Next Article in Journal
Comparative Proteomics Reveals the Difference in Root Cold Resistance between Vitis. riparia × V. labrusca and Cabernet Sauvignon in Response to Freezing Temperature
Next Article in Special Issue
Isolation and Comprehensive in Silico Characterisation of a New 3-Hydroxy-3-Methylglutaryl-Coenzyme A Reductase 4 (HMGR4) Gene Promoter from Salvia miltiorrhiza: Comparative Analyses of Plant HMGR Promoters
Previous Article in Journal
Factors Affecting Seed Germination of the Invasive Species Symphyotrichum lanceolatum and Their Implication for Invasion Success
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Machine Learning for Plant Stress Modeling: A Perspective towards Hormesis Management

by
Amanda Kim Rico-Chávez
1,
Jesus Alejandro Franco
2,
Arturo Alfonso Fernandez-Jaramillo
3,
Luis Miguel Contreras-Medina
1,
Ramón Gerardo Guevara-González
1,* and
Quetzalcoatl Hernandez-Escobedo
2,*
1
Unidad de Ingeniería en Biosistemas, Facultad de Ingeniería Campus Amazcala, Universidad Autónoma de Querétaro, Carretera Chichimequillas, s/n km 1, El Marqués CP 76265, Mexico
2
Escuela Nacional de Estudios Superiores Unidad Juriquilla, UNAM, Querétaro CP 76230, Mexico
3
Unidad Académica de Ingeniería Biomédica, Universidad Politécnica de Sinaloa, Carretera Municipal Libre Mazatlán Higueras km 3, Col. Genaro Estrada, Mazatlán CP 82199, Mexico
*
Authors to whom correspondence should be addressed.
Plants 2022, 11(7), 970; https://doi.org/10.3390/plants11070970
Submission received: 7 March 2022 / Revised: 28 March 2022 / Accepted: 31 March 2022 / Published: 2 April 2022
(This article belongs to the Special Issue Plant Computational Biology)

Abstract

:
Plant stress is one of the most significant factors affecting plant fitness and, consequently, food production. However, plant stress may also be profitable since it behaves hormetically; at low doses, it stimulates positive traits in crops, such as the synthesis of specialized metabolites and additional stress tolerance. The controlled exposure of crops to low doses of stressors is therefore called hormesis management, and it is a promising method to increase crop productivity and quality. Nevertheless, hormesis management has severe limitations derived from the complexity of plant physiological responses to stress. Many technological advances assist plant stress science in overcoming such limitations, which results in extensive datasets originating from the multiple layers of the plant defensive response. For that reason, artificial intelligence tools, particularly Machine Learning (ML) and Deep Learning (DL), have become crucial for processing and interpreting data to accurately model plant stress responses such as genomic variation, gene and protein expression, and metabolite biosynthesis. In this review, we discuss the most recent ML and DL applications in plant stress science, focusing on their potential for improving the development of hormesis management protocols.

1. Introduction

Stress is a defensive state of an organism resulting from deviations of its optimal developmental conditions [1]. Environmental challenges destabilize fundamental biological functions in plants, and this is often perceived as constrained crop growth and development. In agricultural systems, the detrimental effects of plant stress are a significant cause of productivity loss, threatening food security, especially in the current context of climate change [2]. Nevertheless, aside from being deleterious, stress responses can also induce desirable traits in crops and therefore be considered favorable [3]. In that case, stress is often called eustress, a term derived from the Greek prefix eu that means good or well [4].
Whether a stressor will harm or benefit an organism depends entirely on the intensity of its incidence [5]. This observation derives from the fact that plant defensive responses to stress are biphasic, meaning that high doses of a stressor tend to be unfavorable, whereas low doses are beneficial [6] (Figure 1). The phenomenon that explains such biphasic behavior is called hormesis, and it demonstrates how some level of stress is necessary for a plant to achieve optimal fitness [7].
The acquaintance of hormetic responses of plants to stressors is the basis for implementing hormesis management, which refers to the deliberate exposure of crops to eustress for eliciting desirable attributes [8]. This practice is also called controlled elicitation, and it can considerably increase crop yield, growth, quality, pest resistance, and overall stress tolerance [9,10]. Nevertheless, the design of hormesis management protocols still faces many limitations, mainly due to the complexity of plant stress responses and the lack of consideration of the hormesis model in plant stress research [7].
Several factors shape the defensive responses of plants, including the type of stressor, the genetic identity of the individual, its developmental stage, its nutritional stage, and the responding tissue or cell [11]. Consequently, any given dose of a specific stressor will fall within a different dose-response range according to the responding organism and the observed output variable [4]. Therefore, eustress doses must be experimentally determined before proposing their implementation [12,13]. Moreover, until very recently, studies on plant stress responses usually focused on determining a damaging-dose threshold rather than depicting the whole dose-response curve [6,7,13]. For these reasons, applying novel technologies for enhancing the scopes of hormesis management in agriculture is crucial [14].
Artificial intelligence (AI) is an evolving branch of computer science with great potential to solve a variety of complex problems of the modern world. From using advanced fuzzy logic models for wastewater treatment [15], estimating the production of biosurfactants by bacteria with artificial neural networks (ANN) and the fuzzy inference system ANFIS [16], to advanced Deep Learning tools in plant science. In that matter, AI tools are helpful to model plant distribution, identify species, recognize disease and stress, diagnose nutritional deficiencies, and apply agrochemicals in precision agriculture [17]. In particular, Machine Learning (ML) techniques can predict the outcome of various complex biological processes, such as gene function, gene networks, protein interactions, and optimal growing conditions, leading to significant achievements in plant stress research [18,19].
Literature reviews assessing the use of machine learning for plant stress research focus mainly on analyzing the numerous findings on the identification, classification, quantification, and early prediction of deleterious plant stress responses [20,21,22,23]. However, the potential of the most recent modeling techniques for understanding eustress responses for crop improvement remains unexplored. Therefore, this review aims to present the most recent findings in plant stress research and propose machine learning to model dose-response for implementing hormesis management in agriculture.

2. Hormesis and Plant Stress Science

The hormesis model is rapidly gaining recognition and acceptance within academia for depicting the responses to external stimuli of any given biological system [24]. Nevertheless, the concept of hormesis is not recent. The first reports to describe the biphasic dose-response date from the late 19th century, whereas the term hormesis appeared in the literature for the first time in 1943 [14]. Despite its early description, the scientific community rejected the hormesis phenomenon until the last decades of the 20th century because it was mistakenly associated with homeopathy [25]. Consequently, most of the research around dose-response falls under the assumption of linearity of such responses and overlooks low-dose stimulation, which impacts how scientists, regulatory agencies, and society conceive stress [6]. However, hormesis appears in stress studies at such a high frequency that it has quickly regained consideration in the design of research projects [26].
At present, low-dose stimulation has multiple applications in clinical medicine, environmental risk assessment, ecology, crop management, and sustainable agriculture [27,28]. Several findings suggest that hormetic responses are highly generalizable and occur in all kinds of biological systems [29,30]. Moreover, the quantitative characteristics of hormesis remain constant among models as they correspond to the limits of biological plasticity, meaning that hormesis is related to adaptability and evolution [31]. This latter fact is especially relevant for understanding plants, given that their survival relies entirely on their adaptive responses to stressors due to their sessile nature.
Plants possess complex defensive mechanisms to deal with the many biotic and abiotic challenges they may encounter in the wild. Protected agriculture diminishes these challenges causing plants to underdevelop defensive mechanisms and making them more susceptible to environmental stress and pests or decreasing their production of desirable specialized metabolites [32]. In this scenario, deliberately exposing crops to low-dose stress may enhance plant productivity and stress resistance [3]. However, controlled stimulation of plant defensive mechanisms leads to many different observable outcomes as plants can sense different stressors and respond to them in a specific manner [33].
The specificity of defensive responses depends directly on the plant species, the type of stressor, and the responding tissue. As a result, plant stress responses are diverse, and so are the methods for their analysis [34]. Each stress response is a multilayered molecular process that can be understood as an information transfer between the stressor perception and the expression of a phenotypic trait. Calabrese and Blain [30] assessed more than 3000 hormetic plant responses with numerous response variables, each representing a specific point in the physiological pathway triggered by the stress incidence. Such observations show that, given their complexity, analyzing plant physiological mechanisms can generate a significant amount of data.

3. Data in Plant Hormesis Research

The analysis of a substantial number of pathosystems has permitted the description of many receptor-to-response routes [35]. However, to picture plant immunity as a collection of independent downstream cascades is now thought to be quite simplistic [36]. In contrast, plant immunity should rather be understood as an intricate systemic molecular network capable of simultaneously perceiving biotic and abiotic stressors and ultimately leading to transcription reprogramming and protective physiological responses [37,38,39].
The convenience of choosing one variable or another as the output of hormesis management depends on the target crop, the cultivation conditions, and the productivity objective. For example, increasing the drought tolerance of a food crop cultivated in a controlled environment would be irrelevant, whereas augmenting its yield would be paramount. Therefore, the description of plant defensive responses must consider several endpoints (response variables) to possess technological significance. Fortunately, collecting a considerable amount of data from biological systems is becoming more frequent, thanks to the latest advances in analytical instrumentation and techniques [40].
High-throughput analyses increase the chances to elucidate physiological processes and ecological interactions of plants from the broadened perspective of systems biology [41]. The generation of big data sets from the simultaneous analysis of an extensive collection of biomolecules corresponding to a definite category (genes, transcripts, proteins, and metabolites) has led to the so-called omics approach, which is the primary tool of systems biology [42]. Furthermore, a multi-omics approach makes it possible to obtain a more detailed snapshot of a plant system by simultaneously analyzing its whole genome, proteome, transcriptome, and metabolome [40]. Moreover, the multi-omics approach applied to single-cell functional analyses can simplify data processing and modeling to accurately depict many biological processes in plants [43].
In the following subsections, we will briefly describe the main types of data each omics approach can deliver when applied to plant hormesis research.

3.1. Genomics and Transcriptomics

Genomics refers to the sequencing, assembly, and functional analysis of the genome of a plant, and it has advanced more rapidly than any other omics in plant science [44]. Only in the last two decades, the sequences of more than 100 plant genomes have been published, and further technological advances in genomics have increased our understanding of plant biology leading to substantial agricultural progress [45].
Genome sequencing has several applications in plant stress science. The structural analysis of DNA is not only fundamental for classifying organisms but also for identifying stress-driven mutations, which occur in plants under heat [46], drought [47], and other abiotic stresses [48]. Moreover, DNA structural variations occurring under low-dose stress can be linked to gene function using gene ontology analyses to reveal the genetic basis of hormesis [49]. DNA sequence variations such as Single Nucleotide Polymorphisms (SNPs) are also helpful for understanding the molecular mechanisms underlying hormetic responses when analyzed along with phenotypic traits as in Genome Wide Association Studies (GWAS) [50,51]. GWAS analyses consider big data sets to identify and predict gene candidates and quantitative trait loci accountable for stress responses [52].
SNPs genotyping in combination with other data sets from high-throughput analyses such as phenomics or enviromics has also led to the development of genomic selection for optimizing crop breeding [53]. With this strategy, it is possible to improve physiological traits with hormetic behavior in crops, such as yield, pest resistance, and environmental stress tolerance, to shorten breeding cycles and decrease the need for continuous phenotyping [54]. Additionally, the advent of outstanding new genome-editing techniques, such as the Zinc Finger Nucleases (ZFN), the Transcription Activator-Like Effector Nucleases (TALENs), and the Clustered Regulatory Interspaced Short Palindromic Repeats (CRISPR) systems, implies, along with transcriptomics, the most significant advance in the development of stress-resistant crops [55].
The rapid advances in sequencing technologies and bioinformatics have also substantially impelled RNA analyses [56]. The synthesis of RNA is dynamic, depending on the activation of a gene to occur. Therefore, transcriptomics is the key to investigating gene function in targeted physiological mechanisms qualitatively and quantitatively [57]. Detecting hormetic stimulation at the transcript level can be achieved by analyzing the differential expression of known genes on small (~20) [58,59,60] and very large scales using microarray technology (~50,000) [61] or by completely sequencing the RNA from a sample as in next-generation and third-generation sequencing [62,63,64].
The computational analysis of transcriptomic datasets precedes the reconstruction of gene regulatory networks and the crosstalk by which they interconnect during specific physiological processes [65]. In particular, machine learning algorithms can infer interactions between genes with great accuracy [66]. Nevertheless, gene regulation during hormetic responses involves biological processes other than transcription, such as epigenome dynamics, which depends on chromatin structural changes, namely DNA methylation and histone modifications [67]. Therefore, integrating additional types of datasets and adding spatial and temporal information is fundamental to increasing model resolution and depicting the mechanisms of hormesis truthfully [68].

3.2. Proteomics

Proteins are the main regulatory molecules in every cell process. Therefore, the ensemble of differentially translated proteins in response to a given stimulus is an additional dataset that contains essential information for ascertaining hormetic cellular mechanisms [69]. Moreover, many studies show that RNA quantity does not proportionally relate to protein abundance [70]. The latter occurs mainly due to additional regulation steps between transcription and protein synthesis and the stability of the end products [71]. For that reason, transcriptomics and proteomics, or other high-throughput analyses should be simultaneously conducted to validate and reconstruct entire regulatory networks [72].
Detecting differential changes in plant proteome is especially useful for studying plant stress responses since relatively small variations in the dose of a stressor result in a significant difference in the proteome at both mild and severe stressor incidence [73]. Furthermore, under stress conditions, the plant cell upregulates the expression of proteins associated with primary metabolic processes such as photosynthesis, redox homeostasis, energy metabolism, nitrogen absorption, and the biosynthesis of signaling molecules [74,75]. Therefore, plant proteomics can help researchers detect stress at a molecular level earlier than it would be possible by analyzing changes in observable phenotypic traits and for both stress-susceptible and tolerant genotypes [76].
Proteome analyses make it possible to identify and characterize novel proteins, and along with bioinformatics, proteomics enables tracking variations in protein abundance, form, cellular location, and activity following a stressors incidence [77]. Additionally, proteome research has proven helpful for clarifying cellular organelle function, post-translational modifications, and protein–protein interactions, providing a more in-depth insight into the stress-driven molecular mechanisms of plant cells [78]. Proteomics technologies range from the classic gel-based and the Liquid Chromatography coupled to Mass Spectrometry (LC-MS) approaches to the modern Mass Spectrometry Imaging (MSI), and combined with additional high-throughput analyses, these still underexploited tools are among the most powerful methods for unraveling the molecular mechanism of hormetic stress responses in plants [79].

3.3. Metabolomics

A number of the differentially expressed proteins resulting from stress incidence are regulators that activate and shape specialized metabolic pathways inside the cell [80]. Metabolomics is the study of all the small molecules in a tissue, which, in the case of plants, possess a unique structural and functional complexity [81]. Moreover, due to their sessility, plants depend on chemical signaling to maintain homeostasis and ecological interactions at intra- and interspecies levels. Plant specialized metabolism is evolutionarily shaped by environmental pressures to synthesize chemical compounds with an enormous structural and functional diversity and capable of interacting with living tissues [82].
Many plant specialized metabolites are an active part of plant internal signaling pathways or exert bioactivity on other organisms [83]. Furthermore, every plant organism is a genuine biological factory capable of synthesizing an estimated 30,000 metabolites [84]. Hence, plants are a significant source of chemical compounds with pharmacological properties and are particularly valuable among natural products for drug discovery purposes [85]. In addition, plant metabolites are fundamental for maintaining human health by conferring nutritional, functional, and nutraceutical value to food products [86].
Being an adaptive response, the activation of plant metabolism also exhibits a hormetic behavior [87,88,89], and deliberately exposing crops to low-dose stress is a convenient means for stimulating metabolites accumulation [90]. Moreover, the metabolomics analysis of plant stress response along with bioinformatics makes it possible to find candidate markers for directing crop breeding and predicting crop performance under environmental stress [91].
Given the structural diversity of plant metabolites, the main limitation of metabolomics resides in developing comprehensive extraction techniques and analytical methods to detect a big heterogeneous ensemble of chemical compounds simultaneously. Nevertheless, thanks to the recent advances in coupled analytical technologies and bioinformatics, particularly Mass Spectrometry (MS), Nuclear Magnetic Resonance (NMR), and hybrid MS/NMR methods [92], it is now possible to separate and detect the whole metabolome from a biological sample quickly and affordably [93]. Moreover, many intrinsic experimental conditions for metabolome analysis are compatible with other omics studies, making metabolomics a convenient foundation for designing and fulfilling multi-omics experiments and an effective tool for systems biology research [94].

3.4. Phenomics

Plant phenotyping is the measurement of phenotypic traits either at the cell, organ, or whole plant level for understanding the underlying mechanisms of the interactions of plants with their environment [95]. Molecular responses drive phenotypic change, and for that reason, the developmental traits of plants, such as growth, seed germination, photosynthesis, transpiration, stomatal conductance, and pigmentation, among others, also display hormetic behavior [96]. As a result, various sensors can be used for differentially analyzing physiological plant hormetic responses to stress-related events [97].
Image-based phenotyping is useful to detect leaf morphological variations in plants [98]. Red-green-blue (RGB) imaging uses Charge Coupled Device (CCD) or Complementary Metal Oxide Semiconductor (CMOS) sensors to detect color changes related to plant stress responses. Such sensors work within the visible range of the electromagnetic spectrum and are convenient to diagnose nitrogen (N), phosphorous (P), potassium (K), magnesium (Mg), calcium (Ca), and iron (Fe) deficiency symptoms [99]. Detecting nutrient deficit can also help identify environmental stress incidence. For example, Martinez et al. (2020) [100] reported that water deficit modifies nitrate uptake by altering the expression of genes related to nitrate assimilation in the roots and the shoot. Moreover, changes in pigment content can be related to visible stress symptoms in such a detailed manner, that it is possible to discriminate between their biotic or abiotic origin [101].
Yellowing is the most notable symptom of leaf senescence, and it appears due to seasonal developmental processes, pathogen attack, and abiotic stressors incidence, indicating a decrease in the photosynthetic rate [102]. Chlorophyll metabolism is regulated in a hormetic manner, and therefore it can perform as a biomarker to identify other metabolic changes resulting from low-dose stress incidence [103,104]. Many imaging techniques focus on detecting chlorophyll fluctuations with convenient results for biotic and abiotic stress phenotyping, such as chlorophyll fluorescence. This technique is relevant to determining overall crop fitness, and due to its high sensitivity, it has been extensively applied for the early detection of stress incidence [105].
Imaging techniques can also be used to analyze plant physiological processes and identify stress even in the absence of symptoms unobservable for the unaided eye. Magnetic resonance imaging can be applied to plant systems to elucidate plant-water relationships and as a post-harvest control to determine maturity and mechanical damage of agricultural products [106]. Thermography uses optical sensors that detect radiation outside the visual range of the electromagnetic spectrum, and it has been used to detect plant interactions with biotic and abiotic stressors and monitor environmental stress susceptibility and resistance [107,108]. Mild-stress responses can also be detected using multispectral and hyperspectral imaging. Multispectral imaging considers only specific bands of electromagnetic spectra, whereas hyperspectral imaging increases the resolution of the wavelengths. These technologies can identify plant diseases such as yellow rust and powdery mildew in wheat and leaf rust in sugar beet from early stages [109]. Moreover, multispectral imaging works on large scales by employing uncrewed aerial vehicles and satellites. Therefore, multispectral and hyperspectral imaging, along with other omics techniques, could be used to develop hormesis management protocols at a crop scale.
Given the intricacy of physiological responses, the elucidation of the adaptive mechanisms of plants to low-dose stress must be carried out from a multidimensional approach, utilizing comprehensive analyses for detecting the differential changes stimulated in different layers of the stress response. Understanding such mechanisms and, in particular, characterizing the hormetic dose-response curve allows eustress treatments to be implemented to enhance stress tolerance and increase food production and quality [7]. Nevertheless, integrating multiple-layer datasets gives rise to additional challenges beyond data collection and storage, including data management and processing [110]. Therefore, handling and modeling hormetic responses from multi-omics data requires computational methods for transforming data into knowledge.

4. Artificial Intelligence Applications in Plant Stress Science

High-throughput analyses of functional molecules such as genes, transcripts, proteins, and metabolites, produce a tremendous amount of data from biological systems. However, without proper processing, such data lack biological significance. Therefore, the advances in analytical methods and instrumentation have also generated the need for processing tools capable of describing mechanistic associations and interactions. The persisting escalation of computing power has triggered a diversification of artificial intelligence (AI) tools to address various problems in plant science. AI algorithms are remarkably advantageous to identify and classify individual characteristics within an extensive set of experimental data, and thus they are a promising means for analyzing plant stress mechanisms [104]. Furthermore, if we consider the accumulating evidence on the hormetic behavior of plant stress responses, intelligent algorithm applications in plant stress physiology could be helpful for predicting eustress responses that fall under the low-dose stimulation model.
A considerable number of AI applications on plant stress research implement Machine Learning (ML) and its subtype Deep Learning (DL). Such techniques have been applied in the paradigm of the four categories for analyzing the process of plant stress: identification, classification, quantification, and prediction (ICQP) [23]. Most of the published image-processing phenotyping studies use ML and DL tools to identify and classify stress symptoms, whereas the prediction of phenotypic traits before their expression is the most frequent application when analyzing genomic and transcriptomic datasets [111]. Figure 2 shows the four categories of the ICQP paradigm and the different applications for plant stress research integrating each category.
ML and DL techniques have extended the reach of traditional statistics for modeling non-linear systems such as biological processes [112]. Hence, both tools effectively process data to analyze plant responses to biotic and abiotic stressor incidence. Table 1 comprises recent studies examining plant stress using ML and DL techniques. Table 1 comprises some recent studies examining plant stress using ML and DL techniques. This summary was created by searching the academic databases ScienceDirect, Springer link, IEEE Xplore, and Google scholar. This review is up to date until January 2022, covering the work carried out from 2016 to 2022. The keywords used for this search were “artificial intelligence”, “machine learning”, “deep learning”, “plant stress”, “plant disease”, “plant resistance”, and “plant science”. It includes a classification of current research on plant stress elucidation using ML tools, emphasizing the algorithms used, the ICQP paradigm category on which each report lies, the stressors studied, and the datasets analyzed.
The majority of the studies listed in this literature revision are based on supervised algorithms whose quality depends on the data sources, the feature extraction from the available data, and the selection of the output variables and the learning algorithms [139]. The most frequently reported algorithms for successfully modeling plant stress responses are the Support Vector Machine (SVM), Random Forest (RF), and Convolutional Neural Networks (CNN). These techniques were predominantly used for identifying and classifying stress symptoms from extensive images datasets. Nevertheless, machine learning algorithms have also been used to predict many other various stress dose-dependent traits, such as disease-resistance gene expression [140], transcription factor expression [141], yield [142], growth [143], and specialized metabolites biosynthesis [144].

Deep Learning Platforms and Potential Applications for Plant Hormesis Management

Given that data available for understanding biological systems continues growing, the complexity of AI systems keeps increasing, and thus new high-performance artificial neural networks (ANN) were needed for the cases in which conventional ML techniques have fallen short. Such networks are currently known as deep learning, a technology that consists of the assembly of machine learning algorithms, increasing the number of levels and non-linear transformations in the neural networks and the efficiency of the training process [145].
Before the current heyday of DL, a disinterest towards it existed for several years, mainly due to hardware limitations and lack of funding. However, these techniques were reassessed as soon as more powerful hardware became available, especially Graphics Processing Units (GPUs) [146]. These devices were initially designed to compute three-dimensional graphics in video games and proved to be good performers of parallel computing; therefore, such systems were promptly used for processing DL algorithms. Nowadays, thanks to the multicores of modern GPUs allowing more and faster operations, DL is one of the most powerful AI tools to model complex non-linear systems and process an enormous number of experimental data.
The use of high-performance software allows DL to be used to evaluate many problems in systems biology. Computational models have been applied to simulate protein interactions, a universal process in biological systems, and a key process for understanding the mechanism of physiological responses [147]. Using in silico prediction of protein–protein and protein–ligand interactions, it is possible to estimate the activity of potential effector molecules and effectively test complete libraries with millions of molecules without performing costly and time-consuming experiments [148]. For plant systems, in silico prediction could offer a relevant platform for analyzing receptor binding during stressor perception and identifying new phytohormones or designing elicitors to achieve optimal responses for hormesis management [149].
When it comes to DL, there is an impression that an expensive computational infrastructure is necessary, which is not entirely false considering the cost of high-performance GPUs. However, there are many platforms for training DL models through cloud computing. For example, platforms such as Amazon Web Service (AWS), Google Colab, Microsoft Azure provide CPUs, Hard Disks, GPUs, or TPUs through the command terminal, which allows one to utilize high-performance hardware from an average computer with internet access and sufficient bandwidth. Additionally, some of these platforms offer a free version, meaning that cloud computing can be performed for a few hours without cost. This approach avoids the cost of a local computing cluster, the specialized space to house them, and the required electricity, all considerable limitations to perform this kind of computing due to the long time needed to train a DL model at a high performance. Table 2 shows the platforms or hardware used to train DL models applied to current plant science research.
In addition to exploiting hardware features, it is also essential to take account of the existing software to ensure an adequate running performance when training ML networks. Several platforms enable the implementation of ML and DL algorithms, and many of the available frameworks are open-source software, which has led to the rapid adoption of computer modeling for many agricultural technology [160]. Among the most employed is Caffe, a deep learning framework that has been developed by Berkeley AI Research (BAIR) and community contributors. There is also high-level software developed from C++ and C code, such as Open CV, and an increasing number of programs based on Python, such as Keras, Pytorch, scikit-learn, and Theano [161]. Moreover, the Microsoft Cognitive Toolkit (CNTK), used as a library or standalone ML tool, and TensorFlow, available from Google Brain, are two of the most widespread open-source platforms for executing DL tasks. Finally, it is also indispensable to consider the tools offered by Matlab and the Nvidia CUDA software to implement AI applications in agriculture [162].
Hormesis management increases crop yield and quality, stimulates specialized metabolism, and enhances stress tolerance [8]. Nevertheless, characterizing hormetic curves for several species and evaluating multiple stressor doses to produce an expected physiological response is a slow and expensive process. In addition, controlled elicitation studies show that a considerable crop extension or high technology greenhouses are necessary to evaluate the effect of low-dose stress for developing eustress management protocols [163,164]. Furthermore, it is necessary to consider thousands of individual plant markers such as genotype information, yield performance, and environmental data to propose effective treatments.
DL would be an efficient tool to address hormesis management limitations because it can expose complex non-linear relations between environmental conditions and gene expression to decipher gene networks and signaling pathways [165]. Convolutional Neural Networks (CNN) applied to image analysis is one of the most used biometric techniques in agriculture for evaluating plant identity, morphology [157,158], growth [159], disease [153,154,156], and pollution [150]. The CNN architecture is designed as a matrix for data analysis. It is structured so that, at several stages, filters segment the data and acquire specific information to train the deep neural network [166]. Modern plant analysis techniques can easily detect significant variables related to stress mechanisms with enough sensitivity for characterizing hormetic fluctuations. CNN could be trained with such datasets to model crop performance and predict phenotypic variables in response to low-dose stress (Figure 3).

5. Limitations, Challenges, and Future Outlook

In this review, we have discussed the advantages of ML for assessing research problems in plant-stressor systems from a subcellular to an ecosystem scale. Nevertheless, there are also relevant limitations to consider for proposing the implementation of ML tools, particularly for hormesis modeling.
Firstly, given the vast diversity of ML and DL platforms, selecting an appropriate architecture to carry out the proposed strategy constitutes a significant challenge. Furthermore, every architecture performs differently depending on the number and type of deep networks and the running hardware, complicating, even more, the tool selection. In response to this challenge, the number of scientific publications discussing ML tool-pairs, their performance, and new models designed specifically to perform a given task constantly increases [167].
The second challenge to assess refers to AI’s fundamental limitations. The power of ML methods offers advantages over conventional statistics, but they do not explicitly infer or provide confidence boundaries such as p values. This is a problem since scientists commonly rely on confidence intervals and model interpretation to support decision-making. Moreover, increasing the complexity of the network architecture turns ML systems into “black-boxes.” Consequently, most ML and DL methods do not allow a straightforward interpretation or the basis for the resulting predictions [168]. Therefore, an evaluation method after a training process is crucial. A high network intricacy and cumulative input datasets, central to analyzing plant systems, also require more computing power, significantly increasing the time and cost to complete tasks. The latter mainly affects multifaceted algorithms such as SVM and MLP [169,170].
The third challenge arises from the input data. The success of ML depends on the availability of appropriate databases, that is, extensive collections of data sharing specific features [111]. Nonetheless, although hormesis is increasingly being considered in plant research protocols, it still lacks the attention needed to form public databases to enrich model training. Moreover, data from biological systems are highly heterogeneous, and, as a result, detailed data curation and preprocess must be performed to ensure the accuracy of the training process [171]. Furthermore, even if multilayer data sets from plant hormesis research were available, there are simply too many different plant species interacting with changing environments. As a result, any group of experimentally acquired data results partial and unrelated to others. The use of model species is fundamental to depict basic biological processes, but these findings are not always transposable to evolutionarily distant plants or other species of interest. For these reasons, plant scientists must agree and standardize research methods for describing hormetic responses at all levels in representative plant species far and wide the phylogenetic tree. Considering the state-of-the-art discussion and the challenges that arise from conceiving the present proposal, the flow process depicted in Figure 4 conceptualizes the application of ML to model hormetic responses of plants to controlled stress exposure. However, there are still some constraints regarding the lack of experimentally adequate data to develop a robust model that could facilitate eustress doses determination and ultimately optimize hormesis management implementation for improving crop performance. For these reasons, future work on plant stress should emphasize the hormesis model and the construction of public knowledge databases, including plant phenotyping results and validated tools, models, and platforms.

6. Conclusions

The aim of elucidating plant stress responses is to develop cost-effective methods for producers to manipulate plant systems and obtain desirable phenotypes. Nevertheless, given the diversity of the technologies and methods currently available to measure variables associated with plant stress responses, the standardization of the experimental conditions and the integration of different dataset collections is a significant challenge. Moreover, most of the research around stress focuses on the adverse effects it causes on the plant system and completely ignores eustress and the hormetic behavior of plant defense.
Interestingly, it could be possible to develop robust models of plant responses if we consider that the behavior of stress responses is generalizable and varies within the limits of biological plasticity rather than depending only on the genetic identity or the developmental stage of individual systems [172]. However, even if we assess physiological responses from a hormetic approach, the big data challenge remains. In this respect, hormesis research should capitalize on the strengths of ML and DL for developing models capable of utilizing experimental data to predict which actions are required to improve crops traits. The latter would be especially beneficial when the eustress dose ranges of a stressor are unknown, and datasets from related crops are available.

Author Contributions

Conceptualization, A.K.R.-C. and R.G.G.-G.; writing—original draft preparation, A.K.R.-C., J.A.F., A.A.F.-J. and L.M.C.-M.; writing—review and editing, A.K.R.-C., L.M.C.-M., R.G.G.-G. and Q.H.-E.; visualization, A.K.R.-C., J.A.F. and A.A.F.-J.; supervision, R.G.G.-G. and Q.H.-E. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Autonomous University of Querétaro through the FOPER funding [FOPER-2021-FIN02480].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge Claudia Gutiérrez-Antonio for proofreading this manuscript and Marieke Vanthoor-Koopmans for revising this text for language use, grammar, and syntax. This work was supported by the Mexican federal government through the Consejo Nacional de Ciencia y Tecnología (CONACyT) [grant number 636395].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jansen, M.A.; Potters, G. Stress: The Way of Life. In Plant Stress Physiology, 2nd ed.; CABI: London, UK, 2017; pp. ix–xiv. [Google Scholar]
  2. O’Brien, P.; Kral-O’Brien, K.; Hatfield, J.L. Agronomic Approach to Understanding Climate Change and Food Security. Agron. J. 2021, 113, 4616–4626. [Google Scholar] [CrossRef]
  3. Vázquez-Hernández, M.; Parola-Contreras, I.; Montoya-Gómez, L.; Torres-Pacheco, I.; Schwarz, D.; Guevara-González, R. Eustressors: Chemical and Physical Stress Factors Used to Enhance Vegetables Production. Sci. Hortic. 2019, 250, 223–229. [Google Scholar] [CrossRef]
  4. Bienertova-Vasku, J.; Lenart, P.; Scheringer, M. Eustress and Distress: Neither Good nor Bad, but Rather the Same? BioEssays 2020, 42, 1900238. [Google Scholar] [CrossRef] [PubMed]
  5. Schirrmacher, V. Less Can Be More: The Hormesis Theory of Stress Adaptation in the Global Biosphere and Its Implications. Biomedicines 2021, 9, 293. [Google Scholar] [CrossRef] [PubMed]
  6. Agathokleous, E.; Kitao, M.; Calabrese, E.J. Hormesis: Highly Generalizable and beyond Laboratory. Trends Plant Sci. 2020, 25, 1076–1086. [Google Scholar] [CrossRef]
  7. Agathokleous, E.; Kitao, M.; Calabrese, E.J. Hormesis: A Compelling Platform for Sophisticated Plant Science. Trends Plant Sci. 2019, 24, 318–327. [Google Scholar] [CrossRef]
  8. Vargas-Hernandez, M.; Macias-Bobadilla, I.; Guevara-Gonzalez, R.G.; Romero-Gomez, S.d.J.; Rico-Garcia, E.; Ocampo-Velazquez, R.V.; Alvarez-Arquieta, L.d.L.; Torres-Pacheco, I. Plant Hormesis Management with Biostimulants of Biotic Origin in Agriculture. Front. Plant Sci. 2017, 8, 1762. [Google Scholar] [CrossRef]
  9. Aguirre-Becerra, H.; Vazquez-Hernandez, M.C.; Saenz de la, O.D.; Alvarado-Mariana, A.; Guevara-Gonzalez, R.G.; Garcia-Trejo, J.F.; Feregrino-Perez, A.A. Role of Stress and Defense in Plant Secondary Metabolites Production. In Bioactive Natural Products for Pharmaceutical Applications; Springer: Berlin/Heidelberg, Germany, 2021; pp. 151–195. [Google Scholar] [CrossRef]
  10. Rouphael, Y.; Kyriacou, M.C. Enhancing Quality of Fresh Vegetables through Salinity Eustress and Biofortification Applications Facilitated by Soilless Cultivation. Front. Plant Sci. 2018, 9, 1254. [Google Scholar] [CrossRef]
  11. Erofeeva, E.A. Plant Hormesis and Shelford’s Tolerance Law Curve. J. Res. 2021, 32, 1789–1802. [Google Scholar] [CrossRef]
  12. Jalal, A.; de Oliveira Junior, J.C.; Ribeiro, J.S.; Fernandes, G.C.; Mariano, G.G.; Trindade, V.D.R.; Dos Reis, A.R. Hormesis in Plants: Physiological and Biochemical Responses. Ecotoxicol. Environ. Saf. 2021, 207, 111225. [Google Scholar] [CrossRef]
  13. Agathokleous, E.; Barceló, D.; Calabrese, E.J. US EPA: Is There Room to Open a New Window for Evaluating Potential Sub-Threshold Effects and Ecological Risks? Environ. Pollut. 2021, 284, 117372. [Google Scholar] [CrossRef] [PubMed]
  14. Agathokleous, E.; Calabrese, E.J. Hormesis: The Dose Response for the 21st Century: The Future Has Arrived. Toxicology 2019, 425, 152249. [Google Scholar] [CrossRef] [PubMed]
  15. Mazhar, S.; Ditta, A.; Bulgariu, L.; Ahmad, I.; Ahmed, M.; Nadiri, A.A. Sequential treatment of paper and pulp industrial wastewater: Prediction of water quality parameters by Mamdani Fuzzy Logic model and phytotoxicity assessment. Chemosphere 2019, 227, 256–268. [Google Scholar] [CrossRef] [PubMed]
  16. Ahmad, Z.; Arshad, M.; Crowley, D.; Khoshnevisan, B.; Yousefi, M.; Imran, M.; Hussain, S. Comparative efficacy of ANN and ANFIS models in estimating biosurfactant production produced by Klebseilla sp. FKOD36. Stoch. Environ. Res. Risk Assess. 2016, 30, 353–363. [Google Scholar] [CrossRef]
  17. Soltis, P.S.; Nelson, G.; Zare, A.; Meineke, E.K. Plants Meet Machines: Prospects in Machine Learning for Plant Biology. Appl. Plant Sci. 2020, 8, e11371. [Google Scholar] [CrossRef]
  18. van Dijk, A.D.J.; Kootstra, G.; Kruijer, W.; de Ridder, D. Machine Learning in Plant Science and Plant Breeding. Iscience 2021, 24, 101890. [Google Scholar] [CrossRef]
  19. Mahood, E.H.; Kruse, L.H.; Moghe, G.D. Machine Learning: A Powerful Tool for Gene Function Prediction in Plants. Appl. Plant Sci. 2020, 8, e11376. [Google Scholar] [CrossRef]
  20. Benos, L.; Tagarakis, A.C.; Dolias, G.; Berruto, R.; Kateris, D.; Bochtis, D. Machine Learning in Agriculture: A Comprehensive Updated Review. Sensors 2021, 21, 3758. [Google Scholar] [CrossRef]
  21. Gao, Z.; Luo, Z.; Zhang, W.; Lv, Z.; Xu, Y. Deep Learning Application in Plant Stress Imaging: A Review. AgriEngineering 2020, 2, 430–446. [Google Scholar] [CrossRef]
  22. Singh, A.K.; Ganapathysubramanian, B.; Sarkar, S.; Singh, A. Deep Learning for Plant Stress Phenotyping: Trends and Future Perspectives. Trends Plant Sci. 2018, 23, 883–898. [Google Scholar] [CrossRef] [Green Version]
  23. Singh, A.; Ganapathysubramanian, B.; Singh, A.K.; Sarkar, S. Machine Learning for High-Throughput Stress Phenotyping in Plants. Trends Plant Sci. E 2016, 21, 110–124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Calabrese, E.J.; Agathokleous, E. Hormesis: Transforming Disciplines That Rely on the Dose Response. IUBMB Life 2022, 74, 8–23. [Google Scholar] [CrossRef] [PubMed]
  25. Calabrese, E.J. Historical Foundations of Hormesis. Homeopathy 2015, 104, 83–89. [Google Scholar] [CrossRef] [PubMed]
  26. Calabrese, E.J. Hormesis: Path and Progression to Significance. Int. J. Mol. Sci. 2018, 19, 2871. [Google Scholar] [CrossRef] [Green Version]
  27. Sun, H.; Calabrese, E.J.; Lin, Z.; Lian, B.; Zhang, X. Similarities between the Yin/Yang Doctrine and Hormesis in Toxicology and Pharmacology. Trends Pharm. Sci. 2020, 41, 544–556. [Google Scholar] [CrossRef]
  28. Agathokleous, E.; Calabrese, E.J. Hormesis Can Enhance Agricultural Sustainability in a Changing World. Glob. Food Secur. 2019, 20, 150–155. [Google Scholar] [CrossRef]
  29. Calabrese, E.J.; Mattson, M.P. Hormesis Provides a Generalized Quantitative Estimate of Biological Plasticity. J. Cell Commun. Signal. 2011, 5, 25–38. [Google Scholar] [CrossRef] [Green Version]
  30. Calabrese, E.J.; Blain, R.B. Hormesis and Plant Biology. Environ. Pollut. 2009, 157, 42–48. [Google Scholar] [CrossRef]
  31. Calabrese, E.J. Hormetic Mechanisms. Crit. Rev. Toxicol. 2013, 43, 580–606. [Google Scholar] [CrossRef]
  32. Mitchell, C.; Brennan, R.M.; Graham, J.; Karley, A.J. Plant Defense against Herbivorous Pests: Exploiting Resistance and Tolerance Traits for Sustainable Crop Protection. Front. Plant Sci. 2016, 7, 1132. [Google Scholar] [CrossRef] [Green Version]
  33. Lamers, J.; Van Der Meer, T.; Testerink, C. How Plants Sense and Respond to Stressful Environments. Plant Physiol. 2020, 182, 1624–1635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Jez, J.M.; Topp, C.N.; Silva, G.; Tomlinson, J.; Onkokesung, N.; Sommer, S.; Mrisho, L.; Legg, J.; Adams, I.P.; Gutierrez-Vazquez, Y. Plant Pest Surveillance: From Satellites to Molecules. Emerg. Top. Life Sci. 2021, 5, 275–287. [Google Scholar] [CrossRef] [PubMed]
  35. Ngou, B.P.M.; Jones, J.D.; Ding, P. Plant Immune Networks. Trends Plant Sci. 2021, 27, 255–273. [Google Scholar] [CrossRef] [PubMed]
  36. Yuan, M.; Ngou, B.P.M.; Ding, P.; Xin, X.-F. PTI-ETI Crosstalk: An Integrative View of Plant Immunity. Curr. Opin. Plant Biol. 2021, 62, 102030. [Google Scholar] [CrossRef] [PubMed]
  37. Zarattini, M.; Farjad, M.; Launay, A.; Cannella, D.; Soulié, M.-C.; Bernacchia, G.; Fagard, M. Every Cloud Has a Silver Lining: How Abiotic Stresses Affect Gene Expression in Plant-Pathogen Interactions. J. Exp. Bot. 2021, 72, 1020–1033. [Google Scholar] [CrossRef]
  38. Aerts, N.; Pereira Mendes, M.; Van Wees, S.C. Multiple Levels of Crosstalk in Hormone Networks Regulating Plant Defense. Plant J. 2021, 105, 489–504. [Google Scholar] [CrossRef] [PubMed]
  39. Saijo, Y.; Loo, E.P. Plant Immunity in Signal Integration between Biotic and Abiotic Stress Responses. New Phytol. 2020, 225, 87–104. [Google Scholar] [CrossRef] [Green Version]
  40. Jamil, I.N.; Remali, J.; Azizan, K.A.; Nor Muhammad, N.A.; Arita, M.; Goh, H.-H.; Aizat, W.M. Systematic Multi-Omics Integration (MOI) Approach in Plant Systems Biology. Front. Plant Sci. 2020, 11, 944. [Google Scholar] [CrossRef]
  41. Naithani, S.; Tripathi, J.N.; Kumar, D. Systems Biology Approach for Improving and Sustaining Agriculture. Curr. Plant Biol. 2021, 28, 100230. [Google Scholar] [CrossRef]
  42. Argueso, C.T.; Assmann, S.M.; Birnbaum, K.D.; Chen, S.; Dinneny, J.R.; Doherty, C.J.; Eveland, A.L.; Friesner, J.; Greenlee, V.R.; Law, J.A. Directions for Research and Training in Plant Omics: Big Questions and Big Data. Plant Direct 2019, 3, e00133. [Google Scholar] [CrossRef] [Green Version]
  43. Libault, M.; Pingault, L.; Zogli, P.; Schiefelbein, J. Plant Systems Biology at the Single-Cell Level. Trends Plant Sci. 2017, 22, 949–960. [Google Scholar] [CrossRef] [PubMed]
  44. Mir, R.R.; Reynolds, M.; Pinto, F.; Khan, M.A.; Bhat, M.A. High-throughput phenotyping for crop improvement in the genomics era. Plant Sci. 2019, 282, 60–72. [Google Scholar] [CrossRef] [PubMed]
  45. Purugganan, M.D.; Jackson, S.A. Advancing Crop Genomics from Lab to Field. Nat. Genet. 2021, 53, 595–601. [Google Scholar] [CrossRef] [PubMed]
  46. Lu, Z.; Cui, J.; Wang, L.; Teng, N.; Zhang, S.; Lam, H.-M.; Zhu, Y.; Xiao, S.; Ke, W.; Lin, J. Genome-Wide DNA Mutations in Arabidopsis Plants after Multigenerational Exposure to High Temperatures. Genome Biol. 2021, 22, 1–27. [Google Scholar] [CrossRef] [PubMed]
  47. Hou, S.; Zhu, G.; Li, Y.; Li, W.; Fu, J.; Niu, E.; Li, L.; Zhang, D.; Guo, W. Genome-Wide Association Studies Reveal Genetic Variation and Candidate Genes of Drought Stress Related Traits in Cotton (Gossypium Hirsutum L.). Front. Plant Sci. 2018, 9, 1276. [Google Scholar] [CrossRef]
  48. Hu, H.; Scheben, A.; Verpaalen, B.; Tirnaz, S.; Bayer, P.E.; Hodel, R.G.; Batley, J.; Soltis, D.E.; Soltis, P.S.; Edwards, D. Amborella Gene Presence/Absence Variation Is Associated with Abiotic Stress Responses That May Contribute to Environmental Adaptation. New Phytol. 2022, 233, 1548–1555. [Google Scholar] [CrossRef]
  49. Mo, F.; Wang, M.; Li, H.; Li, Y.; Li, Z.; Deng, N.; Chai, R.; Wang, H. Biological Effects of Silver Ions to Trifolium Pratense L. Revealed by Analysis of Biochemical Indexes, Morphological Alteration and Genetic Damage Possibility with Special Reference to Hormesis. Environ. Exp. Bot. 2021, 186, 104458. [Google Scholar] [CrossRef]
  50. Sertse, D.; You, F.M.; Ravichandran, S.; Soto-Cerda, B.J.; Duguid, S.; Cloutier, S. Loci Harboring Genes with Important Role in Drought and Related Abiotic Stress Responses in Flax Revealed by Multiple GWAS Models. Theor. Appl. Genet. 2021, 134, 191–212. [Google Scholar] [CrossRef]
  51. Luo, Z.; Szczepanek, A.; Abdel-Haleem, H. Genome-Wide Association Study (GWAS) Analysis of Camelina Seedling Germination under Salt Stress Condition. Agronomy 2020, 10, 1444. [Google Scholar] [CrossRef]
  52. Xiao, Q.; Bai, X.; Zhang, C.; He, Y. Advanced High-Throughput Plant Phenotyping Techniques for Genome-Wide Association Studies: A Review. J. Adv. Res. 2022, 35, 215–230. [Google Scholar] [CrossRef]
  53. Crossa, J.; Fritsche-Neto, R.; Montesinos-Lopez, O.A.; Costa-Neto, G.; Dreisigacker, S.; Montesinos-Lopez, A.; Bentley, A.R. The Modern Plant Breeding Triangle: Optimizing the Use of Genomics, Phenomics, and Enviromics Data. Front. Plant Sci. 2021, 12, 651480. [Google Scholar] [CrossRef] [PubMed]
  54. Fugeray-Scarbel, A.; Bastien, C.; Dupont-Nivet, M.; Lemarié, S. R2D2 Consortium Why and How to Switch to Genomic Selection: Lessons from Plant and Animal Breeding Experience. Front. Genet. 2021, 12, 1185. [Google Scholar]
  55. Zhan, X.; Lu, Y.; Zhu, J.; Botella, J.R. Genome Editing for Plant Research and Crop Improvement. J. Integr. Plant Biol. 2021, 63, 3–33. [Google Scholar] [CrossRef] [PubMed]
  56. Zappia, L.; Theis, F.J. Over 1000 tools reveal trends in the single-cell RNA-seq analysis landscape. Genome Biol 2021, 22, 1–8. [Google Scholar] [CrossRef]
  57. Imadi, S.R.; Kazi, A.G.; Ahanger, M.A.; Gucel, S.; Ahmad, P. Plant Transcriptomics and Responses to Environmental Stress: An Overview. J. Genet. 2015, 94, 525–537. [Google Scholar] [CrossRef]
  58. Gorbatova, I.V.; Kazakova, E.A.; Podlutskii, M.S.; Pishenin, I.A.; Bondarenko, V.S.; Dontsova, A.A.; Dontsov, D.P.; Snegirev, A.S.; Makarenko, E.S.; Bitarishvili, S.V. Studying Gene Expression in Irradiated Barley Cultivars: PM19L-like and CML31-like Expression as Possible Determinants of Radiation Hormesis Effect. Agronomy 2020, 10, 1837. [Google Scholar] [CrossRef]
  59. Duarte-Sierra, A.; Nadeau, F.; Angers, P.; Michaud, D.; Arul, J. UV-C Hormesis in Broccoli Florets: Preservation, Phyto-Compounds and Gene Expression. Postharvest Biol. Technol. 2019, 157, 110965. [Google Scholar] [CrossRef]
  60. Scott, G.; Dickinson, M.; Shama, G.; Rupar, M. A Comparison of the Molecular Mechanisms Underpinning High-Intensity, Pulsed Polychromatic Light and Low-Intensity UV-C Hormesis in Tomato Fruit. Postharvest Biol. Technol. 2018, 137, 46–55. [Google Scholar] [CrossRef]
  61. Volkova, P.Y.; Duarte, G.T.; Soubigou-Taconnat, L.; Kazakova, E.A.; Pateyron, S.; Bondarenko, V.S.; Bitarishvili, S.V.; Makarenko, E.S.; Churyukin, R.S.; Lychenkova, M.A. Early Response of Barley Embryos to Low-and High-dose Gamma Irradiation of Seeds Triggers Changes in the Transcriptional Profile and an Increase in Hydrogen Peroxide Content in Seedlings. J. Agron. Crop Sci. 2020, 206, 277–295. [Google Scholar] [CrossRef]
  62. Guo, J.; Ma, Z.; Peng, J.; Mo, J.; Li, Q.; Guo, J.; Yang, F. Transcriptomic Analysis of Raphidocelis Subcapitata Exposed to Erythromycin: The Role of DNA Replication in Hormesis and Growth Inhibition. J. Hazard. Mater. 2021, 402, 123512. [Google Scholar] [CrossRef]
  63. He, Y.; Wang, Y.; Hu, Y.; Chen, W.; Yan, Z. Superconducting Electrode Capacitor Based on Double-Sided YBCO Thin Film for Wireless Power Transfer Applications. Supercond. Sci. Technol. 2018, 32, 015010. [Google Scholar] [CrossRef]
  64. Arisha, M.H.; Ahmad, M.Q.; Tang, W.; Liu, Y.; Yan, H.; Kou, M.; Wang, X.; Zhang, Y.; Li, Q. RNA-Sequencing Analysis Revealed Genes Associated Drought Stress Responses of Different Durations in Hexaploid Sweet Potato. Sci. Rep. 2020, 10, 12573. [Google Scholar] [CrossRef]
  65. García-Gómez, M.L.; Castillo-Jiménez, A.; Martínez-García, J.C.; Álvarez-Buylla, E.R. Multi-Level Gene Regulatory Network Models to Understand Complex Mechanisms Underlying Plant Development. Curr. Opin. Plant Biol. 2020, 57, 171–179. [Google Scholar] [CrossRef] [PubMed]
  66. Haque, S.; Ahmad, J.S.; Clark, N.M.; Williams, C.M.; Sozzani, R. Computational Prediction of Gene Regulatory Networks in Plant Growth and Development. Curr. Opin. Plant Biol. 2019, 47, 96–105. [Google Scholar] [CrossRef]
  67. Wang, J.; Chen, B.; Ali, S.; Zhang, T.; Wang, Y.; Zhang, H.; Wang, L.; Zhang, Y.; Xie, L.; Jiang, T. Epigenetic Modification Associated with Climate Regulates Betulin Biosynthesis in Birch. J. Res. 2021, 1–15. [Google Scholar] [CrossRef]
  68. Qian, Y.; Huang, S.C. Improving Plant Gene Regulatory Network Inference by Integrative Analysis of Multi-Omics and High Resolution Data Sets. Curr. Opin. Syst. Biol. 2020, 22, 8–15. [Google Scholar] [CrossRef]
  69. Smith-Sonneborn, J. The Role of the ”Stress Protein Response” in Hormesis. In Biological Effects of Low Level Exposures to Chemicals and Radiation; CRC Press: Boca Raton, FL, USA, 2017; pp. 41–52. ISBN 1-315-15028-X. [Google Scholar]
  70. Koussounadis, A.; Langdon, S.P.; Um, I.H.; Harrison, D.J.; Smith, V.A. Relationship between Differentially Expressed MRNA and MRNA-Protein Correlations in a Xenograft Model System. Sci. Rep. 2015, 5, 10775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Sahoo, J.P.; Behera, L.; Sharma, S.S.; Praveena, J.; Nayak, S.K.; Samal, K.C. Omics Studies and Systems Biology Perspective towards Abiotic Stress Response in Plants. Am. J. Plant Sci. 2020, 11, 2172. [Google Scholar] [CrossRef]
  72. Buccitelli, C.; Selbach, M. MRNAs, Proteins and the Emerging Principles of Gene Expression Control. Nat. Rev. Genet. 2020, 21, 630–644. [Google Scholar] [CrossRef]
  73. Kosová, K.; Vítámvás, P.; Urban, M.O.; Prášil, I.T.; Renaut, J. Plant Abiotic Stress Proteomics: The Major Factors Determining Alterations in Cellular Proteome. Front. Plant Sci. 2018, 9, 122. [Google Scholar] [CrossRef] [Green Version]
  74. Mehmood, S.S.; Lu, G.; Luo, D.; Hussain, M.A.; Raza, A.; Zafar, Z.; Zhang, X.; Cheng, Y.; Zou, X.; Lv, Y. Integrated Analysis of Transcriptomics and Proteomics Provides Insights into the Molecular Regulation of Cold Response in Brassica Napus. Environ. Exp. Bot. 2021, 187, 104480. [Google Scholar] [CrossRef]
  75. Frukh, A.; Siddiqi, T.O.; Khan, M.I.R.; Ahmad, A. Modulation in Growth, Biochemical Attributes and Proteome Profile of Rice Cultivars under Salt Stress. Plant Physiol. Biochem. 2020, 146, 55–70. [Google Scholar] [CrossRef] [PubMed]
  76. Chawade, A.; Alexandersson, E.; Bengtsson, T.; Andreasson, E.; Levander, F. Targeted Proteomics Approach for Precision Plant Breeding. J. Proteome Res. 2016, 15, 638–646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Al-Obaidi, J.R. Proteoinformatics and Agricultural Biotechnology Research: Applications and Challenges. In Essentials of Bioinformatics; Springer: Berlin/Heidelberg, Germany, 2019; Volume III, pp. 1–27. [Google Scholar]
  78. Komatsu, S. Plant Proteomic Research 2.0: Trends and Perspectives. Int. J. Mol. Sci. 2019, 20, 2495. [Google Scholar] [CrossRef] [Green Version]
  79. Jorrin-Novo, J.V. What Is New in (Plant) Proteomics Methods and Protocols: The 2015–2019 Quinquennium. In Plant Proteomics; Springer: Berlin/Heidelberg, Germany, 2020; pp. 1–10. [Google Scholar] [CrossRef]
  80. Jan, R.; Asaf, S.; Numan, M.; Kim, K.-M. Plant Secondary Metabolite Biosynthesis and Transcriptional Regulation in Response to Biotic and Abiotic Stress Conditions. Agronomy 2021, 11, 968. [Google Scholar] [CrossRef]
  81. Kosmacz, M.; Sokołowska, E.M.; Bouzaa, S.; Skirycz, A. Towards a Functional Understanding of the Plant Metabolome. Curr. Opin. Plant Biol. 2020, 55, 47–51. [Google Scholar] [CrossRef]
  82. Weng, J.-K.; Lynch, J.H.; Matos, J.O.; Dudareva, N. Adaptive Mechanisms of Plant Specialized Metabolism Connecting Chemistry to Function. Nat. Chem. Biol. 2021, 17, 1037–1045. [Google Scholar] [CrossRef]
  83. Rinschen, M.M.; Ivanisevic, J.; Giera, M.; Siuzdak, G. Identification of Bioactive Metabolites Using Activity Metabolomics. Nat. Rev. Mol. Cell Biol. 2019, 20, 353–367. [Google Scholar] [CrossRef]
  84. Verpoorte, R.; Choi, Y.H.; Kim, H.K. Metabolomics: Will It Stay? Phytochem. Anal. PCA 2010, 21, 2–3. [Google Scholar] [CrossRef]
  85. Lautie, E.; Russo, O.; Ducrot, P.; Boutin, J.A. Unraveling Plant Natural Chemical Diversity for Drug Discovery Purposes. Front. Pharm. 2020, 11, 397. [Google Scholar] [CrossRef]
  86. Sharma, D.; Kumar, S.; Kumar, V.; Thakur, A. Comprehensive Review on Nutraceutical Significance of Phytochemicals as Functional Food Ingredients for Human Health Management. J. Pharm. Phytochem. 2019, 8, 385–395. [Google Scholar] [CrossRef] [Green Version]
  87. Pishenin, I.; Gorbatova, I.; Kazakova, E.; Podobed, M.; Mitsenyk, A.; Shesterikova, E.; Dontsova, A.; Dontsov, D.; Volkova, P. Free Amino Acids and Methylglyoxal as Players in the Radiation Hormesis Effect after Low-Dose γ-Irradiation of Barley Seeds. Agriculture 2021, 11, 918. [Google Scholar] [CrossRef]
  88. Mengdi, X.; Wenqing, C.; Haibo, D.; Xiaoqing, W.; Li, Y.; Yuchen, K.; Hui, S.; Lei, W. Cadmium-Induced Hormesis Effect in Medicinal Herbs Improves the Efficiency of Safe Utilization for Low Cadmium-Contaminated Farmland Soil. Ecotoxicol. Environ. Saf. 2021, 225, 112724. [Google Scholar] [CrossRef] [PubMed]
  89. Corrado, G.; Vitaglione, P.; Giordano, M.; Raimondi, G.; Napolitano, F.; Di Stasio, E.; Di Mola, I.; Mori, M.; Rouphael, Y. Phytochemical Responses to Salt Stress in Red and Green Baby Leaf Lettuce (Lactuca Sativa L.) Varieties Grown in a Floating Hydroponic Module. Separations 2021, 8, 175. [Google Scholar] [CrossRef]
  90. Alvarado, A.M.; Aguirre-Becerra, H.; Vázquez-Hernández, M.; Magaña-Lopez, E.; Parola-Contreras, I.; Caicedo-Lopez, L.H.; Contreras-Medina, L.M.; Garcia-Trejo, J.F.; Guevara-Gonzalez, R.G.; Feregrino-Perez, A.A. Influence of Elicitors and Eustressors on the Production of Plant Secondary Metabolites. In Natural Bio-Active Compounds; Springer: Berlin/Heidelberg, Germany, 2019; pp. 333–388. [Google Scholar] [CrossRef]
  91. Villate, A.; San Nicolas, M.; Gallastegi, M.; Aulas, P.-A.; Olivares, M.; Usobiaga, A.; Etxebarria, N.; Aizpurua-Olaizola, O. Metabolomics as a Prediction Tool for Plants Performance under Environmental Stress. Plant Sci. 2021, 303, 110789. [Google Scholar] [CrossRef]
  92. Miggiels, P.; Wouters, B.; van Westen, G.J.; Dubbelman, A.-C.; Hankemeier, T. Novel Technologies for Metabolomics: More for Less. TrAC Trends Anal. Chem. 2019, 120, 115323. [Google Scholar] [CrossRef]
  93. Hong, J.; Yang, L.; Zhang, D.; Shi, J. Plant Metabolomics: An Indispensable System Biology Tool for Plant Science. Int. J. Mol. Sci. 2016, 17, 767. [Google Scholar] [CrossRef]
  94. Pinu, F.R.; Beale, D.J.; Paten, A.M.; Kouremenos, K.; Swarup, S.; Schirra, H.J.; Wishart, D. Systems Biology and Multi-Omics Integration: Viewpoints from the Metabolomics Research Community. Metabolites 2019, 9, 76. [Google Scholar] [CrossRef] [Green Version]
  95. Pieruschka, R.; Schurr, U. Plant Phenotyping: Past, Present, and Future. Plant Phenomics 2019, 2019, 7507131. [Google Scholar] [CrossRef]
  96. Arif, Y.; Singh, P.; Siddiqui, H.; Bajguz, A.; Hayat, S. Salinity Induced Physiological and Biochemical Changes in Plants: An Omic Approach towards Salt Stress Tolerance. Plant Physiol. Biochem. 2020, 156, 64–77. [Google Scholar] [CrossRef]
  97. Singh, V.; Sharma, N.; Singh, S. A Review of Imaging Techniques for Plant Disease Detection. Artif. Intell. Agric. 2020, 4, 229–242. [Google Scholar] [CrossRef]
  98. Zheng, C.; Abd-Elrahman, A.; Whitaker, V. Remote Sensing and Machine Learning in Crop Phenotyping and Management, with an Emphasis on Applications in Strawberry Farming. Remote Sens. 2021, 13, 531. [Google Scholar] [CrossRef]
  99. Li, D.; Li, C.; Yao, Y.; Li, M.; Liu, L. Modern Imaging Techniques in Plant Nutrition Analysis: A Review. Comput. Electron. Agric. 2020, 174, 105459. [Google Scholar] [CrossRef]
  100. Martinez, H.E.; de Souza, B.P.; Caixeta, E.T.; de Carvalho, F.P.; Clemente, J.M. Water Deficit Changes Nitrate Uptake and Expression of Some Nitrogen Related Genes in Coffee-Plants (Coffea Arabica L.). Sci. Hortic. 2020, 267, 109254. [Google Scholar] [CrossRef]
  101. Susič, N.; Žibrat, U.; Širca, S.; Strajnar, P.; Razinger, J.; Knapič, M.; Vončina, A.; Urek, G.; Stare, B.G. Discrimination between Abiotic and Biotic Drought Stress in Tomatoes Using Hyperspectral Imaging. Sens. Actuators B Chem. 2018, 273, 842–852. [Google Scholar] [CrossRef] [Green Version]
  102. Mayta, M.L.; Hajirezaei, M.-R.; Carrillo, N.; Lodeyro, A.F. Leaf Senescence: The Chloroplast Connection Comes of Age. Plants 2019, 8, 495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Agathokleous, E.; Feng, Z.; Peñuelas, J. Chlorophyll Hormesis: Are Chlorophylls Major Components of Stress Biology in Higher Plants? Sci. Total Environ. 2020, 726, 138637. [Google Scholar] [CrossRef]
  104. Fenu, G.; Malloci, F.M. Forecasting Plant and Crop Disease: An Explorative Study on Current Algorithms. Big Data Cogn. Comput. 2021, 5, 2. [Google Scholar] [CrossRef]
  105. Pérez-Bueno, M.L.; Pineda, M.; Barón, M. Phenotyping Plant Responses to Biotic Stress by Chlorophyll Fluorescence Imaging. Front. Plant Sci. 2019, 1135. [Google Scholar] [CrossRef]
  106. Jakusch, P.; Kocsis, T.; Székely, I.K.; Hatvani, I.G. The Application of Magnetic Resonance Imaging (Mri) to the Examination of Plant Tissues and Water Barriers. Acta Biol. Hung. 2018, 69, 423–436. [Google Scholar] [CrossRef]
  107. Pineda, M.; Barón, M.; Pérez-Bueno, M.-L. Thermal Imaging for Plant Stress Detection and Phenotyping. Remote Sens. 2021, 13, 68. [Google Scholar] [CrossRef]
  108. Benavente, E.; García-Toledano, L.; Carrillo, J.; Quemada, M. Thermographic Imaging: Assessment of Drought and Heat Tolerance in Spanish Germplasm of Brachypodium Distachyon. Procedia Environ. Sci. 2013, 19, 262–266. [Google Scholar] [CrossRef] [Green Version]
  109. Lowe, A.; Harrison, N.; French, A.P. Hyperspectral Image Analysis Techniques for the Detection and Classification of the Early Onset of Plant Disease and Stress. Plant Methods 2017, 13, 80. [Google Scholar] [CrossRef] [PubMed]
  110. Großkinsky, D.K.; Svensgaard, J.; Christensen, S.; Roitsch, T. Plant Phenomics and the Need for Physiological Phenotyping across Scales to Narrow the Genotype-to-Phenotype Knowledge Gap. J. Exp. Bot. 2015, 66, 5429–5440. [Google Scholar] [CrossRef] [Green Version]
  111. Singh, A.; Jones, S.; Ganapathysubramanian, B.; Sarkar, S.; Mueller, D.; Sandhu, K.; Nagasubramanian, K. Challenges and Opportunities in Machine-Augmented Plant Stress Phenotyping. Trends Plant Sci. 2021, 26, 53–69. [Google Scholar] [CrossRef] [PubMed]
  112. Osama, K.; Mishra, B.N.; Somvanshi, P. Machine Learning Techniques in Plant Biology. In PlantOmics: The Omics of Plant Science; Springer: Berlin/Heidelberg, Germany, 2015; pp. 731–754. [Google Scholar]
  113. Chandel, N.S.; Chakraborty, S.K.; Rajwade, Y.A.; Dubey, K.; Tiwari, M.K.; Jat, D. Identifying Crop Water Stress Using Deep Learning Models. Neural Comput. Appl. 2021, 33, 5353–5367. [Google Scholar] [CrossRef]
  114. Yu, K.; Fang, S.; Zhao, Y. Heavy Metal Hg Stress Detection in Tobacco Plant Using Hyperspectral Sensing and Data-Driven Machine Learning Methods. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2021, 245, 118917. [Google Scholar] [CrossRef]
  115. Rahman, C.R.; Arko, P.S.; Ali, M.E.; Khan, M.A.I.; Apon, S.H.; Nowrin, F.; Wasif, A. Identification and Recognition of Rice Diseases and Pests Using Convolutional Neural Networks. Biosyst. Eng. 2020, 194, 112–120. [Google Scholar] [CrossRef] [Green Version]
  116. Blumenthal, J.; Megherbi, D.B.; Lussier, R. Unsupervised Machine Learning via Hidden Markov Models for Accurate Clustering of Plant Stress Levels Based on Imaged Chlorophyll Fluorescence Profiles & Their Rate of Change in Time. Comput. Electron. Agric. 2020, 174, 105064. [Google Scholar] [CrossRef]
  117. Esgario, J.G.; Krohling, R.A.; Ventura, J.A. Deep Learning for Classification and Severity Estimation of Coffee Leaf Biotic Stress. Comput. Electron. Agric. 2020, 169, 105162. [Google Scholar] [CrossRef] [Green Version]
  118. Das, B.; Manohara, K.; Mahajan, G.; Sahoo, R.N. Spectroscopy Based Novel Spectral Indices, PCA-and PLSR-Coupled Machine Learning Models for Salinity Stress Phenotyping of Rice. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 229, 117983. [Google Scholar] [CrossRef] [PubMed]
  119. Moghimi, A.; Yang, C.; Marchetto, P.M. Ensemble Feature Selection for Plant Phenotyping: A Journey from Hyperspectral to Multispectral Imaging. IEEE Access 2018, 6, 56870–56884. [Google Scholar] [CrossRef]
  120. Barbedo, J.G.A. Plant Disease Identification from Individual Lesions and Spots Using Deep Learning. Biosyst. Eng. 2019, 180, 96–107. [Google Scholar] [CrossRef]
  121. Dao, P.D.; He, Y.; Proctor, C. Plant Drought Impact Detection Using Ultra-High Spatial Resolution Hyperspectral Images and Machine Learning. Int. J. Appl. Earth Obs. Geoinf. 2021, 102, 102364. [Google Scholar] [CrossRef]
  122. Karthickmanoj, R.; Sasilatha, T.; Padmapriya, J. Automated Machine Learning Based Plant Stress Detection System. Mater. Today Proc. 2021, 47, 1887–1891. [Google Scholar] [CrossRef]
  123. Ghosal, S.; Blystone, D.; Singh, A.K.; Ganapathysubramanian, B.; Singh, A.; Sarkar, S. An Explainable Deep Machine Vision Framework for Plant Stress Phenotyping. Proc. Natl. Acad. Sci. USA 2018, 115, 4613–4618. [Google Scholar] [CrossRef] [Green Version]
  124. Zahid, A.; Dashtipour, K.; Abbas, H.T.; Mabrouk, I.B.; Al-Hasan, M.; Ren, A.; Imran, M.A.; Alomainy, A.; Abbasi, Q.H. Machine Learning Enabled Identification and Real-Time Prediction of Living Plants’ Stress Using Terahertz Waves. Def. Technol. 2022; in press. [Google Scholar] [CrossRef]
  125. Niu, Y.; Han, W.; Zhang, H.; Zhang, L.; Chen, H. Estimating Fractional Vegetation Cover of Maize under Water Stress from UAV Multispectral Imagery Using Machine Learning Algorithms. Comput. Electron. Agric. 2021, 189, 106414. [Google Scholar] [CrossRef]
  126. Kang, D.; Ahn, H.; Lee, S.; Lee, C.-J.; Hur, J.; Jung, W.; Kim, S. Identifying Stress-Related Genes and Predicting Stress Types in Arabidopsis Using Logical Correlation Layer and CMCL Loss through Time-Series Data. In Proceedings of the 2018 IEEE International Conference on Bioinformatics and Biomedicine (BIBM), Madrid, Spain, 3–6 December 2018; IEEE: Piscataway, NJ, USA, 2018; pp. 399–404. [Google Scholar] [CrossRef]
  127. Azimi, S.; Kaur, T.; Gandhi, T.K. A Deep Learning Approach to Measure Stress Level in Plants Due to Nitrogen Deficiency. Measurement 2021, 173, 108650. [Google Scholar] [CrossRef]
  128. Fürtauer, L.; Pschenitschnigg, A.; Scharkosi, H.; Weckwerth, W.; Nägele, T. Combined Multivariate Analysis and Machine Learning Reveals a Predictive Module of Metabolic Stress Response in Arabidopsis Thaliana. Mol. Omics 2018, 14, 437–449. [Google Scholar] [CrossRef] [Green Version]
  129. Khanna, R.; Schmid, L.; Walter, A.; Nieto, J.; Siegwart, R.; Liebisch, F. A Spatio Temporal Spectral Framework for Plant Stress Phenotyping. Plant Methods 2019, 15, 13. [Google Scholar] [CrossRef] [Green Version]
  130. Naik, H.S.; Zhang, J.; Lofquist, A.; Assefa, T.; Sarkar, S.; Ackerman, D.; Singh, A.; Singh, A.K.; Ganapathysubramanian, B. A Real-Time Phenotyping Framework Using Machine Learning for Plant Stress Severity Rating in Soybean. Plant Methods 2017, 13, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Pereira, D.R.; Papa, J.P.; Saraiva, G.F.R.; Souza, G.M. Automatic Classification of Plant Electrophysiological Responses to Environmental Stimuli Using Machine Learning and Interval Arithmetic. Comput. Electron. Agric. 2018, 145, 35–42. [Google Scholar] [CrossRef] [Green Version]
  132. Mondal, M.; Edida, M.; Sharma, N.; Lall, B.; Raju, D. Plants Stress Response Detection by Selecting Minimal Bands of Hyperspectral Images. In Proceedings of the 2019 9th International Conference on Advances in Computing and Communication (ICACC), Kochi, India, 6–8 November 2019; IEEE: Piscataway, NJ, USA, 2019; pp. 230–233. [Google Scholar] [CrossRef]
  133. Venal, M.C.A.; Fajardo, A.C.; Hernandez, A.A. Plant Stress Classification for Smart Agriculture Utilizing Convolutional Neural Network-Support Vector Machine. In Proceedings of the 2019 International Conference on ICT for Smart Society (ICISS), Bandung, Indonesia, 19–20 November 2019; IEEE: Piscataway, NJ, USA, 2019; Volume 7, pp. 1–5. [Google Scholar] [CrossRef]
  134. González-Camacho, J.M.; Crossa, J.; Pérez-Rodríguez, P.; Ornella, L.; Gianola, D. Genome-Enabled Prediction Using Probabilistic Neural Network Classifiers. BMC Genom. 2016, 17, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Vakilian, A.K. Machine Learning Improves Our Knowledge about MiRNA Functions towards Plant Abiotic Stresses. Sci. Rep. 2020, 10, 3041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Shikha, M.; Kanika, A.; Rao, A.R.; Mallikarjuna, M.G.; Gupta, H.S.; Nepolean, T. Genomic Selection for Drought Tolerance Using Genome-Wide SNPs in Maize. Front. Plant Sci. 2017, 8, 550. [Google Scholar] [CrossRef] [PubMed]
  137. Montesinos-López, A.; Montesinos-López, O.A.; Gianola, D.; Crossa, J.; Hernández-Suárez, C.M. Multi-Environment Genomic Prediction of Plant Traits Using Deep Learners with Dense Architecture. G3 Genes Genomes Genet. 2018, 8, 3813–3828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Ly, D.; Huet, S.; Gauffreteau, A.; Rincent, R.; Touzy, G.; Mini, A.; Jannink, J.-L.; Cormier, F.; Paux, E.; Lafarge, S. Whole-Genome Prediction of Reaction Norms to Environmental Stress in Bread Wheat (Triticum Aestivum L.) by Genomic Random Regression. Field Crops Res. 2018, 216, 32–41. [Google Scholar] [CrossRef]
  139. Silva, J.C.F.; Teixeira, R.M.; Silva, F.F.; Brommonschenkel, S.H.; Fontes, E.P. Machine Learning Approaches and Their Current Application in Plant Molecular Biology: A Systematic Review. Plant Sci. 2019, 284, 37–47. [Google Scholar] [CrossRef]
  140. Hiddar, H.; Rehman, S.; Lakew, B.; Verma, R.P.S.; Al-Jaboobi, M.; Moulakat, A.; Kehel, Z.; Filali-Maltouf, A.; Baum, M.; Amri, A. Assessment and Modeling Using Machine Learning of Resistance to Scald (Rhynchosporium Commune) in Two Specific Barley Genetic Resources Subsets. Sci. Rep. 2021, 11, 15967. [Google Scholar] [CrossRef]
  141. Song, Q.; Lee, J.; Akter, S.; Rogers, M.; Grene, R.; Li, S. Prediction of Condition-Specific Regulatory Genes Using Machine Learning. Nucleic Acids Res. 2020, 48, e62. [Google Scholar] [CrossRef] [Green Version]
  142. Shook, J.; Gangopadhyay, T.; Wu, L.; Ganapathysubramanian, B.; Sarkar, S.; Singh, A.K. Crop Yield Prediction Integrating Genotype and Weather Variables Using Deep Learning. PLoS ONE 2021, 16, e0252402. [Google Scholar] [CrossRef] [PubMed]
  143. Yasrab, R.; Zhang, J.; Smyth, P.; Pound, M.P. Predicting Plant Growth from Time-Series Data Using Deep Learning. Remote Sens. 2021, 13, 331. [Google Scholar] [CrossRef]
  144. García-Pérez, P.; Zhang, L.; Miras-Moreno, B.; Lozano-Milo, E.; Landin, M.; Lucini, L.; Gallego, P.P. The Combination of Untargeted Metabolomics and Machine Learning Predicts the Biosynthesis of Phenolic Compounds in Bryophyllum Medicinal Plants (Genus Kalanchoe). Plants 2021, 10, 2430. [Google Scholar] [CrossRef] [PubMed]
  145. Dargan, S.; Kumar, M.; Ayyagari, M.R.; Kumar, G. A Survey of Deep Learning and Its Applications: A New Paradigm to Machine Learning. Arch. Comput. Methods Eng. 2020, 27, 1071–1092. [Google Scholar] [CrossRef]
  146. Nicodeme, C. Build Confidence and Acceptance of AI-Based Decision Support Systems-Explainable and Liable AI. In Proceedings of the 2020 13th International Conference on Human System Interaction (HSI), Tokyo, Japan, 6–8 June 2020; IEEE: Piscataway, NJ, USA, 2020; pp. 20–23. [Google Scholar] [CrossRef]
  147. Pavlopoulou, A.; Karaca, E.; Balestrazzi, A.; Georgakilas, A.G. In Silico Phylogenetic and Structural Analyses of Plant Endogenous Danger Signaling Molecules upon Stress. Oxidative Med. Cell. Longev. 2019, 2019, 8683054. [Google Scholar] [CrossRef] [Green Version]
  148. Stokes, J.M.; Yang, K.; Swanson, K.; Jin, W.; Cubillos-Ruiz, A.; Donghia, N.M.; MacNair, C.R.; French, S.; Carfrae, L.A.; Bloom-Ackermann, Z. A Deep Learning Approach to Antibiotic Discovery. Cell 2020, 180, 688–702. [Google Scholar] [CrossRef] [Green Version]
  149. Wang, Y.; Zhou, M.; Zou, Q.; Xu, L. Machine Learning for Phytopathology: From the Molecular Scale towards the Network Scale. Brief. Bioinform. 2021, 22, bbab037. [Google Scholar] [CrossRef]
  150. Mayr, A.; Klambauer, G.; Unterthiner, T.; Hochreiter, S. DeepTox: Toxicity Prediction Using Deep Learning. Front. Environ. Sci. 2016, 3, 80. [Google Scholar] [CrossRef] [Green Version]
  151. Prilianti, K.R.; Setiyono, E.; Kelana, O.H.; Brotosudarmo, T.H.P. Deep Chemometrics for Nondestructive Photosynthetic Pigments Prediction Using Leaf Reflectance Spectra. Inf. Processing Agric. 2021, 8, 194–204. [Google Scholar] [CrossRef]
  152. Mu, C.; Yuan, Z.; Ouyang, X.; Sun, P.; Wang, B. Non-destructive Detection of Blueberry Skin Pigments and Intrinsic Fruit Qualities Based on Deep Learning. J. Sci. Food Agric. 2021, 101, 3165–3175. [Google Scholar] [CrossRef]
  153. Durmuş, H.; Güneş, E.O.; Kırcı, M. Disease Detection on the Leaves of the Tomato Plants by Using Deep Learning. In Proceedings of the 2017 6th International Conference on Agro-Geoinformatics, Fairfax, VA, USA, 7–10 August 2017; IEEE: Piscataway, NJ, USA, 2017; pp. 1–5. [Google Scholar] [CrossRef]
  154. Ferentinos, K.P. Deep Learning Models for Plant Disease Detection and Diagnosis. Comput. Electron. Agric. 2018, 145, 311–318. [Google Scholar] [CrossRef]
  155. Too, E.C.; Yujian, L.; Njuki, S.; Yingchun, L. A Comparative Study of Fine-Tuning Deep Learning Models for Plant Disease Identification. Comput. Electron. Agric. 2019, 161, 272–279. [Google Scholar] [CrossRef]
  156. Wang, G.; Sun, Y.; Wang, J. Automatic Image-Based Plant Disease Severity Estimation Using Deep Learning. Comput. Intell. Neurosci. 2017, 2017, 2917536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Wu, C.; Zeng, R.; Pan, J.; Wang, C.C.; Liu, Y.-J. Plant Phenotyping by Deep-Learning-Based Planner for Multi-Robots. IEEE Robot. Autom. Lett. 2019, 4, 3113–3120. [Google Scholar] [CrossRef]
  158. Ma, W.; Qiu, Z.; Song, J.; Li, J.; Cheng, Q.; Zhai, J.; Ma, C. A Deep Convolutional Neural Network Approach for Predicting Phenotypes from Genotypes. Planta 2018, 248, 1307–1318. [Google Scholar] [CrossRef]
  159. Tausen, M.; Clausen, M.; Moeskjær, S.; Shihavuddin, A.; Dahl, A.B.; Janss, L.; Andersen, S.U. Greenotyper: Image-Based Plant Phenotyping Using Distributed Computing and Deep Learning. Front. Plant Sci. 2020, 11, 1181. [Google Scholar] [CrossRef]
  160. Khan, A.I.; Al-Badi, A. Open Source Machine Learning Frameworks for Industrial Internet of Things. Procedia Comput. Sci. 2020, 170, 571–577. [Google Scholar] [CrossRef]
  161. Bresilla, K.; Perulli, G.D.; Boini, A.; Morandi, B.; Corelli Grappadelli, L.; Manfrini, L. Single-Shot Convolution Neural Networks for Real-Time Fruit Detection within the Tree. Front. Plant Sci. 2019, 10, 611. [Google Scholar] [CrossRef] [Green Version]
  162. Thomas, G.; Balocco, S.; Mann, D.; Simundsson, A.; Khorasani, N. Intelligent Agricultural Machinery Using Deep Learning. IEEE Instrum. Meas. Mag. 2021, 24, 93–100. [Google Scholar] [CrossRef]
  163. Valencia-Hernandez, J.-A.; Solano-Alvarez, N.; Rico-Rodriguez, M.-A.; Rodriguez-Ontiveros, A.; Torres-Pacheco, I.; Rico-Garcia, E.; Guevara-Gonzalez, R.-G. Eustressic Dose of Cadmium in Soil Induces Defense Mechanisms and Protection Against Clavibacter Michiganensis in Tomato (Solanum Lycopersicum L.). J. Plant Growth Regul. 2022, 1–8. [Google Scholar] [CrossRef]
  164. Sáenz-de la, O.D.; Morales, L.O.; Strid, Å.; Torres-Pacheco, I.; Guevara-González, R.G. Ultraviolet-B Exposure and Exogenous Hydrogen Peroxide Application Lead to Cross-tolerance toward Drought in Nicotiana Tabacum L. Physiol. Plant. 2021, 173, 666–679. [Google Scholar] [CrossRef] [PubMed]
  165. Khaki, S.; Wang, L. Crop Yield Prediction Using Deep Neural Networks. Front. Plant Sci. 2019, 10, 621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. LeCun, Y.; Bengio, Y.; Hinton, G. Deep Learning. Nature 2015, 521, 436–444. [Google Scholar] [CrossRef] [PubMed]
  167. Naser, M.Z.; Kodur, V.; Thai, H.T.; Hawileh, R.; Abdalla, J.; Degtyarev, V.V. StructuresNet and FireNet: Benchmarking databases and machine learning algorithms in structural and fire engineering domains. J. Build. Eng. 2021, 44, 102977. [Google Scholar] [CrossRef]
  168. Liang, Y.; Li, S.; Yan, C.; Li, M.; Jiang, C. Explaining the black-box model: A survey of local interpretation methods for deep neural networks. Neurocomputing 2021, 419, 168–182. [Google Scholar] [CrossRef]
  169. Nalepa, J.; Kawulok, M. Selecting training sets for support vector machines: A review. Artif. Intell. Rev. 2019, 52, 857–900. [Google Scholar] [CrossRef] [Green Version]
  170. Sarker, I.H. Machine Learning: Algorithms, Real-World Applications and Research Directions. SN Comput. Sci. 2021, 2, 160. [Google Scholar] [CrossRef]
  171. Xu, C.; Jackson, S.A. Machine learning and complex biological data. Genome Biol. 2019, 20, 76. [Google Scholar] [CrossRef]
  172. Harfouche, A.L.; Jacobson, D.A.; Kainer, D.; Romero, J.C.; Harfouche, A.H.; Mugnozza, G.S.; Moshelion, M.; Tuskan, G.A.; Keurentjes, J.J.; Altman, A. Accelerating Climate Resilient Plant Breeding by Applying Next-Generation Artificial Intelligence. Trends Biotechnol. 2019, 37, 1217–1235. [Google Scholar] [CrossRef]
Figure 1. The hormetic behavior of plant stress responses. At low doses, an overcompensation of the damage caused by the stressor increases plant fitness, whereas, at high doses, the stressors disrupt the homeostasis of the organism.
Figure 1. The hormetic behavior of plant stress responses. At low doses, an overcompensation of the damage caused by the stressor increases plant fitness, whereas, at high doses, the stressors disrupt the homeostasis of the organism.
Plants 11 00970 g001
Figure 2. The ICQP paradigm of the four categories for analyzing the stress process of plants. The uses of ML and DL in plant science are summarized in these four general applications. A wide range of datasets can be used for the design of the intelligent algorithms.
Figure 2. The ICQP paradigm of the four categories for analyzing the stress process of plants. The uses of ML and DL in plant science are summarized in these four general applications. A wide range of datasets can be used for the design of the intelligent algorithms.
Plants 11 00970 g002
Figure 3. Hormesis characterization through Deep Learning. Plant science uses highly sensitive techniques for detecting variations in gene expression, phenotype, and metabolism caused by environmental interactions. Deep learning, particularly through the implementation of Convolutional Neural Networks (CNN), decision trees, and Support Vector Machine (SVM) algorithms, allows big data processing and interpretation for modeling non-linear biological processes, such as hormesis.
Figure 3. Hormesis characterization through Deep Learning. Plant science uses highly sensitive techniques for detecting variations in gene expression, phenotype, and metabolism caused by environmental interactions. Deep learning, particularly through the implementation of Convolutional Neural Networks (CNN), decision trees, and Support Vector Machine (SVM) algorithms, allows big data processing and interpretation for modeling non-linear biological processes, such as hormesis.
Plants 11 00970 g003
Figure 4. Process of ML implementation for improving hormesis management. Analyzing plant stress responses generates many data, and ML integrates data to model complex systems. Considering the hormetic behavior of plant responses, ML could be used to model dose-response and predict eustress doses, simplifying controlled elicitation in agriculture.
Figure 4. Process of ML implementation for improving hormesis management. Analyzing plant stress responses generates many data, and ML integrates data to model complex systems. Considering the hormetic behavior of plant responses, ML could be used to model dose-response and predict eustress doses, simplifying controlled elicitation in agriculture.
Plants 11 00970 g004
Table 1. Machine learning-based studies in plant stress under the Identification, Classification, Quantification, and Prediction (ICQP) paradigm.
Table 1. Machine learning-based studies in plant stress under the Identification, Classification, Quantification, and Prediction (ICQP) paradigm.
Artificial
Intelligence Technique
Algorithms(ICQP)
Application
DatasetsModel Plant
Reported
StressorReference
Deep Learning (image)Convolutional neural networks (CNN), AlexNet, GoogLeNet, and Inception V3Identification1200 images acquired by camera under stress and non-stress conditionsMaize (Zea mays), okra (Abelmoschus esculentus), and soybean (Glycine max)Water stressChandel et al. (2020) [113]
Unsupervised Machine learningLeast squares discriminant analysis (PLS-DA) and least-squares support vector machine (LS-SVM)IdentificationHyperspectral images of the canopy of tobacco plantsTobaccoHeavy metal stress HgYu et al. (2021) [114]
Deep Learning (image)CNNIdentification1426 images of rice diseases and pests from paddy fieldsRiceBiotic stressRahman et al. (2020) [115]
Unsupervised Machine learning (video imaging)Hidden Markov models (HMMs)Identification and classificationChlorophyll fluorescence (ChlF) digital profiles from GrowTech Inc.Phaseolus vulgaris L. (Snap bean)Stressor “level” groups (low, medium, and high stressed) and three stressor “type” categories (drought, nutrient, and chemical stress)Blumenthal et al. (2020) [116]
Deep Learning (image)CNNIdentification and Quantification1747 smartphones images of arabica coffee leaves.Arabica coffeeBiotic stress; leaf miner, rust, brown leaf spot, and Cercospora leaf spotEsgario et al. (2020) [117]
Supervised Machine Learning, Partial Least Square Regression, Principal Component Analysis, and combined modelsK-nearest neighbors (KNN)Identification and classificationSpectral signature of leaf samples obtained with a visible, near-infrared spectrometerRiceSalt stressDas et al. (2020) [118]
Supervised Machine LearningReliefF, support vector machine (SVM), recursive feature elimination (RFE), and random forest (RF)Identification and classificationHyperspectral images from four wheat linesWheatSalt stressMoghimi et al. (2018) [119]
Deep Learning (image)CNNIdentification and classification1575 images (smartphones, compact cameras, DSLRDifferent plant specimensBiotic stressArnal Barbedo (2019) [120]
Deep LearningRF, SVM, multilayer perceptron (MLP)Identification and classificationHyperspectral imagesBromus inermisDrought stressDao et al. (2021) [121]
Supervised Machine LearningSVMIdentification and classificationRGB leave images from the Kaggle databaseBrinjal leavesBiotic stressKarthickmanoj et al. (2021) [122]
Deep Learning (image)Deep convolutional neural network (DCNN)Identification, classification, and quantificationCollection of images of stressed and healthy soybean leaflets in the fieldSoybean [Glycine max (L.) Merr.]Bacterial blight (Pseudomonas savastanoi pv. glycinea), bacterial pustule (Xanthomonas axonopodis pv. glycines), sudden death syndrome (Fusarium virguliforme), septoria brown spot (Septoria glycines), frogeye leaf spot (Cercospora sojina), iron deficiency chlorosis, potassium deficiency, and herbicide injuryGhosal et al. (2018) [123]
Supervised Machine LearningRF, SVM, KNNClassification and predictionReal time terahertz time-domain spectroscopic data (THz-TDS)Basil, coriander, parsley, baby-leaf, coffee, pea-Water StressZahid et al. (2022) [124]
Supervised Machine LearningRF, artificial neural networks (ANN), andClassificationMultispectral imagesMaizeWater stressNiu et al. (2021) [125]
Supervised Machine LearningConfident multiple-choice learningIdentification and predictionGene expression time-series datasetsArabidopsis thalianaHeat, cold, salt, and droughtKang et al. (2018) [126]
Deep Learning (image)CNNClassificationImages of Sorghum plant shoot from the Donald Danforth Plant Science Center.Sorghum plantsNitrogen deficiencyAzimi et al. (2021) [127]
Supervised Machine LearningDecision tree (DT), SVM, and Naïve Bayes (NB)ClassificationMetabolite and protein contentArabidopsis thalianaMetabolic stressFürtauer et al. (2018) [128]
Supervised Machine LearningSVMClassificationBiweekly RGB, stereo and hyperspectral spatio-temporal imagesSugar beet plantsAbiotic stress conditions (drought and nitrogen deficiency) and one biotic stressor (weed)Khanna et al. (2019) [129]
Supervised Machine LearningHierarchical modelsClassification5916 RGB images (493 plots including Plant Introduction (PI) accessions in different time points)Soybean (Glycine max (L.) Merr.)Iron deficiency chlorosisNaik et al. (2017) [130]
Supervised Machine LearningANN, CNN, optimum-path forest, KNN, and SVMClassificationElectrical signal under cold, low light and osmotic stimuli.Soybean plantsCold, low light, and osmotic stimuli.Pereira et al. (2018) [131]
Supervised Machine LearningRFClassificationHyperspectral dataset acquired from the Indian Agricultural Research Institute (IARI)WheatWater stressMondal et al. (2019) [132]
Deep Learning (image)CNN, SVMClassification65,184 labeled images from Github resourcesSoybeanBiotic (fungal and bacterial diseases) and abiotic (nutrient deficiency and chemical injury) stressesVenal et al. (2019) [133]
Supervised Machine LearningMLP and probabilistic neural network (PNN)Classification16 maize and 17 wheat genomic and phenotypic datasets with different trait-environment combinationsMaize and WheatDroughtGonzález-Camacho et al. (2016) [134]
Supervised Machine LearningDecision tree (DT), SVM, and NBPredictionmiRNA concentration.Arabidopsis thaliana plantsDrought, salinity, cold, and heatVakilian (2020) [135]
Supervised Machine LearningRidge regression, LASSO, elastic net, RF, reproducing kernel Hilbert space, Bayes A and Bayes BPredictionA set of 29,619 cured Single Nucleotide Polymorphisms, genotyped across a panel of 240 maize inbred linesMaizeDrought stressShikha et al. (2017) [136]
Deep Learning CNNPredictionThree maize and six wheat data sets.Maize and wheatEnvironmental stressMontesinos-López et al. (2018) [137]
Supervised Machine LearningGenomic random regressionPredictionComplete genotypes, molecular markers, and phenotypic traits of stressed and control groups.WheatEnvironmental stressLy et al. (2018) [138]
Table 2. Deep Learning architecture, hardware, and applications.
Table 2. Deep Learning architecture, hardware, and applications.
DL Architecture ApplicationHardwareReference
Deep Neural NetworksToxicity PredictionNvidia Tesla K40Mayr et al. (2016) [150]
Convolutional Neural NetworkPhotosynthetic pigments PredictionCPU core i5 1.6 GHz, 8 GB DDR3 RAM, GPU not specifiedPrilianti et al. (2020) [151]
Convolutional Neural NetworkPigments PredictionNvidia GTX 1020Ti, Intel Xeon W-2133, 32 GBMu et al. (2020) [152]
AlexNet and SqueezeNetPlant Disease Detection Nvidia Jetson TX1Durmus et al. (2017) [153]
Convolutional Neural NetworkPlant Disease DetectionNvidia GTX1080Ferentinos (2018) [154]
Convolutional Neural NetworkPlant Disease DetectionNvidia Tesla K40cToo et al. (2019) [155]
Deconvolutional Neural NetworkPlant Disease DetectionNvidia GeForce Titan X, Intel Core I7 3.5 GHzWang et al. (2017) [156]
Point Completion NetworkPlant PhenotypingNvidia Titan V. Xeon Gold 6146 3.20 GHz, 128 GB RAMWu et al. (2019) [157]
Deep Convolutional Neural NetworkPredicting Phenotypes from GenotypesNvidia GeForce TITAN-XGPUMa et al. (2018) [158]
U-netPhenotyping and Plant GrowthNvidia Tela V100Tausen et al. (2020) [159]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rico-Chávez, A.K.; Franco, J.A.; Fernandez-Jaramillo, A.A.; Contreras-Medina, L.M.; Guevara-González, R.G.; Hernandez-Escobedo, Q. Machine Learning for Plant Stress Modeling: A Perspective towards Hormesis Management. Plants 2022, 11, 970. https://doi.org/10.3390/plants11070970

AMA Style

Rico-Chávez AK, Franco JA, Fernandez-Jaramillo AA, Contreras-Medina LM, Guevara-González RG, Hernandez-Escobedo Q. Machine Learning for Plant Stress Modeling: A Perspective towards Hormesis Management. Plants. 2022; 11(7):970. https://doi.org/10.3390/plants11070970

Chicago/Turabian Style

Rico-Chávez, Amanda Kim, Jesus Alejandro Franco, Arturo Alfonso Fernandez-Jaramillo, Luis Miguel Contreras-Medina, Ramón Gerardo Guevara-González, and Quetzalcoatl Hernandez-Escobedo. 2022. "Machine Learning for Plant Stress Modeling: A Perspective towards Hormesis Management" Plants 11, no. 7: 970. https://doi.org/10.3390/plants11070970

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop