1458. DOI: 10.1007/s00425-012-1695-x
  165. 165. Khan MS, Kanwal B, Nazir S. Metabolic engineering of the chloroplast genome reveals that the yeast ArDH gene confers enhanced tolerance to salinity and drought in plants. Frontiers in Plant Science. 2015;6:725. DOI: 10.3389/fpls.2015.00725
  166. 166. Bansal K, Singh A, Wani S. Plastid transformation for abiotic stress tolerance in plants. In: Shabala S, Cuin T, editors. Plant Salt Tolerance. Totowa, NJ: Humana Press; 2012. pp. 351-358. DOI: 10.1007/978-1-61779-986-0_23
  167. 167. Lu Y, Rijzaani H, Karcher D, Ruf S, Bock R. Efficient metabolic pathway engineering in transgenic tobacco and tomato plastids with synthetic multigene operons. Proceedings of the National Academy of Sciences. 2013;110(8):E623-E632. DOI: 10.1073/pnas.1216898110
  168. 168. Poage M, Le Martret B, Jansen MAK, Nugent GD, Dix PJ. Modification of reactive oxygen species scavenging capacity of chloroplasts through plastid transformation. Plant Molecular Biology. 2011;76(3):371-384. DOI: 10.1007/s11103-011-9784-y
  169. 169. Stavridou E, Michailidis M, Gedeon S, Ioakeim A, Kostas S, Chronopoulou E, et al. Tolerance of transplastomic tobacco plants overexpressing a theta class glutathione transferase to abiotic and oxidative stresses. Frontiers in Plant Science. 2019;9:1861. DOI: 10.3389/fpls.2018.01861
  170. 170. Pelletier G, Budar F. The molecular biology of cytoplasmically inherited male sterility and prospects for its engineering. Current Opinion in Biotechnology. 2007;18(2):121-125. DOI: 10.1016/j.copbio.2006.12.002
  171. 171. Ruiz ON, Daniell H. Engineering cytoplasmic male sterility via the chloroplast genome by expression of β-ketothiolase. Plant Physiology. 2005;138(3):1232-1246. DOI: 10.1104/pp.104.057729
  172. 172. Saini JK, Saini R, Tewari L. Lignocellulosic agriculture wastes as biomass feedstocks for second-generation bioethanol production: concepts and recent developments. 3 Biotech. 2015;5(4):337-353. DOI: 10.1007/s13205-014-0246-5
  173. 173. Uday USP, Choudhury P, Bandyopadhyay TK, Bhunia B. Classification, mode of action and production strategy of xylanase and its application for biofuel production from water hyacinth. International Journal of Biological Macromolecules. 2016;82:1041-1054. DOI: 10.1016/j.ijbiomac.2015.10.086
  174. 174. Yan XL, Yanan W, Hongxing L, Zhigang L, Lin Z, Yinbo Q , et al. Cow manure as a lignocellulosic substrate for fungal cellulase expression and bioethanol production. AMB Express. 2018;8(1):190. DOI: 10.1186/s13568-018-0720-2
  175. 175. Bhurat KS, Banerjee T, Pandey JK, Belapurkar P. Fermentative bio-hydrogen production using lignocellulosic waste biomass: A review. Waste Disposal & Sustainable Energy. 2020;2:249-264. DOI: 10.1007/s42768-020-00054-9
  176. 176. OECD/FAO. OECD-FAO Agricultural Outlook 2021-2030. Paris: OECD Publishing; 2021. p. 337. DOI: 10.1787/19428846-en
  177. 177. Bajpai P. Pretreatment of lignocelluloses biomass for bioethanol production. In: Bajpai P, editor. Developments in Bioethanol. Singapore: Springer Singapore; 2021. pp. 111-144. DOI: 10.1007/978-981-15-8779-5_6
  178. 178. Benedetti M, Vecchi V, Betterle N, Natali A, Bassi R, Dall’Osto L. Design of a highly thermostable hemicellulose-degrading blend from Thermotoga neapolitana for the treatment of lignocellulosic biomass. Journal of Biotechnology. 2019;296:42-52. DOI: 10.1016/j.jbiotec.2019.03.005
  179. 179. Sanderson K. Lignocellulose: A chewy problem. Nature. 2011;474(7352):S12-S14. DOI: 10.1038/474S012a
  180. 180. Mota TR, de Oliveira DM, Marchiosi R, Ferrarese-Filho O, dos Santos WD. Plant cell wall composition and enzymatic deconstruction. AIMS Bioengineering. 2018;5(1):63. DOI: 10.3934/bioeng.2018.1.63
  181. 181. Zhang L, Loh KC, Zhang J. Enhanced biogas production from anaerobic digestion of solid organic wastes: Current status and prospects. Bioresource Technology Reports. 2019;5:280-296. DOI: 10.1016/j.biteb.2018.07.005
  182. 182. Oliveira DM, Mota TR, Grandis A, de Morais GR, de Lucas RC, Polizeli ML, et al. Lignin plays a key role in determining biomass recalcitrance in forage grasses. Renewable Energy. 2020;147:2206-2217. DOI: 10.1016/j.renene.2019.10.020
  183. 183. Park SH, Ong RG, Sticklen M. Strategies for the production of cell wall-deconstructing enzymes in lignocellulosic biomass and their utilization for biofuel production. Plant Biotechnology Journal. 2015;14(6):1329-1344. DOI: 10.1111/pbi.12505
  184. 184. Hemansi GR, Kuhad RC, Saini JK. Cost effective production of complete cellulase system by newly isolated aspergillus Niger RCKH-3 for efficient enzymatic saccharification: Medium engineering by overall evaluation criteria approach (OEC). Biochemical Engineering Journal. 2018;132:182-190. DOI: 10.1016/j.bej.2018.01.019
  185. 185. Scheller HV, Ulvskov P. Hemicelluloses. Annual Reviews. 2010;61(1):263-289. DOI: 10.1146/annurev-arplant-042809-112315
  186. 186. Momeni MH, Ubhayasekera W, Sandgren M, Ståhlberg J, Hansson H. Structural insights into the inhibition of cellobiohydrolase Cel7A by xylo-oligosaccharides. The FEBS Journal. 2015;282(11):2167-2177. DOI: 10.1111/febs.13265
  187. 187. Verma D, Kanagaraj A, Jin S, Singh ND, Kolattukudy PE, Daniell H. Chloroplast-derived enzyme cocktails hydrolyse lignocellulosic biomass and release fermentable sugars. Plant biotechnology journal. 2010;8(3):332-350. DOI: 10.1111/j.1467-7652.2009.00486.x
  188. 188. Espinoza-Sánchez EA, Álvarez-Hernández MH, Torres-Castillo JA, Rascón-Cruz Q , Gutiérrez-Díez A, Zavala-García F, et al. Stable expression and characterization of a fungal pectinase and bacterial peroxidase genes in tobacco chloroplast. Electronic Journal of Biotechnology. 2015;18(3):161-168. DOI: 10.1016/j.ejbt.2015.03.002
  189. 189. Verma D, Jin S, Kanagaraj A, Singh ND, Daniel J, Kolattukudy PE, et al. Expression of fungal cutinase and swollenin in tobacco chloroplasts reveals novel enzyme functions and/or substrates. PLoS One. 2013;8(2):e57187. DOI: 10.1371/journal.pone.0057187
  190. 190. Davarpanah S, Ahn J-W, Ko S, Jung S, Park Y-I, Liu J, et al. Stable expression of a fungal laccase protein using transplastomic tobacco. Plant Biotechnology Reports. 2012;6(4):305-312. DOI: 10.1007/s11816-012-0225-4
  191. 191. Leelavathi S, Gupta N, Maiti S, Ghosh A, Siva RV. Overproduction of an alkali-and thermo-stable xylanase in tobacco chloroplasts and efficient recovery of the enzyme. Molecular Breeding. 2003;11:59-67
  192. 192. Kim JY, Kavas M, Fouad WM, Nong G, Preston JF, Altpeter F. Production of hyperthermostable GH10 xylanase Xyl10B from Thermotoga maritima in transplastomic plants enables complete hydrolysis of methylglucuronoxylan to fermentable sugars for biofuel production. Plant Molecular Biology. 2011;76:357-369. DOI: 10.1007/s11103-010-9712-6
  193. 193. Kolotilin I, Kaldis A, Pereira EO, Laberge S, Menassa R. Optimization of transplastomic production of hemicellulases in tobacco: Effects of expression cassette configuration and tobacco cultivar used as production platform on recombinant protein yields. Biotechnology for Biofuels. 2013;6(1):1-15
  194. 194. Castiglia D, Sannino L, Marcolongo L, Ionata E, Tamburino R, De Stradis A, et al. High-level expression of thermostable cellulolytic enzymes in tobacco transplastomic plants and their use in hydrolysis of an industrially pretreated Arundo donax L. biomass. Biotechnology for Biofuels. 2016;9:1-16. DOI: 10.1186/s13068-016-0569-z
  195. 195. Yu L-X, Gray BN, Rutzke CJ, Walker LP, Wilson DB, Hanson MR. Expression of thermostable microbial cellulases in the chloroplasts of nicotine-free tobacco. Journal of Biotechnology. 2007;131(3):362-369. DOI: 10.1016/j.jbiotec.2007.07.942
  196. 196. Jin S, Kanagaraj A, Verma D, Lange T, Daniell H. Release of hormones from conjugates: Chloroplast expression of β-glucosidase results in elevated phytohormone levels associated with significant increase in biomass and protection from aphids or whiteflies conferred by sucrose esters. Plant Physiology. 2011;155(1):222-235. DOI: 10.1104/pp.110.160754
  197. 197. Petersen K, Bock R. High-level expression of a suite of thermostable cell wall-degrading enzymes from the chloroplast genome. Plant Molecular Biology. 2011;76:311-321
  198. 198. Gray BN, Yang H, Ahner BA, Hanson MR. An efficient downstream box fusion allows high-level accumulation of active bacterial beta-glucosidase in tobacco chloroplasts. Plant molecular biology. 2011;76:345-355. DOI: 10.1007/s11103-011-9743-7
  199. 199. Nakahira Y, Ishikawa K, Tanaka K, Tozawa Y, Shiina T. Overproduction of hyperthermostable β-1, 4-endoglucanase from the archaeon Pyrococcus horikoshii by tobacco chloroplast engineering. Bioscience, Biotechnology, and Biochemistry. 2013;77(10):2140-2143. DOI: 10.1271/bbb.130413
  200. 200. Longoni P, Leelavathi S, Doria E, Reddy VS, Cella R. Production by tobacco transplastomic plants of recombinant fungal and bacterial cell-wall degrading enzymes to be used for cellulosic biomass saccharification. BioMed Research International. 2015;2015:289759. DOI: 10.1155/2015/289759
  201. 201. Martinez D, Berka RM, Henrissat B, Saloheimo M, Arvas M, Baker SE, et al. Genome sequencing and analysis of the biomass-degrading fungus Trichoderma reesei (syn. Hypocrea jecorina). Nature Biotechnology. 2008;26(5):553-560. DOI: 10.1038/nbt1403
  202. 202. Beguin P, Aubert JP. The biological degradation of cellulose. FEMS Microbiology Reviews. 1994;13(1):25-58. DOI: 10.1016/0168-6445(94)90099-X
  203. 203. Wilson DB. Cellulases and biofuels. Current Opinion in Biotechnology. 2009;20(3):295-299. DOI: 10.1016/j.copbio.2009.05.007
  204. 204. Liu G, Zhang J, Bao J. Cost evaluation of cellulase enzyme for industrial-scale cellulosic ethanol production based on rigorous Aspen plus modeling. Bioprocess and Biosystems Engineering. 2016;39(1):133-140. DOI: 10.1007/s00449-015-1497-1
  205. 205. Gaobotse G, Venkataraman S, Mmereke KM, Moustafa K, Hefferon K, Makhzoum A. Recent Progress on vaccines produced in transgenic plants. Vaccine. 2022;10(11):1861. DOI: 10.3390/vaccines10111861
  206. 206. Kwon K-C, Daniell H. Low-cost oral delivery of protein drugs bioencapsulated in plant cells. Plant Biotechnology Journal. 2015;13(8):1017-1022. DOI: 10.1111/pbi.12462
  207. 207. Walsh G. Biopharmaceutical benchmarks 2014. Nature Biotechnology. 2014;32(10):992-1000. DOI: 10.1038/nbt.3040
  208. 208. BIOPHARMA. Biopharmaceutical Products in the U.S. and European Markets [Internet]. 2023. Available from: http://www.biopharma.com/approvals.html. [Accesed: April 12, 2023]
  209. 209. Kumar AU, Ling APK. Gene introduction approaches in chloroplast transformation and its applications. Journal of Genetic Engineering and Biotechnology. 2021;19(1):1-10. DOI: 10.1186/s43141-021-00255-7
  210. 210. Fernández-San Millán A, Mingo-Castel A, Miller M, Daniell H. A chloroplast transgenic approach to hyper-express and purify Human serum albumin, a protein highly susceptible to proteolytic degradation. Plant Biotechnology Journal. 2003;1(2):71-79. DOI: 10.1046/j.1467-7652.2003.00008.x
  211. 211. Shil PK, Kwon K-C, Zhu P, Verma A, Daniell H, Li Q. Oral delivery of ACE2/Ang-(1-7) bioencapsulated in plant cells protects against experimental uveitis and autoimmune uveoretinitis. Molecular Therapy. 2014;22(12):2069-2082. DOI: 10.1038/mt.2014.179
  212. 212. Daniell H, Lee S-B, Panchal T, Wiebe PO. Expression of the native cholera toxin B subunit gene and assembly as functional oligomers in transgenic tobacco chloroplasts. Journal of Molecular Biology. 2001;311(5):1001-1009. DOI: 10.1006/jmbi.2001.4921
  213. 213. Rigano MM, Manna C, Giulini A, Pedrazzini E, Capobianchi M, Castilletti C, et al. Transgenic chloroplasts are efficient sites for high-yield production of the vaccinia virus envelope protein A27L in plant cells. Plant Biotechnology Journal. 2009;7(6):577-591. DOI: 10.1111/j.1467-7652.2009.00425.x
  214. 214. Lakshmi PS, Verma D, Yang X, Lloyd B, Daniell H. Low cost tuberculosis vaccine antigens in capsules: Expression in chloroplasts, bio-encapsulation, stability and functional evaluation in vitro. PLoS One. 2013;8(1):e54708. DOI: 10.1371/journal.pone.0054708
  215. 215. Saba K, Gottschamel J, Younus I, Syed T, Gull K, Lössl AG, et al. Chloroplast-based inducible expression of ESAT-6 antigen for development of a plant-based vaccine against tuberculosis. Journal of Biotechnology. 2019;305:1-10. DOI: 10.1016/j.jbiotec.2019.08.016
  216. 216. Sherman A, Su J, Lin S, Wang X, Herzog RW, Daniell H. Suppression of inhibitor formation against FVIII in a murine model of hemophilia a by oral delivery of antigens bioencapsulated in plant cells. Blood, The Journal of the American Society of Hematology. 2014;124(10):1659-1668. DOI: 10.1182/blood-2013-10-528737
  217. 217. Morgenfeld M, Lentz E, Segretin ME, Alfano EF, Bravo-Almonacid F. Translational fusion and redirection to thylakoid lumen as strategies to enhance accumulation of Human papillomavirus E7 antigen in tobacco chloroplasts. Molecular Biotechnology. 2014;56(11):1021-1031. DOI: 10.1007/s12033-014-9781-x
  218. 218. Gottschamel J, Lössl A, Ruf S, Wang Y, Skaugen M, Bock R, et al. Production of dengue virus envelope protein domain III-based antigens in tobacco chloroplasts using inducible and constitutive expression systems. Plant Molecular Biology. 2016;91:497-512. DOI: 10.1007/s11103-016-0484-5
  219. 219. Rosales-Mendoza S, Monreal-Escalante E, González-Ortega O, Hernández M, Fragoso G, Garate T, et al. Transplastomic plants yield a multicomponent vaccine against cysticercosis. Journal of Biotechnology. 2018;266:124-132. DOI: 10.1016/j.jbiotec.2017.12.012
  220. 220. Hoelscher M, Tiller N, Teh AY-H, Wu G-Z, Ma JK, Bock R. High-level expression of the HIV entry inhibitor griffithsin from the plastid genome and retention of biological activity in dried tobacco leaves. Plant Molecular Biology. 2018;97:357-370. DOI: 10.1007/s11103-018-0744-7
  221. 221. Tien N-Q-D, Huy N-X, Kim M-Y. Improved expression of porcine epidemic diarrhea antigen by fusion with cholera toxin B subunit and chloroplast transformation in Nicotiana tabacum. Plant Cell, Tissue and Organ Culture (PCTOC). 2019;137:213-223. DOI: 10.1007/s11240-019-01562-1
  222. 222. Davoodi-Semiromi A, Samson N, Daniell H. The green vaccine: A global strategy to combat infectious and autoimmune diseases. Human Vaccines. 2009;5(7):488-493. DOI: 10.4161/hv.8247
  223. 223. Chan HT, Daniell H. Plant-made oral vaccines against human infectious diseases-are we there yet? Plant Biotechnol Journal. 2015;13(8):1056-1070. DOI: 10.1111/pbi.12471
  224. 224. Su J, Sherman A, Doerfler PA, Byrne BJ, Herzog RW, Daniell H. Oral delivery of acid alpha glucosidase epitopes expressed in plant chloroplasts suppresses antibody formation in treatment of Pompe mice. Plant Biotechnology Journal. 2015;13(8):1023-1032. DOI: 10.1111/pbi.12413

Written By

Brenda Julian Chávez, Stephanie Solano Ornelas, Quintín Rascón Cruz, Carmen Daniela González Barriga, Sigifredo Arévalo Gallegos, Blanca Flor Iglesias Figueroa, Luis Ignacio Siañez Estrada, Tania Siqueiros Cendón, Sugey Ramona Sinagawa García and Edward Alexander Espinoza Sánchez

Submitted: 28 April 2023 Reviewed: 11 May 2023 Published: 30 May 2023