Next Article in Journal
High Planting Density Combined with Delayed Topping Improves Short Fruiting Branch Cotton Yield by Enhancing Biomass Accumulation, Canopy Light Interception and Delaying Leaf Senescence
Next Article in Special Issue
Intercropping of Cereals with Lentil: A New Strategy for Producing High-Quality Animal and Human Food
Previous Article in Journal
Vvmrp1, Vvmt1, and Vvmt2 Co-Expression Improves Cadmium Tolerance and Reduces Cadmium Accumulation in Rice
Previous Article in Special Issue
CdGLK1 Transcription Factor Confers Low-Light Tolerance in Bermudagrass via Coordinated Upregulation of Photosynthetic Genes and Enhanced Antioxidant Enzyme Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Genome Era of Forage Selection: Current Status and Future Directions for Perennial Ryegrass Breeding and Evaluation

1
Faculty of Science, The University of Melbourne, Parkville, VIC 3052, Australia
2
Agriculture Victoria Research, AgriBio Centre, 5 Ring Road, Bundoora, VIC 3083, Australia
3
Agriculture Victoria Research, 915 Mount Napier Road, Hamilton, VIC 3300, Australia
4
School of Applied Systems Biology, La Trobe University, Bundoora, VIC 3086, Australia
5
Agriculture Victoria Research, 1301 Hazeldean Road, Ellinbank, VIC 3821, Australia
*
Author to whom correspondence should be addressed.
Agronomy 2025, 15(6), 1494; https://doi.org/10.3390/agronomy15061494
Submission received: 14 May 2025 / Revised: 10 June 2025 / Accepted: 18 June 2025 / Published: 19 June 2025

Abstract

Perennial ryegrass (Lolium perenne L.) is a cornerstone forage species in temperate dairy systems worldwide, valued for its high yield potential, nutritive quality, and grazing recovery. However, current regional evaluation systems face challenges in accurately assessing complex traits like seasonal dry matter yield due to polygenic nature, environmental variability, and lengthy evaluation cycles. This review examines the evolution of perennial ryegrass evaluation systems, from regional frameworks—like Australia’s Forage Value Index (AU-FVI), New Zealand’s Forage Value Index (NZ-FVI), and Ireland’s Pasture Profit Index (PPI)—to advanced genomic prediction (GP) approaches. We discuss prominent breeding frameworks—F2 family, Half-sib family, and Synthetic Population—and their integration with high-throughput genotyping technologies. Statistical models for GP are compared, including marker-based, kernel-based, and non-parametric approaches, highlighting their strengths in capturing genetic complexity. Key research efforts include representative genotyping approaches for heterozygous populations, disentangling endophyte–host interactions, extending prediction to additional economically important traits, and modeling genotype-by-environment (G × E) interactions. The integration of multi-omics data, advanced phenotyping technologies, and environmental modeling offers promising avenues for enhancing prediction accuracy under changing environmental conditions. By discussing the combination of regional evaluation systems with GP, this review provides comprehensive insights for enhancing perennial ryegrass breeding and evaluation programs, ultimately supporting sustainable productivity of the dairy industry in the face of climate challenges.

Graphical Abstract

1. Perennial Ryegrass

Perennial ryegrass (Lolium perenne L.) is a cornerstone forage species in the dairy industry. It has fibrous roots and narrow blades and is densely tillered to form swards under grazing [1,2]. Native to Europe, Asia, and northern Africa, the species has been introduced worldwide and thrives under diverse climatic conditions [1,3,4,5,6,7]. Coupling a high yield [8,9,10,11], sufficient nutritive profile [12,13,14,15,16], and rapid recovery rate post grazing [17,18], perennial ryegrass has been the primary perennial grass species sown in grazed dairy systems across temperate regions of the world.

2. Regional Evaluation Systems

The assessment of perennial ryegrass agronomic performance is primarily on traits that directly impact livestock productivity and farm profitability. Dry matter yield (DMY) determines feed availability, while nutritive traits such as metabolizable energy (ME) [12,13], crude protein [14,19], water-soluble carbohydrates [14,16,19], and digestibility [14,15,19,20] influence milk production and feed efficiency.
In Australia, perennial ryegrass forms the backbone of the forage supply for dairy farms, particularly in the temperate region of southeastern Australia (including Victoria, Tasmania, and South Australia), collectively contributing 80.4% of the national milk production [5,10,21]. However, over 60 commercial cultivars are available on the market and exhibit significant variations in performance [13,22]. As such, an Australian Forage Value Index (AU-FVI) was developed to evaluate cultivar performance independently and guide farmers in selecting the most profitable options. The AU-FVI assesses pasture performance across five seasonal periods—Autumn, Winter, Early Spring, Late Spring, and Summer—and incorporates regional economic values to provide ratings [10,23]. Recent developments have expanded the AU-FVI to include ME, enhancing its ability to evaluate nutritional value [12].
New Zealand’s Forage Value Index (NZ-FVI) adopts a similar economic index approach to rate perennial ryegrass cultivars across four dairy regions based on seasonal DMY [24], wherein farm system modeling revealed substantial economic differences between highest and lowest ranked cultivars, ranging from NZD 556–863/ha/year depending on dairy regions [11]. A recent study investigated environmental variation and identified two distinct mega-environments—the upper North Island and the rest of New Zealand—providing a more accurate estimation of the regional economic values [11]. However, translating performance advantages demonstrated in small-plot trials to commercial farm settings remains challenging, as NZ farmlet studies showed such differences cannot consistently result in measurable gains in pasture yields, suggesting the need for more representative farm-scale data [11,25].
Ireland’s Pasture Profit Index (PPI) provides a more comprehensive evaluation framework by assigning economic values (EUR/ha/year) to traits including seasonal DMY, silage production, and nutritive parameters [9]. The economic values had proven remarkably consistent when tested across different farming systems, milk prices, and management intensities, with rank correlations between scenarios ranging from 0.90 to 1.0. This high consistency allows farmers to select cultivars that will deliver stable economic returns in farm-scale pastures while guiding breeders on economically important traits [9].
Notably, perennial ryegrass DMY is the primary focus in all regional evaluation systems because it represents the most economically important trait for farmers and serves as the fundamental selection criterion in breeding programs aimed at improving pasture productivity. However, the reliable evaluation of such traits depends heavily on the breeding frameworks.

3. Beeding Frameworks

Evaluation varies based on the agronomic traits of interest and local field trial designs employed within breeding approaches. Three primary frameworks—F2 family, Half-Sib family, and Synthetic Population frameworks—are widely adopted in perennial ryegrass breeding, each offering distinct advantages when evaluating diverse traits.

3.1. F2 Family Framework

The F2 family framework follows a structured evaluation process [19,26,27,28]. This approach begins with the selection of two superior parents based on multi-year spaced plant trials. The selected parents are crossed to produce F1 progeny, which remains heterozygous due to the outcrossing nature of ryegrass. These F1 individuals undergo random mating to generate F2 families, which are evaluated in sward trials for key agronomic traits such as yield and forage quality.
The F2 family evaluation framework is effective in assessing polygenic traits, and the ability to develop synthetic varieties from selected F2 families allows for faster cultivar development [19,26,27,28].

3.2. Half-Sib Family Framework

The half-sib framework provides an alternative evaluation strategy by focusing on maternal lines. This approach involves controlled polycrossing, where maternal plants are pollinated by multiple surrounding pollen donors [16,29,30], resulting in each plant within a half-sib family sharing the same maternal genetic material but with different paternal contributions.
The framework allows for identifying elite maternal plants through multi-staged family row trials to select superior half-sib families via progeny testing to assess traits such as herbage accumulation and nutritive quality [15,29]. For instance, Faville et al. evaluated 517 half-sib families in row trials, enabling precise evaluation of maternal lines [29].

3.3. Synthetic Population Framework

The synthetic population framework (Figure 1) manages genetic diversity by intercrossing multiple parents, ensuring high levels of heterozygosity, which is beneficial for long-term genetic improvement and trait stability across different environments [31].
However, this framework requires long evaluation cycles [31,32,33]. The process begins with controlled crosses between two elite ryegrass cultivars to generate an F1 population, which is then intercrossed to produce the F2 generation. These F2 plants undergo evaluation through spaced plant trials and clonal row trials to assess key agronomic traits. The most promising plants then undergo multiple rounds of polycrossing to generate synthetic populations, which may ultimately be released as commercial cultivars or advanced further in the breeding cycle.

3.4. Limitation of Phenotypic Evaluation

Each evaluation framework used in forage breeding presents distinct advantages in measuring plant performance. However, assessing complex traits remains challenging that are controlled by many genes (i.e., polygenic) and susceptible to environmental fluctuations, making phenotypic estimates biased to local trial conditions [10,11,22,24,25,29,31,32,33,34,35,36,37].
Differences in locations, climatic conditions, soil fertility, trial plot size, harvest intervals, and scoring methodologies often lead to inconsistent evaluation across different breeding programs. This inconsistency complicates the direct comparison of trait evaluation across environments and protocols [10,11,24,25,29,33,36].
Furthermore, traditional phenotypic selection remains resource-intensive, requiring repeated manual measurements across seasons and years, with high labor and financial costs [32,33,36]. This restricts the scale of evaluation and slows down genetic gain per breeding cycle.
These limitations inherent in the traditional phenotypic evaluation emphasize the need for integrating high throughput genotyping and advanced statistical models that account for environmental variation to select superior genotypes across diverse environments [38,39].

4. Technological Advancements in Genetic-Phenotypic Association Studies

4.1. Marker-Assisted Selection

The late 20th century witnessed the emergence of marker-assisted selection (MAS), by exploiting linkage disequilibrium (LD) patterns between markers and quantitative trait loci (QTLs), such as genome-wide association studies (GWAS) and QTL mapping [40,41,42,43,44,45,46,47,48,49].
The evolution of marker technologies, including restriction fragment length polymorphism (RFLP) markers [50,51], polymerase chain reaction (PCR)-based random amplified polymorphic deoxyribonucleic acid (DNA) markers [52], amplified fragment length polymorphism (AFLP) markers [53], and simple sequence repeats (SSRs) [54] established a foundation for molecular-based evaluation. Recently, single nucleotide polymorphisms (SNPs) [55,56,57] have enabled polymorphism to have a greater chance of being the causal agent of QTLs and make it possible to evaluate traits controlled by multiple QTLs [44,58,59].
Despite the advances, MAS faced challenges in estimating the traits controlled by numerous markers with individually non-significant but collectively significant effects; perennial ryegrass DMY is such a trait [20,33,46,58,60,61,62,63], necessitating the development of high-throughput genotyping approaches.

4.2. Advancements in High-Throughput Genotyping Technologies

High-throughput sequencing technologies have revolutionized genotyping strategy. The application of SNP arrays represented an early breakthrough [64,65], while restriction enzyme-based genotyping-by-sequencing (GBS) improved genotyping resolution through denser marker discovery by fragmenting the entire genome [29,57,66,67].
Further, GBS transcriptomics (GBS-t) targets expressed genomic regions rather than random fragments. This approach provides higher marker density in functionally relevant regions and more consistent coverage across samples with diverse genetic backgrounds. GBS-t has proven effective, with acceptable missing data rates, while focusing on biologically relevant genomic regions in several crop species [66].
Costs of high-throughput genotyping have decreased alongside technical improvements, enabling practical implementation in large-scale genotyping studies. A primary reduction was through multiplexing processes like sample pooling [29,64,65,66]. This could reduce costs by up to 60% while maintaining genotyping accuracy [65,66]. This advance builds upon reducing the use of sequencing reagents and other consumables. Streamlined library preparation methods further improve efficiency by optimizing laboratory workflows and minimizing processing time [29,64,65,66,68].

4.3. Genomic Prediction

Genomic prediction (GP) represents a transformative advancement in genetic evaluation, initially proposed by Meuwissen et al. [69]. Unlike conventional marker-assisted selection, which relies on the identification and utilization of a few genetic loci associated with major QTLs, GP leverages genome-wide molecular markers to capture both major and minor genetic effects. This approach allows for the prediction of complex polygenic traits, which are governed by many loci with small effects, without the need to identify specific causal variants.
The accuracy and robustness of GP hinge on its underlying statistical assumptions. Fundamentally, GP models assume additive genetic effects, whereby each marker contributes independently and additively to the phenotype, and the cumulative genetic value is expressed as the linear sum of these individual effects. A prediction can be implemented based on either pedigree information or genome-wide marker similarity. Pedigree-based genetic evaluations mainly assume related individuals share genetic materials inherited from common ancestors (Identity by Descent, IBD), where offspring phenotypes are predicted as approximately half of the parents’ genetic potential due to the random inheritance of genome segments during meiosis [63]. In contrast, similarity-based GP measures genetic similarity between individuals based on markers across the whole genome (Identity by State, IBS), regardless of their ancestral origin [36,63,69,70,71,72]. By capturing more precise relationship coefficients than the generalized assumptions in pedigree-based methods and combining with advanced statistical methodology, GP has enabled more accurate prediction than pedigree-based methods [36,63,69,70,71,72].
The effectiveness of GP in trait estimation has been demonstrated in various species. In dairy cattle, the USDA demonstrated its reliability in evaluating genetic merit [73,74]. In crop species, GP proved effective, as example studies in soybean [75] and wheat [76] showed high accuracy in trait prediction. The development of reference-free approaches has enabled GP application in non-model species like switchgrass [77]. These achievements across diverse species demonstrate the versatility of GP, suggesting strong potential for effective application in perennial ryegrass.
The applications of GP in perennial ryegrass have been reported in several traits. For traits with high heritability, such as heading date, GP yielded high predictive accuracies, ranging from 0.75 to 0.90 [14,26]. In addition, GP has shown strong potential in predicting nutritive traits such as crude protein content, water-soluble carbohydrates, and digestibility when leveraging optimal numbers of genome-wide markers [14,15,32,36,78].
Notably, recent research has extended the application of GP to the complex trait of DMY, which has traditionally posed challenges due to its low-to-moderate heritability and high sensitivity to environmental variation. In a recent study, GP was employed alongside large-scale, multi-harvest, multi-site (MHMS) field trial data and dense genotyping datasets comprising over 85 k high-quality SNPs to assess DMY across diverse Australian environments [38].
Nevertheless, achieving reliable GPs in perennial ryegrass requires continued efforts to identify optimal statistical models, incorporate non-additive effects, and refine training population design to maximize prediction accuracy and genetic gain.

5. Statistical Models for Genomic Prediction

Statistical models for GP vary in their approaches to estimating genetic effects and predicting trait performance. These models (Table 1 and Figure 2) can be broadly categorized into parametric approaches (marker-based and kernel-based regression models) and non-parametric approaches that adapt to data patterns during the training process.

5.1. Marker-Based Models

Marker-based regression models directly associate genetic markers with the traits in a linear way:
y = μ + M α + ϵ
where μ denotes the overall mean, M α accounts for marker effects, and ϵ represents residuals.
Marker-based models vary in their assumptions about marker effect distributions from normal distribution to flexible mixed distributions. For instance, the assumptions can be connected through L2 regularization in ridge regression, which is equivalent to assuming normally distributed marker effects in a Best Linear Unbiased Prediction (BLUP) framework (called rrBLUP) [20,29,30,31,47,57,68]. Furthermore, the LASSO and Elastic Net approaches are alternative regularization strategies, with LASSO utilizing L1 regularization and Elastic Net combining both L1 and L2 penalties. These methods have proven effective in performance estimation when handling polygenic traits with sparse marker associations [68,85]. Bayesian approaches, such as BayesA, BayesB, etc., provide probabilistic frameworks by incorporating flexible prior distributions for marker effects [15,68,79,84,86]. For example, Arojju et al. demonstrated that the Bayesian Lasso method achieved a predictive ability of 0.52 for crown rust resistance in perennial ryegrass [68].

5.2. Kernel-Based Models

Another type of prediction model estimates agronomic performances based on variance components through kernel functions f K derived from genetic markers M . Wherein, BLUP, using the kernel derived from genomic data, i.e., GBLUP, has emerged as a powerful approach for predicting random genotypic coefficients [28,29,68,79,86]. In general, these models follow a Linear Mixed Model (LMM) framework:
y = X β + Z u + ϵ
where X β accounts for known fixed effects (e.g., environmental or year-specific effects), Z u represents random genetic effects, and ϵ denotes residual effects following a normal distribution ( ϵ N 0 , σ ε 2 ). The variance components are estimated utilizing restricted maximum likelihood [87], and the prediction can be calculated via Henderson’s mixed model equations [80].
The genetic variance components V a r u are often approximated using the Genomic Relationship Matrix (GRM) among genotypes. The calculation of GRM varies based on marker assumptions but commonly follows G = M M / c , where c is a generic scaling factor [63]. When using the VanRaden GRM [83], which assumes Hardy–Weinberg equilibrium of the populations, GBLUP is mathematically equivalent to rrBLUP, as both assume independent and identically distributed (IID) marker effects with zero mean and constant variance [63]. For instance, Zhu et al. applied a GBLUP model using a GRM derived from ~86 k SNPs to predict dry matter yield of perennial ryegrass across 23 multi-environment trials, achieving a 12.7% improvement in predictive accuracy over the baseline model without genomic data [38].
The robustness of kernel-based models can be further enhanced by incorporating non-linear environmental kernels [19]. Taking Reproducing Kernel Hilbert Space (RKHS) as an example, it could capture non-additive genetic effects, such as epistatic effects, to enhance the predictive ability [16,79,88].

5.3. Non-Parametric Models

In addition, models that do not pre-define but dynamically adjust parameters during the modeling process are featured as non-parametric models. Random Forest (RF) is one such model and has emerged as an alternative in GP. Its successful implementation was reported in several perennial ryegrass prediction studies due to its ability to handle non-linear relationships and perform automatic feature selection [20,28,29,31,57,68]. For instance, in a Lolium perenne breeding study, RF achieved a prediction accuracy of 0.45 for dry matter digestibility, outperforming BLUP [20]. Other non-parametric approaches, such as Gradient Boosting Machines and K-Nearest Neighbors, were also explored for perennial ryegrass GP, demonstrating comparable efficiency [20].

5.4. Model Selection and Optimization Strategies

Implementing GP for perennial ryegrass requires a strategic alignment of statistical models with the biological complexity. One of the core decisions involves selecting appropriate statistical models. Among various GP models, GBLUP remains a preferred choice due to its high predictive accuracy and computational efficiency [15,16,20,26,27,28,29,30,32,38,39,78,79,81,82].
Equally important is the choice of cross-validation strategy, which influences the assessment of model adaptability. Random K-fold Cross-Validation is widely used for initial model evaluation, wherein the dataset is randomly divided into K subsets, with K-1 subsets used as training data and the remaining subset as validation data [26,27,31,32,78,89,90,91,92]. While this approach provides a general assessment of model robustness, it may overestimate prediction accuracy when the same populations are represented in both the training and validation sets [38,39,78,89,93].
Population-specific cross-validation addresses such limitations by partitioning data based on population structure. This strategy has been reported to provide more reliable genomic estimated breeding values (GEBVs) in untested populations [15,20,29,38,39,94].
Cross-validation can also be tailored to the traits of interest, particularly when considering multi-trait GP. For instance, different cross-validation schemes were employed to evaluate scenarios of primary and secondary traits [16], highlighting how the inclusion of correlated secondary traits can influence predictive ability.
Therefore, effective GP requires integrating both appropriate statistical models and validation techniques adapted to the prediction scenarios to ensure reliable estimation of perennial ryegrass [39].

5.5. Refinement of Genomic Prediction Models

Although GP models are promising, several factors should be considered for practical implementation. The effectiveness of GP is often influenced by genomic dissimilarity between training and target populations, with prediction accuracy declining when applied across distinct populations or breeding cycles. This challenge is particularly relevant for perennial ryegrass, given its diverse genetic background and complex breeding history [65,95,96,97].
Studies have shown accuracy decay between distinct populations or over generations [20,26,30,68,82,98]. The impact can be quantified by [99,100]:
r = N h 2 N h 2 + M e
where N is the number of individuals in the training population; h 2 is the heritability of the trait; and M e is the number of independent chromosome segments, which is related to effective population size and genome length.
This formula was later extended to account for more complex scenarios. For instance, Wientjes et al. [101] modified it to account for cross-population prediction: r g = c o r g r , where c o r g represents the genetic correlation between the reference population and the selection candidates. However, Brard et al. [102] argued that various formulas only provide upper bounds of accuracy, and actual performance is usually lower due to factors not considered in the theoretical frameworks.
Training population size remains another factor affecting prediction accuracy, where increasing the training population size could improve prediction performance [99,100]. Additionally, trait heritability and selection intensity interact to determine the model’s long-term predictive ability [32,82,99,100,103]. For perennial ryegrass breeding programs utilizing synthetic populations [31], regular model updating and careful management of selection pressure would be necessary to maintain prediction accuracy across breeding cycles.
Alternative models may be considered when non-additive genetic effects are significant, such as RF for non-linear interactions or Bayesian models when prior genetic knowledge is available [20,31,68,79]. Additionally, reducing SNP sets, while maintaining predictive ability, is required to develop cost-effective genotyping assays to make the routine implementation of GS practical [15,29].
For practical implementation, comprehensive consideration of appropriate statistical models, cross-validation strategies, and genotyping data quality [14,16,27,30,31,32,79,84,86,104] are essential for reliable genomic estimation of perennial ryegrass performance.

6. Research Gaps and Future Research Directions

6.1. Requirements for Representative Genotyping Approaches

Perennial ryegrass is a self-incompatible, outcrossing species that relies on polycrossing and lacks breeding history documentation, leading to heterozygosity bias and complicating the application of GP. Without representative genotyping strategies, the genotypic data obtained may fail to accurately capture the allelic spectrum within breeding populations, thereby reducing the reliability of GP [64,72,83,84,93,98,99,105,106,107]. Polyploidy further complicates marker-trait associations and makes conventional discrete encoded genotyping less effective [14,78,84].
Recent technological advances offer promising alternatives. One such approach is the integration of pooled sequencing with target capture assay [65,66,78,97]. Pooled sequencing allows for genotyping at the population level from multiple individuals into fewer sequencing libraries, thus reducing sequencing costs and mitigating individual-level sampling bias [65,66,78,97]. Furthermore, target capture, which selectively amplifies specific genomic regions using pre-designed probes [66], increases the coverage of genetic variants across heterozygous loci. When combined, these methods shift genotyping from discrete genotype calls to continuous allele frequency, reducing errors associated with variant misclassification.
Despite these methodological improvements, cross-platform incompatibility and the restricted availability of raw genomic data—often due to proprietary or commercial constraints—remain significant barriers to data integration and reuse. Addressing these challenges requires the adoption of advanced computational tools, including machine learning-based algorithms capable of harmonizing heterogeneous datasets, thereby supporting a comprehensive understanding of genomic relationships among ryegrass cultivars across studies [97].

6.2. Endophyte Symbiotic Impacts

Endophytes impact host agronomic performance through complex biological interactions, such as host compatibility, metabolic trade-offs, and alkaloid biosynthesis pathways [108,109,110,111]. These compounds can enhance plant stress resistance but may impose metabolic costs or negatively affect animal health [108,109,110]. As such, the selection of endophytes in ryegrass breeding must consider both plant productivity and animal safety [97].
Recent insights from Zhu et al. emphasized the need to improve the fairness of endophyte-ryegrass performance evaluations, particularly under the constraints of imbalanced datasets [97]. Due to commercial restrictions, certain endophytes are typically tested only within proprietary host cultivars, making it difficult to model endophyte impacts on host genetic effects. To address this, Zhu demonstrated that treating endophytes as fixed effects in statistical models provided distinct estimates of their contributions to DMY. To enable more balanced evaluations, they also suggested that future research explore biologically separating and recombining endophytes and ryegrass lines and conducting independent assessments of endophytes and ryegrasses across breeding programs. These approaches could help reduce confounding and support more robust prediction of untested combinations.
Together, these findings underscore the importance of integrating both statistical and biological perspectives to improve the understanding of endophyte–ryegrass interactions and optimize pasture productivity.

6.3. Extension to Broader Agronomic Trait Profiling

While the current studies mainly focused on DMY, extending the GP framework to other economically important traits would facilitate a more comprehensive evaluation system for perennial ryegrass. Nutritive characteristics [13,15], persistence [7,19], and stress tolerance [68,112,113,114,115] are critical factors that collectively determine the profitability of the dairy industry and other livestock systems where perennial ryegrass is a key pasture species. Multi-trait GP has been shown to improve modeling efficiency by leveraging correlations between traits, such as incorporating water-soluble carbohydrates to improve prediction accuracy for DMY [16].
Persistence can influence long-term productivity in perennial ryegrass pastures. Bornhofen showed that random regression models effectively capture temporal genetic variations across multiple harvests and years [19]. The models revealed how persistence traits change as plants age, enabling the estimation of both persistence and productivity [19].
Stress tolerance, including resistance to abiotic and biotic factors, remains another evaluation objective and has been studied using GP to improve resistance to crown rust [68]. By integrating multiple traits into a single prediction framework, evaluation systems can better align with industry needs, ultimately enhancing the sustainability of ryegrass-based dairy systems.

6.4. Integration of Multi-Omics Data to Improve Performance Estimation

Multi-omics approaches could facilitate trait estimation by integrating diverse biological layers: genomics serves as the base layer when associating QTLs with phenotypic variation [27,31,38,39,47,79,97,116], transcriptomics might reveal the regulation mechanisms [117,118,119,120,121], and proteomics could examine functional molecules such as cellular activities linked to photosynthetic efficiency and carbon allocation [122,123], while metabolomics might offer insights into the plant’s physiological state by profiling molecules in biomass-controlling pathways [124,125]. Such integrated multi-omics studies could provide a systematic understanding of how genetic potential translates into phenotypic performance under varying conditions.

6.5. Advancement in Phenotypic Data Collection

Phenotyping is a crucial step in trait estimation for perennial ryegrass, and its advancements may greatly improve the efficiency of trait data collection.
Sensing technologies using drones and specialized sensors could enable high-throughput measurements of plant growth, canopy structure, and physiological status [126,127,128,129]. Implementing these approaches also requires advanced image processing and time-series analysis methods to extract valid phenotypic traits from these high-dimensional data [128,130,131]. Additionally, the development of non-destructive techniques would allow for more frequent measurements without compromising plant performance [132,133]. These advancements would be particularly promising for capturing seasonal variations more accurately [129,134].

6.6. Integrating Environmental Data and Modeling Climate Adaptation

Genotype-by-environment (G × E) interactions are widespread in perennial ryegrass and significantly affect cultivar rankings across dairy farming environments, presenting a major challenge to the reliability of regional evaluation systems [11,22,37,38,39,135,136]. These interactions are driven by locational, temporal, and management-related environmental variability. For example, it was reported that water content could influence root development and overall plant growth across different locations [135]. Substantial cultivar ranking changes across soil moisture levels have also been observed [11], emphasizing the importance of accounting for environmental factors that present locational variation. Temporal variation often occurs on seasonal timescales and is linked to temperature patterns [22,37,38,39,136,137,138]. This variability can alter performance, complicating consistent estimation [22,37]. Similarly, management practices, such as irrigation [139,140] or grazing intensity [18], could also introduce performance variation.
Addressing the complex sources of phenotypic variation in perennial ryegrass requires comprehensive MHMS field trial data to capture temporal and spatial variability [37,38,39,141]. Recent research has demonstrated that the integration of advanced statistical models and MHMS data significantly improves prediction accuracy and underscores the potential of GP for environment-specific cultivar selection [37,38,39].
Several modeling frameworks have been developed to account for G × E interactions. The Factor Analytic (FA) model simplifies the variance–covariance structure across sites, capturing heterogeneous genetic variances and improving predictive performance in large-scale trials [37,38,39,142,143]. The Additive Main Effects and Multiplicative Interaction Model (AMMI) represents another well-accepted approach that models the variances of additive main effects using principal component analysis as interaction variance components. This enables separating systematic variation from noise while providing visual interpretation through GGE (genotype plus genotype-by-environment) biplots, making it valuable for identifying stable genotypes across environments [61,144,145]. The Reaction Norm Model extends GP by modeling phenotypic responses as functions of environmental gradients using higher-order covariates [19]. This approach directly incorporates environmental variabilities into genomic models, allowing for a more precise prediction under variable environmental conditions.
Furthermore, emerging machine learning techniques, such as recurrent neural networks (RNNs) and long short-term memory (LSTM) architectures, offer the potential for capturing temporal dependencies in multi-harvest datasets. These models maintain the memory of prior states, making them well-suited for modeling seasonal fluctuations [146,147,148].
Furthermore, the crop growth model (CGM) offers a mechanistic approach to simulate plant development by incorporating genetic and environmental variables [149]. Different from fitting environmental variation as variance components, CGM accounts for G × E interactions by mechanistically modeling how plants respond to the environments [149,150,151,152,153]. This is a process-based simulation that translates external factors into physiological responses within the plant. The simulated physiological processes may include photosynthesis, respiration, biomass partitioning and accumulation, and phenological development [149,150,154,155,156,157,158,159]. These processes allow CGMs to predict crop performance, such as perennial ryegrass, across diverse environments based on genotype-specific parameters.
As climate change introduces increasing unpredictability, including temperature extremes, droughts, and salinity stress [160,161,162,163], there is a pressing need to extend prediction models beyond historical data. The integration of G × E-enhanced GP with CGMs could facilitate cultivar performance predictions, supporting anticipatory breeding strategies [149,150,159,164]. CGMs can simulate physiological responses to environmental variables that include projected climate data [165,166]. By using G × E-enhanced GP results as response variables and CGM simulations as predictors, breeders could develop a framework that predicts cultivar performance under future climate scenarios and before implementing actual field trials [39,150,165,166].
This integrated approach could also refine mega-environment classification, which may shift with ongoing climate change. Incorporating time-series environmental data into the prediction could enable dynamic updates to mega-environment boundaries and help identify climate-resilient cultivars for future conditions. Ultimately, combining genomic, environmental, and physiological data will enable the optimization of forage breeding under increasingly variable agricultural conditions.

6.7. Effective Communication and Implementation

Translating cutting-edge research into practical farm decision-making requires strategic implementation through existing evaluation frameworks (Figure 3). While GP offers powerful capabilities for developing superior varieties, successful adoption requires effective communication that bridges the gap between researchers and industry stakeholders.
Understanding GP and statistical models presents initial barriers to adoption. Clarification of both capabilities and limitations is essential, particularly regarding G × E interactions and environmental data requirements for robust predictions. Industry implementation must address forage-specific challenges including representative genotyping protocols [65,66], endophyte interaction complexities [108,109,110], and commercial data sensitivity concerns [97].
Integration of GP outputs with regional evaluation systems including AU-FVI, NZ-FVI, and PPI would provide a viable pathway for assessing cultivar performance [9,10,11]. This integration would enable stakeholders to evaluate performance within a genomic-unified system, facilitating more informed decision-making.
Developing user-friendly interface tools that simplify complex modeling outputs into actionable breeding decisions would enhance industry uptake. Advanced visualization tools are suggested to present statistical outputs through intuitive graphics or interactive dashboards, allowing stakeholders to explore data without requiring deep statistical knowledge. However, the effective utilization of these tools will still require structured training opportunities to ensure that stakeholders can confidently interpret and apply GP outputs in their decision-making processes. Recently, multimodal large language model-based AI systems have emerged as promising interfaces for helping industry stakeholders understand GP principles and interpret complex outputs. For example, farmers and breeders could use AI-powered tools to quickly grasp technical concepts and statistical interpretations, such as understanding GGE biplots through conversational interfaces that explain the relationships between genotypes, environments, and their interactions in plain language. These AI assistants could serve as accessible entry points for stakeholders to engage with GP concepts without requiring extensive statistical training.
The economic implications of implementing GP can be better understood by articulating them to industry stakeholders through cost-benefit analyses. These analyses may demonstrate how GP reduces traditional evaluation cycle length and accelerates genetic gain (Figure 1 and [32,33,36,104]), thereby justifying the initial investment required for GP infrastructure. Comparative studies showing the economic advantages of genomic selection over conventional breeding approaches would help build a compelling case for industry adoption.
Fostering collaborative networks among researchers, breeders, and practitioners is vital for successful GP implementation. These collaborative networks would facilitate knowledge exchange and create feedback loops to improve prediction models continuously and could be formalized through industry-led initiatives or public-private partnerships to ensure the sustainability and effectiveness of GP implementation in perennial ryegrass breeding programs.

7. Conclusions

This review traced the evolution from traditional phenotypic evaluation frameworks to genomic prediction approaches in perennial ryegrass breeding systems. The cornerstone role of perennial ryegrass in global dairy production necessitates reliable performance evaluation, particularly for economically important traits like seasonal dry matter yield.
High-throughput genotyping technologies combined with advanced statistical models represents a fundamental shift in cultivar evaluation paradigms. Representative genotyping data that accounts for the self-incompatible and heterogeneous nature of perennial ryegrass have significantly advanced prediction ability while potentially reducing evaluation timeframes [32,33,36,38,39,70,104].
The comparative review of statistical models, from marker-based approaches to kernel-based and non-parametric methods, provides guidance for model selection aligned with ryegrass genetic complexities. Recent critical explorations include an improved understanding of endophyte impacts on performance estimation, extension to other economically important traits beyond DMY, and sophisticated G × E interaction modeling. However, research gaps remain in multi-omics integration, automated phenotyping methodologies, and environmental modeling enhancement—all essential for maintaining prediction accuracy under evolving climate conditions.
The pathway to practical implementation lies in integrating GP with established cultivar evaluation systems, particularly through AU-FVI, NZ-FVI, and PPI frameworks. Success requires coordinated collaboration among research institutions, breeding companies, and stakeholders to achieve continued model refinement and strategic communication. Through this approach, GP technology can fundamentally enhance perennial ryegrass breeding effectiveness, ultimately contributing to more sustainable and productive global pasture systems under climate variability challenges.

Author Contributions

Conceptualization: J.Z., K.F.S., N.O.C., K.G. and J.L.J.; Methodology: J.Z.; Formal analysis: J.Z.; Investigation: J.Z.; Writing—original draft preparation: J.Z.; Writing—review and editing: J.Z., K.G., N.O.C., K.F.S. and J.L.J.; Visualization: J.Z.; Supervision: K.F.S., N.O.C., J.L.J. and K.G.; Project administration: K.F.S., J.L.J. and N.O.C.; Validation: J.Z., K.F.S., N.O.C., K.G. and J.L.J. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Agriculture Victoria, Dairy Australia, and the Gardiner Dairy Foundation as part of the Victorian Dairy Innovation Agreement. Jiashuai Zhu was supported by a Melbourne Research Scholarship through the University of Melbourne and a Studentship (Stipend) from Agriculture Victoria Research through the Centre for Agricultural Innovation.

Data Availability Statement

Not applicable.

Acknowledgments

We acknowledge the advisory committee involved in the Forage Value Index (FVI) development, including members from the Australian Seed Federation and its member companies.

Conflicts of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Hannaway, D.; Fransen, S.; Cropper, J.; Teel, M.; Chaney, M.; Griggs, T.; Halse, R.; Hart, J.; Cheeke, P.; Klinger, R.; et al. Perennial Ryegrass (Lolium perenne L.). 1999. Available online: https://ir.library.oregonstate.edu/downloads/6108vb53s (accessed on 1 March 2025).
  2. Hunt, W.; Field, T. Growth characteristics of perennial ryegrass. J. N. Z. Grassl. 1978, 104–113. [Google Scholar] [CrossRef]
  3. Gilliland, T.B.T.; Hennessy, D. Opportunities and challenges for breeding perennial ryegrass cultivars with improved livestock production potential. Ir. J. Agric. Food Res. 2021, 59, 233–245. [Google Scholar] [CrossRef]
  4. Schubiger, F.X.; Baert, J.; Bayle, B.; Bourdon, P.; Cagas, B.; Cernoch, V.; Czembor, E.; Eickmeyer, F.; Feuerstein, U.; Hartmann, S.; et al. Susceptibility of European cultivars of Italian and perennial ryegrass to crown and stem rust. Euphytica 2010, 176, 167–181. [Google Scholar] [CrossRef]
  5. Leddin, C.; Giri, K.; Smith, K. Application and Analysis of a Composite Sampling Strategy to Cost-Effectively Compare Nutritive Characteristics of Perennial Ryegrass Cultivars in Field Trials. Agronomy 2020, 10, 1152. [Google Scholar] [CrossRef]
  6. Wilkins, P.W.; Rognli, O.A. Dry matter yield, herbage quality and persistency of equivalent populations of perennial ryegrass with and without reduced flowering. Plant Breed. 2002, 121, 425–428. [Google Scholar] [CrossRef]
  7. Waller, R.; Sale, P. Persistence and productivity of perennial ryegrass in sheep pastures in South-Western victoria: A review. Anim. Prod. Sci. 2001, 41, 117–144. [Google Scholar] [CrossRef]
  8. O’Donovan, M.; McHugh, N.; McEvoy, M.; Grogan, D.; Shalloo, L. Combining seasonal yield, silage dry matter yield, quality and persistency in an economic index to assist perennial ryegrass variety selection. J. Agric. Sci. 2016, 155, 556–568. [Google Scholar] [CrossRef]
  9. McEvoy, M.; O’Donovan, M.; Shalloo, L. Development and application of an economic ranking index for perennial ryegrass cultivars. J. Dairy Sci. 2011, 94, 1627–1639. [Google Scholar] [CrossRef]
  10. Leddin, C.; Jacobs, J.; Smith, K.; Giri, K.; Malcolm, B.; Ho, C. Development of a system to rank perennial ryegrass cultivars according to their economic value to dairy farm businesses in south-eastern Australia. Anim. Prod. Sci. 2018, 58, 1552–1558. [Google Scholar] [CrossRef]
  11. Chapman, D.F.; Bryant, J.R.; Olayemi, M.E.; Edwards, G.R.; Thorrold, B.S.; McMillan, W.H.; Kerr, G.A.; Judson, G.; Cookson, T.; Moorhead, A.; et al. An economically based evaluation index for perennial and short-term ryegrasses in New Zealand dairy farm systems. Grass Forage Sci. 2017, 72, 1–21. [Google Scholar] [CrossRef]
  12. Lewis, C.D.; Smith, K.F.; Jacobs, J.L.; Ho, C.K.M.; Leddin, C.M.; Moate, P.J.; Malcolm, B. Using a two-price market value framework to value differences in metabolizable energy concentration of pasture across seasons. Agric. Syst. 2024, 217, 103939. [Google Scholar] [CrossRef]
  13. Leddin, C.; Giri, K.; Smith, K. Variation in the Nutritive Characteristics of Modern Perennial Ryegrass Cultivars in South-Eastern Australian Dairy Environments and Prospects for Inclusion in the Australian Forage Value Index (FVI). Agronomy 2022, 12, 136. [Google Scholar] [CrossRef]
  14. Malmberg, M.; Smith, C.; Thakur, P.; Drayton, M.; Wilson, J.; Shinozuka, M.; Clayton, W.; Inch, C.; Spangenberg, G.; Smith, K.; et al. Developing an integrated genomic selection approach beyond biomass for varietal protection and nutritive traits in perennial ryegrass (Lolium perenne L.). Theor. Appl. Genet. 2023, 136, 44. [Google Scholar] [CrossRef] [PubMed]
  15. Arojju, S.K.; Cao, M.; Zulfi Jahufer, M.Z.; Barrett, B.A.; Faville, M.J. Genomic Predictive Ability for Foliar Nutritive Traits in Perennial Ryegrass. G3 Genes|Genomes|Genet. 2020, 10, 695–708. [Google Scholar] [CrossRef]
  16. Arojju, S.K.; Cao, M.; Trolove, M.; Barrett, B.A.; Inch, C.; Eady, C.; Stewart, A.; Faville, M.J. Multi-Trait Genomic Prediction Improves Predictive Ability for Dry Matter Yield and Water-Soluble Carbohydrates in Perennial Ryegrass. Front. Plant Sci. 2020, 11, 1197. [Google Scholar] [CrossRef]
  17. Rivero, M.; Oscar, B.; Fabián, L.N.; Siebald, J.A. Grazing Preference of Dairy Cows and Pasture Productivity for Different Cultivars of Perennial Ryegrass under Contrasting Managements. Animals 2019, 9, 253. [Google Scholar] [CrossRef] [PubMed]
  18. Rouquettes, F., Jr.; Bransby, D.; Riewe, M.E. Grazing management and use of ryegrass. Ecol. Prod. Manag. Lolium Forage USA 1997, 24, 79–99. [Google Scholar]
  19. Bornhofen, E.; Fè, D.; Lenk, I.; Greve, M.; Didion, T.; Jensen, C.S.; Asp, T.; Janss, L. Leveraging spatiotemporal genomic breeding value estimates of dry matter yield and herbage quality in ryegrass via random regression models. Plant Genome 2022, 15, e20255. [Google Scholar] [CrossRef]
  20. Grinberg, N.F.; Lovatt, A.; Hegarty, M.; Lovatt, A.; Skøt, K.P.; Kelly, R.; Blackmore, T.; Thorogood, D.; King, R.D.; Armstead, I.; et al. Implementation of Genomic Prediction in Lolium perenne (L.) Breeding Populations. Front. Plant Sci. 2016, 7, 133. [Google Scholar] [CrossRef]
  21. DairyAustralia. Perennial Ryegrass|Dairy Australia. 2025. Available online: https://www.dairyaustralia.com.au/ (accessed on 1 March 2025).
  22. Giri, K.; Chia, K.; Chandra, S.; Smith, K.F.; Leddin, C.M.; Ho, C.K.M.; Jacobs, J.L. Modelling and prediction of dry matter yield of perennial ryegrass cultivars sown in multi-environment multi-harvest trials in south-eastern Australia. Field Crops Res. 2019, 243, 107614. [Google Scholar] [CrossRef]
  23. Lewis, C.D.; Smith, K.F.; Jacobs, J.L.; Ho, C.K.M.; Leddin, C.M.; Malcolm, B. Using a two-price market value method to value extra pasture DM in different seasons. Agric. Syst. 2020, 178, 102729. [Google Scholar] [CrossRef]
  24. Chapman, D.F.; Edwards, G.R.; Stewart, A.V.; McEvoy, M.; O’Donovan, M.; Waghorn, G.C. Valuing forages for genetic selection: What traits should we focus on? Anim. Prod. Sci. 2015, 55, 869–882. [Google Scholar] [CrossRef]
  25. Crush, J.R.; Woodward, S.L.; Eerens, J.P.J.; MacDonald, K.A. Growth and milk solids production in pastures of older and more recent ryegrass and white clover cultivars under dairy grazing. N. Z. J. Agric. Res. 2006, 49, 119–135. [Google Scholar] [CrossRef]
  26. Fè, D.; Cericola, F.; Byrne, S.; Lenk, I.; Ashraf, B.; Pedersen, M.G.; Roulund, N.; Asp, T.; Janss, L.; Jensen, C.S.; et al. Genomic Dissection and Prediction of Heading Date in Perennial Ryegrass. BMC Genom. 2015, 16, 921. [Google Scholar] [CrossRef]
  27. Fè, D.; Ashraf, B.H.; Pedersen, M.G.; Janss, L.; Byrne, S.; Roulund, N.; Lenk, I.; Didion, T.; Asp, T.; Jensen, C.S.; et al. Accuracy of Genomic Prediction in a Commercial Perennial Ryegrass Breeding Program. Plant Genome 2016, 9, 1–12. [Google Scholar] [CrossRef] [PubMed]
  28. Cericola, F.; Lenk, I.; Fè, D.; Byrne, S.; Jensen, C.S.; Pedersen, M.G.; Asp, T.; Jensen, J.; Janss, L. Optimized Use of Low-Depth Genotyping-by-Sequencing for Genomic Prediction Among Multi-Parental Family Pools and Single Plants in Perennial Ryegrass (Lolium perenne L.). Front. Plant Sci. 2018, 9, 369. [Google Scholar] [CrossRef]
  29. Faville, M.J.; Ganesh, S.; Cao, M.; Jahufer, M.Z.Z.; Bilton, T.P.; Easton, H.S.; Ryan, D.L.; Trethewey, J.A.K.; Rolston, M.P.; Griffiths, A.G.; et al. Predictive ability of genomic selection models in a multi-population perennial ryegrass training set using genotyping-by-sequencing. Theor. Appl. Genet. 2018, 131, 703–720. [Google Scholar] [CrossRef]
  30. Jahufer, M.Z.Z.; Arojju, S.K.; Faville, M.J.; Ghamkhar, K.; Luo, D.; Arief, V.; Yang, W.-H.; Sun, M.; DeLacy, I.H.; Griffiths, A.G.; et al. Deterministic and stochastic modelling of impacts from genomic selection and phenomics on genetic gain for perennial ryegrass dry matter yield. Sci. Rep. 2021, 11, 13265. [Google Scholar] [CrossRef]
  31. Faville, M.J.; Ganesh, S.; Moraga, R.; Easton, H.S.; Jahufer, M.Z.Z.; Elshire, R.E.; Asp, T.; Barrett, B.A. Development of Genomic Selection for Perennial Ryegrass. In Breeding in a World of Scarcity; Springer International Publishing: Cham, Switzerland, 2016; pp. 139–143. [Google Scholar]
  32. Lin, Z.; Cogan, N.O.I.; Pembleton, L.W.; Spangenberg, G.C.; Forster, J.W.; Hayes, B.J.; Daetwyler, H.D. Genetic Gain and Inbreeding from Genomic Selection in a Simulated Commercial Breeding Program for Perennial Ryegrass. Plant Genome 2016, 9, 1–12. [Google Scholar] [CrossRef]
  33. Hayes, B.J.; Cogan, N.O.I.; Pembleton, L.W.; Goddard, M.E.; Wang, J.; Spangenberg, G.C.; Forster, J.W. Prospects for genomic selection in forage plant species. Plant Breed. 2013, 132, 133–143. [Google Scholar] [CrossRef]
  34. Conaghan, P.; Casler, M.D. A theoretical and practical analysis of the optimum breeding system for perennial ryegrass. Ir. J. Agric. Food Res. 2011, 50, 47–63. [Google Scholar]
  35. Casler, M.D.; Brummer, E.C. Theoretical Expected Genetic Gains for Among-and-Within-Family Selection Methods in Perennial Forage Crops. Crop Sci. 2008, 48, 890–902. [Google Scholar] [CrossRef]
  36. Barre, P.; Asp, T.; Byrne, S.; Casler, M.; Faville, M.; Rognli, O.A.; Roldan-Ruiz, I.; Skøt, L.; Ghesquière, M. Genomic Prediction of Complex TraitsComplex traits in Forage Plants Species: Perennial Grasses Case. In Genomic Prediction of Complex Traits: Methods and Protocols; Ahmadi, N., Bartholomé, J., Eds.; Springer: New York, NY, USA, 2022; pp. 521–541. [Google Scholar] [CrossRef]
  37. Zhu, J.; Giri, K.; Cogan, N.O.; Smith, K.F.; Jacobs, J.L. Genotype-by-environment interaction analysis of dry matter yield of perennial ryegrass cultivars across south-eastern Australia using factor analytic models. Field Crops Res. 2023, 303, 109143. [Google Scholar] [CrossRef]
  38. Zhu, J.; Giri, K.; Lin, Z.; Cogan, N.O.; Jacobs, J.L.; Smith, K.F. Estimation of ryegrass (Lolium) dry matter yield using genomic prediction considering genotype by environment interaction across south-eastern Australia. Front. Plant Sci. 2025, 16, 1579376. [Google Scholar] [CrossRef]
  39. Zhu, J. Utilising Genomic Relationships During the Estimation of Perennial Ryegrass Performance; The University of Melbourne: Parkville, VIC, Australia, 2025. [Google Scholar]
  40. Collard, B.C.Y.; Mackill, D.J. Marker-assisted selection: An approach for precision plant breeding in the twenty-first century. Philos. Trans. R. Soc. B Biol. Sci. 2008, 363, 557–572. [Google Scholar] [CrossRef]
  41. Unêda-Trevisoli, S.H.; da Silva, F.M.; Di Mauro, A.O. Marker-assisted selection and genomic selection. In Soybean Breeding, 1st ed.; da Silva, F.L., Borém, A., Sediyama, T., Ludke, W.H., Eds.; Springer: Cham, Switzerland, 2017; pp. 275–291. [Google Scholar] [CrossRef]
  42. Sim, S.; Diesburg, K.; Casler, M.; Jung, G. Mapping and Comparative Analysis of QTL for Crown Rust Resistance in an Italian × Perennial Ryegrass Population. Phytopathology 2007, 97, 767–776. [Google Scholar] [CrossRef]
  43. Schejbel, B.; Jensen, L.B.; Xing, Y.; Lübberstedt, T. QTL analysis of crown rust resistance in perennial ryegrass under conditions of natural and artificial infection. Plant Breed. 2007, 126, 347–352. [Google Scholar] [CrossRef]
  44. Hasan, N.; Choudhary, S.; Naaz, N.; Sharma, N.; Laskar, R.A. Recent advancements in molecular marker-assisted selection and applications in plant breeding programmes. J. Genet. Eng. Biotechnol. 2021, 19, 128. [Google Scholar] [CrossRef] [PubMed]
  45. Paina, C.; Byrne, S.; Studer, B.; Rognli, O.A.; Asp, T. Using a Candidate Gene-Based Genetic Linkage Map to Identify QTL for Winter Survival in Perennial Ryegrass. PLoS ONE 2016, 11, e0152004. [Google Scholar] [CrossRef]
  46. Emebiri, L.C.; Moody, D.B. Heritable basis for some genotype–environment stability statistics: Inferences from QTL analysis of heading date in two-rowed barley. Field Crops Res. 2006, 96, 243–251. [Google Scholar] [CrossRef]
  47. Keep, T.; Sampoux, J.-P.; Blanco-Pastor, J.L.; Dehmer, K.J.; Hegarty, M.J.; Ledauphin, T.; Litrico, I.; Muylle, H.; Roldán-Ruiz, I.; Roschanski, A.M.; et al. High-Throughput Genome-Wide Genotyping To Optimize the Use of Natural Genetic Resources in the Grassland Species Perennial Ryegrass (Lolium perenne L.). G3 Genes|Genomes|Genet. 2020, 10, 3347–3364. [Google Scholar] [CrossRef] [PubMed]
  48. Jaškūnė, K.; Aleliūnas, A.; Statkevičiūtė, G.; Kemešytė, V.; Studer, B.; Yates, S. Genome-Wide Association Study to Identify Candidate Loci for Biomass Formation Under Water Deficit in Perennial Ryegrass. Front. Plant Sci. 2020, 11, 570204. [Google Scholar] [CrossRef] [PubMed]
  49. Fois, M.; Bellucci, A.; Malinowska, M.; Greve, M.; Ruud, A.K.; Asp, T. Genome-wide association mapping of crown and brown rust resistance in perennial Ryegrass. Genes 2021, 13, 20. [Google Scholar] [CrossRef] [PubMed]
  50. de Souza, N. Get out the map. Nat. Rev. Genet. 2007, 8, S10. [Google Scholar] [CrossRef]
  51. Powell, W.; Morgante, M.; Andre, C.; Hanafey, M.; Vogel, J.; Tingey, S.; Rafalski, A. The comparison of RFLP, RAPD, AFLP and SSR (microsatellite) markers for germplasm analysis. Mol. Breed. 1996, 2, 225–238. [Google Scholar] [CrossRef]
  52. Niklas, A.; Olszewska, D. Application of the RAPD technique to identify genetic diversity in cultivated forms of Capsicum annuum L. Biotechnologia 2021, 102, 209–223. [Google Scholar] [CrossRef]
  53. Salgotra, R.K.; Stewart, C.N., Jr. Functional Markers for Precision Plant Breeding. Int. J. Mol. Sci. 2020, 21, 4792. [Google Scholar] [CrossRef]
  54. Gonçalves-Vidigal, M.C.; Rubiano, L.B. Development and application of microsatellites in plant breeding. Crop Breed. Appl. Biotechnol. 2011, 11, 66–72. [Google Scholar] [CrossRef]
  55. Birrer, M.; Kölliker, R.; Manzanares, C.; Asp, T.; Studer, B. A DNA Marker Assay Based on High-Resolution Melting Curve Analysis for Distinguishing Species of the Festuca–Lolium Complex. Mol. Breed. 2014, 34, 421–429. [Google Scholar] [CrossRef]
  56. Cogan, N.O.I.; Ponting, R.C.; Vecchies, A.C.; Drayton, M.C.; George, J.; Dracatos, P.M.; Dobrowolski, M.P.; Sawbridge, T.I.; Smith, K.F.; Spangenberg, G.C.; et al. Gene-associated single nucleotide polymorphism discovery in perennial ryegrass (Lolium perenne L.). Mol. Genet. Genom. 2006, 276, 101–112. [Google Scholar] [CrossRef]
  57. Byrne, S.L.; Conaghan, P.; Barth, S.; Arojju, S.K.; Casler, M.; Michel, T.; Velmurugan, J.; Milbourne, D. Using variable importance measures to identify a small set of SNPs to predict heading date in perennial ryegrass. Sci. Rep. 2017, 7, 3566. [Google Scholar] [CrossRef]
  58. Degen, B.; Müller, N.A. A simulation study comparing advanced marker-assisted selection with genomic selection in tree breeding programs. G3 Genes|Genomes|Genet. 2023, 13, jkad164. [Google Scholar] [CrossRef] [PubMed]
  59. Kushanov, F.N.; Turaev, O.S.; Ernazarova, D.K.; Gapparov, B.M.; Oripova, B.B.; Kudratova, M.K.; Rafieva, F.U.; Khalikov, K.K.; Erjigitov, D.S.; Khidirov, M.T.; et al. Genetic Diversity, QTL Mapping, and Marker-Assisted Selection Technology in Cotton (Gossypium spp.). Front. Plant Sci. 2021, 12, 779386. [Google Scholar] [CrossRef]
  60. Ebdon, J.S.; Gauch, H. Additive Main Effect and Multiplicative Interaction Analysis of National Turfgrass Performance Trials: I. Interpretation of Genotype × Environment Interaction. Crop Sci. 2002, 42, 489–496. [Google Scholar] [CrossRef]
  61. Fois, M.; Malinowska, M.; Schubiger, F.X.; Asp, T. Genomic Prediction and Genotype-by-Environment Interaction Analysis of Crown and Stem Rust in Ryegrasses in European Multi-Site Trials. Agronomy 2021, 11, 1119. [Google Scholar] [CrossRef]
  62. Conaghan, P.; Casler, M.D.; McGilloway, D.A.; O’Kiely, P.; Dowley, L.J. Genotype × environment interactions for herbage yield of perennial ryegrass sward plots in Ireland. Grass Forage Sci. 2008, 63, 107–120. [Google Scholar] [CrossRef]
  63. Isik, F.; Holland, J.; Maltecca, C. Genomic Relationships and GBLUP. In Genetic Data Analysis for Plant and Animal Breeding; Isik, F., Holland, J., Maltecca, C., Eds.; Springer International Publishing: Cham, Switzerland, 2017; pp. 311–354. [Google Scholar] [CrossRef]
  64. Wang, J.; Pembleton, L.W.; Baillie, R.C.; Drayton, M.C.; Hand, M.L.; Bain, M.; Sawbridge, T.I.; Spangenberg, G.C.; Forster, J.W.; Cogan, N.O.I. Development and implementation of a multiplexed single nucleotide polymorphism genotyping tool for differentiation of ryegrass species and cultivars. Mol. Breed. 2014, 33, 435–451. [Google Scholar] [CrossRef]
  65. Pembleton, L.W.; Drayton, M.C.; Bain, M.; Baillie, R.C.; Inch, C.; Spangenberg, G.C.; Wang, J.; Forster, J.W.; Cogan, N.O. Targeted genotyping-by-sequencing permits cost-effective identification and discrimination of pasture grass species and cultivars. Theor. Appl. Genet. 2016, 129, 991–1005. [Google Scholar] [CrossRef]
  66. Malmberg, M.M.; Pembleton, L.W.; Baillie, R.C.; Drayton, M.C.; Sudheesh, S.; Kaur, S.; Shinozuka, H.; Verma, P.; Spangenberg, G.C.; Daetwyler, H.D.; et al. Genotyping-by-sequencing through transcriptomics: Implementation in a range of crop species with varying reproductive habits and ploidy levels. Plant Biotechnol. J. 2018, 16, 877–889. [Google Scholar] [CrossRef]
  67. He, J.; Zhao, X.; Laroche, A.; Lu, Z.-X.; Liu, H.; Li, Z. Genotyping-by-sequencing (GBS), an ultimate marker-assisted selection (MAS) tool to accelerate plant breeding. Front. Plant Sci. 2014, 5, 484. [Google Scholar] [CrossRef]
  68. Arojju, S.K.; Conaghan, P.; Barth, S.; Milbourne, D.; Casler, M.D.; Hodkinson, T.R.; Michel, T.; Byrne, S.L. Genomic prediction of crown rust resistance in Lolium perenne. BMC Genet. 2018, 19, 35. [Google Scholar] [CrossRef] [PubMed]
  69. Meuwissen, T.H.; Hayes, B.J.; Goddard, M.E. Prediction of total genetic value using genome-wide dense marker maps. Genetics 2001, 157, 1819–1829. [Google Scholar] [CrossRef]
  70. Meuwissen, T.; Hayes, B.; Goddard, M. Genomic selection: A paradigm shift in animal breeding. Anim. Front. 2016, 6, 6–14. [Google Scholar] [CrossRef]
  71. Budhlakoti, N.; Kushwaha, A.K.; Rai, A.; Chaturvedi, K.K.; Kumar, A.; Pradhan, A.K.; Kumar, U.; Kumar, R.R.; Juliana, P.; Mishra, D.C.; et al. Genomic Selection: A Tool for Accelerating the Efficiency of Molecular Breeding for Development of Climate-Resilient Crops. Front. Genet. 2022, 13, 832153. [Google Scholar] [CrossRef] [PubMed]
  72. Alemu, A.; Åstrand, J.; Montesinos-López, O.; Sánchez, J.; Fernández-Gónzalez, J.; Tadesse, W.; Vetukuri, R.; Carlsson, A.; Ceplitis, A.; Crossa, J.; et al. Genomic selection in plant breeding: Key factors shaping two decades of progress. Mol. Plant 2024, 17, 552–578. [Google Scholar] [CrossRef] [PubMed]
  73. Wiggans, G.R.; Carrillo, J.A. Genomic selection in United States dairy cattle. Front Genet 2022, 13, 994466. [Google Scholar] [CrossRef]
  74. Wiggans, G.R.; Cole, J.B.; Hubbard, S.M.; Sonstegard, T.S. Genomic Selection in Dairy Cattle: The USDA Experience. Annu. Rev. Anim. Biosci. 2017, 5, 309–327. [Google Scholar] [CrossRef]
  75. Chen, Y.; Xiong, Y.; Hong, H.; Li, G.; Gao, J.; Guo, Q.; Sun, R.; Ren, H.; Zhang, F.; Wang, J.; et al. Genetic dissection of and genomic selection for seed weight, pod length, and pod width in soybean. Crop J. 2023, 11, 832–841. [Google Scholar] [CrossRef]
  76. Shahi, D.; Guo, J.; Pradhan, S.; Khan, J.; Avci, M.; Khan, N.; McBreen, J.; Bai, G.; Reynolds, M.; Foulkes, J.; et al. Multi-trait genomic prediction using in-season physiological parameters increases prediction accuracy of complex traits in US wheat. BMC Genom. 2022, 23, 298. [Google Scholar] [CrossRef]
  77. Lu, F.; Lipka, A.E.; Glaubitz, J.; Elshire, R.; Cherney, J.H.; Casler, M.D.; Buckler, E.S.; Costich, D.E. Switchgrass genomic diversity, ploidy, and evolution: Novel insights from a network-based SNP discovery protocol. PLoS Genet. 2013, 9, e1003215. [Google Scholar] [CrossRef]
  78. Guo, X.; Cericola, F.; Fè, D.; Pedersen, M.G.; Lenk, I.; Jensen, C.S.; Jensen, J.; Janss, L.L. Genomic Prediction in Tetraploid Ryegrass Using Allele Frequencies Based on Genotyping by Sequencing. Front. Plant Sci. 2018, 9, 1165. [Google Scholar] [CrossRef] [PubMed]
  79. Konkolewska, A.; Phang, S.; Conaghan, P.; Milbourne, D.; Lawlor, A.; Byrne, S. Genomic prediction of seasonal forage yield in perennial ryegrass. Grassl. Res. 2023, 2, 167–181. [Google Scholar] [CrossRef]
  80. Henderson, C.R. Best linear unbiased estimation and prediction under a selection model. Biometrics 1975, 31, 423–447. [Google Scholar] [CrossRef] [PubMed]
  81. Faville, M.; Schmidt, J.; Trolove, M.; Moran, P.; Hong, W.; Cao, M.; Ganesh, S.; George, R.; Barrett, B. Empirical assessment of a genomic breeding strategy in perennial ryegrass. J. N. Z. Grassl. 2021, 83, 115–122. [Google Scholar] [CrossRef]
  82. Lin, Z.; Wang, J.; Cogan, N.O.; Pembleton, L.W.; Badenhorst, P.; Forster, J.W.; Spangenberg, G.C.; Hayes, B.J.; Daetwyler, H.D. Optimizing resource allocation in a genomic breeding program for perennial ryegrass to balance genetic gain, cost, and inbreeding. Crop Sci. 2017, 57, 243–252. [Google Scholar] [CrossRef]
  83. VanRaden, P.M. Efficient Methods to Compute Genomic Predictions. J. Dairy Sci. 2008, 91, 4414–4423. [Google Scholar] [CrossRef]
  84. Pembleton, L.W.; Inch, C.; Baillie, R.C.; Drayton, M.C.; Thakur, P.; Ogaji, Y.O.; Spangenberg, G.C.; Forster, J.W.; Daetwyler, H.D.; Cogan, N.O.I. Exploitation of data from breeding programs supports rapid implementation of genomic selection for key agronomic traits in perennial ryegrass. Theor. Appl. Genet. 2018, 131, 1891–1902. [Google Scholar] [CrossRef]
  85. Lourenço, V.M.; Ogutu, J.O.; Rodrigues, R.A.P.; Posekany, A.; Piepho, H.P. Genomic prediction using machine learning: A comparison of the performance of regularized regression, ensemble, instance-based and deep learning methods on synthetic and empirical data. BMC Genom. 2024, 25, 152. [Google Scholar] [CrossRef]
  86. Esfandyari, H.; Fè, D.; Tessema, B.B.; Janss, L.L.G.; Jensen, J. Effects of Different Strategies for Exploiting Genomic Selection in Perennial Ryegrass Breeding Programs. G3 Genes|Genome|Genet. 2020, 10, 3783–3795. [Google Scholar] [CrossRef]
  87. Robinson, G.K. That BLUP is a Good Thing: The Estimation of Random Effects. Stat. Sci. 1991, 6, 15–32. [Google Scholar] [CrossRef]
  88. Gianola, D.; van Kaam, J.B. Reproducing kernel hilbert spaces regression methods for genomic assisted prediction of quantitative traits. Genetics 2008, 178, 2289–2303. [Google Scholar] [CrossRef] [PubMed]
  89. Runcie, D.; Cheng, H. Pitfalls and Remedies for Cross Validation with Multi-trait Genomic Prediction Methods. G3 (Bethesda Md.) 2019, 9, 3727–3741. [Google Scholar] [CrossRef] [PubMed]
  90. Shi, S.; Li, X.; Fang, L.; Liu, A.; Su, G.; Zhang, Y.; Luobu, B.; Ding, X.; Zhang, S. Genomic Prediction Using Bayesian Regression Models With Global–Local Prior. Front. Genet. 2021, 12, 628205. [Google Scholar] [CrossRef]
  91. Zhu, S.; Guo, T.; Yuan, C.; Liu, J.; Li, J.; Han, M.; Zhao, H.; Wu, Y.; Sun, W.; Wang, X.; et al. Evaluation of Bayesian alphabet and GBLUP based on different marker density for genomic prediction in Alpine Merino sheep. G3 Genes|Genomes|Genet. 2021, 11, jkab206. [Google Scholar] [CrossRef] [PubMed]
  92. Schrauf, M.F.; de Los Campos, G.; Munilla, S. Comparing Genomic Prediction Models by Means of Cross Validation. Front. Plant Sci. 2021, 12, 734512. [Google Scholar] [CrossRef]
  93. Resende, M.F., Jr.; Muñoz, P.; Resende, M.D.; Garrick, D.J.; Fernando, R.L.; Davis, J.M.; Jokela, E.J.; Martin, T.A.; Peter, G.F.; Kirst, M. Accuracy of genomic selection methods in a standard data set of loblolly pine (Pinus taeda L.). Genetics 2012, 190, 1503–1510. [Google Scholar] [CrossRef]
  94. Werner, C.R.; Gaynor, R.C.; Gorjanc, G.; Hickey, J.M.; Kox, T.; Abbadi, A.; Leckband, G.; Snowdon, R.J.; Stahl, A. How Population Structure Impacts Genomic Selection Accuracy in Cross-Validation: Implications for Practical Breeding. Front. Plant Sci. 2020, 11, 592977. [Google Scholar] [CrossRef]
  95. Guan, X.; Yuyama, N.; Stewart, A.; Ding, C.; Xu, N.; Kiyoshi, T.; Cai, H. Genetic Diversity and Structure of Lolium Species Surveyed on Nuclear Simple Sequence Repeat and Cytoplasmic Markers. Front. Plant Sci. 2017, 8, 584. [Google Scholar] [CrossRef]
  96. Blackmore, T.; Thorogood, D.; Skøt, L.; McMahon, R.; Powell, W.; Hegarty, M. Germplasm dynamics: The role of ecotypic diversity in shaping the patterns of genetic variation in Lolium perenne. Sci. Rep. 2016, 6, 22603. [Google Scholar] [CrossRef]
  97. Zhu, J.; Malmberg, M.M.; Shinozuka, M.; Retegan, R.M.; Cogan, N.O.; Jacobs, J.L.; Giri, K.; Smith, K.F. Machine learning solutions for integrating partially overlapping genetic datasets and modelling host–endophyte effects in ryegrass (Lolium) dry matter yield estimation. Front. Plant Sci. 2025, 16, 1543956. [Google Scholar] [CrossRef]
  98. Daetwyler, H.D.; Kemper, K.E.; van der Werf, J.H.J.; Hayes, B.J. Components of the accuracy of genomic prediction in a multi-breed sheep population. J. Anim. Sci. 2012, 90, 3375–3384. [Google Scholar] [CrossRef] [PubMed]
  99. Daetwyler, H.D.; Villanueva, B.; Woolliams, J.A. Accuracy of predicting the genetic risk of disease using a genome-wide approach. PLoS ONE 2008, 3, e3395. [Google Scholar] [CrossRef] [PubMed]
  100. Daetwyler, H.D.; Pong-Wong, R.; Villanueva, B.; Woolliams, J.A. The impact of genetic architecture on genome-wide evaluation methods. Genetics 2010, 185, 1021–1031. [Google Scholar] [CrossRef]
  101. Wientjes, Y.C.J.; Veerkamp, R.F.; Bijma, P.; Bovenhuis, H.; Schrooten, C.; Calus, M.P.L. Empirical and deterministic accuracies of across-population genomic prediction. Genet. Sel. Evol. 2015, 47, 5. [Google Scholar] [CrossRef]
  102. Brard, S.; Ricard, A. Is the use of formulae a reliable way to predict the accuracy of genomic selection? J. Anim. Breed. Genet. 2015, 132, 207–217. [Google Scholar] [CrossRef]
  103. Newell, M.A.; Jannink, J.L. Genomic Selection in Plant Breeding. In Crop Breeding; Methods in Molecular Biology; Fleury, D., Whitford, R., Eds.; Humana Press: New York, NY, USA, 2014; Volume 1145, pp. 117–130. [Google Scholar] [CrossRef]
  104. Barrett, B.; Jahufer, Z.; Arojju, S.; Sise, J.; Faville, M. Forecasting the genetic and economic impacts of genomic selection in perennial ryegrass. J. N. Z. Grassl. 2022, 83, 92–98. [Google Scholar] [CrossRef]
  105. Habier, D.; Fernando, R.L.; Dekkers, J.C.M. The Impact of Genetic Relationship Information on Genome-Assisted Breeding Values. Genetics 2007, 177, 2389–2397. [Google Scholar] [CrossRef] [PubMed]
  106. Goddard, M. Genomic selection: Prediction of accuracy and maximisation of long term response. Genetica 2008, 136, 245–257. [Google Scholar] [CrossRef]
  107. Yang, J.; Benyamin, B.; McEvoy, B.P.; Gordon, S.; Henders, A.K.; Nyholt, D.R.; Madden, P.A.; Heath, A.C.; Martin, N.G.; Montgomery, G.W.; et al. Common SNPs explain a large proportion of the heritability for human height. Nat. Genet. 2010, 42, 565–569. [Google Scholar] [CrossRef]
  108. Vassiliadis, S.; Reddy, P.; Hemsworth, J.; Spangenberg, G.C.; Guthridge, K.M.; Rochfort, S.J. Quantitation and Distribution of Epichloë-Derived Alkaloids in Perennial Ryegrass Tissues. Metabolites 2023, 13, 205. [Google Scholar] [CrossRef]
  109. Popay, A.J.; Hume, D.E. Endophytes for Improving Ryegrass Performance: Current Status and Future Possibilities. In Proceedings of the 22nd International Grassland Congress, Sydney, Australia, 15–19 September 2013; pp. 1625–1626. [Google Scholar]
  110. Eady, C. The Impact of Alkaloid-Producing Epichloë Endophyte on Forage Ryegrass Breeding: A New Zealand Perspective. Toxins 2021, 13, 158. [Google Scholar] [CrossRef] [PubMed]
  111. Karpyn Esqueda, M.; Yen, A.L.; Rochfort, S.; Guthridge, K.M.; Powell, K.S.; Edwards, J.; Spangenberg, G.C. A Review of Perennial Ryegrass Endophytes and Their Potential Use in the Management of African Black Beetle in Perennial Grazing Systems in Australia. Front. Plant Sci. 2017, 8, 3. [Google Scholar] [CrossRef] [PubMed]
  112. Song, X.; Wang, S.-m.; Jiang, Y. Genotypic Variations in Plant Growth and Nutritional Elements of Perennial Ryegrass Accessions under Salinity Stress. J. Am. Soc. Hortic. Sci. 2017, 142, 476–483. [Google Scholar] [CrossRef]
  113. Miao, C.; Zhang, Y.; Bai, X.; Qin, T. Insights into the Response of Perennial Ryegrass to Abiotic Stress: Underlying Survival Strategies and Adaptation Mechanisms. Life 2022, 12, 860. [Google Scholar] [CrossRef] [PubMed]
  114. Kemesyte, V.; Statkeviciute, G.; Brazauskas, G. Perennial Ryegrass Yield Performance under Abiotic Stress. Crop Sci. 2017, 57, 1935–1940. [Google Scholar] [CrossRef]
  115. Fuchun, X.; Rahul, D.; Dong, Q. Plant Growth and Morphophysiological Modifications in Perennial Ryegrass under Environmental Stress. In Abiotic Stress in Plants; Shah, F., Shah, S., Yajun, C., Chao, W., Depeng, W., Eds.; IntechOpen: Rijeka, Croatia, 2020; p. Ch. 17. [Google Scholar] [CrossRef]
  116. Nagy, I.; Veeckman, E.; Liu, C.; Bel, M.V.; Vandepoele, K.; Jensen, C.S.; Ruttink, T.; Asp, T. Chromosome-scale assembly and annotation of the perennial ryegrass genome. BMC Genom. 2022, 23, 505. [Google Scholar] [CrossRef]
  117. Li, Y.; Chen, L.; Chen, W.; Zhu, J.; Chen, Y.; Li, D. Transcriptomic analysis of the metabolic regulatory mechanism of Schizochytrium limacinum B4D1 using sodium acetate to produce DHA. Biochem. Eng. J. 2023, 197, 108963. [Google Scholar] [CrossRef]
  118. Luo, D.; Zhou, Q.; Wu, Y.; Chai, X.; Liu, W.; Wang, Y.; Yang, Q.; Wang, Z.; Liu, Z. Full-length transcript sequencing and comparative transcriptomic analysis to evaluate the contribution of osmotic and ionic stress components towards salinity tolerance in the roots of cultivated alfalfa (Medicago sativa L.). BMC Plant Biol. 2019, 19, 32. [Google Scholar] [CrossRef]
  119. Jinqiu, Y.; Bing, L.; Tingting, S.; Jinglei, H.; Zelai, K.; Lu, L.; Wenhua, H.; Tao, H.; Xinyu, H.; Zengqing, L.; et al. Integrated Physiological and Transcriptomic Analyses Responses to Altitude Stress in Oat (Avena sativa L.). Front. Genet. 2021, 12, 638683. [Google Scholar] [CrossRef]
  120. Varoquaux, N.; Cole, B.; Gao, C.; Pierroz, G.; Baker, C.R.; Patel, D.; Madera, M.; Jeffers, T.; Hollingsworth, J.; Sievert, J.; et al. Transcriptomic analysis of field-droughted sorghum from seedling to maturity reveals biotic and metabolic responses. Proc. Natl. Acad. Sci. USA 2019, 116, 27124–27132. [Google Scholar] [CrossRef]
  121. Lu, A.; Zeng, S.; Pi, K.; Long, B.; Mo, Z.; Liu, R. Transcriptome analysis reveals the key role of overdominant expression of photosynthetic and respiration-related genes in the formation of tobacco (Nicotiana tabacum L.) biomass heterosis. BMC Genom. 2024, 25, 598. [Google Scholar] [CrossRef]
  122. Faizi, M.; Steuer, R. Optimal proteome allocation strategies for phototrophic growth in a light-limited chemostat. Microb. Cell Factories 2019, 18, 165. [Google Scholar] [CrossRef] [PubMed]
  123. Chen, Q.; Xiao, Y.; Ming, Y.; Peng, R.; Hu, J.; Wang, H.-B.; Jin, H.-L. Quantitative proteomics reveals redox-based functional regulation of photosynthesis under fluctuating light in plants. J. Integr. Plant Biol. 2022, 64, 2168–2186. [Google Scholar] [CrossRef] [PubMed]
  124. Wang, S.; Zhou, X.; Wu, S.; Zhao, M.; Hu, Z. Transcriptomic and metabolomic analyses revealed regulation mechanism of mixotrophic Cylindrotheca sp. glycerol utilization and biomass promotion. Biotechnol. Biofuels Bioprod. 2023, 16, 84. [Google Scholar] [CrossRef]
  125. Gao, B.; Wang, F.; Huang, L.; Liu, H.; Zhong, Y.; Zhang, C. Biomass, lipid accumulation kinetics, and the transcriptome of heterotrophic oleaginous microalga Tetradesmus bernardii under different carbon and nitrogen sources. Biotechnol. Biofuels 2021, 14, 4. [Google Scholar] [CrossRef]
  126. Pranga, J.; Borra-Serrano, I.; Aper, J.; De Swaef, T.; Ghesquiere, A.; Quataert, P.; Roldán-Ruiz, I.; Janssens, I.A.; Ruysschaert, G.; Lootens, P. Improving Accuracy of Herbage Yield Predictions in Perennial Ryegrass with UAV-Based Structural and Spectral Data Fusion and Machine Learning. Remote Sens. 2021, 13, 3459. [Google Scholar] [CrossRef]
  127. Tanaka, T.S.T.; Wang, S.; Jørgensen, J.R.; Gentili, M.; Vidal, A.Z.; Mortensen, A.K.; Acharya, B.S.; Beck, B.D.; Gislum, R. Review of Crop Phenotyping in Field Plot Experiments Using UAV-Mounted Sensors and Algorithms. Drones 2024, 8, 212. [Google Scholar] [CrossRef]
  128. Gebremedhin, A.; Badenhorst, P.; Wang, J.; Shi, F.; Breen, E.; Giri, K.; Spangenberg, G.C.; Smith, K. Development and Validation of a Phenotyping Computational Workflow to Predict the Biomass Yield of a Large Perennial Ryegrass Breeding Field Trial. Front. Plant Sci. 2020, 11, 689. [Google Scholar] [CrossRef]
  129. Wang, J.; Badenhorst, P.; Phelan, A.; Pembleton, L.; Shi, F.; Cogan, N.; Spangenberg, G.; Smith, K. Using Sensors and Unmanned Aircraft Systems for High-Throughput Phenotyping of Biomass in Perennial Ryegrass Breeding Trials. Front. Plant Sci. 2019, 10, 1381. [Google Scholar] [CrossRef]
  130. Parasurama, S.; Banan, D.; Yun, K.; Doty, S.; Kim, S.-H. Bridging Time-series Image Phenotyping and Functional–Structural Plant Modeling to Predict Adventitious Root System Architecture. Plant Phenomics 2023, 5, 0127. [Google Scholar] [CrossRef]
  131. Wang, H.; Singh, K.D.; Poudel, H.P.; Natarajan, M.; Ravichandran, P.; Eisenreich, B. Forage Height and Above-Ground Biomass Estimation by Comparing UAV-Based Multispectral and RGB Imagery. Sensors 2024, 24, 5794. [Google Scholar] [CrossRef] [PubMed]
  132. Rahaman, M.M.; Chen, D.; Gillani, Z.; Klukas, C.; Chen, M. Advanced phenotyping and phenotype data analysis for the study of plant growth and development. Front. Plant Sci. 2015, 6, 619. [Google Scholar] [CrossRef] [PubMed]
  133. Ludovisi, R.; Tauro, F.; Salvati, R.; Khoury, S.; Mugnozza Scarascia, G.; Harfouche, A. UAV-Based Thermal Imaging for High-Throughput Field Phenotyping of Black Poplar Response to Drought. Front. Plant Sci. 2017, 8, 1681. [Google Scholar] [CrossRef]
  134. Nguyen, P.T.; Shi, F.; Wang, J.; Badenhorst, P.E.; Spangenberg, G.C.; Smith, K.F.; Daetwyler, H.D. Within and combined season prediction models for perennial ryegrass biomass yield using ground- and air-based sensor data. Front. Plant Sci. 2022, 13, 950720. [Google Scholar] [CrossRef]
  135. Wedderburn, M.E.; Crush, J.R.; Pengelly, W.J.; Walcroft, J.L. Root growth patterns of perennial ryegrasses under well-watered and drought conditions. N. Z. J. Agric. Res. 2010, 53, 377–388. [Google Scholar] [CrossRef]
  136. Colas, V.; Barre, P.; van Parijs, F.; Wolters, L.; Quitté, Y.; Ruttink, T.; Roldán-Ruiz, I.; Escobar Gutiérrez, A.J.; Muylle, H. Seasonal differences in structural and genetic control of digestibility in perennial ryegrass. Front. Plant Sci. 2022, 12, 801145. [Google Scholar] [CrossRef]
  137. Piepho, H.-P.; Eckl, T. Analysis of series of variety trials with perennial crops. Grass Forage Sci. 2014, 69, 431–440. [Google Scholar] [CrossRef]
  138. Förster, L.; Grant, J.; Michel, T.; Ng, C.; Barth, S. Growth under cold conditions in a wide perennial ryegrass panel is under tight physiological control. PeerJ 2018, 6, e5520. [Google Scholar] [CrossRef]
  139. Waldron, B.; Asay, K.; Jensen, K. Stability and Yield of Cool-Season Pasture Grass Species Grown at Five Irrigation Levels. Crop Sci. 2002, 42, 890–896. [Google Scholar] [CrossRef]
  140. Olsen, C.; Cain, A.; Gould, M.; Mattox, C.; Kowalewski, A. Optimizing Irrigation Rates and Frequency for Perennial Ryegrass in Western Oregon. Crop Forage Turfgrass Manag. 2019, 5, 180094. [Google Scholar] [CrossRef]
  141. Basford, K.E.; Cooper, M. Genotype×environment interactions and some considerations of their implications for wheat breeding in Australia. Aust. J. Agric. Res. 1998, 49, 153–174. [Google Scholar] [CrossRef]
  142. Smith, A.B.; Stringer, J.K.; Wei, X.; Cullis, B.R. Varietal selection for perennial crops where data relate to multiple harvests from a series of field trials. Euphytica 2007, 157, 253–266. [Google Scholar] [CrossRef]
  143. Smith, A.B.; Ganesalingam, A.; Kuchel, H.; Cullis, B.R. Factor analytic mixed models for the provision of grower information from national crop variety testing programs. Theor. Appl. Genet. 2015, 128, 55–72. [Google Scholar] [CrossRef] [PubMed]
  144. Yue, H.; Gauch, H.G.; Wei, J.; Xie, J.; Chen, S.; Peng, H.; Bu, J.; Jiang, X. Genotype by Environment Interaction Analysis for Grain Yield and Yield Components of Summer Maize Hybrids across the Huanghuaihai Region in China. Agriculture 2022, 12, 602. [Google Scholar] [CrossRef]
  145. Annicchiarico, P. Additive main effects and multiplicative interaction (AMMI) analysis of genotype-location interaction in variety trials repeated over years. Theor. Appl. Genet. 1997, 94, 1072–1077. [Google Scholar] [CrossRef]
  146. Jung, M.; Quesada-Traver, C.; Roth, M.; Aranzana, M.J.; Muranty, H.; Rymenants, M.; Guerra, W.; Holzknecht, E.; Pradas, N.; Lozano, L.; et al. Integrative multi-environmental genomic prediction in apple. Hortic. Res. 2024, 12, uhae319. [Google Scholar] [CrossRef]
  147. Mienye, I.D.; Swart, T.G.; Obaido, G. Recurrent Neural Networks: A Comprehensive Review of Architectures, Variants, and Applications. Information 2024, 15, 517. [Google Scholar] [CrossRef]
  148. Khaki, S.; Wang, L.; Archontoulis, S.V. A CNN-RNN Framework for Crop Yield Prediction. Front. Plant Sci. 2019, 10, 1750. [Google Scholar] [CrossRef]
  149. Robert, P.; Le Gouis, J.; BreedWheat Consortium; Rincent, R. Combining Crop Growth Modeling With Trait-Assisted Prediction Improved the Prediction of Genotype by Environment Interactions. Front. Plant Sci. 2020, 11, 827. [Google Scholar] [CrossRef]
  150. Boote, K.J.; Jones, J.W.; Hoogenboom, G. Incorporating realistic trait physiology into crop growth models to support genetic improvement. In Silico Plants 2021, 3, diab002. [Google Scholar] [CrossRef]
  151. Holzworth, D.P.; Huth, N.I.; deVoil, P.G.; Zurcher, E.J.; Herrmann, N.I.; McLean, G.; Chenu, K.; van Oosterom, E.J.; Snow, V.; Murphy, C.; et al. APSIM—Evolution towards a new generation of agricultural systems simulation. Environ. Model. Softw. 2014, 62, 327–350. [Google Scholar] [CrossRef]
  152. Holzworth, D.; Huth, N.I.; Fainges, J.; Brown, H.; Zurcher, E.; Cichota, R.; Verrall, S.; Herrmann, N.I.; Zheng, B.; Snow, V. APSIM Next Generation: Overcoming challenges in modernising a farming systems model. Environ. Model. Softw. 2018, 103, 43–51. [Google Scholar] [CrossRef]
  153. Chapman, S.C.; Cooper, M.; Hammer, G.L. Using crop simulation to generate genotype by environment interaction effects for sorghum in water-limited environments. Aust. J. Agric. Res. 2002, 53, 379–389. [Google Scholar] [CrossRef]
  154. Onogi, A. Integration of Crop Growth Models and Genomic PredictionGenomic predictions (GP). In Genomic Prediction of Complex Traits: Methods and Protocols; Ahmadi, N., Bartholomé, J., Eds.; Springer: New York, NY, USA, 2022; pp. 359–396. [Google Scholar] [CrossRef]
  155. Adnan, A.A.; Diels, J.; Jibrin, J.M.; Kamara, A.Y.; Shaibu, A.S.; Craufurd, P.; Menkir, A. CERES-Maize model for simulating genotype-by-environment interaction of maize and its stability in the dry and wet savannas of Nigeria. Field Crops Res. 2020, 253, 107826. [Google Scholar] [CrossRef]
  156. Li, T.; Angeles, O.; Marcaida, M.; Manalo, E.; Manalili, M.P.; Radanielson, A.; Mohanty, S. From ORYZA2000 to ORYZA (v3): An improved simulation model for rice in drought and nitrogen-deficient environments. Agric. For. Meteorol. 2017, 237–238, 246–256. [Google Scholar] [CrossRef] [PubMed]
  157. Stöckle, C.O.; Donatelli, M.; Nelson, R. CropSyst, a cropping systems simulation model. Eur. J. Agron. 2003, 18, 289–307. [Google Scholar] [CrossRef]
  158. Basso, B.; Liu, L.; Ritchie, J.T. A Comprehensive Review of the CERES-Wheat, -Maize and -Rice Models’ Performances. In Advances in Agronomy; Sparks, D.L., Ed.; Academic Press: Cambridge, MA, USA, 2016; Volume 136, pp. 27–132. [Google Scholar]
  159. Technow, F.; Messina, C.D.; Totir, L.R.; Cooper, M. Integrating Crop Growth Models with Whole Genome Prediction through Approximate Bayesian Computation. PLoS ONE 2015, 10, e0130855. [Google Scholar] [CrossRef]
  160. McGranahan, D.A.; Yurkonis, K.A. Variability in grass forage quality and quantity in response to elevated CO2 and water limitation. Grass Forage Sci. 2018, 73, 517–521. [Google Scholar] [CrossRef]
  161. Elgersma, A.; Smith, K.F. Editorial—Topics from the XXIV International Grassland Congress held in 2021. Grass Forage Sci. 2022, 77, 107–110. [Google Scholar] [CrossRef]
  162. Marinoni, L.d.R.; Zabala, J.M.; Taleisnik, E.L.; Schrauf, G.E.; Richard, G.A.; Tomas, P.A.; Giavedoni, J.A.; Pensiero, J.F. Wild halophytic species as forage sources: Key aspects for plant breeding. Grass Forage Sci. 2019, 74, 321–344. [Google Scholar] [CrossRef]
  163. Smith, K.F.; Elgersma, A. Editorial: Grass and Forage Science—75 years of impact and service to the science of grasslands. Grass Forage Sci. 2020, 75, 351–356. [Google Scholar] [CrossRef]
  164. Fei, Y.F.; Yang, A.L.; Li, W.J.; Yuan, X.Q.; Fenech, A. Forecasting Crop Yield Under Climate Change Using Crop Growth Models in China: A Review. In Proceedings of the 14th International Conference on Environmental Science and Development (ICESD 2023), Xiamen, China, 25–27 May 2024; Springer Nature: Cham, Switzerland, 2024. [Google Scholar]
  165. Paleari, L.; Tondelli, A.; Cattivelli, L.; Igartua, E.; Casas, A.M.; Visoni, A.; Schulman, A.H.; Rossini, L.; Waugh, R.; Russell, J.; et al. Extending genomic prediction to future climates through crop modelling. A case study on heading time in barley. Agric. For. Meteorol. 2025, 368, 110560. [Google Scholar] [CrossRef]
  166. Jighly, A.; Weeks, A.; Christy, B.; O’Leary, G.J.; Kant, S.; Aggarwal, R.; Hessel, D.; Forrest, K.L.; Technow, F.; Tibbits, J.F.G.; et al. Integrating biophysical crop growth models and whole genome prediction for their mutual benefit: A case study in wheat phenology. J. Exp. Bot. 2023, 74, 4415–4426. [Google Scholar] [CrossRef] [PubMed]
Figure 1. A comparative illustration of evaluation schemes for perennial ryegrass. Both evaluation schemes begin with a base population, followed by multiple crossing and evaluation stages, ultimately leading to the development of breeding lines (BLs); the cycles then repeat. The Phenotypic Estimation Scheme for Synthetic Population (left) follows a traditional phenotypic evaluation through spaced plant trials and clonal row trials (F2-SYN0). The Evaluation Scheme using Genomic Prediction (GP) (right) enables genomic estimated breeding values and accelerates the evaluation cycle through recurrent selection. Both strategies contribute to cultivar evaluation for commercial release.
Figure 1. A comparative illustration of evaluation schemes for perennial ryegrass. Both evaluation schemes begin with a base population, followed by multiple crossing and evaluation stages, ultimately leading to the development of breeding lines (BLs); the cycles then repeat. The Phenotypic Estimation Scheme for Synthetic Population (left) follows a traditional phenotypic evaluation through spaced plant trials and clonal row trials (F2-SYN0). The Evaluation Scheme using Genomic Prediction (GP) (right) enables genomic estimated breeding values and accelerates the evaluation cycle through recurrent selection. Both strategies contribute to cultivar evaluation for commercial release.
Agronomy 15 01494 g001
Figure 2. Common genomic prediction models. The models include Least Absolute Shrinkage and Selection Operator (LASSO), Elastic Net, ridge-regression Best Linear Unbiased Prediction (rrBLUP), Genomic Best Linear Unbiased Prediction using genomic relationship matrix kernel (GBLUP), genomic Random Regression Model (gRRM), Reproducing Kernel Hilbert Space (RKHS), BayesA, BayesB, BayesC, BayesD, BayesR, Bayesian Neural Networks (BNNs), Bayesian Kernel and Multikernel Models, Bayesian Kernel variants (Bayesian EX), Bayesian Gaussian Process, Random Forest (RF), Gradient Boosting Machines (GBM), and K-Nearest Neighbors (KNN).
Figure 2. Common genomic prediction models. The models include Least Absolute Shrinkage and Selection Operator (LASSO), Elastic Net, ridge-regression Best Linear Unbiased Prediction (rrBLUP), Genomic Best Linear Unbiased Prediction using genomic relationship matrix kernel (GBLUP), genomic Random Regression Model (gRRM), Reproducing Kernel Hilbert Space (RKHS), BayesA, BayesB, BayesC, BayesD, BayesR, Bayesian Neural Networks (BNNs), Bayesian Kernel and Multikernel Models, Bayesian Kernel variants (Bayesian EX), Bayesian Gaussian Process, Random Forest (RF), Gradient Boosting Machines (GBM), and K-Nearest Neighbors (KNN).
Agronomy 15 01494 g002
Figure 3. Graphical summary for translating genomic prediction research into practical applications in perennial ryegrass breeding. The diagram illustrates the three critical aspects required for successful GP implementation: future advancing research priorities (left), addressing key barriers to adoption (center), and meeting stakeholder needs (right). Success requires coordinated progress across all three areas, with research directions ensuring alignment with industry evaluation frameworks and the needs of farmers.
Figure 3. Graphical summary for translating genomic prediction research into practical applications in perennial ryegrass breeding. The diagram illustrates the three critical aspects required for successful GP implementation: future advancing research priorities (left), addressing key barriers to adoption (center), and meeting stakeholder needs (right). Success requires coordinated progress across all three areas, with research directions ensuring alignment with industry evaluation frameworks and the needs of farmers.
Agronomy 15 01494 g003
Table 1. Summary of genomic prediction models promising for perennial ryegrass performance estimation.
Table 1. Summary of genomic prediction models promising for perennial ryegrass performance estimation.
ModelStructuresAssumptionsFrameworksApplications
LASSO+ L1 penalty: y μ M α 2 + λ α 1 Linear additive marker effects;
unified variances across markers
Sparse Marker-Based;
Frequentist
High-dimensional additive genotyping data, feature selection [79]
Ridge regression+ L2 penalty: y μ M α 2 + λ α 2 2 ; this is mathematically equivalent to α N 0 , σ 2 Linear additive marker effects;
unified variances across markers
Dense Marker-Base;
Frequentist
Traits with additive effects per marker, also called rrBLUP [20,29,30,31,47,57,68]; when markers are SNPs, also called snpBLUP
Elastic Net+ L1 and L2 penaltiesLinear additive marker effects;
unified variances across markers
Sparse Marker-Base;
Frequentist
Correlated markers with additive effects per marker
GBLUPVariance components are in a genomic relationship matrix ( G ).
Coefficients are solved by Henderson mixed model equation [80].
Linear association between genetic markers and phenotypesKernel-Based;
Frequentist
Traits with linear additive effects; computational efficiency [15,26,28,29,30,32,38,79,81,82]. The basic GBLUP uses a genomic relationship matrix (GRM) [83].
When the GRM is expanded with Legendre polynomials to model environmental variances, it becomes a genomic random regression model (gRRM) [19].
RKHSAkin to GBLUP but uses non-linear kernel K .Non-linear association between markers and phenotypesKernel-Based;
Frequentist
Traits with non-linear effects and epistatic interactions [16,79]
BayesA/B/C P α k = 0 = π
P α k N 0 , σ k 2 = 1 π
Linear additive marker effects;
Variances can be marker-specific ( σ k χ 2 ν , s ) or unified
Marker-Based: Dense when π = 0 , Sparse when π 0 ;
Bayesian
Traits with flexible genetic architecture [14,15,68,79,84]
BayesR/D P α k = 0 = π 0
P α k N 0 , σ i 2 = π i ,
where i = 0 π i = 1
Linear additive marker effects with flexible shrinkage across multiple effect sizes;
marker-specific variances
Sparse Marker-Based;
Bayesian
Traits with mixed levels of effect sizes
Bayesian EXGeneral-purpose Bayesian framework: α P r i o r ( α ) , whereLinear additive marker effects;
prior choice determines shrinkage or sparsity level
Marker-Based;
Bayesian
Flexible prior choices for traits with additive effects per marker. The prior could be Gaussian (Bayesian Ridge) [79], Laplace (Bayesian LASSO) [68,79], or other distributions.
Bayesian Neural Networks (BNNs)Network learns f ( M ; W ) y , mapping from markers M to phenotypes y via weights W .Flexible marker effects;
Implicit variance components
Marker-Based;
Bayesian
Flexible genetic associations [79]
Bayesian Kernel ModelsFlexible kernel K is derived from the markers M to account for genetic variance components, f ( K ) then estimates genetic effects.Flexible genotype similarities;
predefined variance components
Kernel-Based;
Bayesian
Traits with non-linear genetic associations. When combining multiple kernels: f ( K ) = m f m ( K m ) , it becomes a Bayesian Multikernel Model.
Bayesian Gaussian Process f α G P μ α , K Flexible genotype similarities;
stochastic variance components
Non-parametric;
Kernel-Based;
Bayesian
Traits requiring uncertainty quantification in genetic relationships.
Random Forest (RF) y ^ t y , M is trained on bootstrapped samples based on certain splitting criteria.
RF parallelly aggregates tree predictions: y ^ RF = 1 T t = 1 T y ^ t .
Similarity-based prediction: patterns among similar data points provide higher prediction reliabilityNon-parametric;
Tree-Based Splitter
Non-linear additive prediction; noisy datasets with missing data; marker interactions [20,28,29,31,57,68]
Gradient Boosting Machines (GBM)GBM sequentially corrects tree predictions: y ^ G B M , t = y ^ G B M , t 1 + η h ^ t , where h ^ t is negative gradient of the loss function in each tree, η is the learning rate.Similarity-based predictionNon-parametric;
Tree-Based Splitter
Non-linear prediction [20]
K-Nearest Neighbours (KNN)Predicts based on k closest training samples in the marker feature space: y ^ K N N = 1 k i = 1 K y i where i indexes the K nearest neighbours.Similarity-based predictionNon-parametric;
Splitter
Non-linear prediction [20]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhu, J.; Smith, K.F.; Cogan, N.O.; Giri, K.; Jacobs, J.L. The Genome Era of Forage Selection: Current Status and Future Directions for Perennial Ryegrass Breeding and Evaluation. Agronomy 2025, 15, 1494. https://doi.org/10.3390/agronomy15061494

AMA Style

Zhu J, Smith KF, Cogan NO, Giri K, Jacobs JL. The Genome Era of Forage Selection: Current Status and Future Directions for Perennial Ryegrass Breeding and Evaluation. Agronomy. 2025; 15(6):1494. https://doi.org/10.3390/agronomy15061494

Chicago/Turabian Style

Zhu, Jiashuai, Kevin F. Smith, Noel O. Cogan, Khageswor Giri, and Joe L. Jacobs. 2025. "The Genome Era of Forage Selection: Current Status and Future Directions for Perennial Ryegrass Breeding and Evaluation" Agronomy 15, no. 6: 1494. https://doi.org/10.3390/agronomy15061494

APA Style

Zhu, J., Smith, K. F., Cogan, N. O., Giri, K., & Jacobs, J. L. (2025). The Genome Era of Forage Selection: Current Status and Future Directions for Perennial Ryegrass Breeding and Evaluation. Agronomy, 15(6), 1494. https://doi.org/10.3390/agronomy15061494

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