Open access peer-reviewed chapter

New Age of Common Bean

Written By

Monika Vidak, Boris Lazarević, Jerko Gunjača and Klaudija Carović-Stanko

Submitted: 12 December 2022 Reviewed: 08 February 2023 Published: 13 March 2023

DOI: 10.5772/intechopen.110437

From the Edited Volume

Production and Utilization of Legumes - Progress and Prospects

Edited by Mirza Hasanuzzaman

Chapter metrics overview

155 Chapter Downloads

View Full Metrics

Abstract

Common bean (Phaseolus vulgaris L.) is a plant with high nutritional value that occupies an important place in human nutrition worldwide. Its yields vary widely and are often below the genetic potential of the species, given the diversity of cropping systems and climatic changes. Common bean landraces are a source of great genetic variability and provide excellent material for the selection and improvement of numerous agronomic traits and the creation of modern cultivars. It is also important to use high quality seed of high-yielding cultivars in production, because in common bean, in addition to yield and resistance to abiotic and biotic stress factors, traits such as nutritional value and digestibility are also the focus of interest. The success of common bean production depends to a large extent on the quality of the seed, the production approach and new breeding programs.

Keywords

  • breeding
  • common bean production
  • climate changes
  • genetic variability
  • landraces

1. Introduction

Grain legumes production have been neglected regardless of their potential to provide nutrition and food security [1]. They are at the crossroads of many societal challenges affecting agriculture, such as climate change, sustainability and food security [2]. Due to their high content of proteins, fibers, carbohydrates, vitamins and minerals, they play a crucial role in the development of a plant-based diet and are important nutritional components to eliminate hunger and malnutrition [1, 2]. In addition, they improve soil fertility, for example by fixing nitrogen through symbiosis with rhizobia, and at the same time keep crop yields high [3]. Nowadays, however, the trend is changing and many consumers are demanding local food for economic reasons (increasing farmers’ income, adding more value to local stakeholders, etc.), social reasons (i.e. maintaining the population in the area), environmental reasons (reducing traffic and gas emissions, landscape conservation and biodiversity, etc.) and because local products are perceived to be fresher or of better quality [4]. The COVID-19 pandemic and the Russian-Ukrainian war have led to drastic fluctuations in energy prices and disruptions in energy and food supply chains, access to fertilizers is limited, and future harvests are uncertain [5, 6, 7, 8]. As a result, the availability and supply of an extensive range of food commodities and end products are under threat, with notable implications for sourcing, production, processing and logistics, and world markets have recently experienced an increase in food prices [6]. Both crises initially seemed to be an opportunity for a low-carbon energy transition: the pandemic because it accelerated the transition from carbon-intensive energy to modern renewables (such as solar and wind) and illustrated the scale of changes in lifestyle and behavior in a short period of time, and the war because it highlighted the need for greater diversification of energy supply and reliance on local, renewable energy sources [6, 7]. However, early indications suggest that policymakers around the world are focusing on short-term, seemingly quicker solutions, such as supporting the established energy industry in the post-pandemic period to save the economy and finding new fossil fuel supply routes to increase energy security after the war [7]. Accordingly, interest in the use of grain legumes and their components in food is growing in many developed countries. Factors contributing to this trend include the fact that legumes are grown in almost all climatic conditions, as well as their nutritional and health benefits [9].

The aim of this chapter is therefore to highlight the importance of common bean as one of the most important legumes in the world and to point out the possibility of creating new cultivars with desirable traits using new technologies (GWAS and high-throughput phenotyping).

Advertisement

2. Common bean importance

Although its Latin and English names (Phaseolus vulgaris L.; common bean) suggest that it is an ordinary plant species, considering the nutritional properties and genetic structure of the common bean, it can be concluded that, on the contrary, it is an exceptional species that represents a potential crop for future food and nutrient security [10, 11]. It originated in Mexico and, through later diversification and spread throughout the Americas and the world, has become the most ecologically adapted species of the genus Phaseolus, quickly becoming popular for its nutritional qualities [12]. The common bean is the most widely cultivated legume in the world for direct human consumption and a staple food that does not require industrial processing [2, 13]. It is mainly grown as a grain (i.e. dried beans) and as a fresh vegetable (i.e. snap beans, green beans) [2]. In 2021, global production of dry common bean is estimated at more than 27 million tons on more than 34 million ha [14], feeding more than 300 million people linked to the global agricultural economy [15]. An outstanding feature of the germplasm of common bean is its particularly high diversity [16]. It is grown all over the world in different local environments and climates, with extremely diverse cultivation methods, uses and range of environments to which it is adapted, which has contributed to the great diversity of common bean in terms of growth type, seed characteristics and maturity period [16, 17], but unfavorable environmental conditions, especially drought and salinity in soils, affect its overall performance and reduce productivity and harvest but also consequently the nutritional value [10, 18]. As a food source, it can help reduce global food shortages in the coming years [11]. Common bean is a food with high nutritional value, but also with medicinal properties, which is why it plays an important role in human nutrition and is valued as a functional food [19, 20]. It contains all amino acids but is low in sulfur-containing amino acids such as methionine, cysteine and tryptophan, and is an excellent substitute for meat when combined with cereals, which contain plenty of them [21]. Although the nutrient composition of common bean seeds depends on factors such as origin, genotype and environmental conditions [22], it has the highest content of minerals in its seeds of all legumes [23]. In fact, the common bean is among the most nutrient-dense foods available and is often referred to as the “poor man’s meat”, “the near-perfect food” and “the grain of hope” for poor communities [11, 24, 25, 26]. It is considered a potential food to address malnutrition [27]. In addition, the common bean properties are also recognized in other areas. For example, the use of common bean protein as a fish meal substitute and the production of functional fermented beverages from germinated bean seeds have recently been demonstrated [23, 28]. In addition, the N2-fixing capacity of this crop is well known to minimize the need for synthetic N fertilizer to increase yield and quality [29].

Advertisement

3. Common bean breeding

It is of great importance to make synergistic efforts to advance the efficiency and accuracy of common bean breeding and to develop genetic gain opportunities by integrating common bean genetic and genomic resources and improved phenotyping methods into breeding activities [11, 30].

Breeding of common beans is often done locally and focuses on improving response to biotic and abiotic stresses, which are particularly challenging in certain locations [30]. Among abiotic stressors, drought is the number one environmental stress because of temperature dynamics, lighting intensity and lack of rain, affecting 60% of total crop production worldwide [31, 32]. Farmers’ preferences in seed selection and seed lot management had a significant role in the evolutionary development of domesticated beans, their genetic diversity, population structure and chemical composition, which change over time and also depend on agroecological growing conditions [9]. In this context, various researchers around the world use local populations or samples of common bean landraces as reference sets to study their genetic diversity and population structure [9]. Although conventional plant breeding and a collection of global germplasm were the primary sources of improvements in common bean to produce cultivars with greater yield potential [30, 33], in the cases where the study samples come from gene banks, this diversity has remained static over the years [9]. Conventional plant breeding is also designed to address limited requirements and the specific needs of farmers and certain growing environments [30].

In common beans, in addition to yield and resistance to abiotic and biotic stress factors, traits such as nutritional value and digestibility are also the focus of interest [13, 30, 34]. For years, a variety of breeding activities have been carried out to improve several key traits [11] but researchers recognize that current breeding projects would not be sufficient to meet expected future food needs under current climatic conditions [35]. However, to improve the efficiency and accuracy of bean breeding and increase genetic gain, there are tremendous opportunities such as the use of genomic tools and improved phenotyping methods [30]. Thus, common bean variability and a large number of local populations can be used for breeding purposes to create new cultivars with desirable traits (high yielding, adapted to abiotic stresses and with increased nutritional value), and new technologies such as genome-wide association studies (GWAS) and high-throughput phenotyping (HTP) can help to quickly select for these desirable traits.

Advertisement

4. GWAS for biofortification

Hidden hunger is generally a nutritional deficiency that occurs when the quality of food is inadequate for normal growth and development, i.e. as a result of an energy-rich but nutrient-poor diet [36]. It is estimated that more than two billion people worldwide are affected, with young children and women of reproductive age living in low-income countries most at risk [36, 37].

Biofortification is a multidisciplinary strategy to improve staple foods in terms of mineral or vitamin content as a means to combat malnutrition in developing countries [24]. It can be achieved through a variety of approaches such as fertilizer application to the soil or foliage, conventional plant breeding or genetic engineering with genetic modification and transgenesis, using expertise from different fields [34]. Biofortification of common bean is an important strategy to reduce mineral deficiencies, especially in regions of the world where this crop plays a key role in nutrition [38]. Since iron, phosphorus and zinc deficiencies are among the most important nutrient deficiencies in the human diet, research on the genetic control of seed composition focuses mainly on the study of these minerals [10, 39, 40, 41, 42]. Iron is essential for the prevention of anemia and for the proper functioning of many metabolic processes, while zinc is essential for proper growth and resistance to gastrointestinal and respiratory infections, especially in children [24]. In recent years, many efforts have been made to achieve Fe biofortification of common beans with two main objectives: to increase the Fe concentration in common bean seeds and to reduce the content of phytic acid (PA), which is known to reduce the absorption of dietary iron [24, 38, 43]. Finally, research on Fe nutrition has shown that biofortified Fe in common beans can improve the nutritional status of the target population [44].

Recent advances in molecular markers, sequencing technologies and the finishing of the common bean genome sequence have opened up numerous opportunities for fine mapping and characterization of genes [9, 45, 46, 47]. The application of marker-assisted selection (MAS) for more complex traits, such as yield, has recently shifted to genomic selection approaches that are based on genome-wide association studies (GWAS) [33]. Genome-wide association studies (GWAS) have become a widely accepted strategy for studying traits of importance to agriculture, thanks to the introduction of NGS-based SNP markers to decipher genotype–phenotype associations in many species [48]. Recently, a number of GWAS studies on diseases [49, 50, 51, 52], abiotic stress [53, 54], agronomic traits [55, 56, 57], cooking time and culinary quality traits [58] and root traits [59, 60] have been conducted on common beans. In addition, the GWAS results can serve as a basis for understanding the genetic architecture of the nutritional properties of bean seeds, with the aim of increasing the macro- and micronutrient content in the bean breeding program [61].

4.1 Case study: Croatian common bean landraces

Although the common bean is an important food crop in Croatia, production is almost exclusively based on landraces, as there are no current breeding programs that would create new varieties [62]. On the other hand, in the course of the long tradition of bean cultivation in Croatia, many landraces with great genetic and morphological diversity have developed (Figure 1), known by their vernacular names, which are mainly based on the morphological characteristics of the seeds, i.e. the color and mosaic of the seed coat [62, 63, 64]. Landraces are an important source of genes for adaptation to local growing conditions and disease resistance [60, 65]. Furthermore, compared to modern cultivars, landraces are essential sources of key nutritional components for food security and a healthy food supply [66]. However, due to complex socio-economic changes in rural communities in recent decades, such as the low profitability of smaller farms and the aging of farmers who grow modern common bean cultivars and/or other more profitable crops instead of landraces, there is a risk of genetic erosion of landraces [67]. In addition, current abiotic and biotic stress factors are also affecting Croatia, with a focus on drought, reducing agricultural production [68].

Figure 1.

Diversity of Croatian common bean landraces.

Accordingly, the aim of our studies was first to collect landraces of common bean throughout Croatia, in such a way as to include the most cultivated landraces that could be clearly distinguished based on seed morphological characteristics and accordingly divided into 10 morphotypes [62]. Subsequently, by combining phaseolin genotyping, analysis of SSR and SNP markers and morphological traits, 174 accessions of Croatian common bean landraces were evaluated for their origin, genetic diversity, population structure and morphological diversity, and a set of true-type morphogenetic groups was created. The 122 accessions were classified into 14 morphogenetic groups: (1) Mesoamerican (H1A) (‘Biser’, ‘Kukuruzar’, ‘Tetovac’, ‘Trešnjevac’), (2) Andean indeterminate type (H2B1) (‘Dan noć’, ‘Sivi’, ‘Puter’, ‘Sivi prošarani’, ‘Trešnjevac’) and (3) Andean determinate type (H3B2) (‘Bijeli’, ‘Dan noć’, ‘Puter’, ‘Trešnjevac’, ‘Zelenčec’). Fifty-two accessions are putative hybrids between morphogenetic groups [69]. As published in Carović-Stanko et al. [62], the STRUCTURE analysis based on 26 SSRs identified K = 2 as the most likely number of clusters (ΔK = 20,533, 24) and assigned the accessions of Mesoamerican origin (phaseolin type ‘S’) to cluster A, while the accessions of Andean origin (phaseolin type ‘H’/'C’ or ‘T’) formed cluster B, which split into two clusters (B1 and B2) at K = 3 (ΔK = 1935.93) and separated the vast majority of phaseolin type ‘H’/'C’ accessions from those with phaseolin type ‘T’. Thus, at K = 3, 48 (27.59%) accessions were assigned to cluster A, 29 (16.67%) to cluster B1 and 80 (45.96%) to cluster B2. For 17 accessions (9.77%), the membership probabilities Q < 75% for any of the clusters and they were therefore considered as “mixed origin”. The Q values of each accession obtained at K = 3 were used to control for genetic background in the GWAS.

The created panel of accessions was then used for GWAS based on DArTseq-derived SNP markers with the aim of identifying quantitative trait nucleotides (QTNs) associated with variation in seed nutrient content (N, P, K, Ca, Mg, Fe, Zn and Mn) for which phenotypic data on nutrient content were collected from a broader panel of 226 accessions in the research of Palčić et al. [70].

DArTseq analysis was carried out by Diversity Arrays Technology Pty Ltd., Bruce, Australia (https://www.diversityarrays.com/). The quality of the SNP markers derived from DarTseq was determined using the parameters ‘reproducibility’ (percentage of technical replicate pairs that score identically for a given marker), ‘call-rate’ (percentage of samples for which a given marker was scored) and ‘MAF’ (minor- allele frequency) [71]. The marker sequences were aligned against the reference genome of P. vulgaris [46] using BLASTN [72]. By excluding all SNPs with MAF < 0.05 and all SNPs with >0.05 heterozygotes, a final quality control of the SNP data was performed, resulting in the final set of 6311 high-quality DArTseq-derived SNPs. The missing SNP data were imputed using the Beagle 5.1 genotype imputation method [73]. The imputed dataset was then used to construct a kinship matrix using four methods implemented in the software TASSEL 5 [74]: (1) centered IBS [75], (2) normalized IBS [76], (3) dominance-centered IBS [77] and (4) dominance-normalized IBS [78]. In addition, as suggested by Diniz et al. [79], we have used the corrected kinship matrix.

Linkage disequilibrium, the random association between alleles at different loci was measured by the squared value of the coefficient of determination (r2). Bias caused by relatedness and/or population structure was removed by correcting r2: (a) for relatedness using different relatedness matrices (rV2), (b) for population structure using Q values obtained with STRUCTURE (rS2), or (c) for both (rVS2) [80]. The Hill and Weir model [81] was used to represent the decline of LD as a function of distance between loci. According to the uncorrected r2 estimate, the strength of LD did not decrease at all even at a distance of 10 Mbp, and the value of r2 remained above 0.3, even for pairs of loci at opposite ends of the chromosome. The bias caused by consanguinity is stronger than the bias caused by population structure. There was almost no difference between the correction for consanguinity alone and for consanguinity and population structure, in both cases the r2 value fell below 0.1 at about 1 Mbp. Although the differences between the curves for the different kinship matrices were not so pronounced, the centralized IBS matrix was used for GWAS as it gave a slightly better result.

Before performing GWAS, missing phenotypic data were imputed with the method PHENIX which was implemented in the R package of the same name [82]. Before imputation, outliers with the option “trim” in “phenix” were removed (trim.sds = 1.96). GWAS was performed using single-locus models fitted in TASSEL 5 [83] and multi-locus models used in the R package MLMM [84]. In both cases, mixed linear models were fitted with corrections for population structure and genetic relatedness (Q and K matrices). TASSEL “raw” p-values were subjected to adjustment for multiple testing using the “qvalue” package for R [85], with a q-value of 0.2 chosen as the significance threshold. The distribution of TASSEL “raw” p-values was visualized using Manhattan plots created with the “CMplot” package for R [86]. In creating the Manhattan plots, an approximate threshold was calculated for each trait as the p-value of a hypothetical SNP that would have a q-value of 0.2. A similar approximate significance threshold was calculated for MLMM, by using zero-step p-values to estimate the p-value of a hypothetical SNP that would have a q-value of approximately 0.2. Violin plots were created to visualize the distribution of alleles across subpopulations for each QTN.

On chromosomes Pv01, Pv02, Pv03, Pv05, Pv07, Pv08 and Pv10 were detected 22 QTNs which were associated with nitrogen content of the seeds (Figure 2). A total of five QTNs were associated with seed phosphorus content, four on chromosome Pv07 and one on Pv08. On chromosome Pv09, one QTN was found for seed calcium content and on chromosome Pv08 one for seed magnesium content. On chromosome Pv06, two QTNs were found for the zinc content of the seeds.

Figure 2.

Circular Manhattan plot for significant markers detected by TASSEL for N.

As expected, fitting the multilocus model to the MLMM resulted in significantly fewer discoveries of marker-trait associations. Of the 22 QTNs found by TASSEL for N, the MLMM confirmed only two: one of four on chromosome Pv01 and the first of two QTNs on Pv10. Similarly, only one of the four QTNs found by TASSEL for P was on chromosome Pv07. An additional discovery was a QTN found by MLMM for N on chromosome Pv05.

Regarding the relationship between the sizes of the different variance component estimates by MLMM, the comparison of the residual sum of squares (RSS) plots for N and P can be summarized in two key points: (1) population structure explained 40% of the total variability for N and 0% of the total variability for P; (2) the error variability was similar to the genetic variability for N and twice as large for P. Consequently, MLMM discovered three QTNs with p-values below the threshold for N and only one for P despite the similar relative size of genetic variability for N and P.

The largest proportion of total phenotypic variability is explained by QTN Mg_8, which explains 13% of the total phenotypic variability for Mg. Associated markers were distributed throughout the genome, except on chromosomes Pv04 and Pv11, where none were found. N is the trait associated with the greatest number of markers, but individual marker effects were smaller than for other traits. Most markers were located closer to the ends of the chromosomes and only a few were closer to the centromeric region.

The strong effect of population structure on N may be related to the effect of allelic substitution at the QTN loci. In all subpopulations, the reference allele was always present for all QTNs, and the mean N content of individuals carrying the reference allele in subpopulation A (Mesoamerican origin) always lies somewhere between the mean values of subpopulations B1 and B2 (Andean origin). There are three possible scenarios for the distribution of the SNP alleles. They could be present only in the subpopulations of Andean origin, but their positive effect, which is visible in B1 and B2, has almost disappeared at the level of the total population hidden by the effect of population structure. In the second scenario, the SNP allele is only present in subpopulation A (Mesoamerican origin) and has an obvious negative effect that is attenuated by the effect of population structure. Finally, if an SNP allele is present in all subpopulations, its effect varies from one subpopulation to another and becomes almost invisible at the population level. The same scenarios occur with other elements, e.g., with P (Figure 3).

Figure 3.

P seed content distribution for different allele classes within subpopulations (A Mesoamerican; B1 Andean; B2 Andean). Diamonds designate subpopulation means for reference allele homozygotes (gray) and SNP homozygotes (yellow).

This result will serve as a basis for breeding and improving common beans for nutrient content.

Advertisement

5. Phenotyping

Climate change and growing populations have led to the need to develop integrated biotechnological approaches to increase agricultural production while coping with environmental threats. This has led to the concept of developing “climate-proof crop varieties” [31]. In plant breeding and quantitative genetics, a large number of measurements are usually made to select superior individuals or identify regions in the genome that control a trait, which requires high-throughput phenotyping [87]. The tools and applied methods of phenotyping differ in the various -omics disciplines, but basically, they are all used to assess and measure complex traits related to growth, yield, quality and adaptation to various biotic and abiotic stress factors [88]. Therefore, it is necessary to monitor the plant during its growth and development for the early detection of plant responses to stress in each growing season (because each season is different) [89]. This can be done by using new non-destructive automated phenotyping techniques with integrative and simultaneous quantification of multiple morphological and physiological traits, allowing early detection and quantification of different stress factors on a whole-plant basis, i.e. timely identification of how certain stress factors such as drought or e.g. nutrient deficiencies affect the plant [90]. That is, high-throughput phenotyping (HTP) technology plays a crucial role in developing new or better crops through traditional or molecular breeding using marker-assisted selection or genetic selection [35]. The most widely used methods for the non-destructive investigation of phenotypic traits of plants under stress conditions, combining different techniques for measuring gas exchange and techniques for imaging and analyzing the images obtained, are multispectral imaging and multispectral 3D scanning, and chlorophyll fluorescence imaging [90]. These methods provide precise insight into the physiological state of plants under specific environmental conditions, excellently detect morphological and biochemical changes such as light utilization by the photosystem II (PSII) and the underlying biochemical processes, leaf pigment content, chemical composition of leaves, morphological and architectural features of leaves and shoots, etc., and enable rapid data collection and processing [90, 91]. HTP enables objective, fast and precise quantification of morphological, anatomical, physiological and biochemical properties of plants and modeling of ideotypes of agricultural crops adapted to growing in specific agroecological conditions [92]. By growing plants in controlled conditions of growth chambers that enable the management and control of environmental factors such as temperature, duration, spectral composition and intensity of light, availability of nutrients and water, in combination with the latest available methods of spectral analysis (VIS, NIR, IR), chlorophyll fluorescence and measurements of gas exchange, the phenotypic properties of plants, i.e. the complex interaction of genotypes with their environment, are analyzed in an innovative way [88].

In recent years, HTP technology has revolutionized phenotyping and accelerated plant breeding in screening large numbers of plants at different phenological stages [35]. To increase the accuracy and efficiency of plant trait evaluation non-destructive and high-throughput methods have been developed [93]. By using advanced sensors and data acquisition systems, HTP platforms can take full advantage of monitoring, quantifying and evaluating specific phenotypes for large-scale agricultural experiments [94]. Platforms can be used in the laboratory and in the field under controlled and natural conditions and are not necessary to wait for the plants to mature in the field, as the desired traits can be studied quickly in the early stages [35]. This is a crucial step in breeding to select better performing cultivars in terms of yield, abiotic and biotic stress tolerance to accelerate crop improvement programs [35].

To improve the production and quality of common beans, it is necessary to possess in-depth knowledge of its genetic diversity, the genome and the functions of the genes, but also to be familiar with the new phenotyping techniques [35, 95].

As drought is one of the significant environmental stressors due to its significant detrimental effects, there is an increasing need to create tolerant genotypes of agricultural crops [32]. The analysis of gas exchange is based on the fact that drought stress causes rapid closure of stomata [96]. By closing the leaves, the plant saves water, but also reduces the diffusion of CO2 needed for photosynthesis from the atmosphere into the leaf. By measuring stomatal conductance (gas exchange) it is possible to quantify drought stress and select tolerant genotypes [97]. The analysis of chlorophyll fluorescence is based on the fact that the light energy absorbed by the chlorophyll molecules in the photosystems can undergo one of three processes: It can be used to initiate photosynthesis (photochemical reactions), it can be released as heat or it can be re-emitted as long-wave light radiation, i.e., fluorescence. These three processes are interdependent, i.e., any increase in one process leads to a decrease in the value of the other two. Photosystem two (PSII), located in the thylakoid membranes of chloroplasts, is responsible for the uptake of light energy and the initiation of photosynthesis, and at the same time is very sensitive to abiotic stress. Therefore, the measurement of chlorophyll fluorescence, which provides information on changes in the efficiency of photosynthesis, is one of the most commonly used methods for stress assessment in plants [98]. Multispectral analyses are based on the reflection of light of different wavelengths. Many physiological and chemical properties of plants affect the way their tissues absorb and reflect light. When a plant is exposed to stress, these properties can change and thus the intensity of the light reflected by the leaves also changes [99]. The spectral reflectance data of the leaves are used to calculate vegetation indices. Some of the commonly used vegetation indices to assess abiotic stress are the normalized differential vegetation index (NDVI), anthocyanin index (ARI) and chlorophyll index (CHI) [100, 101]. Since drought stress leads to physical and biochemical changes (reduction of leaf area, wilting of the plant, closure of stomata, closure of PSII, disruption of gas exchange, decrease in the intensity of photosynthesis, changes in the composition of pigments, etc.), these parameters can be used to assess the tolerance of genotypes to drought.

Also, for the successful production of beans and obtaining a high yield, a good supply of nutrients to the plant is necessary [102]. Their level can be determined by analyzing the soil, but the nutritional status of beans can be determined by analyzing plant material (leaves). Nutrient deficiency in plants leads to specific symptoms that can be easily detected, for example, by 3D scans with PlantEye F500 multispectral 3D scanner (Figure 4) and chlorophyll fluorescence measurements with CropReporter™ (PhenoVation B.V.,Wageningen, The Netherlands) (Figure 5) [103].

Figure 4.

Color [red, green, and blue (RGB)] and pseudo-color [near infra-red (NIR) and normalized differential vegetation index (NDVI)] images of 3D common bean plants grown for 9 days (MT3) in treatment solutions [12 modified Hoagland’s solution (control), and solutions without nitrogen (N), phosphorus (P), potassium (K), magnesium (Mg), and iron (Fe)] [103].

Figure 5.

Color and pseudo-color images of common bean plants with maximum quantum yield of PSII (Fv/Fm), effective quantum yield of PSII (Fq’/Fm′) and non-photochemical quenching (NPQ) taken during four measurements (MT1-MT4), for 12 days every 3 days of growth in the control [12 modified Hoagland solution (Cont)] and solutions without nitrogen (N), phosphorus (P), potassium (K), magnesium (Mg) and iron (Fe) [103].

Current ground-based phenotyping platforms are likely to be replaced by new and specialized UAVs (drones) and will facilitate next-generation breeding programs to develop improved varieties [35].

The combination of HTP methods with advanced high-throughput genotyping techniques in genome-wide association studies (GWAS) will allow the identification of gene regions and genes associated with specific phenotypic traits (such as drought resistance, increased efficiency of nutrient utilization or disease resistance). The implementation of the results of this research into breeding programs through marker-assisted breeding will enable a faster and more efficient breeding process that will produce new, more efficient and more productive crop genotypes.

Advertisement

6. Conclusion and future perspectives

In these challenging times (climate change, sustainability and food security), the success of common bean production depends not only on seed quality and production approach, but also to a large extent on breeding programs. The implementation of the results of GWAS and phenotyping research into breeding programs through marker-assisted breeding will enable a faster and more efficient breeding process that will produce new, more efficient and more productive common bean genotypes. The combination of advanced high-throughput genotyping techniques in genome-wide association studies (GWAS) will enable the identification of gene regions and genes associated with specific phenotypic traits (e.g. mineral content, drought resistance, increased nutrient utilization or disease resistance).

Advertisement

Acknowledgments

This paper was supported by the project KK.01.1.1.01.0005 Biodiversity and Molecular Plant Breeding, Centre of Excellence for Biodiversity and Molecular Plant Breeding (CoE CroPBioDiv), Zagreb, Croatia and by Croatian Science Foundation under the project UIP-11-2013-3290.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Mecha E, Figueira ME, Vaz Patto MC, do Rosário Bronze M. Two sides of the same coin: The impact of grain legumes on human health: Common bean (Phaseolus vulgaris L.) as a case study. In: Jimenez-Lopez JC, Clemente A, editors. Legume Seed Nutraceutical Research [Internet]. London, UK: IntechOpen; 2018. p. 18. Available from: https://www.intechopen.com/chapters/62359. DOI: 10.5772/intechopen.78737
  2. 2. Murube E, Beleggia R, Pacetti D, Nartea A, Frascarelli G, Lanzavecchia G, et al. Characterization of nutritional quality traits of a common bean germplasm collection. Foods. 2021;10(7):1-26
  3. 3. Allito BB, Ewusi-Mensah N, Logah V. Legume-rhizobium strain specificity enhances nutrition and nitrogen fixation in faba bean (Vicia faba L.). Agronomy. 2020;10(6):1-21
  4. 4. Mallor C, Barberán M, Aibar J. Recovery of a common bean landrace (Phaseolus vulgaris L.) for commercial purposes. Frontiers in Plant Science. 2018;871(1440):1-7
  5. 5. Ben Hassen T, El Bilali H. Impacts of the Russia-Ukraine war on global food security: Towards more sustainable and resilient food systems? Food. 2022;11(15):1-17
  6. 6. Jagtap S, Trollman H, Trollman F, Garcia-Garcia G, Parra-López C, Duong L, et al. The Russia-Ukraine conflict: Its implications for the global food supply chains. Food. 2022;11(14):1-23
  7. 7. Zakeri B, Paulavets K, Barreto-Gomez L, Echeverri LG, Pachauri S, Boza-Kiss B, et al. Pandemic, war, and global energy transitions. Energies. 2022;15(17):1-23
  8. 8. Li K, Qi S, Shi X. The COVID-19 pandemic and energy transitions: Evidence from low-carbon power generation in China. Journal of Cleaner Production. 2022;368(132994):1-9
  9. 9. Chávez-Servia JL, Heredia-García E, Mayek-Pérez N, Aquino-Bolaños EN, Hernández-Delgado S, Carrillo-Rodríguez JC, et al. Diversity of common bean (Phaseolus vulgaris L.) landraces and the nutritional value of their grains. In: Goyal AK, editor. Grain Legumes. London, UK: IntechOpen; 2016. pp. 1-33
  10. 10. Losa A, Vorster J, Cominelli E, Sparvoli F, Paolo D, Sala T, et al. Drought and heat affect common bean minerals and human diet—What we know and where to go. Food and Energy Security. 2022;11(1):1-28
  11. 11. Nadeem MA, Yeken MZ, Shahid MQ , Habyarimana E, Yılmaz H, Alsaleh A, et al. Common bean as a potential crop for future food security: An overview of past, current and future contributions in genomics, transcriptomics, transgenics and proteomics. Biotechnology & Biotechnological Equipment [Internet]. 2021;35(1):758-786. DOI: 10.1080/13102818.2021.1920462
  12. 12. Bitocchi E, Rau D, Bellucci E, Rodriguez M, Murgia ML, Gioia T, et al. Beans (Phaseolus ssp.) as a model for understanding crop evolution. Frontiers in Plant Science. 2017;8(722):1-21
  13. 13. Castro-Guerrero NA, Isidra-Arellano MC, Mendoza-Cozatl DG, Valdes-Lopez O. Common bean: A legume model on the rise for unraveling responses and adaptations to iron, zinc, and phosphate deficiencies. Frontiers in Plant Science. 2016;7(600):1-7
  14. 14. Food and Agricultural Organization of the United Nations. Crops and livestock products [Internet]. 2023. Available from: https://www.fao.org/faostat/en/#data
  15. 15. Nasar S, Shaheen H, Murtaza G, Tinghong T, Arfan M, Idrees M. Socioeconomic evaluation of common bean (Phaseolus vulgaris L.) cultivation in providing sustainable livelihood to the mountain populations of Kashmir Himalayas. Plants. 2023;12(1):1-12
  16. 16. Sedlar A, Marjetka K, Šuštar-Vozlič J, Pipan B, Zadražnik T, Meglič V. Drought stress response in agricultural plants: A case study of common bean (Phaseolus vulgaris L.). In: Ondrasek G, editor. Drought- Detection and Solutions [Internet]. London, UK: IntechOpen; 2019. Available from: https://www.intechopen.com/books/advanced-biometric-technologies/liveness-detection-in-biometrics
  17. 17. Caldas DGG, Konzen ER, Recchia GH, Pereira ACVZ, Tsai SM, Caldas DGG, et al. Functional genomics of biotic and abiotic stresses in Phaseolus vulgaris. In: Abiotic and Biotic Stress in Plants - Recent Advances and Future Perspectives. London, UK: IntechOpen; 2016. pp. 121-151. Available from: https://www.intechopen.com/chapters/49750
  18. 18. Mazón-Suástegui MJ, Ojeda-Silvera CM, García-Bernal M, Avilés-Quevedo AM, Abasolo-Pacheco F, Batista-Sánchez D, et al. Agricultural homoeopathy: A new insight into organics. In: Multifunctionality and Impacts of Organic and Conventional Agriculture. London, UK: IntechOpen; 2016. pp. 1-18. DOI: 10.5772/intechopen.84482
  19. 19. Câmara C, Urrea C, Schlegel V. Pinto beans (Phaseolus vulgaris L.) as a functional food: Implications on human health. Agriculture. 2013;3(1):90-111
  20. 20. Angeles JGC, Villanueva JC, Uy LYC, Mercado SMQ , Tsuchiya MCL, Lado JP, et al. Legumes as functional food for cardiovascular disease. Applied Sciences. 2021;11(12):1-39
  21. 21. Bennetau-Pelissero C. Plant proteins from legumes. In: Reference Series in Phytochemistry [Internet]. Cham: Springer; 2019 [cited 2022 Dec 8]. pp. 223-265. DOI: 10.1007/978-3-319-78030-6_3
  22. 22. Gouveia CSS, Freitas G, de Brito JH, Slaski JJ, de Carvalho MAÂP. Nutritional and mineral variability in 52 accessions of common bean varieties (Phaseolus vulgaris L.) from Madeira Island. Agricultural Sciences. 2014;5(4):317-329
  23. 23. Veber A, Zaręba D, Ziarno M. Functional fermented beverage prepared from germinated white kidney beans (Phaseolus vulgaris L.). In: Ziarno M, editor. Milk Substitutes - Selected Aspects [Internet]. London, UK: IntechOpen; 2021. Available from: https://www.intechopen.com/books/advanced-biometric-technologies/liveness-detection-in-biometrics
  24. 24. Blair MW. Mineral biofortification strategies for food staples: The example of common bean. Journal of Agricultural and Food Chemistry. 2013;61(35):8287-8294
  25. 25. Kotue T, Marlyne J, Wirba L, Amalene S, Nkenmeni D, Kwuimgoin I, et al. Nutritional properties and nutrients chemical analysis of common beans seed. MedCrave Online Journal of Biology and Medicine. 2018;3(2):41-47
  26. 26. CIAT. Common bean: The nearly perfect food. The importance of common bean. CIAT in Focus Crop Commitment. 2019
  27. 27. Jannat S, Shah AH, Sabir SM. Nutraceutical characterisation of common bean (Phaseolus vulgaris L.) germplasm from Pakistan. International Food Research Journal. 2019;26(2):1835-1843
  28. 28. Delgado E, Reyes-Jaquez D. Extruded aquaculture feed: A review. In: Qamar SZ, editor. Extrusion of Metals, Polymers, and Food Products. [Internet]. London, UK: IntechOpen; 2018. p. 21. Available from: https://www.intechopen.com/books/advanced-biometric-technologies/liveness-detection-in-biometrics
  29. 29. Karavidas I, Ntatsi G, Vougeleka V, Karkanis A, Ntanasi T, Saitanis C, et al. Agronomic practices to increase the yield and quality of common bean (Phaseolus vulgaris L.): A systematic review. Agronomy [Internet]. 2022;12(2):271. Available from: https://www.mdpi.com/2073-4395/12/2/271/htm
  30. 30. Assefa T, Assibi Mahama A, Brown AV, Cannon EKS, Rubyogo JC, Rao IM, et al. A review of breeding objectives, genomic resources, and marker-assisted methods in common bean (Phaseolus vulgaris L.). Molecular Breeding. 2019;39(2):1-23
  31. 31. Zargar SM, Mahajan R, Nazir M, Nagar P, Kim ST, Rai V, et al. Common bean proteomics: Present status and future strategies. Journal of Proteomics. [Internet]. 2017;169:239-248. DOI: 10.1016/j.jprot.2017.03.019
  32. 32. Seleiman MF, Al-Suhaibani N, Ali N, Akmal M, Alotaibi M, Refay Y, et al. Drought stress impacts on plants and different approaches to alleviate its adverse effects. Plants. 2021;10(2):1-25
  33. 33. Reinprecht Y, Qi Y, Shahmir F, Smith TH, Pauls KP. Yield and antiyield genes in common bean (Phaseolus vulgaris L.). Legume Science. 2021;3(3):1-17
  34. 34. Cominelli E, Rodiño AP, De Ron AM, Sparvoli F. Genetic approaches to improve common bean nutritional quality: Current knowledge and future perspectives. In: Quality Breeding in Field Crops. Cham: Springer; 2019. pp. 109-138. DOI: 10.1007/978-3-030-04609-5_5
  35. 35. Jangra S, Chaudhary V, Yadav RC, Yadav NR. High-throughput phenotyping: A platform to accelerate crop improvement. Phenomics [Internet]. 2021;1(2):31-53. DOI: 10.1007/s43657-020-00007-6
  36. 36. Lowe NM. The global challenge of hidden hunger: Perspectives from the field. The Proceedings of the Nutrition Society. 2021;80(3):283-289
  37. 37. Muthayya S, Rah JH, Sugimoto JD, Roos FF, Kraemer K, Black RE. The global hidden hunger indices and maps: An advocacy tool for action. PLoS One. 2013;8(6):1-12
  38. 38. Caproni L, Raggi L, Talsma EF, Wenzl P, Negri V. European landrace diversity for common bean biofortification: A genome-wide association study. Scientific Reports [Internet]. 2020;10(1):1-13. DOI: 10.1038/s41598-020-76417-3
  39. 39. Blair MW, Astudillo C, Grusak MA, Graham R, Beebe SE. Inheritance of seed iron and zinc concentrations in common bean (Phaseolus vulgaris L.). Molecular Breeding. 2009 Feb;23(2):197-207
  40. 40. Cichy KA, Caldas GV, Snapp SS, Blair MW. QTL analysis of seed iron, zinc, and phosphorus levels in an Andean Bean population. Crop Science. 2009;49(5):1742-1750
  41. 41. Gregory PJ, Wahbi A, Adu-Gyamfi J, Heiling M, Gruber R, Joy EJM, et al. Approaches to reduce zinc and iron deficits in food systems. Global Food Security. Dec 2017;1(15):1-10
  42. 42. Bindraban PS, Dimkpa CO, Pandey R. Exploring phosphorus fertilizers and fertilization strategies for improved human and environmental health. [cited 2022 Dec 12]; DOI: 10.1007/s00374-019-01430-2
  43. 43. Blair MW, Hurtado N, Sharma P. New gene-derived simple sequence repeat markers for common bean (Phaseolus vulgaris L.). Molecular Ecology Resources. 2012;12(4):661-668
  44. 44. Bouis HE, Saltzman A. Improving nutrition through biofortification: A review of evidence from HarvestPlus, 2003 through 2016. Global Food Security. [Internet]. 2017;12:49-58. DOI: 10.1016/j.gfs.2017.01.009
  45. 45. Gioia T, Logozzo G, Attene G, Bellucci E, Benedettelli S, Negri V, et al. Evidence for introduction bottleneck and extensive inter-gene pool (Mesoamerica x Andes) hybridization in the European common bean (Phaseolus vulgaris L.) germplasm. PLoS One. 2013;8(10):1-14
  46. 46. Schmutz J, McClean PE, Mamidi S, Wu GA, Cannon SB, Grimwood J, et al. A reference genome for common bean and genome-wide analysis of dual domestications. Nature Genetics [Internet]. 2014;46(7):707-713. DOI: 10.1038/ng.3008
  47. 47. Valdisser PAMR, Pereira WJ, Almeida Filho JE, Müller BSF, Coelho GRC, de Menezes IPP, et al. In-depth genome characterization of a Brazilian common bean core collection using DArTseq high-density SNP genotyping. BMC Genomics. 2017;18(1):1-19
  48. 48. Xiao Y, Liu H, Wu L, Warburton M, Yan J. Genome-wide association studies in maize: Praise and stargaze. Molecular Plant [Internet]. 2017;10(3):359-374. DOI: 10.1016/j.molp.2016.12.008
  49. 49. Perseguini J, Oblessuc PR, Rosa J, Gomes KA, Chiorato AF, Carbonell S. Genome-wide association studies of anthracnose and angular leaf spot resistance in common bean (Phaseolus vulgaris L.). PLoS One. 2016;11(3):150506
  50. 50. Wu J, Zhu J, Wang L, Wang S. Genome-wide association study identifies NBS-LRR-encoding genes related with anthracnose and common bacterial blight in the common bean. Frontiers in Plant Science. 2017;8(August):1-15
  51. 51. Campa A, García-Fernández C, Ferreira JJ. Genome-wide association study (GWAS) for resistance to sclerotinia sclerotiorum in common bean. Genes (Basel). 2020;11(12):1-16
  52. 52. Dramadri IO, Amongi W, Kelly JD, Mugisha MC. Genome-wide association analysis of resistance to Pythium ultimum in common bean (Phaseolus vulgaris). Plant Breeding. 2020;139(6):1168-1180
  53. 53. Oladzad A, Porch T, Rosas JC, Moghaddam SM, Beaver J, Beebe SE, et al. Single and multi-trait GWAS identify genetic factors associated with production traits in common bean under abiotic stress environments. G3 Genes, Genomes, Genetics. 2019;9(6):1881-1892
  54. 54. Valdisser PAMR, Müller BSF, de Almeida Filho JE, Morais Júnior OP, Guimarães CM, Borba TCO, et al. Genome-wide association studies detect multiple QTLs for productivity in Mesoamerican diversity panel of common bean under drought stress. Frontiers in Plant Science. 2020;11:1-22
  55. 55. Kamfwa K, Cichy KA, Kelly JD. Genome-wide association study of agronomic traits in common bean. Plant Genome [Internet]. 2015;8(2):1-12. Available from: https://onlinelibrary.wiley.com/doi/full/10.3835/plantgenome2014.09.0059
  56. 56. Moghaddam SM, Mamidi S, Osorno JM, Lee R, Brick M, Kelly J, et al. Genome-wide association study identifies candidate loci underlying agronomic traits in a Middle American diversity panel of common bean. Plant Genome [Internet]. 2016;9(3):1-21. Available from: https://onlinelibrary.wiley.com/doi/full/10.3835/plantgenome2016.02.0012
  57. 57. Nkhata W, Shimelis H, Melis R, Chirwa R, Mzengeza T, Mathew I, et al. Genome-wide association analysis of bean fly resistance and agro-morphological traits in common bean. PLoS One [Internet]. 2021;16(4):e0250729. Available from: https://journals.plos.org/plosone/article?id=10.1371/journal.pone.0250729
  58. 58. Cichy KA, Wiesinger JA, Mendoza FA. Genetic diversity and genome-wide association analysis of cooking time in dry bean (Phaseolus vulgaris L.). Theoretical and Applied Genetics. 2015;128(8):1555-1567
  59. 59. Ambachew D, Blair MW. Genome wide association mapping of root traits in the Andean genepool of common bean (Phaseolus vulgaris L.) grown with and without Aluminum toxicity. Front. Plant Science. 2021;12:1-14
  60. 60. Wu L, Chang Y, Wang L, Wu J, Wang S. Genetic dissection of drought resistance based on root traits at the bud stage in common bean. Theoretical and Applied Genetics [Internet]. 2021;134(4):1047-1061. Available from: https://link.springer.com/article/10.1007/s00122-020-03750-6
  61. 61. Gunjača J, Carović-Stanko K, Lazarević B, Vidak M, Petek M, Liber Z, et al. Genome-wide association studies of mineral content in common bean. Frontiers in Plant Science [Internet]. 2021;12. Available from: https://www.frontiersin.org/articles/10.3389/fpls.2021.636484/full
  62. 62. Carović-Stanko K, Liber Z, Vidak M, Barešić A, Grdiša M, Lazarević B, et al. Genetic diversity of Croatian common bean landraces. Frontiers in Plant Science. 2017;8:1-8
  63. 63. Čupić T, Gantner R, Popović S, Tucak M, Sudar R, Stjepanović M. Widespread annual legumes in Croatia. In: Stipešević B, Sorić R, editors. Proceedings & Abstracts 5th International Scientific/Professional Conference. Osijek, Croatia: Glas Slavonije d.d; 2012. pp. 220-225
  64. 64. Vidak M, Malešević S, Grdiša M, Šatović Z, Lazarević B, Carović-Stanko K. Phenotypic diversity among Croatian common bean (Phaseolus vulgaris L.) landraces. Agriculturae Conspectus Scientificus. 2015;80(3):133-137
  65. 65. Azeez MA, Adubi AO, Durodola FA. Landraces and crop genetic improvement. In: Grillo O, editor. Rediscovery of Landraces as a Resource for the Future [Internet]. London: IntechOpen; 2018. p. 1-19. Available from: http://www.intechopen.com/books/trends-in-telecommunications-technologies/gps-total-electron-content-tec-prediction-at-ionosphere-layer-over-the-equatorial-region%0AInTec
  66. 66. Celmeli T, Sari H, Canci H, Sari D, Adak A, Eker T, et al. The nutritional content of common bean (Phaseolus vulgaris L.) landraces in comparison to modern varieties. Agronomy. 2018;8(9):1-9
  67. 67. Yin L, Zhang H, Tang Z, Xu J, Yin D, Yuan X, et al. rMVP: A memory-efficient, visualization-enhanced, and parallel-1 accelerated tool for genome-wide association study. Genomics Proteomics Bioinforma. 2021;19(4):619-628
  68. 68. Pandžić K, Likso T, Curić O, Mesić M, Pejić I, Pasarić Z. Drought indices for the Zagreb-Grič observatory with an overview of drought damage in agriculture in Croatia. Theoretical and Applied Climatology. 2020;142(1-2):555-567
  69. 69. Vidak M, Šatović Z, Liber Z, Grdiša M, Gunjača J, Kilian A, et al. Assessment of the origin and diversity of Croatian common bean germplasm using Phaseolin type, SSR and SNP markers and morphological traits. Plants. 2021;10(4):665, 1-23. DOI: 10.3390/plants10040665
  70. 70. Palčić I, Karažija T, Petek M, Lazarević B, Herak Ćustić M, Gunjača J, et al. Relationship between origin and nutrient content of Croatian common bean landraces. Journal of Central European Agriculture. 2018;19(3):490-502
  71. 71. Wenzl P, Carling J, Kudrna D, Jaccoud D, Huttner E, Kleinhofs A, et al. Diversity arrays technology (DArT) for whole-genome profiling of barley. Proceedings of the National Academy of Sciences of the United States of America. 2004;101(26):9915-9920
  72. 72. Zhang Z, Schwartz S, Wagner L, Miller W. A greedy algorithm for aligning DNA sequences. Journal of Computational Biology. 2000;7(1-2):203-214
  73. 73. Browning BL, Zhou Y, Browning SR. A one-penny imputed genome from next-generation reference panels. American Journal of Human Genetics. 2018;103(3):338-348
  74. 74. Bradbury PJ, Zhang Z, Kroon DE, Casstevens TM, Ramdoss Y, Buckler ES. TASSEL: Software for association mapping of complex traits in diverse samples. Bioinformatics. 2007;23(19):2633-2635
  75. 75. Endelman JB, Jannink JL. Shrinkage estimation of the realized relationship matrix. G3-Genes Genomes Genetics. 2012;2(11):1405-1413
  76. 76. Yang J, Lee SH, Goddard ME, Visscher PM. GCTA: A tool for genome-wide complex trait analysis. American Journal of Human Genetics. 2011;88(1):76-82
  77. 77. Muñoz PR, Resende MFR, Gezan SA, Resende MDV, de los Campos G, Kirst M, et al. Unraveling additive from nonadditive effects using genomic relationship matrices. Genetics. 2014;198(4):1759-1768
  78. 78. Zhu Z, Bakshi A, Vinkhuyzen AAE, Hemani G, Lee SH, Nolte IM, et al. Dominance genetic variation contributes little to the missing heritability for human complex traits. American Journal of Human Genetics. 2015;96(3):377-385
  79. 79. Diniz WJS, Mazzoni G, Coutinho LL, Banerjee P, Geistlinger L, Cesar ASM, et al. Detection of co-expressed pathway modules associated with mineral concentration and meat quality in nelore cattle. Frontiers in Genetics. 2019;10:1-12
  80. 80. Mangin B, Siberchicot A, Nicolas S, Doligez A, This P, Cierco-Ayrolles C. Novel measures of linkage disequilibrium that correct the bias due to population structure and relatedness. Heredity (Edinb). 2012;108(3):285-291
  81. 81. Hill WG, Weir BS. Variances and covariances of squared linkage disequilibria in finite populations. Theoretical Population Biology. 1988;33(1):54-78
  82. 82. Dahl A, Iotchkova V, Baud A, Johansson S, Gyllensten U, Soranzo N, et al. A multiple-phenotype imputation method for genetic studies. Nature Genetics. 2016;48(4):466-472
  83. 83. Yu J, Pressoir G, Briggs WH, Bi IV, Yamasaki M, Doebley JF, et al. A unified mixed-model method for association mapping that accounts for multiple levels of relatedness. Nature Genetics. 2006;38(2):203-208
  84. 84. Segura V, Vilhjálmsson BJ, Platt A, Korte A, Seren Ü, Long Q , et al. An efficient multi-locus mixed-model approach for genome-wide association studies in structured populations. Nature Genetics. 2012;44(7):825-830
  85. 85. Storey JD, Bass AJ, Dabney A, Robinson D. qvalue: Q-value estimation for false discovery rate control version 2.22.0 from bioconductor [Internet]. 2020. Available from: https://rdrr.io/bioc/qvalue/
  86. 86. Yin L, Zhang H, Tang Z, Xu J, Yin D. rMVP: A memory-efficient, visualization-enhanced, and parallel-accelerated tool for genome-wide association study. Genomics Proteomics Bioinforma. 2020;19(4):619-628
  87. 87. Walter A, Liebisch F, Hund A. Plant phenotyping: From bean weighing to image analysis. Plant Methods. 2015;11(1):1-11
  88. 88. Kondić-Špika A, Mikić S, Mirosavljević M, Trkulja D, Marjanović Jeromela A, Rajković D, et al. Crop breeding for a changing climate in the Pannonian region: Towards integration of modern phenotyping tools. Journal of Experimental Botany. 2022;73(15):5089-5110. Available from: https://academic.oup.com/jxb/article/73/15/5089/6583379
  89. 89. Yao Y, Yang Y, Li C, Huang D, Zhang J, Wang C, et al. Plants research progress on the functions of gasotransmitters in plant responses to abiotic stresses. 2019 [cited 2022 Dec 12]; Available from: www.mdpi.com/journal/plants
  90. 90. Lazarević B, Šatović Z, Nimac A, Vidak M, Gunjača J, Politeo O, et al. Application of phenotyping methods in detection of drought and salinity stress in basil (Ocimum basilicum L.). Frontiers in Plant Science. 2021;12:1-13
  91. 91. Brestic M, Zivcak M. PSII fluorescence techniques for measurement of drought and high temperature stress signal in crop plants: Protocols and applications. In: Molecular Stress Physiology of Plants [Internet]. 2013; [cited 2022 Dec 12]. pp. 87-131. Available from: https://link.springer.com/chapter/10.1007/978-81-322-0807-5_4
  92. 92. Roth L, Barendregt C, Bétrix CA, Hund A, Walter A. High-throughput field phenotyping of soybean: Spotting an ideotype. Remote Sensing of Environment. 2022;269:1-13
  93. 93. Sankaran S, Quirós JJ, Miklas PN. Unmanned aerial system and satellite-based high resolution imagery for high-throughput phenotyping in dry bean. Computers and Electronics in Agriculture. [Internet]. 2019;165:1-38. DOI: 10.1016/j.compag.2019.104965
  94. 94. Li D, Quan C, Song Z, Li X, Yu G, Li C, et al. High-throughput plant phenotyping platform (HT3P) as a novel tool for estimating agronomic traits from the lab to the field. Frontiers in Bioengineering and Biotechnology. 2021;8:1-24
  95. 95. Zargar SM, Nazir M, Rai V, Hajduch M, Agrawal GK, Rakwal R. Towards a common bean proteome atlas: Looking at the current state of research and the need for a comprehensive proteome. Frontiers in Plant Science. 2015;6:1-4
  96. 96. Taiz L, Zeiger E. Plant Physiology. 3rd ed. Massachusetts, USA: Sinauer Associates; 2002
  97. 97. Farooq M, Hussain M, Wahid A, Siddique KHM. Drought stress in plants: An overview. In: Plant Responses to Drought Stress: From Morphological to Molecular Features [Internet]. Berlin Heidelberg: Springer-Verlag; 2012. [cited 2022 Dec 12]. pp. 1-33. Available from: https://link.springer.com/chapter/10.1007/978-3-642-32653-0_1
  98. 98. Maxwell K, Johnson GN. Chlorophyll fluorescence—A practical guide. Journal of Experimental Botany [Internet]. 2000;51(345):659-668. Available from: https://academic.oup.com/jxb/article/51/345/659/652534
  99. 99. Zubler AV, Yoon JY. Proximal methods for plant stress detection using optical sensors and machine learning. Biosens [Internet]. 2020;10(12):193. Available from: https://www.mdpi.com/2079-6374/10/12/193/htm
  100. 100. Gitelson AA, Merzlyak MN, Chivkunova OB. Optical properties and nondestructive estimation of anthocyanin content in plant leaves. Photochemistry and Photobiology. 2001;74(1):38
  101. 101. Gitelson AA, Gritz Y, Merzlyak MN. Relationships between leaf chlorophyll content and spectral reflectance and algorithms for non-destructive chlorophyll assessment in higher plant leaves. Journal of Plant Physiology. 2003;160(3):271-282
  102. 102. Kandel H, Endres G. Dry Bean Production Guide. North Dakota: NDSU Extension; 2019. pp. 1-128
  103. 103. Lazarević B, Carović-Stanko K, Živčak M, Vodnik D, Javornik T, Safner T. Classification of high-throughput phenotyping data for differentiation among nutrient deficiency in common bean. Frontiers in Plant Science. 2022;13(931877):1-17

Written By

Monika Vidak, Boris Lazarević, Jerko Gunjača and Klaudija Carović-Stanko

Submitted: 12 December 2022 Reviewed: 08 February 2023 Published: 13 March 2023