Next Article in Journal
Typification of Names in Iris (Iridaceae) Described by Victor Janka with Taxonomic Considerations
Next Article in Special Issue
Dissection of Hyperspectral Reflectance to Estimate Photosynthetic Characteristics in Upland Cotton (Gossypium hirsutum L.) under Different Nitrogen Fertilizer Application Based on Machine Learning Algorithms
Previous Article in Journal
Multilocus Data Analysis Reveal the Diversity of Cryptic Species in the Tillandsia ionantha (Bromeliaceae: Tillansiodeae) Complex
Previous Article in Special Issue
Optimal Deployment of WSN Nodes for Crop Monitoring Based on Geostatistical Interpolations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Utilization of Spectral Indices for High-Throughput Phenotyping

1
Department of Applied Biosciences, Kyungpook National University, Daegu 41566, Korea
2
Horticultural and Herbal Crop Environment Division, National Institute of Horticultural and Herbal Science, Rural Development Administration, Wanju 55365, Korea
3
Department of Engineering Technology, University of Houston, Texas, TX 77204, USA
4
Crop Production Technology Research Division, National Institute of Crop Science, Rural Development Administration, Miryang 50424, Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Plants 2022, 11(13), 1712; https://doi.org/10.3390/plants11131712
Submission received: 30 May 2022 / Revised: 20 June 2022 / Accepted: 24 June 2022 / Published: 28 June 2022

Abstract

:
The conventional plant breeding evaluation of large sets of plant phenotypes with precision and speed is very challenging. Thus, consistent, automated, multifaceted, and high-throughput phenotyping (HTP) technologies are becoming increasingly significant as tools to aid conventional breeding programs to develop genetically improved crops. With rapid technological advancement, various vegetation indices (VIs) have been developed. These VI-based imaging approaches, linked with artificial intelligence and a variety of remote sensing applications, provide high-throughput evaluations, particularly in the field of precision agriculture. VIs can be used to analyze and predict different quantitative and qualitative aspects of vegetation. Here, we provide an overview of the various VIs used in agricultural research, focusing on those that are often employed for crop or vegetation evaluation, because that has a linear relationship to crop output, which is frequently utilized in crop chlorophyll, health, moisture, and production predictions. In addition, the following aspects are here described: the importance of VIs in crop research and precision agriculture, their utilization in HTP, recent photogrammetry technology, mapping, and geographic information system software integrated with unmanned aerial vehicles and its key features. Finally, we discuss the challenges and future perspectives of HTP technologies and propose approaches for the development of new tools to assess plants’ agronomic traits and data-driven HTP resolutions for precision breeding.

1. Introduction

By 2050, the human population is projected to reach 9.7 billion, and the increasing food demand for the rising population would require a 25% to 70% increase in crop production compared to its current levels [1]. Achieving this target, however, poses a challenge to plant breeders, because food production will have to grow at a rate of 2.4% each year, but the actual growth rate is only 1.3%, with yields stagnating in up to 40% of cereal fields [2]. To increase crop production while reducing its impact on the environment, sustainable intensification in agriculture needs to be developed by adopting modern techniques [3]. One of the easiest ways to maximize the efficiency of major food and feed crops worldwide is to boost plant breeding and propagation.
In recent years, high-throughput phenotyping (HTP) has attracted considerable attention, leading to the creation of many new protocols for documenting different plant traits of particular interest [4]. The crop improvement research field, however, must adapt to address the growing threats posed by climate change [5]. The effects of climate change vary depending on geographical areas, demanding targeted strategies for the world’s numerous agroecological regions. Due to inadequate access to capital and the adoption of less productive farming methods, farmers from low-income countries are highly vulnerable to climate change [6]. As the growing crop production will increase the environmental footprint of agriculture, a sustainable intensification of agricultural practices will be required. Agricultural practices including modern plant breeding, crop cultivation, crop management [3], and HTP methods are ideally suited for improving crop productivity, as well as developing climate-resilient crops. Due to constant environmental changes and their impact on plant growth, it is necessary to gather data from more plant samples to identify common structural and physiological features. For this reason, HTP aims to characterize a large number of populations accurately [7]. Together with HTP, modern plant breeding or new genomics tools, such as genomic selection technology, have been considered the frontiers of crop breeding programs [8]. The current progress in sensors, aeronautics, and computing systems are paving the way for the development of effective field-based phenotyping platforms [9].
Most common HTP platforms consist of the utilization of unmanned vehicles containing various image sensors which are: (i) red-green-blue (RGB) imaging, (ii) infrared (IR) imaging, (iii) hyperspectral imaging, (iv) thermal imaging, (v) fluorescence imaging, (vi) light detection and ranging (LiDAR), and (vii) satellite imaging [10] (Figure 1). Among them, RGB and hyperspectral imaging tools are widely used to evaluate the quantitative and qualitative attributes of plants. In most cases, raw image data need to be converted to other indexes, because plants show different reflectance values at specific wavelengths depending on their growth stages and conditions (Kim et al., 2021). Typically, vegetation indices (VIs) are broadly used to measure plant health and growth conditions (soil property, water or moisture, and nutrient content).
VIs represents the statistical transformation of the initial spectral reflectance, and they are obtained using two or more wavebands to measure and interpret the incidence and condition of vegetation, such as the biomass and canopy attributes [11,12,13,14]. These indices allow reliable spatial and temporal intercomparisons of terrestrial photosynthetic activity and canopy structural variations [14]. The VIs obtained from remote sensing-based canopies are simple and effective algorithms for the quantitative and qualitative evaluation of vegetation cover, robustness, and growth dynamics. As they are simple transformations of spectral bands, they are computed directly without any bias or assumptions regarding land cover class, soil type, or climatic conditions. VIs assist researchers in monitoring seasonal, interannual, and long-term structural variations, as well as the phenological and biophysical parameters of vegetation cover [14]. As previously mentioned, VIs can predict plant growth conditions using numerous spectral reflectance bands, which can easily be acquired by spectral cameras, therefore they are particularly adequate in HTP. However, there is limited information about the use of VIs in the agricultural sector for multiple traits (physiological, biotic and abiotic), replicated multi environment and large populations evaluation. Thus, it is of great interest to understand different VI and HTP applications for crop improvement. In this study, we provide an overview of VIs and HTP and summarize the utilization of VIs for HTP in the agricultural research on the basis of an intensive literature survey.

2. Types of Vegetative Indices in Crop Research

In general, most substances including plants have three interactions/responses to electromagnetic energy (for example light) that is reflected, transmitted, or absorbed. In other words, each material or vegetation has its spectral signature/reflectance which can be measured in three key wavelength bands (Figure 1). In the case of plants, the unique interaction with solar radiation distinguishes them from other natural materials. Plants use photosynthesis to generate energy and, as a result, absorb a lot more blue and red light. On the other hand, they strongly reflect green and near-infrared (NIR) light [10]. Plants consist of various materials—such as water, nutrients, and pigments—therefore, they present variations across the spectrum, which provide important information related to water, nutrient, and pigment contents [10,16]. The spectrum variation is often described as VIs for the prediction of plant growth conditions [11].
Over the years VI research has evolved, and numerous new VIs have been developed to understand the aerial imagery included in the literature [17,18,19,20]. The chronological information details of most of the vegetation indices developed and reported by researchers are presented in (Table 1). Among them, 25 are broadly used in plant research because they are very effective in determining various features, such as leaf greenness, light use efficiency, leaf pigment, and water content [10,20]. Most VIs are calculated by two or more reflectance wavelengths, which are normally involved in the concentration of photosynthetic pigments [10]. However, among these VIs, very few have been critically compared or tested. Therefore, here we describe several selected indices that are most often used for crop or vegetation evaluation, and that have the linear link with crop yield, pigments, health, and canopy water content, and are also often used in agricultural production estimation.

2.1. Determination of Leaf Greenness

The VIs associated with leaf greenness determine photosynthetic pigments, which are present in green vegetation. They record chlorophyll and utilize the red and NIR wavelength spectrum to assess green vegetation. The following two VIs are widely used by researchers as leaf greenness determinants.
i. Normalized difference vegetation index (NDVI): Typically a simple and suitable indicator, first proposed by Rouse Jr. et al. [97], this index calculates and differentiates vegetation from non-vegetation zones based on reflectance in the red and NIR wavelengths. NDVI is sensitive to chlorophyll content. The values of this index are in the range (−1.0–1.0). The common range for healthy green vegetation is 0.2 to 0.9. The following Equation (1) is used for its calculation:
NDVI = R N I R R R E D R N I R + R R E D
where RNIR and RRED represent NIR wavebands and spectral reflectance in the red respectively.
ii. Modified chlorophyll absorption in reflectance index (MCARI): MCARI is an alternative index that was originally defined by [98], and was mainly developed to study variations in chlorophyll content. However, researchers consider MCARI as very sensitive to low chlorophyll content, as it is influenced by non-photosynthetic materials/pigments, background noise, and reflectance. The value of this index ranges from −1 to 1. The common range for healthy green vegetation is 0.2 to 0.7. The following Equation (2) is used to define this index:
MCARI = [ ( R 700 R 670 ) 0.2   ×   ( R 700 R 550 ) ]   ×   ( R 700 R 670 )
where R denotes the reflectance for chlorophyll absorption at the 670 nm waveband relative to the reflectance at the 550 nm and 700 nm wavebands.

2.2. Determination of Light Use Efficiency

The VIs used to determine light use efficiency are indicators of how well vegetation can absorb incident light to perform photosynthesis. These VIs are useful in precision agriculture, as they assist in the estimation of growth and productivity. In addition, they utilize visible spectrum reflectance measurements to assess the vegetation’s total light consumption efficiency by taking into account interactions between different pigments. A couple of well-established light use efficiency VIs are mentioned below.
i. Photochemical reflectance index (PRI): PRI is a reflectance assessment of green vegetation that is sensitive to variations in carotenoid pigments, particularly xanthophyll. Changes in carotenoid pigments are indicative of per unit energy absorbed during photosynthesis as light use efficiency, or of the rate of carbon dioxide absorbed by green vegetation. The range for a PRI is −1 to 1, where healthy vegetation generally often ranges between −0.2 and 0.2. PRI was originally known as the “physiological reflectance index” [99,100,101] and the equation below is used for its calculation:
PRI = R 531 R 570 R 531 + R 570
where R denotes the reflectance at the waveband of concern, particularly light absorption for xanthophyll at the 531 nm waveband, and to minimize the effects of chloroplast mobility, a reference waveband was employed at 570 nm.
ii. Red-green ratio index (RGRI): This index uses the broad wavelengths called red band (600–699 nm), and green band (500–599 nm), and is a reflectance measurement that compares anthocyanin-induced leaf redness to chlorophyll. The red-green ratio has been used to forecast the development of vegetation in canopies, and also to determine leaf production and stress and, in some instances, flowering. The range of an RGRI is from 0.1 to >8, where healthy green vegetation usually falls between values of 0.7 to 3. This index is defined by the equation below proposed by [78]:
R G R I = R RED : R GREEN
where R is the reflectance, RED refers to a broad range red band at 600–699 nm, and GREEN to a broad range green band at 500–599 nm.

2.3. Determination of Leaf Pigment

Leaf pigment VIs are used to determine pigments (anthocyanin and carotenoid), which are present in higher amounts in stressed vegetation, and they do not record chlorophyll. The following two VIs are widely used as leaf pigment determinants.
i. Anthocyanin reflectance index 1 (AR1): This index is based on reflectance measurements in the visible spectrum and quantifies stressed vegetation. Typically, alterations of, or increases in ARI1 suggest canopy modification caused by either new growth or death. The values of this index ranged from 0 to 0.2; however, it can vary with the plant species from −∞ to ∞. The estimate of anthocyanin accumulation is calculated by the equation below described in [21]:
ARI 1 = 1     R 550 1 R 700
where R is the reciprocal reflectance at the waveband of concern, 550 nm and 700 nm.
ii. Carotenoid reflectance index 1 (CRI1): this index is considered as a reflectance measure sensitive to the carotenoid pigments found in plant leaves and is mostly applicable in the measurement of stressed vegetation. Typically, stressed or weak vegetation presents a greater accumulation of carotenoids. The values of this index range from 0 to >15. The common range for green vegetation is between values of 1 and 12. CRI1 is calculated based on the well-defined equation mentioned below [24]:
CRI 1   =   1 R 510 1 R 550  
where R is the reciprocal reflectance at the waveband of concern; at 510 nm both carotenoids and chlorophyll affect reflectance, and at 550 nm, only chlorophyll affects reflectance.

2.4. Determination of Canopy Water Content

The VIs that determines canopy water content use reflectance measurements in the near-infrared and shortwave infrared ranges. These VIs provide an indication of how much water is contained within the leafy canopy. The presence of a high water content in vegetation suggests that plants are healthier and more likely to grow faster and tolerate extreme conditions (high temperature, fires, and drought). The following VIs are widely applicable for the measurement of canopy water content.
i. Moisture stress index (MSI): This index is linearly correlated to the relative water content in leaves. Unlike in other water VIs, the relationship between values of the MSI is reversed; higher values indicate a higher water stress and a lower water content. The values of this index range from 0 to >3. The common range for green vegetation is 0.4 to 2. The following equation is used for the calculation of this index [60,61]:
MSI = ( R 1599 R 819 )
where R is the reflectance at the waveband of concern; water absorbs shortwave infrared light at wavebands around 1599 nm. Absorption at 819 nm is used as a reference since it does not get affected by changes in water content.
ii. Normalized multiband drought index (NMDI): This index was initially proposed for monitoring soil and vegetation moisture from space, as was described in [71]. Index values range from 0.7 to 1 for dry soil, 0.6 to 0.7 for soil with intermediate moisture, and less than 0.6 for wet soil. It was developed based on three different wavelengths (NIR, centered around 860 nm, and two others in the shortwave infrared, centered around 1640 nm and 2130 nm), and is calculated as follows:
NMDI = R 860 ( R 1640 ρ 2130 ) R 860 + ( R 1640   + ρ 2130 )
where R is the reflectance at the waveband of concern, particularly reflectance observed by a satellite sensor at the 860 nm, 1640 nm, and 2130 nm bands, respectively. It uses the difference (slope) between two absorption bands (1640 nm and 2130 nm), as the soil and vegetation water-sensitive band.
Apart from the above-mentioned widely used VIs, with recent technological advancement, several additional VIs have also been developed. Examples include the ratio spectral index and normalized difference spectral index, which are used for recognizing vegetation stress caused by hydrocarbon pollution [102], the enhanced vegetation index (EVI) [14,103], the perpendicular vegetation index, and the atmospherically resistant vegetation index (ARVI) [104]. Furthermore, these VIs improve sensitivity in high biomass environments by reducing soil background (brightness, color) and other atmospheric impacts (cloud, leaf canopy shadow, and cloud shadow). However, different VIs have their own distinct set of spectra that reveal information regarding specific plant properties. Depending on individual research objectives or interests, these VIs can be integrated into agriculture for both quantitative and qualitative evaluation of vegetation to obtain precise information about crops. However, most of these indices have indeed been evaluated only for a few species, so it is unclear whether they will be applicable across a wide range of plant species with different leaf characteristics and phenotypic, or morphological, expression. Thus, further comparative and validation studies of the numerous VIs developed for crop assessment are needed.

3. Utilization of Vegetation Indices in Crop Research

3.1. Importance in Breeding Programs

The continuous advancement of breeding techniques allows an increase in the pace of genetic enhancement [105]. The selection of desired plants based on phenotypes has been performed by farmers and plant breeders long before DNA and molecular markers were discovered. Crop breeding is a game of numbers: the more crosses and conditions used for selection, the higher the chance of superior varieties to be found. Plant breeders must be able to determine the best progeny easily and specifically, by phenotyping vast numbers of lines. There is a need to improve breeding productivity to satisfy potential future needs. Advances in high-throughput genotyping have produced fast and inexpensive genomic knowledge, and this technology has paved the way for the creation of thousands of recombinant inbred lines for phenotyping broad mapping populations and diversity panels [106]. As molecular breeding strategies (such as marker-assisted recurrent selection and genomic selection) concentrate mostly on genotypic information-based choices, phenotypic information is still needed [107]. Similarly, to classify promising events in transgenic experiments, phenotyping is also needed [108,109]. To capitalize on advances in traditional, molecular, and transgenic breeding, and ensure the genetic enhancement of crops, successful phenotyping is likely to be necessary. The demand for efficient phenotyping methods has been introduced in numerous domains. In many cases, phenotypes are robust predictors of important biological traits, such as disease and mortality [110]. Molecular biologists and breeders believe that advanced molecular techniques can only be useful in breeding if the collection of quantitative traits is based on reliable phenotyping techniques [111]. High-performance phenotyping techniques can transform the plant breeding process by accelerating the generation advancing process [8].

3.2. Importance in Precision Agriculture

Overall, yield and plant performance are linked to traditional phenotypes—such as seedling vigor, days to flowering time, and terminal plant height—and a considerable amount of research tried to measure these phenotypes by HTP [8,9,16,112]. Therefore, more advancements in phenotyping techniques are needed to enhance breeding programs, accelerate genetic gains, and simplify the tracking of plants’ health status to minimize qualitative and quantitative losses during crop development. For phenotyping approaches to be approved by both breeders and producers, cost-effectiveness is a very important criterion for adopting the technique. Sustainable intensification in cultivation allows bred cultivars to obtain an optimum yield capacity, ensure more accurate crop management, and reduce the use of fertilizers, chemicals, and irrigation [112]. Phenotyping also focuses on optimizing predictable yields and maintaining the efficiency of crops [3]. In comparison to plant breeding, precise agriculture is used to enhance management methods, which results in the collection of suitable genotypes [113]. The main challenges that can be addressed by field phenotyping are irrigation, fertilization, and disease control to secure crop production. In phenotyping, various monitoring techniques for the evaluation, severity, and recording of plant health are essential. Research studies on the development of a suitable combination of sensors, vehicles, and analysis methods to customize phenotyping, based on specific interests, are promising in this sector [4].
In plants, phenotypic variation can be the result of a complex web of interactions between genotypes and the environment [114]. Therefore, precision phenotyping can determine which component in the set of phenotypes can predict specific parameters in the field, and the acquisition of phenotypic data allows a more precise description of G × E interactions [7]. Precise phenotyping facilitates the identification of quantitative trait loci, which can govern yield across different environmental conditions; but within the context of a more effective transition from the results of stress-related studies to improved cultivars, the difficulty of implementing accurate phenotyping is a widely recognized obstacle [115]. Thus, phenotyping techniques capable of revealing plant responses to environmental challenges, in both in-lab and field experimental conditions, are required [116]. In a more general sense, as they represent the technical basis of plant phenomics, precision phenotyping methods are both intensively and extensively required.

4. Image Processing Software

Manually monitoring crop growth stages and vegetation features on a broad scale is a time-consuming and laborious technique, and it is difficult to sample the entire area. Nowadays, unmanned aerial vehicles (UAVs) are being used more than satellites and manned aircraft, because these have a lower spatial and temporal resolution and a higher equipment cost [117]. As a result, UAVs have gained significant popularity, and their use for different applications related to agriculture, to obtain the required information about crops and field status is widely studied [118,119,120,121,122]. Furthermore, imaging sensors characterized by a low weight and high spatial and temporal resolutions have been developed for UAVs in recent years [123,124], and UAV imaging techniques are readily available, due to the upgrading of imaging and computation programs [125]. Table 2 presents recent photogrammetry and mapping software that have been integrated into UAVs, and describes their key parameters, this information was compiled based on the scientific papers published from 2007 to 2020. These techniques, which were previously only limited to other aerial vehicles, are now being used on UAVs, resulting in the promotion of frequent surveying of agricultural fields at a reasonable cost [125].
The expanding capabilities of UAVs, enhanced by the use of geographic information systems (GIS) and by technological breakthroughs in imaging methods, have increased the importance of this type of vehicle in precision agriculture [126]. Table 3 lists some of the recent pieces of GIS software that have been integrated into UAVs and describes their key features based on the prime scientific publications related to UAVs and remote sensing from 2007 to 2020. The change in vegetation features over time and space can be detected with an UAV fitted with a multispectral camera [126]. In healthy plants, a high reflection of infrared light and low reflection of red light were observed, whereas the opposite pattern (i.e., high reflection of red, and low reflection of infrared light) was revealed in unhealthy plants [126]. UAVs equipped with various sensors assist farmers in monitoring the health status of crops, irrigation requirements, insect pest incidence, and time of harvesting, and enable them to instantly take steps toward such issues [127,128]. HTP using a commercial UAV platform requires the selection of phenotypic traits for the breeding of target plants and, based on this, the appropriate sensor and UAV are then selected [118,119,120]. Nuijten et al. (2019) used a drone with high-resolution optical imagery to comprehend the potential of drone data for the evaluation of crop productivity on a large scale.
Remote sensing technology has become one of the most promising HTP technologies, as it provides nondestructive measurements of crop performance in both controlled and field environments [16]. In addition to its high flexibility and easy operation, remote sensing also helps to extract crop phenotypic information quickly and precisely [129]. Thus, remote sensing imaging products with a high spatial and temporal resolution and a light weight, have proven to be a valid technology and a suitable aerial platform for precision agriculture and plant breeding [16].
Finally, novel advances in remote sensing technology are expected to be quickly accepted and implemented by UAV platforms, making them one of the most important applications in HTP in the near future.
Table 3. List of recent geographic information system software used for unmanned aerial vehicles (UAVs).
Table 3. List of recent geographic information system software used for unmanned aerial vehicles (UAVs).
Geographic Information System SoftwareFeaturesManufacturer
ArcGISEasy operation, compatible with the model builder or Python; supports visualization, analysis, and maintenance of data in 2D, 3D, and 4D; wide range of data sharing; users can operate the ArcGIS system via Web GIS ESRI
QGISFeatures are similar to those of ArcGIS; supports both bitmap and vector layers; it is free and open-source QGIS
ERDAS ImagineSimplifies image processing; includes a wide range of tools for the analysis of image data; allows graphical editing; includes hyperspectral and multispectral data tools and LiDAR tools; allows spatial modelingHexagon geospatial [130]
ENVIEasy to use; supports visualization, processing, and analysis of all types of geospatial dataL3Harris geospatial [131]
Grass GISLicense-free use, open-source; compatible with SQL programming language, geocoding, 2D, and 3D raster analysis; provides LiDAR tools, raster, and vector statisticsGrass Development Team [132]

5. Challenges in High Throughput Phenotyping

The major purpose of phenotyping techniques is to incorporate different phenotypic methods for evaluating agronomic traits, abiotic and biotic stress, and also assess factors that contribute to the yield potential of the crop in order to increase the genetic potential and use in crop improvement. In general, HTP in the plant is still challenging, as the development and utilization of techniques for the accurate recording of important agronomic, biotic and abiotic traits, and monitoring of crops have not been promoted to their full potential. In this context, there are several challenges for the HTP, for instance, complex/quantitative traits, root phenotyping, environmental effect, multilocation and replicated trials field plot measurement. Particularly, employing the HTP at the spatial and temporal resolutions of tissue or cellular level, morphologies, micro phenotypes, and below the soil traits is very challenging [133]. The major challenges faced by HTP are the identification of nondestructive, easy, operational, highly repeatable, robust, efficient, low-cost, and fast phenotypic tools. Moreover, a huge quantity of data is generated by HTP so storing, managing, processing these HTP data and making meaningful biological information are also very challenging [94,134]. However, with the emerging new tools, researchers have proposed the utilization of VIs, portable tools, multisensor-based hyperspectral imaging, and their utilization in plant HTP to counter the mentioned challenges. UAVs have a great potential to determine plants’ phenotypic trait differences among crops and, at the same time, they allow the collection of different information about vegetation across wide regions rapidly and cost-effectively, without destroying plants. Vis integrated with UAVs are attracting increasing attention, and researchers are more interested in implementing these technologies for the HTP of diverse plant species. Similarly, several publicly available open-source online databases and repositories (AraPheno, Cleared Leaves DB, PHENOPSIS DB, PhenoFront, and plant genomics and phenomics research data repository) were developed to store, manage, and access HTP big data [133]. Besides, novel imagery-based two-dimensional (2D) and three-dimensional (3D) tools (RootSlice and RootScan) along with X-ray, microcomputed tomography and computational software (RootReader2D, RhizoTubes, WinRHIZO, SmartRoot, and RootAnalyzer) have been introduced to analyze plant root traits [135,136,137,138]. Recent reviews [135,136,137,139] have described the different HTP platforms and the need for a multidomain approach to address the challenges faced by HTP. With the advancement of HTP, there seems to be no doubt that plant phenomics has many challenges, and these need to be addressed in the coming future.

6. Future Prospects

In this study, we reviewed various VIs and UAV platforms used for HTP in agricultural research. Plant phenotyping with the advancement of high throughput is on the verge of entering the big-data era. Individual phenotypic information is not sufficient for association analyses. Therefore, complete phenomics information, multiscale (physiology, structure, omics, genomics, and environmental) interaction will be the foundation of research in the future. In light of the emerging environmental challenges, future research in plant phenotyping needs to develop new cost-effective technologies based on artificial intelligence/remote sensing for the advancement of image-based phenotyping. Deep or machine learning modeling, simulation tools are also needed to consider the development of new applications of HTP. In addition, it is also important to identify a proper automated phenotyping system to accurately work across plant species and draw reliable conclusions based on large sample statistics and the analysis of relationships with agronomic traits. In this context, the importance of collaboration and data sharing at the national and international levels cannot be overstated. Thus, research on the rapid advancement of genetic improvement in breeding programs, and on reliable, automatic, comprehensible, and multifunctional HTP techniques is required. We are already aiming to obtain some of these important techniques in our laboratory, to assist conventional breeding in the restructuring of existing phenotypic breeding processes and allow for a gradual shift to a more high-throughput, data-driven approach.

Author Contributions

R.T., J.Y., and L.L.: wrote the manuscript and prepared illustrations, figures, tables, and references; Y.K., A.L.K., and Y.Y.: conceptualized the overall structure; Y.K. and R.T.: edited the manuscript; Y.K.: contributed critical comments to the draft and approved the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was carried out with the support of “Cooperative Research Program for Agriculture Science and Technology Development (Project No. PJ01567802)” Rural Development Administration, Republic of Korea.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hunter, M.C.; Smith, R.G.; Schipanski, M.E.; Atwood, L.W.; Mortensen, D.A. Agriculture in 2050: Recalibrating targets for sustainable intensification. Bioscience 2017, 67, 386–391. [Google Scholar] [CrossRef] [Green Version]
  2. Ray, D.K.; Ramankutty, N.; Mueller, N.D.; West, P.C.; Foley, J.A. Recent patterns of crop yield growth and stagnation. Nat. Commun. 2012, 3, 1293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Chawade, A.; Armoniené, R.; Berg, G.; Brazauskas, G.; Frostgård, G.; Geleta, M.; Gorash, A.; Henriksson, T.; Himanen, K.; Ingver, A. A transnational and holistic breeding approach is needed for sustainable wheat production in the Baltic Sea region. Physiol. Plant. 2018, 164, 442–451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Chawade, A.; van Ham, J.; Blomquist, H.; Bagge, O.; Alexandersson, E.; Ortiz, R. High-throughput field-phenotyping tools for plant breeding and precision agriculture. Agronomy 2019, 9, 258. [Google Scholar] [CrossRef] [Green Version]
  5. Asseng, S.; Ewert, F.; Martre, P.; Rötter, R.P.; Lobell, D.B.; Cammarano, D.; Kimball, B.A.; Ottman, M.J.; Wall, G.; White, J.W. Rising temperatures reduce global wheat production. Nat. Clim. Change 2015, 5, 143–147. [Google Scholar] [CrossRef]
  6. Altieri, M.A.; Nicholls, C.I. The adaptation and mitigation potential of traditional agriculture in a changing climate. Clim. Change 2017, 140, 33–45. [Google Scholar] [CrossRef]
  7. Lin, Y. LiDAR: An important tool for next-generation phenotyping technology of high potential for plant phenomics? Comput. Electron. Agric. 2015, 119, 61–73. [Google Scholar] [CrossRef]
  8. Cabrera-Bosquet, L.; Crossa, J.; von Zitzewitz, J.; Serret, M.D.; Luis Araus, J. High-throughput phenotyping and genomic selection: The frontiers of crop breeding converge F. J. Integr. Plant Biol. 2012, 54, 312–320. [Google Scholar] [CrossRef] [Green Version]
  9. Araus, J.L.; Kefauver, S.C.; Zaman-Allah, M.; Olsen, M.S.; Cairns, J.E. Translating high-throughput phenotyping into genetic gain. Trends Plant Sci. 2018, 23, 451–466. [Google Scholar] [CrossRef] [Green Version]
  10. Kim, J.; Kim, K.-S.; Kim, Y.; Chung, Y.S. A short review: Comparisons of high-throughput phenotyping methods for detecting drought tolerance. Sci. Agric. 2021, 78. [Google Scholar] [CrossRef]
  11. Kokhan, S. Application of vegetative indexxes derived from satellite images IRS-1D LISS-III for determination of crop status. Space Sci. Technol. 2011, 17, 58–63. [Google Scholar]
  12. He, Y.; Guo, X.; Wilmshurst, J. Studying mixed grassland ecosystems I: Suitable hyperspectral vegetation indices. Can. J. Remote Sens. 2006, 32, 98–107. [Google Scholar] [CrossRef]
  13. Rahman, A.F.; Gamon, J.A.; Sims, D.A.; Schmidts, M. Optimum pixel size for hyperspectral studies of ecosystem function in southern California chaparral and grassland. Remote Sens. Environ. 2003, 84, 192–207. [Google Scholar] [CrossRef]
  14. Huete, A.; Didan, K.; Miura, T.; Rodriguez, E.P.; Gao, X.; Ferreira, L.G. Overview of the radiometric and biophysical performance of the MODIS vegetation indices. Remote Sens. Environ. 2002, 83, 195–213. [Google Scholar] [CrossRef]
  15. Roman, A.; Ursu, T.-M. Multispectral Satellite Imagery and Airborne Laser Scanning Techniques for the Detection of Archaeological Vegetation Marks. In Landscape Archaeology on the Northern Frontier of the Roman Empire at Porolissum; Mega Publishing House: Cluj-Napoca, Romania, 2016; pp. 141–152. [Google Scholar]
  16. Jang, G.; Kim, J.; Yu, J.-K.; Kim, H.-J.; Kim, Y.; Kim, D.-W.; Kim, K.-H.; Lee, C.W.; Chung, Y.S. Review: Cost-Effective Unmanned Aerial Vehicle (UAV) Platform for Field Plant Breeding Application. Remote Sens. 2020, 12, 998. [Google Scholar] [CrossRef] [Green Version]
  17. Cammarano, D.; Fitzgerald, G.J.; Casa, R.; Basso, B. Assessing the Robustness of Vegetation Indices to Estimate Wheat N in Mediterranean Environments. Remote Sens. 2014, 6, 2827–2844. [Google Scholar] [CrossRef] [Green Version]
  18. Cammarano, D.; Fitzgerald, G.; Basso, B.; O’Leary, G.; Chen, D.; Grace, P.; Fiorentino, C. Use of the Canopy Chlorophyl Content Index (CCCI) for Remote Estimation of Wheat Nitrogen Content in Rainfed Environments. Agron. J. 2011, 103, 1597–1603. [Google Scholar] [CrossRef]
  19. Tedesco-Oliveira, D.; Pereira da Silva, R.; Maldonado, W.; Zerbato, C. Convolutional neural networks in predicting cotton yield from images of commercial fields. Comput. Electron. Agric. 2020, 171, 105307. [Google Scholar] [CrossRef]
  20. Yue, J.; Yang, G.; Tian, Q.; Feng, H.; Xu, K.; Zhou, C. Estimate of winter-wheat above-ground biomass based on UAV ultrahigh-ground-resolution image textures and vegetation indices. ISPRS J. Photogramm. Remote Sens. 2019, 150, 226–244. [Google Scholar] [CrossRef]
  21. Gitelson, A.A.; Merzlyak, M.N.; Chivkunova, O.B. Optical properties and nondestructive estimation of anthocyanin content in plant leaves. Photochem. Photobiol. 2001, 74, 38–45. [Google Scholar] [CrossRef]
  22. Kaufman, Y.J.; Tanré, D. Atmospherically Resistant Vegetation Index (ARVI) for EOS-MODIS. IEEE Trans. Geosci. Remote Sens. 1992, 30, 261–270. [Google Scholar] [CrossRef]
  23. Zhou, X.; Huang, W.; Kong, W.; Ye, H.; Dong, Y.; Casa, R. Assessment of leaf carotenoids content with a new carotenoid index: Development and validation on experimental and model data. Int. J. Appl. Earth Obs. Geoinf. 2017, 57, 24–35. [Google Scholar] [CrossRef]
  24. Gitelson, A.A.; Zur, Y.; Chivkunova, O.B.; Merzlyak, M.N. Assessing Carotenoid Content in Plant Leaves with Reflectance Spectroscopy. Photochem. Photobiol. 2002, 75, 272–281. [Google Scholar] [CrossRef]
  25. Gitelson, A.A.; Peng, Y.; Viña, A.; Arkebauer, T.; Schepers, J.S. Efficiency of chlorophyll in gross primary productivity: A proof of concept and application in crops. J. Plant Physiol. 2016, 201, 101–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Ciganda, V.; Gitelson, A.; Schepers, J. Non-destructive determination of maize leaf and canopy chlorophyll content. J. Plant Physiol. 2009, 166, 157–167. [Google Scholar] [CrossRef] [Green Version]
  27. He, R.; Li, H.; Qiao, X.; Jiang, J. Using wavelet analysis of hyperspectral remote-sensing data to estimate canopy chlorophyll content of winter wheat under stripe rust stress. Int. J. Remote Sens. 2018, 39, 4059–4076. [Google Scholar] [CrossRef]
  28. Lilienthal, H.; Panten, K.; Schick, J.; Schroetter, S.; Schnug, E. Potential and Limitations of Hyperspectral Measurements to Determine the Nutritional Status of Maize. In Proceedings of the 16th World Fertilizer Congress of CIEC, Rio de Janeiro, Brazil, 20–24 October 2014; pp. 126–128. [Google Scholar]
  29. Daughtry, C.S.T. Agroclimatology: Discriminating crop residues from soil by shortwave infrared reflectance. Agron. J. 2001, 93, 125–131. [Google Scholar] [CrossRef]
  30. Daughtry, C.S.T.; Hunt, E.R.; McMurtrey, J.E. Assessing crop residue cover using shortwave infrared reflectance. Remote Sens. Environ. 2004, 90, 126–134. [Google Scholar] [CrossRef]
  31. Ghasemian Sorboni, N.; Pahlavani, P.; Bigdeli, B. Vegetation mapping of sentinel-1 and 2 satellite images using convolutional neural network and random forest with the aid of dual-polarized and optical vegetation indexes. Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci.-ISPRS Arch. 2019, 42, 435–440. [Google Scholar] [CrossRef] [Green Version]
  32. Periasamy, S. Significance of dual polarimetric synthetic aperture radar in biomass retrieval: An attempt on Sentinel-1. Remote Sens. Environ. 2018, 217, 537–549. [Google Scholar] [CrossRef]
  33. Anderegg, J.; Aasen, H.; Perich, G.; Roth, L.; Walter, A.; Hund, A. Temporal trends in canopy temperature and greenness are potential indicators of late-season drought avoidance and functional stay-green in wheat. Field Crops Res. 2021, 274, 108311. [Google Scholar] [CrossRef]
  34. Routh, P.K.; Sarkar, N.C.; Das, P.K.; Debnath, D.; Bandyopadhyay, S.; Raj, U. Retrival of Bio-physical paramaters in sunflower crop (Helianthus annus) using field based hyperspectral remote sensing. Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci. 2019, XLII-3/W6, 623–629. [Google Scholar] [CrossRef] [Green Version]
  35. Rustamova, S.; Shrestha, A.; Naz, A.A.; Huseynova, I. Expression profiling of DREB1 and evaluation of vegetation indices in contrasting wheat genotypes exposed to drought stress. Plant Gene 2021, 25, 100266. [Google Scholar] [CrossRef]
  36. Pinty, B.; Verstraete, M.M. GEMI: A non-linear index to monitor global vegetation from satellites. Vegetatio 1992, 101, 15–20. [Google Scholar] [CrossRef]
  37. Gitelson, A.A.; Kaufman, Y.J.; Merzlyak, M.N. Use of a green channel in remote sensing of global vegetation from EOS-MODIS. Remote Sens. Environ. 1996, 58, 289–298. [Google Scholar] [CrossRef]
  38. Gitelson, A.A.; Gritz, Y.; Merzlyak, M.N. Relationships between leaf chlorophyll content and spectral reflectance and algorithms for non-destructive chlorophyll assessment in higher plant leaves. J. Plant Physiol. 2003, 160, 271–282. [Google Scholar] [CrossRef]
  39. Sripada, R.P.; Heiniger, R.W.; White, J.G.; Meijer, A.D. Aerial Color Infrared Photography for Determining Early In-Season Nitrogen Requirements in Corn. Agron. J. 2006, 98, 968–977. [Google Scholar] [CrossRef]
  40. Louhaichi, M.; Borman, M.M.; Johnson, D.E. Spatially Located Platform and Aerial Photography for Documentation of Grazing Impacts on Wheat. Geocarto Int. 2001, 16, 65–70. [Google Scholar] [CrossRef]
  41. Kauth, R.J. Tasselled Cap—A Graphic Description of the Spectral-Temporal Development of Agricultural Crops as Seen by Landsat. LARS Symp. 1976, 41–51. Available online: http://docs.lib.purdue.edu/lars_symp/159 (accessed on 10 July 2021).
  42. Pagani, A.; Echeverría, H.E.; Andrade, F.H.; Sainz Rozas, H.R. Characterization of corn nitrogen status with a greenness index under different availability of sulfur. Agron. J. 2009, 101, 315–322. [Google Scholar] [CrossRef]
  43. Marshall, M.; Thenkabail, P.; Biggs, T.; Post, K. Hyperspectral narrowband and multispectral broadband indices for remote sensing of crop evapotranspiration and its components (transpiration and soil evaporation). Agric. For. Meteorol. 2016, 218–219, 122–134. [Google Scholar] [CrossRef] [Green Version]
  44. Thenkabail, P.S.; Mariotto, I.; Gumma, M.K.; Middleton, E.M.; Landis, D.R.; Huemmrich, K.F. Selection of hyperspectral narrowbands (hnbs) and composition of hyperspectral twoband vegetation indices (HVIS) for biophysical characterization and discrimination of crop types using field reflectance and hyperion/EO-1 data. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2013, 6, 427–439. [Google Scholar] [CrossRef] [Green Version]
  45. Crippen, R.E. Calculating the vegetation index faster. Remote Sens. Environ. 1990, 34, 71–73. [Google Scholar] [CrossRef]
  46. Mistele, B.; Schmidhalter, U. Tractor-based quadrilateral spectral reflectance measurements to detect biomass and total aerial nitrogen in winter wheat. Agron. J. 2010, 102, 499–506. [Google Scholar] [CrossRef]
  47. Boegh, E.; Soegaard, H.; Broge, N.; Hasager, C.B.; Jensen, N.O.; Schelde, K.; Thomsen, A. Airborne multispectral data for quantifying leaf area index, nitrogen concentration, and photosynthetic efficiency in agriculture. Remote Sens. Environ. 2002, 81, 179–193. [Google Scholar] [CrossRef]
  48. Srivastava, A.; Roy, S.; Kimothi, M.; Kumar, P.; Sehgal, S.; Mamatha, S.; Ray, S. Detetction of bacterial wilt disease (Pseudomonas solancearum) in brinjal using hyperspectral remote sensing. Int. Arch. Photogramm. Remote Sens. Spat. Inf. Sci. 2019, XLII-3/W6, 515–520. [Google Scholar] [CrossRef] [Green Version]
  49. He, R.; Qiao, X.; Jiang, J.; Guo, H. Retrieving canopy leaf total nitrogen content of winter wheat by continuous wavelet transform. Trans. Chin. Soc. Agric. Eng. 2015, 31, 141–146. [Google Scholar]
  50. Daughtry, C.S.T.; Hunt, E.R.; Doraiswamy, P.C.; McMurtrey, J.E. Remote sensing the spatial distribution of crop residues. Agron. J. 2005, 97, 864–871. [Google Scholar] [CrossRef]
  51. Zarco-Tejada, P.J.; Ustin, S.; Whiting, M. Temporal and spatial relationships between within-field yield variability in cotton and high-spatial hyperspectral remote sensing imagery. Agron. J. 2005, 97, 641–653. [Google Scholar] [CrossRef] [Green Version]
  52. Abrahão, S.A.; Pinto, F.d.A.d.C.; Queiroz, D.M.d.; Santos, N.T.; Carneiro, J.E.d.S. Determination of nitrogen and chlorophyll levels in bean-plant leaves by using spectral vegetation bands and indices. Rev. Ciência Agronômica 2013, 44, 464–473. [Google Scholar] [CrossRef]
  53. Haboudane, D.; Miller, J.R.; Pattey, E.; Zarco-Tejada, P.J.; Strachan, I.B. Hyperspectral vegetation indices and novel algorithms for predicting green LAI of crop canopies: Modeling and validation in the context of precision agriculture. Remote Sens. Environ. 2004, 90, 337–352. [Google Scholar] [CrossRef]
  54. Yang, Z.; Willis, P.; Mueller, R. Impact of Band-Ratio Enhanced Awifs Image to Crop Classification Accuracy. 2008. Available online: https://www.asprs.org/a/publications/proceedings/pecora17/0041.pdf (accessed on 10 July 2021).
  55. Datt, B. A New Reflectance Index for Remote Sensing of Chlorophyll Content in Higher Plants: Tests using Eucalyptus Leaves. J. Plant Physiol. 1999, 154, 30–36. [Google Scholar] [CrossRef]
  56. Sims, D.A.; Gamon, J.A. Relationships between leaf pigment content and spectral reflectance across a wide range of species, leaf structures and developmental stages. Remote Sens. Environ. 2002, 81, 337–354. [Google Scholar] [CrossRef]
  57. Chen, J.M. Evaluation of Vegetation Indices and a Modified Simple Ratio for Boreal Applications. Can. J. Remote Sens. 1996, 22, 229–242. [Google Scholar] [CrossRef]
  58. Qi, J.; Chehbouni, A.; Huete, A.R.; Kerr, Y.H.; Sorooshian, S. A modified soil adjusted vegetation index. Remote Sens. Environ. 1994, 48, 119–126. [Google Scholar] [CrossRef]
  59. Eitel, J.U.H.; Long, D.S.; Gessler, P.E.; Smith, A.M.S. Using in-situ measurements to evaluate the new RapidEye™ satellite series for prediction of wheat nitrogen status. Int. J. Remote Sens. 2007, 28, 4183–4190. [Google Scholar] [CrossRef]
  60. Hunt, E.R.; Rock, B.N. Detection of changes in leaf water content using Near- and Middle-Infrared reflectances. Remote Sens. Environ. 1989, 30, 43–54. [Google Scholar]
  61. Ceccato, P.; Flasse, S.; Tarantola, S.; Jacquemoud, S.; Grégoire, J.M. Detecting vegetation leaf water content using reflectance in the optical domain. Remote Sens. Environ. 2001, 77, 22–33. [Google Scholar] [CrossRef]
  62. Goel, N.S.; Qin, W. Influences of canopy architecture on relationships between various vegetation indices and LAI and FPAR: A computer simulation. Remote Sens. Rev. 1994, 10, 309–347. [Google Scholar] [CrossRef]
  63. Hardisky, M.A.; Klemas, V.; Smart, R.M. The influence of soil salinity, growth form, and leaf moisture on the spectral radiance of Spartina alterniflora canopies. Photogramm. Eng. Remote Sens. 1983, 49, 77–83. [Google Scholar]
  64. Serrano, L.; Peñuelas, J.; Ustin, S.L. Remote sensing of nitrogen and lignin in Mediterranean vegetation from AVIRIS data. Remote Sens. Environ. 2002, 81, 355–364. [Google Scholar] [CrossRef]
  65. Fourty, T.; Baret, F.; Jacquemoud, S.; Schmuck, G.; Verdebout, J. Leaf optical properties with explicit description of its biochemical composition: Direct and inverse problems. Remote Sens. Environ. 1996, 56, 104–117. [Google Scholar] [CrossRef]
  66. Melillo, J.M.; Aber, J.D.; Muratore, J.F. Nitrogen and Lignin Control of Hardwood Leaf Litter Decomposition Dynamics. Ecology 1982, 63, 621–626. [Google Scholar] [CrossRef]
  67. Liu, M.; Liu, X.; Li, M.; Fang, M.; Chi, W. Neural-network model for estimating leaf chlorophyll concentration in rice under stress from heavy metals using four spectral indices. Biosyst. Eng. 2010, 106, 223–233. [Google Scholar] [CrossRef]
  68. Gao, B.C. A normalized difference water index for remote sensing of vegetation liquid water from space Bo-Cai Gao Joint Center for Earth System Sciences Code 913, NASA Goddard Space Flight Center, Greenbelt, MD 20771. Remote Sens. Environ. 1996, 58, 257–266. [Google Scholar] [CrossRef]
  69. Jackson, T.J.; Chen, D.; Cosh, M.; Li, F.; Anderson, M.; Walthall, C.; Doriaswamy, P.; Hunt, E.R. Vegetation water content mapping using Landsat data derived normalized difference water index for corn and soybeans. Remote Sens. Environ. 2004, 92, 475–482. [Google Scholar] [CrossRef]
  70. Wang, L.; Qu, J.J.; Hao, X. Forest fire detection using the normalized multi-band drought index (NMDI) with satellite measurements. Agric. For. Meteorol. 2008, 148, 1767–1776. [Google Scholar] [CrossRef]
  71. Wang, L.; Qu, J.J. NMDI: A normalized multi-band drought index for monitoring soil and vegetation moisture with satellite remote sensing. Geophys. Res. Lett. 2007, 34, L20405. [Google Scholar] [CrossRef]
  72. Thenot, F.; Méthy, M.; Winkel, T. The Photochemical Reflectance Index (PRI) as a water-stress index. Int. J. Remote Sens. 2002, 23, 5135–5139. [Google Scholar] [CrossRef]
  73. Riedell, W.E.; Blackmer, T.M. Leaf reflectance spectra of cereal aphid-damaged wheat. Crop Sci. 1999, 39, 1835–1840. [Google Scholar] [CrossRef]
  74. Rama Rao, N.; Garg, P.K.; Ghosh, S.K.; Dadhwal, V.K. Estimation of leaf total chlorophyll and nitrogen concentrations using hyperspectral satellite imagery. J. Agric. Sci. 2008, 146, 65–75. [Google Scholar] [CrossRef]
  75. Merzlyak, M.N.; Gitelson, A.A.; Chivkunova, O.B.; Rakitin, V.Y. Non-destructive optical detection of pigment changes during leaf senescence and fruit ripening. Physiol. Plant. 1999, 106, 135–141. [Google Scholar] [CrossRef] [Green Version]
  76. Gitelson, A.; Merzlyak, M.N. Spectral Reflectance Changes Associated with Autumn Senescence of Aesculus hippocastanum L. and Acer platanoides L. Leaves. Spectral Features and Relation to Chlorophyll Estimation. J. Plant Physiol. 1994, 143, 286–292. [Google Scholar] [CrossRef]
  77. Curran, P.J.; Windham, W.R.; Gholz, H.L. Exploring the relationship between reflectance red edge and chlorophyll concentration in slash pine leaves. Tree Physiol. 1995, 15, 203–206. [Google Scholar] [CrossRef]
  78. Gamon, J.A.; Surfus, J.S. Assessing leaf pigment content and activity with a reflectometer. New Phytol. 1999, 143, 105–117. [Google Scholar] [CrossRef]
  79. Roujean, J.-L.; Breon, F.-M. Estimating PAR absorbed by vegetation from bidirectional reflectance measurements. Remote Sens. Environ. 1995, 51, 375–384. [Google Scholar] [CrossRef]
  80. Serrano, L.; Filella, I.; Penuelas, J. Remote sensing of biomass and yield of winter wheat under different nitrogen supplies. Crop Sci. 2000, 40, 723–731. [Google Scholar] [CrossRef] [Green Version]
  81. Ranjan, R.; Chopra, U.K.; Sahoo, R.N.; Singh, A.K.; Pradhan, S. Assessment of plant nitrogen stress in wheat (Triticum aestivum L.) through hyperspectral indices. Int. J. Remote Sens. 2012, 33, 6342–6360. [Google Scholar] [CrossRef]
  82. Huete, A.R. A soil-adjusted vegetation index (SAVI). Remote Sens. Environ. 1988, 25, 295–309. [Google Scholar] [CrossRef]
  83. Penuelas, J.; Baret, F.; Filella, I. Semi-empirical indices to assess carotenoids/chlorophyll a ratio from leaf spectral reflectance. Photosynthetica 1995, 31, 221–230. [Google Scholar]
  84. Genc, H.; Genc, L.; Turhan, H.; Smith, S.; Nation, J. Vegetation indices as indicators of damage by the sunn pest (Hemiptera: Scutelleridae) to field grown wheat. Afr. J. Biotechnol. 2008, 7, 173–180. Available online: http://www.academicjournals.org/AJB (accessed on 10 July 2021).
  85. Lobell, D.B.; Asner, G.P. Hyperion Studies Of Crop Stress In Mexico. In Geography; 2004. Available online: https://ntrs.nasa.gov/api/citations/20050192449/downloads/20050192449.pdf (accessed on 10 July 2021).
  86. Delalieux, S.; Somers, B.; Hereijgers, S.; Verstraeten, W.W.; Keulemans, W.; Coppin, P. A near-infrared narrow-waveband ratio to determine Leaf Area Index in orchards. Remote Sens. Environ. 2008, 112, 3762–3772. [Google Scholar] [CrossRef]
  87. Stuckens, J.; Swennen, R.L.; Coppin, P.; Dzikiti, S.; Verreynne, S.; Verstraeten, W.W. Extracting physiological info from a hyperspectral time series of a citrus orchard. In XXVIII International Horticultural Congress on Science and Horticulture for People (IHC2010): International Symposium on Engineering the Modelling, Monitoring, Mechanization and Automation Tools for Precision Horticulture; International Society for Horticultural Science: Leuven, Belgium, 2011; pp. 11–18. [Google Scholar]
  88. Broge, N.H.; Leblanc, E. Comparing prediction power and stability of broadband and hyperspectral vegetation indices for estimation of green leaf area index and canopy chlorophyll density. Remote Sens. Environ. 2001, 76, 156–172. [Google Scholar] [CrossRef]
  89. Bannari, A.; Asalhi, H.; Teillet, P.M. Transformed Difference Vegetation Index (TDVI) for Vegetation Cover Mapping. In Proceedings of the IEEE International Geoscience and Remote Sensing Symposium, Toronto, ON, Canada, 24–28 June 2002; Volume 5, pp. 3053–3055. [Google Scholar]
  90. Raymond Hunt, E.; Daughtry, C.S.T.; Eitel, J.U.H.; Long, D.S. Remote sensing leaf chlorophyll content using a visible band index. Agron. J. 2011, 103, 1090–1099. [Google Scholar] [CrossRef] [Green Version]
  91. Vogelmann, J.E.; Rock, B.N.; Moss, D.M. Red edge spectral measurements from sugar maple leaves. Int. J. Remote Sens. 1993, 14, 1563–1575. [Google Scholar] [CrossRef]
  92. Penuelas, J.; Filella, I.; Biel, C.; Serrano, L.; Save, R. The reflectance at the 950-970 nm region as an indicator of plant water status. Int. J. Remote Sens. 1993, 14, 1887–1905. [Google Scholar] [CrossRef]
  93. Champagne, C.; Pattey, E.; Bannari, A.; Stratchan, I.B. Mapping Crop Water Status: Issues of Scale in the Detection of Crop Water Stress Using Hyperspectral Indices. In Proceedings of the 8th International Symposium on Physical Measurements and Signatures in Remote Sensing, Aussois, France, 8–12 January 2001; pp. 79–84. [Google Scholar]
  94. Ates, A.M.; Bukowski, M. 2021/22 Global Soybean Stocks Fall in Tandem with South American Production Economic Research Service|Situation and Outlook Report Domestic Outlook International Outlook. 2022. Available online: https://www.ers.usda.gov/webdocs/outlooks/103026/ocs-22a.pdf?v=6071 (accessed on 29 May 2022).
  95. Antonio, F.W. Using WorldView-2 Vis-NIR multispectral imagery to support land mapping and feature extraction using normalized difference index ratios. In Proceedings Volume 8390, Algorithms and Technologies for Multispectral, Hyperspectral, and Ultraspectral Imagery XVIII; SPIE: Bellingham, WA, USA, 2012. [Google Scholar] [CrossRef]
  96. Jain, N.; Ray, S.S.; Singh, J.; Panigrahy, S. Use of hyperspectral data to assess the effects of different nitrogen applications on a potato crop. Precis. Agric. 2007, 8, 225–239. [Google Scholar] [CrossRef]
  97. Rouse, J.W.; Haas, R.H.; Schell, J.A.; Deering, D.W. Monitoring vegetation systems in the Great Plains with ERTS. NASA Spec. Publ. 1974, 351, 309. [Google Scholar]
  98. Daughtry, C.S.T.; Walthall, C.L.; Kim, M.S.; De Colstoun, E.B.; McMurtrey, J.E. Estimating Corn Leaf Chlorophyll Concentration from Leaf and Canopy Reflectance. Remote Sens. Environ. 2000, 74, 229–239. [Google Scholar] [CrossRef]
  99. Peñuelas, J.; Gamon, J.A.; Griffin, K.L.; Field, C.B. Assessing community type, plant biomass, pigment composition, and photosynthetic efficiency of aquatic vegetation from spectral reflectance. Remote Sens. Environ. 1993, 46, 110–118. [Google Scholar] [CrossRef]
  100. Gamon, J.A.; Penuelas, J.; Field, C.B. A narrow-waveband spectral index that tracks diurnal changes in photosynthetic efficiency. Remote Sens. Environ. 1992, 41, 35–44. [Google Scholar] [CrossRef]
  101. Gamon, J.A.; Field, C.B.; Bilger, W.; Björkman, O.; Fredeen, A.L.; Peñuelas, J. Remote sensing of the xanthophyll cycle and chlorophyll fluorescence in sunflower leaves and canopies. Oecologia 1990, 85, 1–7. [Google Scholar] [CrossRef] [PubMed]
  102. Arellano, P.; Stratoulias, D. 20—Hyperspectral vegetation indices to detect hydrocarbon pollution. In Hyperspectral Remote Sensing; Pandey, P.C., Srivastava, P.K., Balzter, H., Bhattacharya, B., Petropoulos, G.P., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 401–425. [Google Scholar] [CrossRef]
  103. Matłok, N.; Basara, O.; Zardzewiały, M.; Gorzelany, J.; Balawejder, M. Effectiveness of a complex fertilisation technology applied to Zea mays, assessed based on normalised difference vegetation index (NDVI) from terra moderate resolution imaging spectroradiometer (MODIS). Agriculture 2021, 11, 754. [Google Scholar] [CrossRef]
  104. Baret, F.; Guyot, G. Potentials and limits of vegetation indices for LAI and APAR assessment. Remote Sens. Environ. 1991, 35, 161–173. [Google Scholar] [CrossRef]
  105. Phillips, R.L. Mobilizing science to break yield barriers. Crop Sci. 2010, 50, S-99–S-108. [Google Scholar] [CrossRef] [Green Version]
  106. McMullen, M.D.; Kresovich, S.; Villeda, H.S.; Bradbury, P.; Li, H.; Sun, Q.; Flint-Garcia, S.; Thornsberry, J.; Acharya, C.; Bottoms, C. Genetic properties of the maize nested association mapping population. Science 2009, 325, 737–740. [Google Scholar] [CrossRef] [Green Version]
  107. Jannink, J.-L.; Lorenz, A.J.; Iwata, H. Genomic selection in plant breeding: From theory to practice. Brief. Funct. Genom. 2010, 9, 166–177. [Google Scholar] [CrossRef] [Green Version]
  108. Gaudin, A.C.; Henry, A.; Sparks, A.H.; Slamet-Loedin, I.H. Taking transgenic rice drought screening to the field. J. Exp. Bot. 2013, 64, 109–117. [Google Scholar] [CrossRef] [Green Version]
  109. Saint Pierre, C.; Crossa, J.L.; Bonnett, D.; Yamaguchi-Shinozaki, K.; Reynolds, M.P. Phenotyping transgenic wheat for drought resistance. J. Exp. Bot. 2012, 63, 1799–1808. [Google Scholar] [CrossRef] [Green Version]
  110. Houle, D.; Govindaraju, D.R.; Omholt, S. Phenomics: The next challenge. Nat. Rev. Genet. 2010, 11, 855–866. [Google Scholar] [CrossRef] [PubMed]
  111. Araus, J.L.; Slafer, G.A.; Royo, C.; Serret, M.D. Breeding for yield potential and stress adaptation in cereals. Crit. Rev. Plant Sci. 2008, 27, 377–412. [Google Scholar] [CrossRef]
  112. Pretty, J.N. The sustainable intensification of agriculture. Nat. Resour. Forum 1997, 21, 247–256. [Google Scholar] [CrossRef]
  113. Mahlein, A.-K. Plant disease detection by imaging sensors–parallels and specific demands for precision agriculture and plant phenotyping. Plant Dis. 2016, 100, 241–251. [Google Scholar] [CrossRef] [Green Version]
  114. Dhondt, S.; Wuyts, N.; Inzé, D. Cell to whole-plant phenotyping: The best is yet to come. Trends Plant Sci. 2013, 18, 428–439. [Google Scholar] [CrossRef]
  115. Tuberosa, R. Phenotyping for drought tolerance of crops in the genomics era. Front. Physiol. 2012, 3, 347. [Google Scholar] [CrossRef] [Green Version]
  116. Fiorani, F.; Schurr, U. Future scenarios for plant phenotyping. Annu. Rev. Plant Biol. 2013, 64, 267–291. [Google Scholar] [CrossRef] [Green Version]
  117. Shafian, S.; Rajan, N.; Schnell, R.; Bagavathiannan, M.; Valasek, J.; Shi, Y.; Olsenholler, J. Unmanned aerial systems-based remote sensing for monitoring sorghum growth and development. PLoS ONE 2018, 13, e0196605. [Google Scholar] [CrossRef] [Green Version]
  118. Hoffmann, A.M.; Noga, G.; Hunsche, M. Fluorescence indices for monitoring the ripening of tomatoes in pre-and postharvest phases. Sci. Hortic. 2015, 191, 74–81. [Google Scholar] [CrossRef]
  119. Xue, L.; Yang, L. Deriving leaf chlorophyll content of green-leafy vegetables from hyperspectral reflectance. ISPRS J. Photogramm. Remote Sens. 2009, 64, 97–106. [Google Scholar] [CrossRef]
  120. Xu, H.; Ying, Y.; Fu, X.; Zhu, S. Near-infrared spectroscopy in detecting leaf miner damage on tomato leaf. Biosyst. Eng. 2007, 96, 447–454. [Google Scholar] [CrossRef]
  121. Tripodi, P.; Massa, D.; Venezia, A.; Cardi, T. Sensing technologies for precision phenotyping in vegetable crops: Current status and future challenges. Agronomy 2018, 8, 57. [Google Scholar] [CrossRef] [Green Version]
  122. Yousaf, M.; Li, J.; Lu, J.; Ren, T.; Cong, R.; Fahad, S.; Li, X. Effects of fertilization on crop production and nutrient-supplying capacity under rice-oilseed rape rotation system. Sci. Rep. 2017, 7, 1270. [Google Scholar] [CrossRef] [PubMed]
  123. Zaman-Allah, M.; Vergara, O.; Araus, J.; Tarekegne, A.; Magorokosho, C.; Zarco-Tejada, P.; Hornero, A.; Albà, A.H.; Das, B.; Craufurd, P. Unmanned aerial platform-based multi-spectral imaging for field phenotyping of maize. Plant Methods 2015, 11, 35. [Google Scholar] [CrossRef] [Green Version]
  124. Haghighattalab, A.; Pérez, L.G.; Mondal, S.; Singh, D.; Schinstock, D.; Rutkoski, J.; Ortiz-Monasterio, I.; Singh, R.P.; Goodin, D.; Poland, J. Application of unmanned aerial systems for high throughput phenotyping of large wheat breeding nurseries. Plant Methods 2016, 12, 35. [Google Scholar] [CrossRef] [Green Version]
  125. Arnold, T.; De Biasio, M.; Fritz, A.; Leitner, R. UAV-based measurement of vegetation indices for environmental monitoring. In 2013 Seventh International Conference on Sensing Technology (ICST); IEEE: Piscataway, NJ, USA, 2013; pp. 704–707. [Google Scholar]
  126. Khan, Z.; Rahimi-Eichi, V.; Haefele, S.; Garnett, T.; Miklavcic, S.J. Estimation of vegetation indices for high-throughput phenotyping of wheat using aerial imaging. Plant Methods 2018, 14, 20. [Google Scholar] [CrossRef]
  127. Hariharan, J.; Fuller, J.; Ampatzidis, Y.; Abdulridha, J.; Lerwill, A. Finite difference analysis and bivariate correlation of hyperspectral data for detecting laurel wilt disease and nutritional deficiency in avocado. Remote Sens. 2019, 11, 1748. [Google Scholar] [CrossRef] [Green Version]
  128. Jiménez-Brenes, F.M.; López-Granados, F.; De Castro, A.; Torres-Sánchez, J.; Serrano, N.; Peña, J. Quantifying pruning impacts on olive tree architecture and annual canopy growth by using UAV-based 3D modelling. Plant Methods 2017, 13, 55. [Google Scholar] [CrossRef] [Green Version]
  129. Deng, L.; Mao, Z.; Li, X.; Hu, Z.; Duan, F.; Yan, Y. UAV-based multispectral remote sensing for precision agriculture: A comparison between different cameras. ISPRS J. Photogramm. Remote Sens. 2018, 146, 124–136. [Google Scholar] [CrossRef]
  130. ERDAS ERDAS IMAGINE: World Class Remote Sensing Software. Available online: https://www.hexagongeospatial.com/products/power-portfolio/erdas-imagine/erdas-imagine-remote-sensing-software-package (accessed on 9 July 2021).
  131. ENVI-L3HARRIS GEOSPATIAL. Available online: https://www.l3harrisgeospatial.com/Software-Technology/ENVI#arcgis (accessed on 10 July 2021).
  132. GrassGIS What Is GRASS GIS? Available online: https://grass.osgeo.org/learn/overview/ (accessed on 10 July 2021).
  133. Zhao, C.; Zhang, Y.; Du, J.; Guo, X.; Wen, W.; Gu, S.; Wang, J.; Fan, J. Crop Phenomics: Current Status and Perspectives. Front. Plant Sci. 2019, 10, 714. [Google Scholar] [CrossRef] [Green Version]
  134. Jo, H.; Lee, J.Y.; Cho, H.; Choi, H.J.; Son, C.K.; Bae, J.S.; Bilyeu, K.; Song, J.T.; Lee, J.-D. Genetic Diversity of Soybeans (Glycine max (L.) Merr.) with Black Seed Coats and Green Cotyledons in Korean Germplasm. Agronomy 2021, 11, 581. [Google Scholar] [CrossRef]
  135. Burton, A.L.; Williams, M.; Lynch, J.P.; Brown, K.M. RootScan: Software for high-throughput analysis of root anatomical traits. Plant and Soil 2012, 357, 189–203. [Google Scholar] [CrossRef]
  136. Chopin, J.; Laga, H.; Huang, C.Y.; Heuer, S.; Miklavcic, S.J. RootAnalyzer: A Cross-Section Image Analysis Tool for Automated Characterization of Root Cells and Tissues. PLoS ONE 2015, 10, e0137655. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Jeudy, C.; Adrian, M.; Baussard, C.; Bernard, C.; Bernaud, E.; Bourion, V.; Busset, H.; Cabrera-Bosquet, L.; Cointault, F.; Han, S.; et al. RhizoTubes as a new tool for high throughput imaging of plant root development and architecture: Test, comparison with pot grown plants and validation. Plant Methods 2016, 12, 31. [Google Scholar] [CrossRef] [PubMed]
  138. Clark, R.T.; Famoso, A.N.; Zhao, K.; Shaff, J.E.; Craft, E.J.; Bustamante, C.D.; McCouch, S.R.; Aneshansley, D.J.; Kochian, L.V. High-throughput two-dimensional root system phenotyping platform facilitates genetic analysis of root growth and development. Plant Cell Environ. 2013, 36, 454–466. [Google Scholar] [CrossRef]
  139. Song, P.; Wang, J.; Guo, X.; Yang, W.; Zhao, C. High-throughput phenotyping: Breaking through the bottleneck in future crop breeding. Crop J. 2021, 9, 633–645. [Google Scholar] [CrossRef]
Figure 1. Different spectral reflectance curves for vegetation, modified from [15]. The main absorption and reflectance characteristics are represented.
Figure 1. Different spectral reflectance curves for vegetation, modified from [15]. The main absorption and reflectance characteristics are represented.
Plants 11 01712 g001
Table 1. Vegetation index types and their equation and utilization in plant phenotyping.
Table 1. Vegetation index types and their equation and utilization in plant phenotyping.
Index/Abbreviations FormulaUtilizationCropsReference
Anthocyanin reflectance index 1 (ARI1) ARI 1 = 1 R 550 1 R 700 Estimates anthocyanin accumulation in leavesMaple, cotoneaster, dogwood, and pelargonium[21]
Anthocyanin reflectance index 2 (ARI2) ARI 2 = R 800 [ 1 R 550 1 R 700 ] Estimates anthocyanin accumulation in leavesMaple, cotoneaster, dogwood, and pelargonium[21]
Atmospherically resistant vegetation index (ARVI) ARVI = R 800   + [ R 680   γ ( R 450 R 680 ) ] R 800 + [ R 680 γ ( R 450 R 680 ) ] Utilized for remote sensing of vegetation and atmospheric effect[22]
Carotenoid index (CARI) CRI 2   =   R 720 R 510 1 Detects leaf carotenoid contentWinter wheat[23]
Carotenoid reflectance index 1 (CRI1) CRI 1   =   1 R 510 1 R 550 Detects leaf carotenoid contentNorway maple, chestnut, and beech[24]
Carotenoid reflectance index 2 (CRI2) CRI 2 = 1 R 510 1 R 700 Determines carotenoid content in leavesNorway maple, chestnut, and beech[24]
Red-edge chlorophyll index CIrededge = R N I R R r e d   e d g e 1 Determines chlorophyll content in both anthocyanin-containing and anthocyanin-free leavesSoybean and maize[25,26]
Carter stress index (CTR1) CTR 1 = R 695 R 420 Utilized to derive the chlorophyll content of winter wheat under stripe rust stressWheat[27]
Carter stress index (CTR2) CTR 2 = R 695 R 760 Detects the nutritional status of cropsMaize[28]
Cellulose absorption (CAI) CAI = 0.5   ( R 2000 + R 2200 ) R 2100 Estimates crop residue cover Corn, soybean, and wheat [29,30]
Dual-polarization SAR vegetation index (DPSVI) DPSVI = σ v h ( i ) [   ( σ v v ( m a x ) σ v h ( i ) σ v v ( i ) σ v h ( i ) + σ 2 v h ( i ) ) + ( σ v v ( m a x ) σ v v ( i ) σ 2 v v ( i ) + σ v h ( i ) + σ v v ( i ) )   ] 2 ×   σ v v ( i ) Utilized to estimate biomass and vegetationElephant foot yam, turmeric, onion, grass, cassava manioc, millet crops, and black gram[31,32]
Enhanced vegetation index (EVI) EVI = 2.5 ×   ( N I R R E D ) ( N I R + 6     ×   R E D 7.5     ×   B l u e + 1 ) Utilized to assess canopy structural variations, leaf area, canopy type, plant physiognomy, and canopy architectureGrass/shrub, savanna, and tropical forest biomes[14]
Growing degree days mid-stay-green index (GDDmidstg,i) GDD midstg,i   =   GDD Heading,i   +   OnSen GDDAS,i GDD Heading,i 2 Utilized for biomass determination, canopy temperature, and greenness indicatorWheat[33]
Gitelson and Merzlyak index (GM1) GM 1 = R 750 R 550 Measures biophysical parameters Sunflower[34]
Gitelson and Merzlyak index (GM2) GM 2 = R 750 R 700 Detects drought stressWheat[35]
Global environmental monitoring index (GEMI) GEMI = eta ( 1     0.25 × eta )     ( R E D 0.125 ) ( 1 R e d )
where   eta   =   2 ( N I R 2 R E D 2 ) + 1.5   ×   N I R + 0.5   ×   R E D N I R + R E D + 0.5
Useful to compare observations under varying atmospheric and illumination conditions; more sensitive to actual surface conditions than SR or NDVI over the bulk of the range of vegetation conditions[36]
Green atmospherically resistant index (GARI) GARI = N I R [ G r e e n γ ( B l u e R e d ) ] N I R + [ G r e e n γ ( B l u e R e d ) ] Measures the rate of photosynthesis and monitors plant stressMaple, chestnut [37]
Green chlorophyll index (GCI) GCI = ( R N I R R G r e e n ) × 1 Measures leaf chlorophyll and carotenoids contentMaple, chestnut, wild vine, and beech[38]
Green difference vegetation index (GDVI) GDVI = N I R G r e e n Utilized to improve in-season estimates of N requirements Corn[39]
Green leaf index (GLI) GLI = ( G r e e n R e d ) + ( G r e e n B l u e ) ( 2 ×   G r e e n ) + R e d + B l u e Utilized to map and document the extent and intensity of goose impacts on wheat fieldsWheat[40]
Green optimized soil-adjusted vegetation index (GOSAVI) GOSAVI = N I R G r e e n N I R + G r e e n + 0.16 Utilized to improve in-season estimates of N requirements Corn[39]
Green ratio vegetation index (GRVI) GRVI = N I R G r e e n Utilized to improve in-season estimates of N requirements Corn[39]
Green soil-adjusted vegetation index (GSAVI) GSAVI = 1.5   ×   ( N I R G r e e n ) ( N I R + G r e e n + 0.5 ) Utilized to improve and predict nitrogen requirementsCorn[39]
Green vegetation index (GVI) GVI = ( 0.2848   ×   T M 1 ) + ( 0.2435   ×   T M 2 ) + ( 0.5436   ×   T M 3 ) + ( 0.7243   ×   T T M 4 ) + ( 0.0840   ×   T M 5 ) + ( 0.1800   ×   T M 7 ) Minimizes the effects of background soil while emphasizing green vegetation to estimate and correct atmospheric haze and moisture effectsWheat[41]
Greenness index (G) G = R 554 R 667 Evaluates corn nitrogen status under different sulfur levelsCorn[42]
Hyperspectral narrow bands (HNB) HNB = R 2 R 1 R 1 + R 2 Utilized to determine biophysical (biomass, leaf area index) and biochemical quantities (leaf nitrogen and plant pigments)Wheat, maize, rice, barley, soybeans, pulses, cotton, and alfalfa[43,44]
Hyperspectral vegetation indices (HVIij) HVIij = R j R i R j + R i Utilized to determine biophysical (biomass, leaf area index) and biochemical quantities (leaf nitrogen and plant pigments)Wheat, maize, rice, barley, soybeans, pulses, cotton, and alfalfa[43,44]
Infrared percentage vegetation index (IPVI) IPVI = N I R N I R R e d Significant in the monitoring of global biomass[45]
Index R780/R700 and R780/R740 I = R 780   R 700     and   R 780   R 740   Determine the above nitrogen and biomassWheat[46]
Leaf area index (LAI) LAI = ( 3.618   × E V I 0.118 ) Utilized to estimate foliage cover and to forecast crop growth and yieldWinter barley and wheat, spring barley, peas, grass, maize, and beets[47]
Lichtenthaler index (LIC1) LIC 1 = R 790 R 680 R 790 + R 680 Detects bacterial wilt disease in BrinjalBrinjal[48]
Lichtenthaler index (LIC2) LIC 2 = R 440 R 690 Measures leaf nitrogen contentWheat[49]
Lignin cellulose absorption index (LCAI) LCAI = 100   ×   [ ( R 2185   t o   2225 R 2145   t o   2185 ) + ( R 2185   t o   2225 R 2295   t o   2365 ) ] Utilized to quantify crop residue cover and classify tillage intensity over diverse landsCorn, soybean, wheat, tall fescue, and alfalfa[50]
Modified chlorophyll absorption in reflectance index (MCARI) MCARI = [ ( R 700 R 670 ) 0.2   ×   ( R 700 R 550 ) ]   ×   ( R 700 R 670 ) Utilized to identify in-field heterogeneity for field segmentation as a feature of plant morphology and chlorophyll stress statusCotton [51]
Modified chlorophyll absorption in reflectance index (MCARI1) MCARI 1 = 1.2   ×     [ 2.5   ×   ( R 790 R 670 ) 1.3   ×     ( R 790 R 550 )]Determines nitrogen and chlorophyll levels in bean-plant leavesBean[52]
Modified chlorophyll absorption ratio index improved (MCARI2) MCARI 2 = 1.5   [ 2.5   ( R 800 R 670 ) 1.3   ( R 800 R 550 ) ] ( 2   × R 800 + 1 ) 2   ( 6   × R 800 R 670 ) 0.5 Predicts the LAI of crop canopies Corn, wheat, and soybean[53]
Modified nonlinear index (MNLI) MNLI = ( N I R 2 R e d ) × ( 1 + L ) N I R 2 + R e d + L Utilized to interpret soil background and improve crop discrimination, crop yield, crop stress, pest/disease surveillance, and disaster management[54]
Modified red-edge
normalized difference vegetation index (MRENDVI)
MRENDVI = R 750 R 705 ρ 750 + ρ 705   2 ×   ρ 445 Utilized to estimate chlorophyll content and determine the variation in reflectance caused by chlorophyll absorptionEucalyptus[55]
Modified red-edge simple ratio (MRESR) MRESR = R 750 R 445 R 705 R 445 Utilized to estimate pigment content and carotenoid/chlorophyll ratios in green leavesWide species [56]
Modified simple ratio (MSR)
MSR = ( N I R R e d ) 1 ( N I R R e d ) + 1 Advantageous in field data evaluation and less sensitivity to canopy optical and geometrical propertiesJackpine and black spruce[57]
Modified soil-adjusted vegetation index 2 (MSAVI2) MSAVI 2 = 2 × N I R + 1 ( 2   × N I R + 1 ) 2 8 ( N I R R e d ) 2 Reduces soil noise and increases the dynamic range of the vegetation signalCotton[58]
Modified triangular vegetation index improved (MTVI2) MTVI 2 = 1.5   [   1.2 ( R 800 R 550 ) 2.5   ( R 670 R 550 ) ] ( 2   ×   R 800   + 1 ) 2 ( 6   ×   R 800 5 ×   R 670   ) 0.5 Determines chlorophyll content variations and linearly related to green LAICorn, wheat, and soybean[53,59]
Modified triangular vegetation index (MTVI) MTVI = 1.2 [ 1.2 ( R 800 R 550 ) 2.5 ( R 670 R 550 ) Determines chlorophyll content variations and linearly related to green LAICorn, wheat, soybean [53]
Moisture stress index (MSI) MSI = R 1599 R 819 Determines leaf and canopy water contentCalifornia live oak, blue spruce, sweetgum, soybean, maple, apricot, mulberry, and cherry laurel[60,61]
Nonlinear index (NLI) NLI = N I R 2 R e d N I R 2 + R e d Linearizes relationships with surface parameters that tend to be nonlinearCorn and aspen [62]
Normalized difference infrared index (NDII) NDII = R 819 R 1649 R 819   + R 1649 Estimates leaf and canopy water content, and correlates spectral response to ambient soil salinitySmooth cordgrass[63]
Normalized difference lignin index (NDLI) NDLI = log ( 1 R 1754 ) log ( 1 R 1680 ) log ( 1 R 1754 ) + log ( 1 R 1680 ) Determines biochemical concentration (nitrogen and lignin) and canopy structural features Wide species [64,65,66]
Normalized difference nitrogen index (NDNI) NDNI = log ( 1 R 1510 ) log ( 1 R 1680 ) log ( 1 R 1510 ) + log ( 1 R 1680 ) Determines biochemical concentration (nitrogen and lignin) and canopy structural featuresWide species[64]
Normalized difference vegetation index
(NDVI)
NDVI = R N I R R R E D R N I R + R R E D Predicts chlorophyll content in rice plants under stress from heavy metal conditionRice[67]
Normalized difference water index (NDWI) NDWI = ( R 857 R 1241 ) ( R 857   +   R 1241 ) Detects vegetation liquid or water content Corn, soybean, and redwood[68,69]
Normalized multiband drought index (NMDI) NMDI = R 860 ( R 1640 R 2130 ) R 860 + ( R 1640   + R 2130 ) Monitors soil and vegetation moisture from space, detects fires[70,71]
Normalized phaeophytization index (NPQI) NPQI = R 415 R 435 R 415 + R 435 Detects mite effects on apple treesApple [72]
Normalized pigment chlorophyll index (NPCI) NPCI = R 680 R 430 R 680 + R 430 Evaluates chlorophyll loss and leaf senescence caused by aphid feedingWheat[73]
Plant biochemical index (PBI) PBI = R 810 R 560 Determine total chlorophyll and nitrogen concentrationsRice, sorghum, mung bean, and pigeon pea[74]
Photochemical reflectance index (PRI) PRI = R 531 R 570 R 531 + R 570 Rapidly evaluates leaf water status to estimate the water stress index of cropsQuinoa [72]
Plant senescence reflectance index (PSRI) PSRI = R 680 R 500 R 750 Estimates pigment content depending on the onset, stage, relative rates, and kinetics of leaves and fruits Maple, chestnut, potato, coleus, lemon, and apple [75]
Reflectance at 1200 nm (Ratio1200) Ratio 1200   =   2   ×   A v g   ( R 1180   t o   1220 ) A v g   ( R 1090   t o   1110 ) + A v g   ( R 1265   t o   1285 ) Determine shoot biomass, phenology, morphology, and canopy structural parametersWheat[33]
Red-edge normalized difference vegetation index (RENDVI) RENDVI = R 750 R 705 R 750 + R 705 Quantitative estimation of pigments Horse chestnut and Norway maple[76]
Red-edge position index (REPI)Estimates chlorophyll concentration in fieldsSlash pine [77]
Red-green ratio index (RGRI) RGRI = i = 600 699 R i j = 500 599 R j or RRED:RGREENDetermines chlorophyll, xanthophyll cycle, anthocyanin contents, and the change in photochemicalSunflower, Douglas, and coast live oak[78]
Renormalized difference vegetation index (RDVI) RDVI = ( N I R R e d ) ( N I R + R e d ) Utilized to investigate healthy vegetation, suitable for larger vegetation coverages and denser canopiesCorn, wheat, and soybean [53,79]
Simple ratio index (SR) SR = R N I R R R E D Utilized to estimate crop growth and to forecast grain yieldWheat[80]
Simple ratio pigment index (SRPI) SRPI = R 430 R 680 Utilized for large-area monitoring of plants’ N status in wheat, allows the precise application of fertilizersWheat[81]
Soil-adjusted vegetation index (SAVI) SAVI = 1.5   × ( N I R R e d ) ( N I R + R e d + 0.5 ) Best used in areas with relatively sparse vegetation where the soil is visible through the canopyCotton and range grass [82]
Structure-insensitive pigment index (SIPI) SIPI = R 800 R 445 R 800 R 680 Detects pest damage on wheat through the identification of chlorophyll loss and estimates carotenoids and chlorophyll content ratioMaize, wheat, tomato, soybean, sunflower, and sugar beet[83,84]
Sum green index (SGI) Detects changes in vegetation greenness and vegetation canopy openingWheat[85]
Standardized LAI determining index (sLAIDI) sLAIDI = ( R 1050 R 1250 R 1050 + R 1250 ) Determine LAI, more accurately extract biochemical parameters Apple, peach, citrus, and orchard [86,87]
Transformed chlorophyll absorption reflectance index (TCARI) TCARI = 3 [ ( R 700 R 670 ) 0.2 ( R 700 R 550 ) ( R 700 R 670 ) ] Predicts the LAI of crop canopies [53]
Triangular vegetation index (TVI) TVI   R 120 ( R 750 R 550 ) 200 ( R 670 R 550 ) 2 Estimates the green leaf area index and canopy chlorophyll density[88]
Transformed difference vegetation index (TDVI) TDVI = 1.5 [ ( N I R R e d ) N I R 2 + R e d + 0.5 ] Useful for monitoring vegetation cover in urban environmentsCotton[89]
Triangular greenness index (TGI) TGI = (   λ R e d λ B l u e   ) ( R R e d R G r e e n ) (   λ R e d λ G r e e n ) ( ρ R e d ρ B l u e ) 2
Detects leaf chlorophyll content and green vegetationCorn, soybean, sorghum, dandelion, sweetgum, tulip tree, and small-leaf linden[90]
Visible atmospherically resistant index (VARI) VARI = G r e e n R e d G r e e n + R e d B l u e Estimates the fraction of vegetation in environments with low sensitivity to atmospheric effectsWheat[24]
Vogelmann red-edge index 1 (VREI1) VREI 1 = R 740 R 720 Quantifies leaf-level chlorophyll contentSugar maple[91]
Vogelmann red-edge index 2 (VREI2) VREI 2 = R 734 R 747 R 715 + R 726 Quantifies leaf-level chlorophyll contentSugar maple[91]
Water band index (WBI) WBI = R 970 R 900 Estimates water status and relative water content Gerbera, pepper, and bean[92,93]
Wide dynamic range vegetation index (WDRVI) WDRVI = ( a   × N I R R e d ) ( a   ×   N I R + R e d ) Aids in the robust characterization of crops physiological and phenological characteristics, dry matter content, water content, leaf mesophyll structure index, and some other less influential factorsWheat and maize [24,94]
World view improved
vegetation index (WV-VI)
WV - VI = ( N I R 2 R e d ) ( N I R 2   + R e d ) Determines moisture content, vegetation health; distinguishes natural features from man-made objects and supports land mapping[95]
Zarco-Tejada and Miller (ZMI) ZMI = R 750 R 710 Estimates leaf nitrogenPotato[96]
Note: ‘—’ Information is not available, all the equation details can be found in the respective publication.
Table 2. List of photogrammetry and mapping software used for unmanned aerial vehicles (UAVs).
Table 2. List of photogrammetry and mapping software used for unmanned aerial vehicles (UAVs).
Photogrammetry and Mapping SoftwareFeaturesManufacturer
Agi Soft photoscan ProSimple interface, easy to learn for beginners; supports Python script; distributed, flexible, nonlinear processing; elaborate model editing for accurate results; includes inbuilt tools to measure distances, areas, and volumes; produces better quality point clouds, digital elevation models, orthoimage generation, export, makes accurate measurements; access to 3D modeling presents a wider range of panorama stitching; processes RGB, NIR, thermal, and multispectral imageryAgisoft
Precision MapperA better option for the agricultural sector; includes tools to analyze crop health, obtain volumetric measurements, generate orthoimages, point clouds, and 3D models; offers NDVI enhancements, a canopy cover calculator, and even finds standing water in a fieldPrecision Hawk
Maps Made EasyFree package, easy to operate; generates 3D models, accurate calculations, and NDVI maps; allows the user to collage, classify, and add a comment to NDVI maps easily DRONES MADE EASY
Drone Deploy Field scannerEasy to operate, offers cloud-based processing; supports DJI drones; allows DTM (digital terrain models), 3D models, and orthomosaic generation; collects NDVI data, and allows the quick view of NDVI maps even without internet connectionDrone Deploy
PiX4dEasy to use; supports cameras with a wider range; provides automated computing such as orthoimage, DSM generation, and cloud service; image data can be processed even without internet connection PiX4d
Sentera Ag VaultFeatures are similar to those of Pix4D and Drone Deploy; allows data collection for NIR, NDVI, and NDRE; allows the instant view of RGB and NIR data after capture; quick tile images can be generated in the field even without internet connectionSENTERA
Botlink MapperCreates high-definition maps, VI maps, and terrain maps to help find wet and dry areas; NIR photography helps to monitor and analyze the crops’ health status; allows easy collage of aerial images into a single, high-definition map; creates stunning digital surface and 3D models to identify high and low points or drainage issuesBotlink
Icaros OneButtonEasy to use, very high-quality processing, and program to monitor UAV flight periodIcaros
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tayade, R.; Yoon, J.; Lay, L.; Khan, A.L.; Yoon, Y.; Kim, Y. Utilization of Spectral Indices for High-Throughput Phenotyping. Plants 2022, 11, 1712. https://doi.org/10.3390/plants11131712

AMA Style

Tayade R, Yoon J, Lay L, Khan AL, Yoon Y, Kim Y. Utilization of Spectral Indices for High-Throughput Phenotyping. Plants. 2022; 11(13):1712. https://doi.org/10.3390/plants11131712

Chicago/Turabian Style

Tayade, Rupesh, Jungbeom Yoon, Liny Lay, Abdul Latif Khan, Youngnam Yoon, and Yoonha Kim. 2022. "Utilization of Spectral Indices for High-Throughput Phenotyping" Plants 11, no. 13: 1712. https://doi.org/10.3390/plants11131712

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