Open access peer-reviewed chapter

Recent Biotechnological Advances in the Improvement of Cassava

Written By

Vincent N. Fondong and Chrissie Rey

Submitted: 10 May 2017 Reviewed: 30 August 2017 Published: 17 January 2018

DOI: 10.5772/intechopen.70758

From the Edited Volume

Cassava

Edited by Viduranga Waisundara

Chapter metrics overview

2,011 Chapter Downloads

View Full Metrics

Abstract

Cassava (Manihot esculenta Crantz) is the fourth most important source of carbohydrates for human consumption in the tropics and thus occupies a uniquely important position as a food security crop for smallholder farmers. Consequently, cassava improvement is of high priority to most national agricultural research institutions in the tropics. With advances in functional genomics and genome editing approaches in this post genomics era, there are unprecedented opportunities and potential to accelerate the improvement of this important crop. These new technologies will need to be directed toward addressing major cassava production constraints, notably virus resistance, protein content, tolerance to drought and reduction of hydrogen cyanide content. Here, we discuss the important role novel functional genomics and genome editing technologies have and will continue to play in cassava improvement efforts. These approaches, including artificial miRNA (amiRNA), trans-acting small interfering RNA (tasiRNA), clustered regularly interspaced short palindromic repeat (CRISPR)-associated protein 9 (Cas9), and Targeting Induced Local Lesions IN Genomes (TILLING), have been shown to be effective in addressing major crop production constraints. In addition to reviewing specific applications of these technologies in cassava improvement, this chapter discusses specific examples being deployed in the amelioration of cassava or of other crops that could be applied to cassava in future.

Keywords

  • AmiRNA
  • cassava
  • CRISPR
  • genetic engineering
  • tasiRNA
  • TILLING

1. Introduction

Cassava, Manihot esculenta Crantz, was transported to Africa by the Portuguese in the sixteenth century, and was initially grown in and around trading posts in the Gulf of Guinea in West Africa; it was subsequently introduced into East Africa from Madagascar in the later part of the eighteenth century [1]. Today, cassava is a staple food to an estimated 800 million people worldwide [2] and is grown almost exclusively by smallholder farmers (Figure 1) and in isolated areas where soils are poor and rainfall is low or unpredictable. Additional attributes of this crop include low-cost and readily planting material, tolerance to acid soils, forms a symbiotic association with soil fungi to help its roots absorb phosphorus and micronutrients. Thus, cassava production requires very low input and gives reasonable harvests where other crops would fail [2]. Cassava is also increasingly being adopted as a source of family income following the fall of coffee and cocoa in the world market. Consequently, improvement of this crop is of high priority to most national agricultural research institutions in Africa. Moreover, the recognition that cassava industrial starch-based products, especially in renewable energy, could enhance food security and livelihoods, makes this crop a potentially valuable source of economic growth on the African continent.

Figure 1.

Africa produces more cassava than any other crop. (A) Cassava is a woody shrub that grows well in marginal lands; (B) a family in Southern Cameroons transporting cassava tuberous root harvest; (C) cassava tuberous roots are processed into many food types, here a family in Southern Cameroons preparing “garri,” a flour produced from cassava tuberous roots.

Cassava is cultivated principally for its tuberous roots, which are a good source of energy; additionally, in some parts of central Africa, leaves are also consumed as a source of protein, vitamins, and minerals. Cassava roots and leaves are deficient in sulfur-containing amino acids (methionine and cysteine) and some nutrients are not optimally distributed within the plant [3], leading to a deficiency in protein content, especially in roots. Cassava also contains antinutrients, most notably cyanogens, that can interfere with nutrient absorption and utilization and may have toxic side effects [4]. There are efforts to add nutritional value to cassava (biofortification) by increasing the contents of protein, minerals, starch, and β-carotene through biotechnological approaches [3]. Thus, nutritional content and production of cassava will benefit greatly from advances in genomics and biotechnological approaches.

Cassava improvement, either through conventional breeding or through genetic engineering, is challenging. The most reliable regeneration system cassava so far is through somatic embryogenesis (Figure 2) [5]. In the case of conventional breeding, which so far is the most routinely used approach to improve this crop is challenging due to several factors associated with several factors, include: (1) Lack of useful genes in the core cassava germplasm collections; (2) Heterozygosity and allopolyploidy of the cassava genomes; (3) Irregular flowering; and (4) Low fertility, seed set, and germination rates. As for genetic engineering and gene transfer, over the past few decades, this approach has been used to complement conventional breeding [6]. Undoubtedly, advances in modern technologies such as transcriptomics, proteomics, and metabolomics are likely to benefit breeding and genetic engineering strategies from an understanding of plant metabolic pathways and the role of key genes associated with their regulation. In this chapter, we identify some of the most important nutritional characteristics of cassava and production constraints that can benefit from advances in genome editing and functional genomics approaches are discussed in this section.

Figure 2.

Regeneration of cassava cultivars from Cameroon [5]. Callus with proembryogenic masses (A); clusters of organized embryogenic structures consisting of globular, heart and torpedo structures, early cotyledonary stage, asynchronous development of somatic embryos (B); organogenic callus with green cotyledons developed clusters of shoot buds (C); shoot buds rooted and developed into whole plantlets in vitro (D).

Advertisement

2. Important characteristics of cassava requiring improvement

2.1. Protein content

Cassava tuberous roots have relatively low protein content, which on the average ranges from 2 to 3% dry weight [7], compared with 9–11% for maize grain [8]. Indeed, a 500-g cassava meal provides only 30% of the daily protein requirement. Added to the low protein content, is the fact that roots are processed and the processed product is essentially protein-free. Consequently, individuals consuming exclusively or predominantly cassava usually suffer from protein-deficiency symptoms [9]. There is evidence suggesting that protein content in the roots can be considerably higher (6–8%) in some landraces [7] and such an important attribute can be introgressed into cassava through classical breeding. However, as indicated above, cassava breeding is rather challenging. Thus, fortification via genetic engineering is a more feasible option in efforts aimed at improving cassava protein content. For example, cyanide derived from linamarin is a major cause of reduced nitrogen for cassava root protein synthesis, thus disruption of linamarin transport from leaves to the roots through gene silencing-mediated inhibition of the two cytochrome P450s genes, CYP79D1/D2, resulted in an increase in nitrogen levels in cassava roots and higher levels of root protein content [10]. There have also been attempts to increase protein root content by producing transgenic cassava expressing genes that enhance protein root accumulation, including an artificial storage protein gene, ASP1 [11]. Thus, advances in genomics and transcriptomics will undoubtedly identify genes and pathways that will provide new opportunities to increase protein content using genetic engineering approaches.

2.2. Hydrogen cyanide content

Consumption of residual cyanogens (linamarin and lotaustralin) in incompletely processed cassava roots can cause various health disorders that render a person unsteady and uncoordinated [12]. Hydroxynitrile lyase (HNL) catalyzes the conversion of acetone cyanohydrin to cyanide and is expressed predominantly in the cell walls and laticifers of leaves, compared with tuberous roots, which exhibit very low [10]. Transgenic cassava over-expressing HNL was shown to display significantly reduced acetone cyanohydrin levels and exhibited increased cyanide volatilization in processed or homogenized roots [12]. It has been shown that the genomic region surrounding the cytochrome P450, CYP79D3, contains all genes required for cyanogenic glucoside biosynthesis in cassava [13]. As indicated above, this provides an additional opportunity to reduce cyanogen content in cassava by for example tissue specific suppression of two P450 genes, CYP79D1/D2, that catalyze the first-dedicated step in cyanogen synthesis [14]. Thus, at the molecular level, cyanogen detoxification can either be achieved by gene overexpression or through gene suppression, either of which can be achieved through genome editing techniques.

2.3. Starch quality

Due to its high starch content, cassava provides a source of dietary carbohydrate to an estimated 800 million people worldwide [2]. Insight in cassava development and starch biosynthesis is necessary to improve cassava starch quality and quantity [15]. Isolation and characterization of cassava gene homologs implicated in processes affecting the conversion of assimilated carbon to sucrose in photosynthetic cells, the phloem transport of sucrose to storage organs, the transition of sucrose to starch, and the degradation of starch into simple sugars, could be exploited to improve starch quality. Molecular and functional characterization of the genes involved in these processes will greatly enhance cassava varietal improvement by altering the gene activities either via genetic manipulation or through gene editing. Also, application of advanced systemic-based computational techniques to understand the physiological regulation and control of starch metabolism in plant plastids would be the basis for understanding these processes in cassava.

Cassava is also a good source of industrial starch and bioethanol [16, 17]; in both situations, the quality of starch is important. Starch consists of two glucan polymers: amylose and amylopectin. Amylopectin is extremely soluble in water whereas amylose has a strong tendency to recrystallize after dispersion in water, a property referred to as retrogradation. Retrogradation is undesirable for many applications of starch in which a defined and stable viscosity is required. Therefore, for industrial purposes, starch is often treated with chemicals in order to make the amylose less sensitive to crystallization [18]. As retrogradation is caused mainly by the amylose fraction in starch, amylose-free starches do not have to be treated with chemicals [19]. There are therefore efforts to generate amylose-free cassava through genetic engineering; for example, starch-free cassava was obtained by silencing GBSSI, the granule-bound starch synthase gene, which is required for the synthesis of amylose [20].

2.4. Postharvest physiological deterioration and storage

Harvested cassava tuberous roots undergo rapid postharvest physiological deterioration (PPD) [21, 22]. PPD is initiated by mechanical damage, which typically occurs during tuberous roots harvesting and progresses from the proximal site of damage to the distal end, making the roots unpalatable within 72 h [22, 23]. Reactive oxygen species (ROS) production has been identified as one of the earliest events in PPD [21, 22, 24]. Under conditions of stress, the equilibrium between the production and scavenging of ROS is disturbed, resulting in a rapid increase in the buildup of ROS known as an oxidative burst [25]. In cassava roots, an oxidative burst occurs within 15 min of harvest [21], resulting therefore in an early PPD. Other early events that result in rapid PPD include, increased activity of enzymes that modulate ROS levels, such as catalase, peroxidase, and superoxide dismutase [22]. Further evidence in support of a role of oxidative stress in PPD comes from the observation that cassava cultivars that have high levels of b-carotene (which quenches ROS) are less susceptible to PPD [26]. Reduction of ROS and PPD was also shown to be induced by cyanogenesis, suggesting that a possible solution to cassava PPD is to reduce the cyanide-induced accumulation of ROS.

2.5. Cassava pathogens

Cassava viruses constitute a major challenge to cassava production; of particular importance are cassava mosaic geminiviruses (CMGs) (Family, Geminiviridae: Genus, Begomovirus), which cause the cassava mosaic disease (CMD) in all cassava growing regions of Africa and the Indian subcontinent. CMGs are transmitted by the whitefly vector, Bemisia tabaci (Gennadius) and through cuttings used routinely for vegetative propagation. Tuberous root losses due to CMD range from 20 to 100% [27]. With the emergence of new molecular and sequencing capabilities, CMGs have been shown to exhibit considerable sequence and biological differences and so far, 11 species have been described in the cassava growing regions of African and the Indian subcontinent [28] and some of these viruses co-infect the same plant resulting in a synergistic interaction, characterized by severe symptoms (Figure 3). Interestingly, cassava was introduced in Africa from South America [29], yet CMGs are not found in South America and therefore these viruses are likely recent descendants of geminiviruses adapted to indigenous uncultivated African plant species [30]. The problem of CMGs has been compounded by the emergence, in eastern Africa, of cassava brown streak disease (CBSD), which is caused by cassava brown streak viruses (CBSVs (Family, Potyviridae: Genus, Ipomovirus). Like CMGs, CBSVs are transmitted by the whitefly vector [31] and through infected stem propagules. For a long time, CBSD was considered to be limited to lowland coastal regions of Tanzania, and to a limited extent in lowland areas of Uganda [32], northwestern Tanzania, southern Uganda [32, 33, 34]. Since 2004, however, the CBSD epidemic has spread around the Great Lakes Region to affect eastern Uganda, western Kenya, the Lake Zone of Tanzania, Rwanda, Burundi and the DRC [35, 36]. The most damaging symptoms of CBSD are found in tuberous roots, including brown, corky necrosis of the starchy tissue, occasional radial constrictions and a reduction in the content of starch and cyanide [32, 34]. Yield losses are estimated to be up to 70% in highly susceptible cultivars [37]. In additional to viral pathogens, is cassava bacterial blight (CBB), caused by Xanthomonas axonopodis pv. manihotis (Xam). CBB is considered to be one of the most relevant plant pathogenic bacteria because of the yield losses, estimated to be 70%, it causes in cassava [38, 39].

Figure 3.

Cassava plant mixed infected by two cassava mosaic geminiviruses displaying severe mosaic and leaf distortion and size reduction, resulting in plant stunting.

Use of resistant varieties has been the most effective in controlling CMD in Africa thanks to the discovery in the 1930s that some of the cassava varieties being grown were less affected by CMD than others. Thus, resistance breeding began in Ghana, Madagascar, Tanzania and elsewhere in Africa [40, 41]. In the last 2 decades, use of genetic engineering to produce virus resistant cassava has gained considerable attention, especially with the discovery of RNA interference pathways [42].

2.6. Tolerance to drought

In most cassava growing regions of the world, the cassava growth cycle is typically interrupted by months of drought, influencing various plant physiological processes and resulting in depressed growth, development and yield [43, 44]. Although cassava is a drought tolerant crop, there is a range of drought-tolerance levels in available germplasm. Thus, growth and productivity of genotypes with a low threshold of drought tolerance in marginal areas are constrained by severe drought stress, especially during the earlier stages of growth [45]. Indeed, experimental data suggest that root production is positively correlated with the life span of individual leaves [46] and increased leaf retention was found to increase root yield under irrigated and stressed conditions [47]. With continuous advances in genome science, there will be opportunities to enhance drought tolerance in the cassava crop. For example, Zhang et al. [46] have shown that transgenic cassava expressing isopentenyltransferase (IPT) gene under the control of senescence-activated promoter (SAG12), delayed leaf senescence under both greenhouse and field conditions, leading an increase in drought resistance. Also, identification of miRNA gene targets involved in post-transcriptional abiotic stress regulation could prove useful in engineering cassava for drought resistance [48].

Advertisement

3. RNA-based functional genomics technologies in cassava improvement

3.1. Hairpin dsRNA, co-suppression and antisense RNA silencing

The hairpin dsRNA (hpRNA), anti-sense silencing and co-suppression strategies have been extensively employed in crop improvement [49, 50, 51]. In cassava improvement, hpRNA and antisense silencing procedures have been employed mostly in virus control [52, 53, 54, 55, 56]. An indirect approach where the hpRNA is used to knockdown the expression of V-ATPase A, an enzyme that provides force for many transport processes, has been used to control whitefly vectors of CMGs and CBSVs [57, 58].

The antisense strategy has also been used to inactivate allergens and toxins in cassava, especially in the inhibition of hydrogen cyanide (HCN), which is the product of linamarase-mediated hydrolysis of linamarin. The presence of residual linamarin and its breakdown product (acetone cyanohydrin) in cassava-based food products has been a cause for concern because of their possible effects on human health. As discussed above, the first committed steps in linamarin biosynthesis is catalyzed by cytochrome P450 genes (CYP79D1 and CYP79D2) and therefore efforts have been made to knockdown CYP79D1 and CYP79D2 using hpRNA-mediated silencing so as to reduce HCN toxicity in cassava. Thus, transgenic cassava lines containing antisense copies of both genes exhibited almost complete absence of linamarin in tuberous roots [10]. Unfortunately, this approach could not be applied extensively as transgenic cassava lines exhibited poor tuberous root development.

In spite of the encouraging early results obtained from the use of hpRNA, co-suppression and antisense RNA silencing in crop improvement, this approaches have been tempered by several disadvantages associated with these approaches, these include poor stability of the transgene in transformed plants, dependence on the expression levels of the antisense strand, and limited penetration of the silencing signal to the appropriate target cells due to target-sequence folding (reviewed in Fondong et al. [59].

3.2. Small RNA (sRNA)-mediated silencing

The limitations of hpRNA, co-suppression and antisense RNA silencing strategies are, to a large extent overcome in sRNA strategies, including especially artificial microRNAs (amiRNAs) and trans-acting siRNA (tasiRNA). microRNAs constitute a well-studied class of sRNAs; their biogenesis starts with the transcription of long primary RNAs (pri-miRNAs) [60, 61]. miRNAs function in a homology-dependent manner against target mRNAs to typically either directly cleave at highly specific sites or to suppress translation. The amiRNA silencing technique exploits the biogenesis and function of endogenous miRNAs to silence genes in plants. In this approach, the endogenous miRNA-miRNA duplex in a native miRNA precursor is replaced with a customized sequence designed from the target gene. Upon processing, the amiRNA redirects the miRNA-induced silencing complex to silence the targeted mRNA, thereby generating a loss-of-function phenotype for the gene of interest [62, 63, 64, 65, 66]. The amiRNA strategy has especially been used in targeting plant viruses (reviewed in Fondong et al. [59]. However, there has been little application in cassava improvement. Indeed, to our knowledge, the only report of use of amiRNA in cassava improvement is the replacement of miR159 precursor with amiRNAs from cassava brown streak viruses in miR159 precursor; transgenic Nicotiana benthamiana lines thus produced were virus resistant [67].

It is important to note that the amiRNA platform has several advantages over the hpRNA strategy, including the fact that amiRNAs are small and thus have a reduced likelihood of off-targeting and the approach can easily be multiplexed via use of polycistronic miRNA backbone. In addition, processing of miRNA is not affected by changes in temperature compared with hpRNA-derived siRNAs whose levels decrease at low temperatures [68]. Thus, it is likely that this platform will prove useful in studying cassava gene function. A major limitation of the amiRNA strategy is that the small size of the amiRNA (21nt) increases opportunities for loss of complementarity between the amiRNA and the target gene, and genes from the same family with variations may not be silenced using a single amiRNA. To reduce these risks, a multimeric amiRNA approach in which multiple amiRNAs targeting different conserved regions of the gene can be adopted as has been reported in plant virus control [69, 70].

A second class of sRNAs used in crop improvement is tasiRNAs, which are produced from noncoding TAS genes, which have been identified in all examined land plants. TAS genes differ from most other genes in that they do not code for a protein, but rather produce long non-coding RNA transcripts, which are subsequently processed into 21nt tasiRNAs. Synthesis of tasiRNA is initiated by miRNA-directed and Argonaute (AGO) protein-mediated cleavage of TAS transcripts, of which four (TAS1, 2, 3, 4) have been extensively studied in Arabidopsis (see reviews Allen and Howell [71] and Yoshikawa [72]. Two models of tasiRNA biogenesis, referred to as “one-hit” and “two-hit”, have been described in Arabidopsis [73]. The tasiRNA strategy is very efficient, highly predictable in processing siRNAs and can easily be multiplexed to target multiple genes, especially genes from the same family; yet it remains an underutilized strategy in plant improvement. It has been used to successfully engineer resistance to plant viruses [74, 75]. In cassava virus control, transgenic N. benthamiana containing Arabidopsis TAS1a gene modified with tasiRNA from the cassava geminivirus, East African cassava mosaic Cameroon virus (EACMCV) exhibits strong resistance to the virus (Fondong et al., unpublished). Fifty-four tasiRNAs and fifteen possible cis-nat- siRNAs were identified in cassava infected with cassava bacterial blight, and many of these loci were induced or repressed in response to Xam infection [76]. A similar transgenic strategy using a TAS gene modified with tasiRNAs from Xam could be promising. This finding emphasizes the potential potency of this strategy in plant virus control.

Advertisement

4. Role of reverse genetics and gene editing techniques in cassava improvement

4.1. TILLING and EcoTILLING

TILLING is a non-recombinant reverse genetics approach used to identify novel sequence variation in genomes, with the aims of investigating gene function and/or developing useful alleles for breeding. TILLING involves induction of mutations in the plant genome using classical mutagenesis approaches followed by traditional or high throughput deep sequencing to identify the mutations in the gene of interest [77, 78, 79]. This technique has been used in allele discovery in different plant species [80, 81, 82, 83]. EcoTILLING, which is an adaptation of the TILLING, is used in detecting rare single nucleotide polymorphism (SNPs) or small INDELs in target genes in natural populations [84]. In EcoTILLING, mismatches formed by hybridization of different genotypes in a test panel are cleaved with CEL I, which is a mismatch-specific endonuclease from celery. A valuable application of EcoTILLING in plants is in the search for variation in disease resistance genes. There are only a few reports of the use of TILLING or EcoTILLING in cassava improvement. Of these, is the recent report of irradiation of seeds of elite cassava lines and wild Manihot species in an effort to broaden the genetic base of the germplasm pool so as to expand the industrial uses of cassava [85]. The study led to the discovery of small granules, which are abnormal amylose starch molecules resulting from a mutation. These small granules are ideal for industrial ethanol production due to the fact that they facilitate the activity of starch-degrading enzymes [86]. Because of the promise of the technique in cassava improvement, the International Institute of Tropical Agriculture (IITA) in Ibadan, Nigeria, is developing a TILLING protocol for discovery of important cassava traits [87].

TILLING has several advantages over other crop improvement techniques: (1) it produces a spectrum of allelic mutations that are useful for genetic analysis, (2) it is applicable to any organism, (3) mutations that are difficult to be detected by forward genetics can be revealed via TILLING since it can focus at on the gene of interest, and (4) it is a non-transgenic method, hence there are no biosafety or environmental concerns [88]. The main disadvantages of TILLING are the requirement of locus-specific polymerase chain reaction (PCR) products (difficult for gene families with very similar sequences and in polyploids) and the inability to detect mutations near simple sequence repeats (SSRs) (because of the flare caused by polymerase slippage-induced deletions) [89].

4.2. Clustered regularly interspaced short palindromic repeat (CRISPR)

As indicated above, cassava transformation and crossing are challenging and thus gene editing is potentially a method that can be used to improve the crop. The clustered regularly interspaced short palindromic repeats (CRISPRs) and associated protein (Cas) approach has recently gained wide application in gene editing. In bacteria and especially archaea, CRISPRs/Cas is a nucleic acid-based adaptive immune system, which confers molecular immunity to foreign nucleic acids, including plasmids and viruses (see review Barrangou [90]. CRISPR genomic loci consist of repeat sequences, typically 20–50 bp in length, separated by variable spacer sequences (or protospacers) of similar length that match a segment of invading nucleic acids. These protospacers serve as a molecular memory of prior infections and together with repeat sequences, constitute CRISPR RNAs in the CRISPR locus [90, 91]. CRISPR RNAs are used as guides by Cas proteins for base-pairing with and degradation of complementary sequences in invading DNAs [90, 91]. The CRISPR/Cas system is functional in eukaryotic systems, for which the Streptococcus pyogenes endonuclease Cas9 (Cas9) has been harnessed for efficient eukaryotic genome editing and gene regulation [92, 93]. The ease of deployment of the CRISPR/Cas9 system is due to its dependence on RNA as the moiety that directs the Cas9 nuclease to a desired DNA sequence [94, 95].

The functionality of CRISPR/Cas9 system in eukaryotes has revolutionized genome editing and in a very short time since its discovery, has become a very useful tool in crop improvement. Successful examples have been reported for several crops with complex genomes (reviewed in Paul and Qi [96]). However, only a few reports of use of CRISPR/Cas9 system in cassava improvement exist and are still in the preliminary stage, these include CRISPR/Cas9-mediated modification of cassava flowering genes to induce flowering in this predominantly clonally propagated crop [97]. Because of the successful development of a modified geminivirus vector based on Cabbage leaf curl virus for a virus-guided delivery of CRISPR/Cas9 [98], it is likely that a similar vector system can be developed for cassava using the cassava geminivirus, African cassava mosaic virus.

There are drawbacks of the CRISPR/Cas system, including: (1) imbalance in stoichiometry between Cas9 and sgRNA ratio that may lead to off-target cleavage [99, 100]. (2) Many protospacer adjacent motif (PAM) sites may lead to undesired cleavage of DNA regions [101]; to resolve this problem, bioinformatics tools are being developed at whole genome sequence level to improve specificity [102]. (3) Codon usage varies across species and may affect Cas9 translation; several codon-optimized versions of Cas9 genes have therefore been harnessed for several individual crops [102] and there may be need for a cassava codon optimized Cas9. (4) CRISPR/Cas9 systems use exogenous promoters for Cas9 and sgRNA expression; for cassava, Cassava vein mosaic virus promoter [103] has been shown to be very efficient. (5) Homology between gene family members may complicate sequence targeting and directing sgRNAs to the 5′ region of the targeted gene has been proposed improve target specificity [102].

Advertisement

5. Natural host resistance to pathogen in cassava

5.1. Natural pathogen resistance and cassava virus control

It is now clear that the cassava geminiviruses and cassava brown streak viruses are the most important constraints to cassava production in the African [104]. Correspondingly, in Latin America, a diverse set of virus species that cause the cassava frog skin disease syndrome has a serious impact on cassava production [105]. Thus, considerable effort will be required to expand sources of resistance to cassava viral diseases and advances in genomics have provided new opportunities to explore sources of natural resistance. An important source of resistance that may be useful in cassava is non-host resistance. Mechanistically, non-host resistance is likely due to an intrinsic lack of susceptibility, which is a multigenic trait. It is now known that natural compounds, such as melatonin that modulate immune responses, such as ROS metabolism, calcium signaling and mitogen-activated protein kinase (MAPK) cascades, can be used to enhance natural resistance [106]. Notably, the recent identification and functional analysis of melatonin synthesis genes in cassava has provided a direct link between melatonin and immune responses [107]. Furthermore, the importance of resistance targets that function as host susceptibility factors, such as translation initiation factors 4E and 4G in RNA viruses, have been studied in model systems and can potentially be exploited for CBSV resistance in cassava [108].

Viruses that successfully infect the host induce changes in host cells by manipulating the host molecular pathways and host responses can provide clues for functional manipulation of resistance traits. It has been shown that CMGs [109] and CBSVs [110] induce global transcriptome reprogramming of cassava. In the case of CBSD, of the 700 overexpressed genes in a resistant cassava variety, none of the genes was identified as a resistance gene, instead most belonged to hormone signaling and metabolic pathway gene classes [110]. Interestingly, three functional genomic studies with South African cassava mosaic virus in three hosts, Arabidopsis [111], cassava [109] and N. benthamiana [112] revealed a small number of common differentially expressed genes at the early infection stage of full systemic symptoms. However, a common theme in all three hosts was virus-induced changes in hormone signaling, and primary and secondary metabolisms. Understanding the roles of host reprogramming and RNA silencing during cassava-virus interactions could be exploited to improve natural immunity in cassava.

5.2. Identification of cassava immunity-related or resistance (R) genes

Dominant and recessive genes have been associated with natural plant virus resistance [113]. Using a combination of genotype-by-sequencing (GBS)-based SNPs and physical mapping of scaffolds from cassava whole genome sequencing (WGS), 1061 cassava immunity-related genes were mapped [114]. Notably, from 105 putative CMD2 genes identified from the CMD2 locus on chromosome 8 [115], 35 were identical to those identified in a RNA-seq study of SACMV-infected cassava genotype TME3 [109]. These genes could be strong candidates contributing to resistance in cassava. Proteins encoded by R gene usually occur as large families of proteins with nucleotide binding-leucine rich repeat (NLR) domains and function as indirect perception sensors of pathogen avirulence (avr) proteins. The determinants of apparent virus R gene-wide specificity lies in the leucine-rich repeat (LRR) domains and sequencing of wild cassava varieties may provide a source for discovery of new cassava virus resistance genes. Recently, 228 NLR and 99 partial NBS genes were mapped to the cassava reference genome (http://phytozome.jgi.doe.gov) and these genes show high sequence similarity to genes found in other plant species [116]. However, involvement of these genes in CMD or CBSD resistance is not known. Furthermore, microRNAs are master regulators that trigger processing of genes coding for NLR into phased small interfering RNAs (phasiRNAs) [117] and are therefore regulators of genes that are the first line of defense. Unveiling the role of miRNAs in cassava virus resistance would provide new tools in the combat against these viruses.

Based on a holistic approach, combining high-throughput transcriptome sequence data, public genomic data from cassava and Arabidopsis, Leal et al. [118] identified predicted immunity related gene (IRG) pathways, which showed that several cellular pathways are strongly related to immune response pathways. We will need to exploit these genomics data to identify evolutionarily diverse resistance or immunity genes in different cassava genotypes for development of durable resistance to cassava viruses.

5.3. Resistance against whitefly, B. tabaci (Gennadius)

Another approach to generate resistance to CMGs and CBSVs is through use of functional genomics to control the whitefly (B. tabaci), vector of both virus groups. Until recently, little was known about the molecular mechanisms of insect defense. Development of B. tabaci type B on Arabidopsis was shown to rely on the concomitant increase of salicylic acid and decline or unchanged levels of jasmonic acid and ethylene defense pathways [119]. Transgenic mediated overexpression or down-regulation of genes involved in lignin or other defenses against insect pests could be exploited to develop insect resistant cassava [120]. Application of functional genomics in insect resistance was recently elucidated by expressing an insecticidal fern protein in cotton, which exhibited resistance to whitefly [121]. Efforts in editing genes that play a role in whitefly resistance in cassava will thus play a role in developing cassava with resistance to the whitefly, vector of many cassava viruses.

Advertisement

6. Conclusion

The future challenge in cassava is the ability to combine desirable traits with different agronomic requirements using molecular breeding, gene editing and RNAi technologies. This is critically important, given that cassava is fundamental to food security in many parts of the world. In this chapter, we have discussed advances in the improvement of this crop, especially with regards to nutrient quality and biotic as well as abiotic constraints. We have also proposed novel genome editing technologies that will likely address some of the challenges faced by this crop. These include technologies such as amiRNA, tasiRNA, TILLING/EcoTILLING and CRISPR-Cas9, which provide enormous potentials in cassava improvement. Also, the increasing reduction in the cost of high-throughput sequencing and lessons from ongoing and past work will continue to provide new insights into additional new genome-editing and functional genomic approaches for the improvement of the crop.

Advertisement

Acknowledgments

The disease resistance work in Vincent N. Fondong’s laboratory is supported by NSF-IOS-1212576.

References

  1. 1. Legg JP, Thresh JM. Cassava mosaic virus disease in East Africa: A dynamic disease in a changing environment. Virus Research. 2000;71(1-2):135-149
  2. 2. FAO. FAOSTAT Statistical Database 2013 Available from: http://faostat.fao.org [Accessed: August 2017]
  3. 3. Montagnac JA, Davis CR, Tanumihardjo SA. Nutritional value of cassava for use as a staple food and recent advances for improvement. Comprehensive Reviews in Food Science and Food Safety. 2009;8:181-194
  4. 4. Montagnac JA, Davis CR, Tanumihardjo SA. Processing techniques to reduce toxicity and antinutrients of cassava for use as a staple food. Comprehensive Reviews in Food Science and Food Safety. 2009;8:17-27
  5. 5. Mongomake K, Doungous O, Khatabi B, Fondong VN. Somatic embryogenesis and plant regeneration of cassava (Manihot esculenta Crantz) landraces from Cameroon. SpringerPlus. 2015;4:477. DOI: 10.1186/s40064-015-1272-4
  6. 6. Sanghera GS, Wani SH, Gill MS, Kashyap PL, Gosal SS. RNA interference: Its concept and application in crop plants. In: Malik CP, Verma A, editors. Biotechnology Cracking New Pastures. New Delhi: MD Publishers; 2010. p. 33-78
  7. 7. Ceballos H, Iglesias CA, Pérez JC, Dixon AG. Cassava breeding: Opportunities and challenges. Plant Molecular Biology. 2004;56(4):503-516
  8. 8. Ensminger ME, Oldfield JE, Heinemann WW. Feeds and Nutrition. 2nd ed. Clovis, CA: Ensminger Publishing Co; 1990 505 p
  9. 9. Sayre R, Beeching JR, Cahoon EB, Egesi C, Fauquet C, Fellman J, Fregene M, Gruissem W, Mallowa S, Manary M, et al. The BioCassava plus program: Biofortification of cassava for sub-Saharan Africa. Annual Review of Plant Biology. 2011;62:251-272. DOI: 10.1146/annurev-arplant-042110-103751
  10. 10. Siritunga D, Arias-Garzon D, White W, Sayre RT. Over-expression of hydroxynitrile lyase in transgenic cassava roots accelerates cyanogenesis and food detoxification. Plant Biotechnology Journal. 2004;2(1):37-43
  11. 11. Zhang P, Jaynes J, Potrykus I, Gruissem W. Transfer and expression of an artificial storage protein (ASP1) gene in cassava (Manihot esculenta Crantz). Transgenic Research. 2003;12(2):243-250
  12. 12. Narayanan NN, Ihemere U, Ellery C, Sayre RT. Overexpression of hydroxynitrile lyase in cassava roots elevates protein and free amino acids while reducing residual cyanogen levels. PLoS One. 2011;6(7):e21996. DOI: 10.1371/journal.pone.0021996
  13. 13. Takos AM, Knudsen C, Lai D, Kannangara R, Mikkelsen L, Motawia MS, Olsen CE, Sato S, Tabata S, Jørgensen K, Møller BL, Rook F. Genomic clustering of cyanogenic glucoside biosynthetic genes aids their identification in Lotus japonicus and suggests the repeated evolution of this chemical defence pathway. The Plant Journal. 2011;68(2):273-286. DOI: 10.1111/j.1365-313X.2011.04685.x
  14. 14. Andersen MD, Busk PK, Svendsen I, Møller BL. Cytochromes P-450 from cassava (Manihot esculenta Crantz) catalyzing the first steps in the biosynthesis of the cyanogenic glucosides linamarin and lotaustralin. Cloning, functional expression in Pichia pastoris, and substrate specificity of the isolated recombinant enzymes. The Journal of Biological Chemistry. 2000;275(3):1966-1975
  15. 15. Karlström A, Calle F, Salazar S, Morante N, Dufour D, Ceballos H. Biological implications in cassava for the production of amylose-free starch: Impact on root yield and related traits. Frontiers in Plant Science. 2016;20(7):604. DOI: 10.3389/fpls.2016.00604
  16. 16. Shetty JK, Chotani G, Duan G, Bates D. Cassava as an alternative feedstock in the production of transportation fuel. International Sugar Journal. 2007;109:3-11
  17. 17. Pandey A, Soccol CR, Nigam P, Soccol VT, Vanderberghe LPS, Mohan R. Biotechnological potential of agro-industrial residue. II: Cassava bagasse. Bioresource Technology. 2000;74:81-87
  18. 18. Bruinenberg PM, Jacobsen E, Visser RGF. Starch from genetically engineered crops. Chemistry and Industry. 1995;95:881-884
  19. 19. Koehorst-van Putten HJ, Sudarmonowati E, Herman M, Pereira-Bertram IJ, Wolters AM, Meima H, de Vetten N, Raemakers CJ, Visser RG. Field testing and exploitation of genetically modified cassava with low-amylose or amylose-free starch in Indonesia. Transgenic Research. 2012;21(1):39-50. DOI: 10.1007/s11248-011-9507-9
  20. 20. Raemakers K, Schreuder M, Suurs L, Furrer-Verhorst H, Vincken JP, de Vetten N, Jacobsen E, Visser RGF. Improved cassava starch by antisense inhibition of granule-bound starch synthase I. Molecular Breeding. 2005;16:163-172
  21. 21. Reilly K, Gómez-Vásquez R, Buschmann H, Tohme J, Beeching JR. Oxidative stress responses during cassava post-harvest physiological deterioration. Plant Molecular Biology. 2004;56(4):625-641
  22. 22. Iyer S, Mattinson DS, Fellman JK. Study of the early events leading to cassava root postharvest deterioration. Tropical Plant Biology. 2010;3:151-165
  23. 23. Buschmann H, Reilly K, Rodriguez MX, Tohme J, Beeching JR. Hydrogen peroxide and flavan-3-ols in storage roots of cassava (Manihot esculenta Crantz) during postharvest deterioration. Journal of Agricultural and Food Chemistry. 2000;48(11):5522-5529
  24. 24. Zidenga T, Leyva-Guerrero E, Moon H, Siritunga D, Sayre R. Extending cassava root shelf life via reduction of reactive oxygen species production. Plant Physiology. 2012;159(4):1396-1407. DOI: 10.1104/pp.112.200345
  25. 25. Apostol I, Heinstein PF, Low PS. Rapid stimulation of an oxidative burst during elicitation of cultured plant cells: Role in defense and signal transduction. Plant Physiology. 1989;90(1):109-116
  26. 26. Sanchez T, Dufour D, Moreno IX, Ceballos H. Comparison of pasting and gel stabilities of waxy and normal starches from potato, maize, and rice with those of a novel waxy cassava starch under thermal, chemical, and mechanical stress. Journal of Agricultural and Food Chemistry. 2010;58(8):5093-5099. DOI: 10.1021/jf1001606
  27. 27. Masinde EA, Ogendo JO, Maruthi MN, Hillocks R, Mulwa RMS, Arama PF. Occurrence and estimated losses caused by cassava viruses in Migori county, Kenya. African Journal of Agricultural Research. 2016;11:2064-2074
  28. 28. Patil BL, Fauquet CM. Cassava mosaic geminiviruses: Actual knowledge and perspectives. Molecular Plant Pathology. 2009;10(5):685-701. DOI: 10.1111/j.1364-3703.2009.00559.x
  29. 29. Olsen K, Schaal B. Evidence on the origin of cassava: Phylogeography of Manihot esculenta. Proceedings of the National Academy of Sciences of the United States of America. 1999;96(10):5586-5591
  30. 30. De Bruyn A, Villemot J, Lefeuvre P, Villar E, Hoareau M, Harimalala M, Abdoul-Karime AL, Abdou-Chakour C, Reynaud B, Harkins GW, et al. East African cassava mosaic-like viruses from Africa to Indian ocean islands: Molecular diversity, evolutionary history and geographical dissemination of a bipartite begomovirus. BMC Evolutionary Biology. 2012;12:228. DOI: 10.1186/1471-2148-12-228
  31. 31. Storey HH. 1939 Report of the Plant Pathologist. East African Agricultural Research Station; 1939. p. 9
  32. 32. Nichols RFW. The brown streak disease of cassava: Distribution, climatic effects and diagnostic symptoms. The East African Agricultural Journal. 1950;15:154-160
  33. 33. Storey HH. Virus diseases of East African plants. VI. A progress report on studies of the disease of cassava. The East African Agricultural Journal. 1936;2:34-39
  34. 34. Hillocks RJ, Jennings DL. Cassava brown streak disease: A review of present knowledge and research needs. International Journal of Pest Management. 2003;49:225-234
  35. 35. Patil BL, Ogwok E, Wagaba H, Mohammed IU, Yadav JS, Bagewadi B, Taylor NJ, Kreuze JF, Maruthi MN, Alicai T, Fauquet CM. RNAi-mediated resistance to diverse isolates belonging to two virus species involved in Cassava brown streak disease. Molecular Plant Pathology. 2011;12(1):31-41. DOI: 10.1111/j.1364-3703.2010.00650.x
  36. 36. Legg JP, Jeremiah SC, Obiero HM, Maruthi MN, Ndyetabula I, Okao-Okuja G, Bouwmeester H, Bigirimana S, Tata-Hangy W, et al. Comparing the regional epidemiology of the cassava mosaic and cassava brown streak virus pandemics in Africa. Virus Research. 2011;159(2):161-170. DOI: 10.1016/j.virusres.2011.04.018
  37. 37. Hillocks RJ, Raya MD, Mtunda K, Kiozia H. Effects of brown streak disease on yield and quality of cassava in Tanzania. Journal of Phytopathology. 2001;149:389-394
  38. 38. Mansfield J, Genin S, Magori S, Citovsky V, Sriariyanum M, Ronald P, Dow M, Verdier V, Beer SV, Machado MA, Toth I, Salmond G, Foster GD. Top 10 plant pathogenic bacteria in molecular plant pathology. Molecular Plant Pathology. 2012;13(6):614-629. DOI: 10.1111/j.1364-3703.2012.00804.x
  39. 39. Wydra K, Verdier V. Occurrence of cassava diseases in relation to environmental, agronomic and plant characteristics. Agriculture, Ecosystems and Environment. 2002;93:211-226
  40. 40. Thresh JM, Cooter RJ. Strategies for controlling cassava mosaic virus disease in Africa. Review article. Plant Pathology. 2005;54:587-614
  41. 41. Fondong VN. The search for resistance to cassava mosaic geminiviruses: How much we have accomplished, and what lies ahead. Frontiers in Plant Science. 2017;8:408. DOI: 10.3389/fpls.2017.00408
  42. 42. Vanderschuren H, Akbergenov R, Pooggin M, Hohn T, Gruissem W, Zhang P. Transgenic cassava resistance to African cassava mosaic virus is enhanced by viral DNA-A bidirectional promoterderived siRNAs. Plant Molecular Biology. 2007;64(5):549-557
  43. 43. Turyagyenda LF, Kizito EB, Ferguson M, Baguma Y, Agaba M, Harvey JJ, Osiru DS. Physiological and molecular characterization of drought responses and identification of candidate tolerance genes in cassava. AoB Plants. 2013;5:plt007. DOI: 10.1093/aobpla/plt007
  44. 44. Bakayoko S, Tschannen A, Nindjin C, Dao D, Girardin O, Assa A. Impact of water stress on fresh tuber yield and dry matter content of cassava (Manihot esculenta Crantz) in Côte d’Ivoire. African Journal of Agricultural Research. 2009;4(1):021-027
  45. 45. Perez JC, Lenis JI, Calle F, Morante N, Sanchez T, Debouck D, Ceballos H. Genetic variability of root peel thickness and its influence in extractable starch from cassava (Manihot esculenta Crantz) roots. Plant Breeding. 2011;130(6):688-693. DOI: 10.1111/j.1439-0523.2011.01873.x
  46. 46. Zhang P, Wang WQ, Zhang GL, Kaminek M, Dobrev P, Xu J, Gruissem W. Senescence-inducible expression of isopentenyl transferase extends leaf life, increases drought stress resistance and alters cytokinin metabolism in cassava. Journal of Integrative Plant Biology. 2010;52(7):653-669. DOI: 10.1111/j.1744-7909.2010.00956.x
  47. 47. Lenis JI, Calle F, Jaramillo G, Perez JC, Ceballos H, Cock JH. Leaf retention and cassava productivity. Field Crops Research. 2006;95:126-134
  48. 48. Ballén-Taborda C, Plata G, Ayling S, Rodríguez-Zapata F, Becerra Lopez-Lavalle LA, Duitama J, et al. Identification of cassava microRNAs under abiotic stress. International Journal of Genomics. 2013;53:257-269. DOI: 10.1155/2013/857986
  49. 49. Wesley SV, Helliwell CA, Smith NA, Wang MB, Rouse DT, Liu Q, Gooding PS, Singh SP, Abbott D, Stoutjesdijk PA, et al. Construct design for efficient, effective and high-throughput gene silencing in plants. The Plant Journal. 2001;27(6):581-590
  50. 50. Helliwell C, Waterhouse P. Constructs and methods of high-throughput gene silencing in plants. Methods. 2003;30(4):289-295
  51. 51. Serio FD, Schob H, Iglesias A, Tarina C, Bouldoires E, Meins FJ. Sense- and antisense-mediated gene silencing in tobacco is inhibited by the same viral suppressors and is associated with accumulation of small RNAs. Proceedings of the National Academy of Sciences of the United States of America. 2001;98(11):6506-6510
  52. 52. Dogar AM. RNAi dependent epigenetic marks on a geminivirus promoter. Virology Journal. 2006;30;3:5
  53. 53. Vanderschuren H, Alder A, Zhang P, Gruissem W. Dose-dependent RNAi-mediated geminivirus resistance in the tropical root crop cassava. Plant Molecular Biology. 2009;70(3):265-272. DOI: 10.1007/s11103-009-9472-3
  54. 54. Ntui VO, Kong K, Khan RS, Igawa T, Janavi GJ, Rabindran R, Nakamura I, Mii M. Resistance to Sri Lankan cassava mosaic virus (SLCMV) in genetically engineered cassava cv. KU50 through RNA silencing. PLoS One. 2015;10(4):e0120551. DOI: 10.1371/journal.pone.0120551
  55. 55. Yadav JS, Ogwok E, Wagaba H, Patil BL, Bagewadi B, Alicai T, Gaitan-Solis E, Taylor NJ, Fauquet CM. RNAi-mediated resistance to Cassava brown streak Uganda virus in transgenic cassava. Molecular Plant Pathology. 2011;12(7):677-687. DOI: 10.1111/j.1364-3703.2010.00700.x
  56. 56. Odipio J, Ogwok E, Taylor NJ, Halsey M, Bua A, Fauquet CM, Alicai T. RNAi-derived field resistance to Cassava brown streak disease persists across the vegetative cropping cycle. GM Crops & Food. 2014;5(1):16-19. DOI: 10.4161/gmcr.26408
  57. 57. Upadhyay SK, Chandrashekar K, Thakur N, Verma PC, Borgio JF, Singh PK, Tuli R. RNA interference for the control of whiteflies (Bemisia tabaci) by oral route. Journal of Biosciences. 2011;36(1):153-161
  58. 58. Malik HJ, Raza A, Amin I, Scheffler JA, Scheffler BE, Brown JK, Mansoor S. RNAi-mediated mortality of the whitefly through transgenic expression of double stranded RNA homologous to acetylcholinesterase and ecdysone receptor in tobacco plants. Scientific Reports. 2016;6:38469. DOI: 10.1038/srep38469
  59. 59. Fondong VN, Nagalakshmi U, Dinesh-Kumar SP. Novel functional genomics approaches: A promising future in the combat against plant viruses. Phytopathology. 2016;106(10):1231-1239
  60. 60. Voinnet O. Origin, biogenesis, and activity of plant microRNAs. Cell. 2009;136(4):669-687. DOI: 10.1016/j.cell.2009.01.046
  61. 61. Xie M, Zhang S, Yu B. microRNA biogenesis, degradation and activity in plants. Cellular and Molecular Life Sciences. 2015;72(1):87-99. DOI: 10.1007/s00018-014-1728-7
  62. 62. Li JF, Chung HS, Niu Y, Bush J, McCormack M, Sheen J. Comprehensive protein-based artificial microRNA screens for effective gene silencing in plants. The Plant Cell. 2013;25(5):1507-1522. DOI: 10.1105/tpc.113.112235
  63. 63. Fahim M, Larkin J. Designing effective amiRNA and multimeric amiRNA against plant viruses. Methods in Molecular Biology. 2013;942:357-377. DOI: 10.1007/978-1-62703-119-6_19
  64. 64. Khraiwesh B, Ossowski S, Weigel D, Reski R, Frank W. Specific gene silencing by artificial MicroRNAs in Physcomitrella patens: An alternative to targeted gene knockouts. Plant Physiology. 2008;148(2):684-693. DOI: 10.1104/pp.108.128025
  65. 65. Ossowski S, Schneeberger K, Clark RM, Lanz C, Warthmann N, Weigel D. Sequencing of natural strains of Arabidopsis thaliana with short reads. Genome Research. 2008;18(12):2024-2033. DOI: 10.1101/gr.080200.108
  66. 66. Schwab R, Ossowski S, Riester M, Warthmann N, Weigel D. Highly specific gene silencing by artificial microRNAs in Arabidopsis. The Plant Cell. 2006;18(5):1121-1133
  67. 67. Wagaba H, Patil BL, Mukasa S, Alicai T, Fauquet CM, Taylor NJ. Artificial microRNA-derived resistance to Cassava brown streak disease. Journal of Virological Methods. 2016;231:38-43. DOI: 10.1016/j.jviromet.2016.02.004
  68. 68. Szittya G, Silhavy D, Molnar A, Havelda Z, Lovas A, Lakatos L, Banfalvi Z, Burgyan J. Low temperature inhibits RNA silencing-mediated defence by the control of siRNA generation. The EMBO Journal. 2003;22(3):633-640
  69. 69. Fahim M, Millar AA, Wood CC, Larkin PJ. Resistance to Wheat streak mosaic virus generated by expression of an artificial polycistronic microRNA in wheat. Plant Biotechnology Journal. 2012;10(2):150-163. DOI: 10.1111/j.1467-7652.2011.00647.x
  70. 70. Kung YJ, Lin SS, Huang YL, Chen TC, Harish SS, Chua NH, Yeh SD. Multiple artificial microRNAs targeting conserved motifs of the replicase gene confer robust transgenic resistance to negative-sense single-stranded RNA plant virus. Molecular Plant Pathology. 2012;13(3):303-317. DOI: 10.1111/j.1364-3703.2011.00747.x
  71. 71. Allen E, Howell MD. miRNAs in the biogenesis of trans-acting siRNAs in higher plants. Seminars in Cell & Developmental Biology. 2010;21(8):798-804. DOI: 10.1016/j.semcdb.2010.03.008
  72. 72. Yoshikawa M. Biogenesis of trans-acting siRNAs, endogenous secondary siRNAs in plants. Genes & Genetic Systems. 2013;88(2):77-84
  73. 73. Fei Q, Xia R, Meyers BC. Phased, secondary, small interfering RNAs in posttranscriptional regulatory networks. The Plant Cell. 2013;25(7):2400-2415. DOI: 10.1105/tpc.113.114652
  74. 74. Singh A, Taneja J, Dasgupta I, Mukherjee SK. Development of plants resistant to tomato geminiviruses using artificial trans-acting small interfering RNA. Molecular Plant Pathology. 2015;16(7):724-734. DOI: 10.1111/mpp.12229
  75. 75. Chen L, Cheng X, Cai J, Zhan L, Wu X, Liu Q, Wu X. Multiple virus resistance using artificial trans-acting siRNAs. Molecular Plant Pathology. 2015;16(7):724-734. DOI: 10.1111/mpp.12229
  76. 76. Quintero A, Pérez-Quintero AL, López C. Identification of ta-siRNAs and cis-nat-siRNAs in cassava and their roles in response to cassava bacterial blight. Genomics, Proteomics & Bioinformatics. 2013;11(3):172-181. DOI: 10.1016/j.gpb.2013.03.001
  77. 77. Henikoff S, Till BJ, Comai L. TILLING. Traditional mutagenesis meets functional genomics. Plant Physiology. 2004;135(2):630-636
  78. 78. Kurowska M, Daszkowska-Golec A, Gruszka D, Marzec M, Szurman M, Szarejko I, Maluszynski M. TILLING: A shortcut in functional genomics. Journal of Applied Genetics. 2011;52(4):371-390. DOI: 10.1007/s13353-011-0061-1
  79. 79. McCallum CM, Comai L, Greene EA, Henikoff S. Targeting induced local lesions IN genomes (TILLING) for plant functional genomics. Plant Physiology. 2000;123:439-442
  80. 80. Colasuonno P, Incerti O, Lozito ML, Simeone R, Gadaleta A, Blanco A. DHPLC technology for high-throughput detection of mutations in a durum wheat TILLING population. BMC Genetics. 2016;17:43. DOI: 10.1186/s12863-016-0350-0
  81. 81. Harloff H, Lemcke S, Mittasch J, Frolov A, Wu J, Dreyer F, et al. A mutation screening platform for rapeseed (Brassica napus L.) and the detection of sinapine biosynthesis mutants. Theoretical and Applied Genetics. 2012;124(5):957-969. DOI: 10.1007/s00122-011-1760-z
  82. 82. Wang N, Wang Y, Tian F, King G, Zhang C, Long Y, et al. A functional genomics resource for Brassica napus: Development of an EMS mutagenized population and discovery of FAE1 point mutations by TILLING. The New Phytologist. 2008;180(4):751-765. DOI: 10.1111/j.1469-8137.2008.02619.x
  83. 83. Stephenson P, Baker D, Girin T, Perez A, Amoah S, King J, et al. A rich TILLING resource for studying gene function in Brassica rapa. BMC Plant Biology. 2010;10:62. DOI: 10.1186/1471-2229-10-62
  84. 84. Comai L, Young K, Till BJ, Reynolds SH, Greene EA, Codomo CA, Enns LC, Johnson J, Burtner C, Oden AR, et al. Efficient discovery of DNA polymorphisms in natural populations by Ecotilling. The Plant Journal. 2004;37(5):778-786
  85. 85. Ceballos H, Okogbenin E, Pérez JC, Becerra LA, Debouck D. Cassava. In: Bradshaw JE, editor. Root and Tuber Crops. New York: Springer Publishers; 2010. p. 53-96
  86. 86. Lehman U, Robin F. Slowly digestible starch – Its structure and health implications: A review. Trends in Food Science and Technology. 2007;18:346-355
  87. 87. Esfeld K, Uauy C, Tadele Z. Application of TILLING for Orphan Crop Improvement. In: Jain SM, Gupta SD, editors. Biotechnology of Neglected and Underutilized Crops. Dordrecht: The Netherlands Springer; 2013. pp. 83-113
  88. 88. Tadele Z, Esfeld K, Plaza S. Applications of high-throughput techniques to the understudied crops of Africa. Aspects of Applied Biology. 2010;96:233-240
  89. 89. Ülker B, Weisshaar B. Resources for reverse genetics approaches in Arabidopsis thaliana. In: Schmidt R, Bancroft I, editors. Genetics and Genomics of the Brassicaceae. Plant Genetics and Genomics: Crops and Models, New York, NY: Springer; 2011. pp. 527-560
  90. 90. Barrangou R. The roles of CRISPR-Cas systems in adaptive immunity and beyond. Current Opinion in Immunology. 2015;32:36-41. DOI: 10.1016/j.coi.2014.12.008
  91. 91. Barrangou R, Marraffini LA. CRISPR-Cas systems: Prokaryotes upgrade to adaptive immunity. Molecular Cell. 2014;54(2):234-244. DOI: 10.1016/j.molcel.2014.03.011
  92. 92. Cong L, Ran FA, Cox D, Lin S, Barretto R, Habib N, Hsu PD, Wu X, Jiang W, Marraffini LA, et al. Multiplex genome engineering using CRISPR/Cas systems. Science. 2013;339(6121):819-823. DOI: 10.1126/science.1231143
  93. 93. Jinek M, East A, Cheng A, Lin S, Ma E, Doudna J. RNA-programmed genome editing in human cells. eLife. 2013;2:e00471. DOI: 10.7554/eLife.00471
  94. 94. Sternberg SH, Doudna JA. Expanding the biologist's toolkit with CRISPR-Cas9. Molecular Cell. 2015;58(4):568-574. DOI: 10.1016/j.molcel.2015.02.032
  95. 95. Wright AV, Nunez JK, Doudna JA. Biology and applications of CRISPR systems: Harnessing nature's toolbox for genome engineering. Cell. 2016;164(1-2):29-44. DOI: 10.1016/j.cell.2015.12.035
  96. 96. Paul JW 3rd, Qi Y. CRISPR/Cas9 for plant genome editing: Accomplishments, problems and prospects. Plant Cell Reports. 2016;35(7):1417-1427. DOI: 10.1007/s00299-016-1985-z
  97. 97. Odipio J, Alicai T, Nusinow D, Bart R, Ingelbrecht I, Taylor N. CRISPR/CAS9 genome editing of putative cassava flowering genes. In: Third International Conference of the Global Cassava Partnership for the 21st Century; January 18-23, 2016; Nanning, China; 2016
  98. 98. Yin K, Han T, Liu G, Chen T, Wang Y, Yu AY, et al. A geminivirus-based guide RNA delivery system for CRISPR/Cas9 mediated plant genome editing. Scientific Reports. 2015;5:14926. DOI: 10.1038/srep14926
  99. 99. Hsu PD, Scott DA, Weinstein JA, Ran FA, Konermann S, Agarwala V, Li Y, Fine EJ, Wu X, Shalem O, Cradick TJ, Marraffini LA, Bao G, Zhang F. DNA targeting specificity of RNA-guided Cas9 nucleases. Nature Biotechnology. 2013;31(9):827-832. DOI: 10.1038/nbt.2647
  100. 100. Pattanayak V, Lin S, Guilinger JP, Ma E, Doudna JA, Liu DR. High-throughput profiling of off-target DNA cleavage reveals RNA-programmed Cas9 nuclease specificity. Nature Biotechnology. 2013;31(9):839-843. DOI: 10.1038/nbt.2673
  101. 101. Jiang Y, Chen B, Duan C, Sun B, Yang J, Yang S. Multigene editing in the Escherichia coli genome via the CRISPR-Cas9 system. Applied and Environmental Microbiology. 2015;81(7):2506-2514. DOI: 10.1128/AEM.04023-14
  102. 102. Samanta MK, Dey A, Gayen S. CRISPR/Cas9: An advanced tool for editing plant genomes. Transgenic Research. 2016;25(5):561-73. DOI: 10.1007/s11248-016-9953-5
  103. 103. Verdaguer B, de Kochko A, Beachy RN, Fauquet C. Isolation and expression in transgenic tobacco and rice plants, of the cassava vein mosaic virus (CVMV) promoter. Plant Molecular Biology. 1996;31(6):1129-1139
  104. 104. Legg JP, Lava Kumar P, Makeshkumar T, Tripathi L, Ferguson M, Kanju E, Ntawuruhunga P, Cuellar W. Cassava virus diseases: Biology, epidemiology, and management. Advances in Virus Research. 2015;91:85-142. DOI: 10.1016/bs.aivir.2014.10.001
  105. 105. Legg JP, Lava Kumar P, Makeshkumar T, Tripathi L, Ferguson M, Kanju E, Ntawuruhunga P, Cuellar W. Cassava virus diseases: Biology, epidemiology, and management. Advances in Virus Research. 2015;91:85-142. DOI: 10.1016/bs.aivir.2014.10.001
  106. 106. Lee W-S, Fu S-F, Li Z, Murphy AM, Dobson EA, Garland L, Chaluvadi SR, Lewsey MG, Nelson RS, Carr JP. Salicylic acid treatment and expression of an RNA-dependent RNA polymerase 1 transgene inhibit lethal symptoms and meristem invasion during tobacco mosaic virus infection in Nicotiana benthamiana. BMC Plant Biology. 2016;16:15. DOI: 10.1186/s12870-016-0705-8
  107. 107. Wei Y, Hu W, Wang Q, Liu W, Wu C, Zeng H, Yan Y, Li X, He C, Shi H. Comprehensive transcriptional and functional analyses of melatonin synthesis genes in cassava reveal their novel role in hypersensitive-like cell death. Scientific Reports. 2016;6:35029. DOI: 10.1038/srep35029
  108. 108. Bastet A, Robaglia C, Gallois JL. eIF4E resistance: Natural variation should guide gene editing. Trends in Plant Science. 2017;22(5):411-419. DOI: 10.1016/j.tplants.2017.01.008
  109. 109. Allie F, Pierce EJ, Okoniewski MJ, Rey C. Transcriptional analysis of South African cassava mosaic virus-infected susceptible and tolerant landraces of cassava highlights differences in resistance, basal defense and cell wall associated genes during infection. BMC Genomics. 2014;15:1006. DOI: 10.1186/1471-2164-15-1006
  110. 110. Maruthi M, Bouvaine S, Tufan HA, Mohammed IU, Hillocks RJ. Transcriptional response of virus-infected cassava and identification of putative sources of resistance for cassava brown streak disease. PLoS One. 2014;9(5):e96642. DOI: 10.1371/journal.pone.0096642
  111. 111. Pierce EJ, Rey MEC. Assessing global transcriptome changes in response to South African cassava mosaic virus [ZA-99] infection in susceptible Arabidopsis thaliana. PLoS One. 2013;8(6):e67534. DOI: 10.1371/journal.pone.0067534
  112. 112. Allie F, Rey MEC. Transcriptional alterations in model host, Nicotiana benthamiana, in response to infection by South African cassava mosaic virus. European Journal of Plant Pathology. 2013;137:765-785
  113. 113. Maule AJ, Caranta C, Boulton MI. Sources of natural resistance to plant viruses: Status and prospects. Molecular Plant Pathology. 2007 Mar;8(2):223-231. DOI: 10.1111/j.1364-3703.2007.00386.x
  114. 114. Soto JC, Ortiz JF, Perlaza-Jiménez L, Vásquez AX, Lopez-Lavalle LA, Mathew B, Léon J, Bernal AJ, Ballvora A, López CE. A genetic map of cassava (Manihot esculenta Crantz) with integrated physical mapping of immunity-related genes. BMC Genomics. 2015;16:190. DOI: 10.1186/s12864-015-1397-4
  115. 115. Wolfe MD, Kulakow P, Rabbi IY, Jannink J-L. Marker-based estimates reveal significant non-additive effects in clonally propagated cassava (Manihot esculenta): Implications for the prediction of total genetic value and the selection of varieties. G3 (Bethesda). 2016;6(11):3497-3506. DOI: 10.1534/g3.116.033332
  116. 116. Lozano R, Hamblin MT, Prochnik S, Jannink JL. Identification and distribution of the NBS-LRR gene family in the Cassava genome. BMC Genomics. 2015;16:360. DOI: 10.1186/s12864-015-1554-9
  117. 117. Zhai J, Jeong DH, De Paoli E, Park S, Rosen BD, Li Y, Gonzalez AJ, Yan Z, Kitto SL, Grusak MA, et al. MicroRNAs as master regulators of the plant NB-LRR defense gene family via the production of phased, trans-acting siRNAs. Genes & Development. 2011;25(23):2540-2553. DOI: 10.1101/gad.177527.111
  118. 118. Leal LG, Perez Á, Quintero A, Bayona Á, Ortiz JF, Gangadharan A, Mackey D, López C, López-Kleine L. Identification of immunity-related genes in Arabidopsis and cassava using genomic data. Genomics, Proteomics & Bioinformatics. 2013;11(6):345-353. DOI: 10.1016/j.gpb.2013.09.010
  119. 119. Walling LL. Avoiding effective defenses: Strategies employed by phloem-feeding insects. Plant Physiology. 2008;146(3):859-866. DOI: 10.1104/pp.107.113142
  120. 120. Wu J, Baldwin IT. New insights into plant responses to the attack from insect herbivores. Scientific Reports. 2017;7:44729. DOI: 10.1038/srep44729
  121. 121. Shukla AK, Upadhyay SK, Mishra M, Saurabh S, Singh R, Singh H, Thakur N, Rai P, Pandey P, Hans AL, et al. Expression of an insecticidal fern protein in cotton protects against whitefly. Nature Biotechnology. 2016;34(10):1046-1051. DOI: 10.1038/nbt.3665

Written By

Vincent N. Fondong and Chrissie Rey

Submitted: 10 May 2017 Reviewed: 30 August 2017 Published: 17 January 2018