Next Article in Journal
Magnesium and the Brain: A Focus on Neuroinflammation and Neurodegeneration
Next Article in Special Issue
Accumulation of Anthocyanins in Detached Leaves of Kalanchoë blossfeldiana: Relevance to the Effect of Methyl Jasmonate on This Process
Previous Article in Journal
Dihydromyricetin Attenuates High-Intensity Exercise-Induced Intestinal Barrier Dysfunction Associated with the Modulation of the Phenotype of Intestinal Intraepithelial Lymphocytes
Previous Article in Special Issue
Study on ZmRPN10 Regulating Leaf Angle in Maize by RNA-Seq
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Multi-Year, Multi-Cultivar Approach to Differential Expression Analysis of High- and Low-Protein Soybean (Glycine max)

1
Agriculture and Agri-Food Canada, 960 Carling Ave, Ottawa, ON K1A 0C6, Canada
2
Department of Biology, Ottawa Institute of Systems Biology, Carleton University, 1125 Colonel By Dr., Ottawa, ON K1S 5B6, Canada
3
Canadian Centre for Computational Genomics, 740 Dr. Penfield Ave, Montréal, QC H3A 0G1, Canada
4
Agriculture and Agri-Food Canada, 2701 Grand Valley Road, Brandon, MB R7A 5Y3, Canada
5
Crop Development Centre, University of Saskatchewan, Saskatoon, SK S7N 5A8, Canada
6
Agriculture and Agri-Food Canada, Morden, MB R6M 1Y5, Canada
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(1), 222; https://doi.org/10.3390/ijms24010222
Submission received: 5 November 2022 / Revised: 12 December 2022 / Accepted: 15 December 2022 / Published: 23 December 2022
(This article belongs to the Collection Feature Papers in Molecular Plant Sciences)

Abstract

:
Soybean (Glycine max (L.) Merr.) is among the most valuable crops based on its nutritious seed protein and oil. Protein quality, evaluated as the ratio of glycinin (11S) to β-conglycinin (7S), can play a role in food and feed quality. To help uncover the underlying differences between high and low protein soybean varieties, we performed differential expression analysis on high and low total protein soybean varieties and high and low 11S soybean varieties grown in four locations across Eastern and Western Canada over three years (2018–2020). Simultaneously, ten individual differential expression datasets for high vs. low total protein soybeans and ten individual differential expression datasets for high vs. low 11S soybeans were assessed, for a total of 20 datasets. The top 15 most upregulated and the 15 most downregulated genes were extracted from each differential expression dataset and cross-examination was conducted to create shortlists of the most consistently differentially expressed genes. Shortlisted genes were assessed for gene ontology to gain a global appreciation of the commonly differentially expressed genes. Genes with roles in the lipid metabolic pathway and carbohydrate metabolic pathway were differentially expressed in high total protein and high 11S soybeans in comparison to their low total protein and low 11S counterparts. Expression differences were consistent between East and West locations with the exception of one, Glyma.03G054100. These data are important for uncovering the genes and biological pathways responsible for the difference in seed protein between high and low total protein or 11S cultivars.

1. Introduction

Soybean (Glycine max (L.) Merr.) is among the world’s most important crops due to the widespread uses for its seed protein and oil. By 2050 the global population is expected to approach 10 billion people, requiring the implementation of more efficient and imperishable food production and storage strategies [1]. Soybean has the capacity to fix inert atmospheric nitrogen into more useful biologically available forms within root nodules via symbiosis with rhizobia in the soil [2]. This unique relationship incentivizes the incorporation of soybean into crop rotation practices because it reduces the need to add nitrogen fertilizers, which are costly and have a role in greenhouse gas emissions of nitrous oxide [3]. Soybean plays an important role in improving soil quality and thereby serves as a pillar in strategic planning of sustainable agriculture.
Seed protein and oil content are complex quantitatively inherited traits that are influenced by genotype, environment and the interaction between the two [4,5]. A strong direct and/or indirect phenotypic correlation is evident in the inverse balance of these two storage components [6]. When considering soybean seed storage, an important balance exists between proteins, lipids and carbohydrates and it is important to consider expression of genes related to each of these macromolecules. Carbohydrate metabolism is the precursor to protein and oil biosynthesis, making it a key decision-making step in downstream seed storage molecule biosynthesis. Sugar transporters have been identified as playing a role in protein content in soybeans [7,8]. Seed protein content accounts for 30–46% (average 40%) and oil accounts for 12–24% (average 20%) of the seed weight [6]. There is a limit to the current understanding of oil and storage protein regulation and accumulation in soybean seeds, as well as the determining factors for the allocation of carbon to the production of proteins or oil [9,10].
Soybean seeds, like many plant seeds, accumulate considerable amounts of storage proteins as a reserve of carbon, nitrogen and sulfur for the germinating/developing of seedlings [5,11]. Seed storage proteins are categorized by their sedimentation coefficient at 0.5 M ionic strength: 2S, 7S, 11S and 15S fractions [12]. The most abundant are the 11S (glycinins) fraction and the 7S (β-conglycinins) fraction, which account for approximately 40% and 30%, respectively [13]. Together the 11S and 7S globulins provide a total of 18 amino acids, which include the 10 essential amino acids, but the amount of sulfur-containing amino acids, cysteine (Cys) and methionine (Met), are not sufficient for the nutrient requirements of monogastric animals [14]. Excess amino acids are not stored by the body like fats and carbohydrates, so they must be obtained from daily food intake.
Glycinins normally form a 360 kDa hexamer made up of both acidic and basic subunits; β-conglycinins typically exist as a 180 kDa trimer of α, α′ and β subunits [4,15]. Glycinins have approximately 3–4 times the amount of sulfur-containing amino acids (Met, Cys) in comparison to β-conglycinins [4]. Evidently, the β subunit of β-conglycinin has poor nutritional value because of the lack of sulfur-containing amino acids. Through the promotion of accumulation of glycinins and simultaneous suppression of β-conglycinin accumulation, it is entirely possible to make valuable improvements to the nutritional quality of soybean by increasing the sulfur-containing amino acid content [4]. The seed protein composition, particularly that of the 11S and 7S globulins, is important for the germinating seed as well as the nutritional and functional quality of soy foods [4]. Subunit composition of both the glycinins and β-conglycinins, as well as the ratio of the two (i.e., 11S:7S ratio), affect the quality of the resultant food product [4]. Studies have shown that the 11S and 7S proteins present in soybean seed affect food sensory attributes including texture, taste, heat stability, tofu gelling and water retention [16]. With this information, the development of high-quality protein composition in seeds must be acknowledged when considering all aspects of improving soybean seed quality.
Currently, there are 15 known genes which encode β-conglycinins (CG1CG15) and five known genes encoding glycinins (Gy1Gy5) [17,18]. Gene mining for sulfur-containing amino acid metabolites in soybean predicted 12 candidate genes in addition to quantitative trait loci (QTLs) for Cys and Met content in 11S and 7S fractions on chromosomes 1, 3, 4, 6, 10, 13, 16, 17, 19 and 20 [19]. In another study, F6-derived recombinant inbred lines (RILs) were used in linkage mapping to identify three QTLs associated with glycinin on chromosomes 17, 19 and 20 and two QTLs associated with β-conglycinin on chromosomes 17 and 16 [15]. One QTL was associated with the α′ subunit of β-conglycinin on chromosome 10 and five QTLs associated with 11S subunits on chromosomes 3, 10, 13 and 19 [20].
Soybean production in Canada has observed a consistently lower (by ~1–5%) seed protein content in Western growing regions in comparison to Eastern growing regions [21]. In 2021, the average Eastern soybean protein content on a dry basis was 40.3%, while the Western-grown soybeans had an average seed protein content of 36.0% [21]. Compounded with the other challenges of growing soybean in western Canada (i.e., less precipitation, cooler temperatures and shorter growing season), the issue with lower-protein soybean contributes to a reduced interest in soybean production and subsequently the loss of a key nitrogen fixing option from Canadian crop rotation practices. Seed protein and oil composition are important factors in soybean seed quality; a deeper foundational understanding of the genes differentially expressed (DE) between high and low protein soybeans across Canada is valuable for tailored, high-quality soybean breeding.
The current understanding of seed storage protein regulation and accumulation in soybean is limited. Here we used RNA-seq and differential expression (DE) analysis of high and low total protein (TP) soybeans and high and low 11S soybeans, grown in multiple regions of Canada over three years, in order to uncover commonly differentially expressed genes which may influence seed protein and oil accumulation.

2. Results

2.1. Seed Content, Composition and Selection of Lines for DE

Soybean seed composition data from each line in each location was collected each year from 2018–2020. The average seed protein content for each line (1–10) was taken across all locations over 3 years in order to narrow down lines for DE analysis. Figure 1A shows the average seed TP and seed oil content for each line over 2018–2020. Supplementary Table S1 includes the 3-year averages for TP, oil, 11S, 7S and 11S:7S for all 10 lines and the standard deviations from the means. Lines 8, 9 and 10 had the lowest average oil content and lines 2, 1, 6 and 3 had the first, second, third and fifth highest oil content, respectively (Figure 1A, Table S1). Lines 1, 2 and 3 had the lowest protein content and lines 8, 9 and 10 had the highest protein content (Figure 1A, Table S1). Based on these results lines 8, 9 and 10 were designated “high TP” and lines 1, 2 and 3 were designated “low TP” and were selected for DE analysis. Table 1 shows the seed protein and seed oil content for high TP (i.e., low oil) soybeans and low TP (i.e., high oil) soybean samples used for DE analysis (difference in TP content between high and low TP soybeans p-value 0.01). The low TP samples served as the control and the high TP are the experimental data; therefore, DE shows up/downregulation in the high TP soybeans.
Seed protein composition analysis was then considered for average 11S and average 7S content, as a percentage of total protein, for all lines. Figure 1B shows the proportion of 11S and 7S of the total seed protein content (primary y axis) and the 11S:7S ratio value (secondary y axis) for each line across all 3 years (2018–2020). Lines 1, 5 and 9 have the highest 7S content and lines 2, 4, 3 and 8 have the 4 lowest 7S content values, respectively (Figure 1B, Table S1). The three lines with highest 11S content are lines 2, 4, 8 and three lines with lowest 11S content are lines 1, 5, 9 (Figure 1B, Table S1). Incidentally, the lines with the lowest 11S:7S ratio were the same lines that had the lowest 11S content (lines 1, 5, 9). Lines 2, 4, 3 and 8 had the top four highest 11S:7S ratios, respectively (Figure 1B, Table S1). Based on these results, lines 1, 5 and 9 were grouped as “low 11S” and lines 2, 4 and 8 were designated the “high 11S” group selected for DE analysis. The average 11S content, 7S content and ratio of 11S:7S for the high and low 11S soybean lines are listed in Table 1 (difference in 11S content between high and low 11S soybeans p-value 0.005). Low 11S soybeans were used as the control and high 11S soybeans were used as the experimental sample; therefore, when discussing DE, the log2FC up- and downregulation describe the DE in the high 11S samples.

2.2. RNA-seq and DE Output Quality

In total, 234 samples (10 year-locations, 8 lines) were used to create 20 DE datasets: 10 high TP vs. low TP and 10 high 11S vs. low 11S. Across all samples, Q30 scores of at least 36 were accepted. In total, across three replicates from lines 1, 2, 3, 4, 5, 8, 9, 10 in each location from 2018–2020 there were 5,267,207,213 reads. The TP analysis included a total of 3,932,862,305 reads across all years and locations combined. Analysis of high TP data came from 1,968,419,408 reads and low TP data came from 1,964,442,897 reads. The 11S analysis included 3,478,643,109 reads across all years and locations. The high 11S samples (lines 2, 4, 8) had a total of 2,030,526,766 reads across all years and locations; the low 11S (lines 1, 5, 9) had a total of 2,002,088,417 reads. The average paired read survival rate was 98.0%, with the lowest survival rate of 90.9%. Figure 2 shows the principal component (PC) analysis of the normalized RNA-seq read data for each high TP vs. low TP DE analysis. Figure 3 shows the PC analysis of the normalized RNA-seq read data for each high 11S vs. low 11S DE analysis. Because both Figure 2 and Figure 3 are built on normalized expression data, phenotypic TP and 11S data are not necessarily always the first PC. For both Figure 2 and Figure 3, orange data points represent high TP or 11S and blue data points represent low TP or 11S, respectively. The line numbers corresponding to each individual datapoint are presented next to each point.

2.3. High vs. Low Seed Protein Content DE

From each of the high vs. low (Lines 8, 9, 10 vs. Lines 1, 2, 3) TP DE analyses, the top 15 most upregulated and the 15 most downregulated genes and their respective log2foldchanges (log2FCs) in expression were extracted for each environment (300 genes total). Genes were sorted based on frequency of occurrence; those that were present in at least 5/10 DE location-year datasets were short-listed. Table 2 shows the shortlist of genes for the cumulative high vs. low TP analyses across all 10 DE datasets. Included in this table is the log2FC difference in expression in high TP soybeans for each candidate gene in each dataset and the total number of datasets out of 10 from which each gene was found to be in the top 15. Eleven shortlisted genes were upregulated in at least 5 of 10 high vs. low TP DE datasets; seven shortlisted genes were downregulated in at least 5 of 10 high vs. low TP DE datasets. Figure 4 shows the individual relative expression heatmaps for the shortlisted genes as they appear in each location-year DE dataset.
Three genes, Glyma.03G057800, Glyma.10G092400 and Glyma.16G081500, were each identified to be upregulated in 8/10 high vs. low TP DE datasets (Table 2). The log2FC upregulation of Glyma.03G057800 across the datasets ranged from 5.14 (Morden 2020) to 27.5 (Ottawa 2019), with an average log2FC of 16.8. In Figure 4, gene 1 in the heatmaps corresponds to Glyma.03G057800. Morden 2018 and Brandon 2019 are the two datasets which did not identify Glyma.03G057800 within the most upregulated (nor downregulated) genes; in Morden 2018 Glyma.03G057800 was the 93rd most upregulated gene (log2FC 0.99; Figure 4D) and in Brandon 2019 Glyma.03G057800 was the 17th most upregulated gene (log2FC 6.75; Figure 4H). The log2FC upregulation of Glyma.10G092400 ranged from 5.08 (Morden 2018) to 25.5 (Ottawa 2020), averaging 13.9 (Table 2). In Figure 4, gene 2 in the heatmaps corresponds to Glyma.10G092400. Glyma.10G092400 is within the top 15 upregulated genes of all datasets with the exception of Ottawa 2018 and Saskatoon 2020; in Ottawa 2018, it was the 32nd most upregulated gene (log2FC 3.46; Figure 4A) in high TP soybeans, while in Saskatoon 2020 it was not identified to be DE (Figure 4J, non-DE shortlisted genes are shown in grey). Glyma.16G081500 was among the topmost prevalent upregulated genes in the high vs. low TP DE analyses. In Figure 4, gene 3 in the heatmaps corresponds to Glyma.16G081500. This gene had a range of log2FC DE from 23.80 (Ottawa 2020; Figure 4C) to 31.83 (Morden 2020; Figure 4F) and an average log2FC of 29.40 in high TP soybeans. Glyma.16G081500 was identified in our study to be among the top 15 genes in 8/10 datasets, only excluding Morden 2018 and Brandon 2018, in neither of which Glyma.16G081500 was identified to be DE (Figure 4D,G).
An additional four genes from the high vs. low TP DE analyses were found to be upregulated in 7 of the 10 DE datasets; Glyma.01G179100, Glyma.02G060600, Glyma.10G092300 and Glyma.19G140200 (Table 2). In Figure 4, gene 4 in the heatmaps corresponds to Glyma.01G179100. The log2FC upregulation of Glyma.01G179100 had a wide range from 2.54 (Saskatoon 2020; Figure 4J) to 24.7 (Brandon 2019; Figure 4H), with an average log2FC 11.6 (Table 2). In Morden 2018 Glyma.01G179100 was the 933rd most upregulated gene, with a log2FC increase of 0.35 (Figure 4D). In Morden 2019 Glyma.01G179100 was the 19th most upregulated gene, with a log2FC increase by 4.16 (Figure 4E). In Ottawa 2019 Glyma.01G179100 was the 18th most upregulated gene, with a log2FC increase of 7.35 (Figure 4B). In Figure 4, gene 5 in the heatmaps corresponds to Glyma.02G060600. The log2FC DE of Glyma.02G060600 in high TP soybeans ranges from 26.7 (Morden 2018; Figure 4D) to 35.9 (Ottawa 2019; Figure 4B) and the average DE was by log2FC 32.9 (Table 2). Glyma.02G060600 was not DE between low and high TP soybeans in Morden 2020, Brandon 2018 and Saskatoon 2019 (Figure 4G,F,I). In Figure 4, gene 6 in the heatmaps corresponds to Glyma.10G092300. Glyma.10G092300 was within the top 15 upregulated genes in 7/10 high TP DE analyses, with a minimum log2FC increase by 4.84 (Morden 2018; Figure 4D), a maximum log2FC by 27.1 (Ottawa 2020; Figure 4C) and an average log2FC of 20.3 (Table 2). Glyma.10G092300 was the 22nd most upregulated gene in high TP soybeans in Brandon 2018 (Figure 4G); it was the 17th most upregulated gene in high TP soybeans in Morden 2019 (Figure 4E); and was not detected to be DE between high and low TP soybeans in Saskatoon 2020 (Figure 4J). In Figure 4, gene 7 in the heatmaps corresponds to Glyma.19G140200. Glyma.19G140200 was upregulated in 7 of the 10 datasets by a minimum log2FC 18.4 (Brandon 2018; Figure 4G), a maximum log2FC 22.8 (Ottawa 2019; Figure 4B) (Table 2), with an average increase by log2FC 21.2. Glyma.19G140200 was not DE in high TP soybeans in Ottawa 2018, Ottawa 2020 and Morden 2018 (Figure 4A,C,D).
Glyma.13G077600 was identified to be within the top 15 downregulated genes in 7/10 high vs. low TP DE datasets (Table 2). In Figure 4, gene 12 in the heatmaps corresponds to Glyma.13G077600. The log2FC DE of Glyma.13G077600 ranged from −13.4 (Brandon 2018) to –35.4 (Ottawa 2019) and an average log2FC of −31.2 across the seven datasets (Table 2). Glyma.13G077600 was not identified to be DE in high TP soybeans grown in Morden 2019, Brandon 2019, nor Saskatoon 2020 (Figure 4E,H,J). Gene 13 in the heatmaps in Figure 4 corresponds to Glyma.15G246500. Glyma.15G246500 was downregulated in 6/10 high vs. low TP DE analyses. The minimum log2FC difference in expression of Glyma.15G246500 ranged from −19.2 (Morden 2018) to −26.3 (Ottawa 2019), averaging at −23.2 (Table 2). Glyma.15G246500 was not identified to be DE in high TP soybeans in Brandon 2018, Brandon 2019, Saskatoon 2019 and Saskatoon 2020 (Figure 4G–J). Gene 16 in the heatmaps in Figure 4 corresponds to Glyma.03G054100. Glyma.03G054100 is downregulated in 5/10 high vs. low TP datasets; downregulation ranges from log2FC –9.52 (Brandon 2019; Figure 4H) to –25.9 (Saskatoon 2019; Figure 4I) and an average of −21.6 (Table 2). Glyma.03G054100 was the 67th most downregulated gene in Morden 2018, by a log2FC −1.08 (Figure 4D); it was not DE in high TP soybeans in Morden 2019, Morden 2020, Brandon 2018 and Saskatoon 2020 (Figure 4E–G,J).

2.4. High vs. Low 11S Content DE

From each of the 10 year-location high vs. low 11S (lines 2, 4, 8, vs. 1, 5, 9) DE analyses, the top 15 most upregulated and the 15 most downregulated genes and their respective log2FCs were extracted (300 genes total). Genes were sorted based on frequency of occurrence and those that were present in at least 5/10 DE datasets were short-listed. Three genes, Glyma.01G016700, Glyma.06G306900, Glyma.19G231100 and Glyma.16G086800, were DE in 5 or 6 of 10 datasets; however, in some instances these genes were upregulated and in some instances they were downregulated. This inconsistency rules them out as a genes of interest in this study. Table 3 shows the shortlist of the most frequently occurring genes across the 10 11S DE datasets. There are eight genes upregulated in at least 5/10 DE datasets; Glyma.19G084500, Glyma.02G077300, Glyma.17G209900, Glyma.01G091300, Glyma.06G287800, Glyma.10G141200, Glyma.14G204900 and Glyma.18G112500 (Table 3). There are six genes downregulated in at least 5/10 datasets; Glyma.13G077600, Glyma.17G261800, Glyma.01G127800, Glyma.03G054100, Glyma.12G156500 and Glyma.18G082700. Figure 5 shows the relative expression heatmaps of each of the shortlisted genes as they appear in each year-location 11S DE analysis.
Glyma.19G084500 was within the top 15 most upregulated genes across 9/10 DE datasets. Gene 1 in the heatmaps in Figure 5 corresponds to Glyma.19G084500. The log2FC DE in high 11S soybeans of Glyma.19G084500 ranged from 21.5 (Saskatoon 2019; Figure 5I) to 27.5 (Morden 2020; Figure 5F), with an average upregulation by a log2FC of 24.8 (Table 3). Brandon 2018 is the only dataset in which Glyma.19G084500 did not fall within the top 15 upregulated (nor downregulated) genes; further, it was not identified to be DE at a p-value < 0.05 in Brandon 2018 (Figure 5G).
Glyma.02G077300 was identified within the top 15 most upregulated genes in 8/10 high vs. low 11S DE analyses (Table 3). In Figure 5, gene 2 corresponds with Glyma.02G077300. Upregulation in high 11S soybeans ranged from 16.1 (Morden 2019; Figure 5E) to 22.2 (Ottawa 2020; Figure 5C), with an average log2FC of 20.3. Morden 2018 and Ottawa 2018 were the only high vs. low 11S datasets that did not identify Glyma.02G077300 within the 15 most up- or downregulated genes (Table 3). In Morden 2018 Glyma.02G077300 was upregulated by a log2FC of 3.28, the 32nd most upregulated gene in the dataset (Figure 5D). Glyma.02G077300 was not identified to be DE (p-value < 0.05) between high and low 11S soybeans in Ottawa in 2018 (Figure 5A).
Glyma.13G077600 was found to be among the most persistently downregulated genes in high 11S soybeans compared to low 11S soybeans (Table 3). In Figure 5, gene 9 corresponds with Glyma.13G077600. Glyma.13G077600 was downregulated in 7/10 of the DE datasets; log2FC DE ranged from −29.6 (Saskatoon 2020; Figure 5J) to −33.5 (Ottawa 2019; Figure 5B) and the average downregulation of Glyma.13G077600 was by a log2FC of −32.0. In Brandon 2018, Glyma.13G077600 was downregulated by a log2FC of −9.95, but not within the top 15 most downregulated genes (Figure 5G). Glyma.13G077600 was not identified to be DE between high and low 11S soybeans from Ottawa 2020 and Morden 2019 (Figure 5C,E).
Glyma.17G261800 was also downregulated in 7/10 DE datasets between high vs. low TP soybeans (Table 3); log2FC DE ranged from −15.0 (Brandon 2018; Figure 5G) to −46.0 (Morden 2019; (Figure 5E) and the average was −38.4. In Figure 5, gene 10 corresponds with Glyma.17G261800. Glyma.17G261800 was not identified within the top 15 most downregulated (nor upregulated) genes in Ottawa 2019, Morden 2020 and Brandon 2019 (Figure 5B,F,H). Glyma.17G261800 is the 17th most downregulated gene (log2FC −4.33) in Brandon 2019 high TP soybeans (Figure 5H). In Ottawa 2019 Glyma.17G261800 was the 18th most downregulated gene (log2FC −5.56; Figure 5B). In Morden 2020 Glyma.17G261800 was downregulated by a log2FC of −1.58, the 67th most downregulated gene in the Morden 2020 dataset (Figure 5F).
The annotated list of up- and downregulated shortlist genes from both differentially expressed in high TP soybeans and high 11S soybeans are listed in order of chromosome position in Supplementary Table S2. Table S2 also includes the potential role of each gene in relation to seed content accumulation, the orientation of DE and the data from which each gene was found to be significant.

2.5. Gene Ontology

Supplementary Table S3 lists all the biological process (BP) and molecular function (MF) gene ontologies (GOs) for all shortlisted genes. In order to gain a broad view of the relationships of the BP GOs associated with all shortlisted genes, the upregulated shortlist and downregulated shortlist were run through the SoyBase GO Term Enrichment Tool [22] (soybase.org (accessed on 17 May 2022)). The GO terms were submitted to Revigo [23,24] (http://revigo.irb.hr/ (accessed on 17 May 2022)), using A. thaliana as a reference species to assess relative similarity between GO terms. Figure 6 shows the relationship between indispensable (dispensability score of 0 on a scale of 0 to 1) BP terms associated with the high TP upregulated (Figure 6A) and downregulated (Figure 6B) short list of genes. Indispensable terms from the upregulated genes in high TP soybeans are oligopeptide transport (GO:0006857), cell population proliferation (GO:0008283), response to xenobiotic stimulus (GO:0009410), regulation of G2/M transition of mitotic cell cycle (GO:0010389) and floral organ formation (GO:0048449). Indispensable terms from the downregulated genes in high TP soybeans are regulation of transcription–DNA-templated (GO:0006355), regulation of gene expression (GO:0010468), phosphatidylinositol biosynthetic process (GO:0006661), embryo development ending in seed dormancy (GO:0009793) and defense response to bacterium (GO:0042742). Figure 7 shows the relationship between BP GO terms associated with the high 11S upregulated (Figure 7A) and downregulated (Figure 7B) shortlisted genes. Supplementary Table S4 provides the Revigo outputs including each GO term and the corresponding frequency, uniqueness and dispensability and PC analysis. Importantly, the high TP soybeans upregulating lipid metabolic process (GO:0006629) and carbohydrate metabolic process (GO:0005975) had the two highest logSize values on the Revigo plot (Figure 7A) of 3.04 and 3.05, respectively and the highest frequency values of 5.08 and 5.19, respectively (Table S4). The dispensability score for lipid metabolic process was 0.102 and the dispensability score for carbohydrate metabolic process is 0.189, indicating removal of any of the multiple terms that are daughter terms under the lipid metabolic process (GO:0006629) and carbohydrate metabolic process (GO:0005975) umbrellas would largely impact the overall relationship structure (and PCs). The indispensable BP GO terms from the genes upregulated in high 11S soybeans are response to wounding (GO:0009611), photorespiration (GO:0009853) and leaf senescence (GO:0010150) (Figure 7A, Table S4). Other important terms upregulated in high 11S soybeans include glucose metabolic process (GO:0006006), starch metabolic process (GO:0005982), maltose catabolic process (GO:0000023) and oxylipin metabolic process (GO:0031407) (Figure 7A, Table S4). The indispensable BP GO terms from the downregulated genes in high 11S soybeans are gibberellic acid mediated signaling pathway (GO:0009740) and seed dormancy process (GO:0010162) (Figure 7B, Table S4).

2.6. East vs. West Analysis

Division of the shortlist data on the basis of location shows some discrepancies in expression of some key genes between soybeans grown in East (Ottawa) versus West (Morden, Brandon, Saskatoon) Canada. The average DE for each shortlisted gene was taken from Ottawa (2018–2020) and the three West locations combined (2018–2020) for both high vs. low TP and high vs. low 11S soybeans samples. Supplementary Table S5 shows the average DE for each short-listed gene in the East and West for both high vs. low TP soybeans and high vs. low 11S soybeans.
Between the high vs. low TP soybeans grown in East and West, the eastern samples showed much higher expression of Glyma.03G057800; eastern-grown soybeans had an average DE of log2FC 25.2 compared to log2FC 11.9 in the West (standard deviation of 9.31). Glyma.10G092400 was also found to be much more upregulated in eastern-grown soybeans than the western-grown soybeans; the average upregulation in the East was log2FC 24.2, whereas the average upregulation in the West was log2FC 10.5 (standard deviation of 9.7).
Among the high vs. low 11S soybean samples, the average DE of each shortlisted gene is fairly similar between East and West, with the exception of Glyma.03G054100. Glyma.03G054100 is downregulated in high 11S soybeans in the East by an average of log2FC −7.29, in the West this gene is downregulated by log2FC −38.9 (standard deviation of 22.4). This gene is also more downregulated in high TP soybeans in the West (log2FC –18.6 in the East; log2FC −25.2 in the West; standard deviation of 4.71).

3. Discussion

Glyma.16G081500 is among the topmost upregulated genes in the high vs. low TP DE analyses (Table 2). Glyma.16G081500 is uncharacterized in G. max, but the top BLASTP hit identified a Subtilisin-like serine protease in Medicago truncatula (Table S2). Subtilase (SBT) family proteins have been identified in plants to play a broad range of biological functions involved in many different aspects of plant life, starting with seed and fruit development, cell wall modification, response to abiotic and biotic stressors, protein turnover, peptide growth factors, epidermal development and programmed cell death [25]. These functions are all favorable during seed development; allowing seed cell growth via cell wall modifications and cell-to-cell communication for choreographed development. Upregulation of Glyma.16G081500 suggests that high TP soybeans may be better adapted for seed development than low TP soybeans, in part due to increased SBT function during seed development.
Glyma.16G060600 is in the shortlist of upregulated genes in high TP soybeans (Table 2). Glyma.16G060600 is uncharacterized in G. max but has the MF GO phospholipase activator activity (GO: 0016004) (Tables S2 and S3). The top BLASTP hit identified the ADP-ribosylation factor (Arf) in M. truncatula as the most closely related protein (Table S2). The Arf family is composed of small GTP-binding proteins that play roles in intracellular trafficking and cargo sorting in yeast, animal cells and plant cells [26]. During seed development, precursors for seed storage proteins (in particular 11S and 7S proteins) are first synthesized in the ER. Arf proteins subsequently transport these protein subunits to protein precursor-accumulating vesicles (PSVs) by the Golgi-dependent trafficking pathway in which they are converted to mature subunits and accumulated in the developing seed. Arf1 is thought to also be involved in cargo protein sorting [26]. Arf1 is specifically shown to function in retrograde trafficking from the Golgi apparatus to the ER, as well as from the trans-Golgi network to the endosome [27]. In pea plants, storage proteins such as vicilin (7S) were identified to be sorted to the storage vacuole at the cis-Golgi [28]. In pumpkin seeds, GTP-binding proteins have been shown to target and/or fuse with the precursor-accumulating vesicles which accumulate and transport insoluble components of storage proteins to vacuoles directly from the ER [29]. The upregulation of Glyma.16G060600 in high TP soybeans suggests increased Arf function in comparison to low TP soybeans, likely indicating increased transportation and sorting of storage protein precursors to PSVs for biosynthesis of storage proteins. This increased expression of Arf-like trafficking proteins for protein storage vacuoles in high TP soybeans is a plausible means for an underlying factor for the difference in TP between experimental soybean samples.
Glyma.16G082200 is upregulated across high TP soybeans (Table 2). The top TAIR10 identity for Glyma.16G082200 is the NAD(P)-binding Rossmann-fold superfamily proteins, also known as enoyl-acyl carrier protein (ACP)-dependent reductases (NCBI gene ID: 815152) (Table S2). Enoyl-ACP reductase proteins are an enzymatic component of the mitochondrial type II fatty acid synthase pathway and plastidic type II fatty acid synthase activities, including acyl precursor synthesis for lipoic acid biosynthesis [30]. Upregulation of genes for Enoyl-ACP reductases suggests that these high TP soybeans are increasing biosynthesis of lipoic acid, essential for oxidation of carbohydrates among other important cell functions. Arabidopsis enoyl-ACP reductase knockdown mutants showed symptoms consistent with plants with deficient mitochondrial fatty acid synthase activities: depleted lipoliation of photorespiratory glycine cleavage system H-protein, glycine hyperaccumulation and reduced growth with very small ariel organs [30]. Glyma.16G082200 was upregulated across high TP soybeans compared to low TP soybeans (Table 2). Reduced activity of enoyl-ACP reductases in low TP soybeans may reflect the symptoms observed in the Arabidopsis mutants reported by [30]. The phenotype of reduced aerial organs results in less photosynthetic tissue and therefore a reduced capacity for energy production and successful maturation. This gene may pinpoint a specific difference in carbohydrate metabolic pathways between high TP and low TP soybeans.
Glyma.10G092300 was upregulated among high TP soybeans (Table 2). The top BLASTP hit for Glyma.10G092300 is the Peptide transporter (PTR) 3-A from M. truncatula (Table S2). PTRs are part of a larger umbrella family of nitrate transporter 1/peptide transporters (NTR/PTR; now known as the NPF family); a family characterized by their role in nitrogen uptake and transportation of nitrates and peptides [31]. Studies in rice have confirmed positive relationships between NPF expression and enhanced nitrogen allocation, nitrogen use efficiency, grain yield, branching, influx/concentration of nitrate and ammonium in the roots and potential kinase involvement [32]. Increased expression of Glyma.10G092300 in high TP soybeans suggests enhanced allocation and use of nitrogen in these plants and subsequently likely attributes of the difference in TP.
Just downstream from Glyma.10G092300, Glyma.10G092400 was also found to be upregulated across high TP soybeans (Table 2). With no lysine (WNK) kinase 3 is the top Arabidopsis homolog for Glyma.10G092400, which is otherwise uncharacterized in soybean (Table S2). GmWNK1, a soybean WNK, is suggested to play a role in ABA signaling and ABA-mediated homeostatic response to osmotic changes in roots which mediates root architecture [33]. Both Glyma.10G092300 and Glyma.10G092400 are upregulated in all the same year-locations: Brandon-2018, Morden-2018, Brandon-2019, Ottawa-2019, Saskatoon-2019, Morden-2020 and Ottawa-2020 (Table 2). The only difference was found in Morden-2019 where Glyma.10G092400 which was upregulated but not Glyma.10G092300 (Table 2). These similar expression patterns may indicate a potential relationship between the roles of Glyma.10G092300 (an NPF-like protein, with potential for kinase involvement) and Glyma.10G092400 (a kinase). A relationship between the functional activities of Glyma.10G092300 and Glyma.10G092400 could be indicative of the crosstalk between ABA signaling and nitrogen availability.
Glyma.09G184300, upregulated in high TP soybeans, is identified as a bZIP transcription factor in G. max and TAIR10 identified a TGACG motif-binding factor 6 as the top Arabidopsis homolog (Table 2 and Table S2). bZip transcription factors influence plant development, drought stress response, defense response and seed development. In Phaseolus bZip factors Regulator of MAT1 (ROM1) moderates lectin and storage protein gene transcription and expression is developmentally regulated (abundant in early embryogenesis and decreasing as maturation progresses) [34]. The increased expression of Glyma.09G184300 in high TP soybeans, particularly during the R5 stage at which sampling took place, may make it a key gene that needs temporal regulation for improved seed storage protein accumulation. Temporal DE analysis of this transcription factor might be of interest to gain further insight to its role in seed development and protein content, including the gene(s) it acts upon.
Glyma.03G057800, a gene found to be highly upregulated in high TP soybeans, is predicted to be a rhodanese-like domain-containing protein in G. max. The top BLASTP hit identified as a Rhodanese-like family protein-like protein in M. truncatula and the TAIR10 hit identifies a Rhodanese/Cell cycle control phosphatase superfamily protein in A. thaliana (Table S2). Proteins with Rhodanese (thiosulfate:cyanide sulfur-transferase) domains are versatile sulfur carriers capable of fulfilling reactive roles in metabolic and regulatory pathways, including senescence [35]. This may suggest that cell cycle control is differently regulated in high TP soybeans compared to low TP soybeans. High TP proteins may be expressing these genes for tightly controlled growth and development and/or low TP soybeans may be deficient in cell cycle control, leading to poorer growth and development under the same environmental conditions.
Glyma.06G205700 was found to be downregulated across high TP soybeans (Table 2). The top BLASTP hit identified this gene as closely related to the Squamosa promoter-binding protein-like (SPL) transcription factor family protein (fragment) in ancestral soybean, Glycine soja (Table S2). Squamosa family transcription factors have important roles regulating plant growth, development, response to stress, architecture and yield [36]. There is an inverse relationship between high protein and yield in soybean [37]; downregulation of Glyma.06G205700 in high TP soybean may be important for improving protein accumulation in the seed and a compensatory response to increased protein content in seeds.
Glyma.15G246500 was among the most commonly downregulated genes in high TP soybeans, present in the topmost downregulated genes in 6/10 high vs. low TP DE datasets. Glyma.15G246500 is predicted as an uncharacterized protein LOC100812621 isoform X3 in G. max; BLASTP and TAIR10 results indicate the protein embryo defective 3012 (EMB3012) in A. thaliana to be the most closely related known protein (Table S2). EMB family proteins have a diverse range of identified functions, but the majority are pentatricopeptide repeat (PPR) proteins, important for regulation of gene expression at the RNA level by facilitating post translational modifications such as splicing, editing and RNA stability [38]. EMB-defective Arabidopsis mutants produced seeds with defective embryo patterning, enlarged endosperm nuclear size, arrested or weak embryos [39]. The downregulation of Glyma.15G246500 in high TP soybeans implies there is a difference in chromosome maintenance and/or regulation of gene expression that may play a role underlying the difference in seed protein content between high TP and low TP soybean varieties. Differential mRNA processing (i.e., splicing) between high TP and low TP soybeans is likely. Low TP soybeans may be compensating for their less adequate storage protein genetics (in comparison to high TP counterparts) by increasing their expression of Glyma.15G246500. Splice variance expression comparisons (differential transcript usage) between Glyma.15G246500 in high TP soybeans and low TP soybeans would make an interesting next step to determine the effects of post translational modifications.
Glyma.19G084500 was the most commonly occurring upregulated gene in high 11S soybeans, present in the top 15 upregulated genes from 9/10 high vs. low 11S DE analyses (Table 3). Glyma.19G084500 is predicted to be a 52 kDa repressor of the inhibitor of the protein kinase-like in G. max with no known GOs (Tables S2 and S3). The top BLASTP hit identifies a hAT family dimerization domain containing protein in Medicago truncatula as the most closely related protein and the TAIR10 Arabidopsis homolog is a general transcription factor 2-related zinc finger protein (Table S2). PFAM results identified this gene to have a hAT family C-terminal dimerization region (Table S3). The Activator superfamily (hAT element superfamily) is a family of small and autonomous transposases with characteristic, highly conserved regions at the C-termini, important for dimerization. It would be of interest to investigate PPI prediction analysis and pulldown assays to uncover interacting partners for the product of Glyma.19G084500; this would be useful in further investigating the role, if any, this gene product plays in seed content accumulation.
Glyma.17G209900 is among the topmost commonly upregulated genes in high 11S soybeans (Table 3). Glyma.17G209900 is uncharacterized in G. max but has a BP GO for lipid metabolic process (GO:0006629) (Tables S2 and S3). The top BLASTP hit identified 12-oxo-phytodienoic acid reductase (OPR) in Zea mays as the most closely related protein and the TAIR10 hit identified 12-oxophytodienoate reductase 1 (OPR1) in Arabidopsis as a homolog (Table S2). 12-oxophytodienoate reductase 1 (OPR1) along with OPR2 and OPR3 are isoenzymes of the 12-oxophytodienoate reductase [40]. These three enzymes are highly related genes of oxylipin 12-oxophytodienoic acid (OPDA) responsible for the synthesis of jasmonic acid (JA) [41]. JA and its derivatives regulate gene expression, influencing a spectrum of developmental processes including seed germination, root development, fertility, fruit ripening, plant defense and senescence. OPR1, OPR2 and OPR3 can all play a role in the synthesis of JA; however OPR3 was identified as the most efficient stereoisomer [40]. OPR3 catalyzes the reduction of 9S,13S-12-oxo-phytodienoate, which is responsible for the plant hormone JA. Mutations in the OPR3 lead to a loss of enzymatic activity, which in turn leads to trouble synthesizing JA and fine-tuning gene expression in plants. Our data indicates high 11S soybeans increased expression of Glyma.17G209900 compared to low 11S soybeans. The effects of JA on protein and amino acid accumulation in oilseed crops is poorly understood, but foliar application of JA during vegetative and flowering stages of soybean development increased the amount of sulfur-containing amino acids in seeds [42]. Increasing expression of genes underlying JA synthesis is a plausible avenue for increased 11S accumulation as a result of increased sulfur-containing amino acid content. Upregulation of JA hormone signaling in high 11S soybeans should be further investigated for a potential positive correlation between JA signaling and 11S accumulation.
Glyma.06G287800 was upregulated among high 11S soybeans (Table 3). Poly-galacturonase in M. truncatula was the top BLASTP identity and the TAIR10 top identity was a Pectin lyase-like superfamily protein (Table S2). Poly-galacturonases are a family of hydrolases with a roll in cell separation by catalyzing α(1–4) linkages between D-galacturonic acid residues in cell wall pectins [43]. As a result of this role, poly-galacturonases are involved in a range of developmental programs including embryo development, organ abscission and pod dehiscence [44]. An increase in expression of poly-galacturonase in high 11S soybeans may attribute to cell wall loosening during embryo development in anticipation of accumulation of seed storage molecules.
Glyma.10G141200 was upregulated in high 11S soybeans and was identified to be closely related to the Disproportionating enzyme 1 (DPE1) found in Phaseolus angularis (Table 3 and Table S2). Plastidic DPE1 in rice has a significant role in the starch synthesis pathway by mediating the transfer of maltooligosyl groups from amylose, as well as amylopectin, to amylopectin [45]. The expression of Glyma.10G141200 in high 11S soybeans, not seen in low 11S soybeans, may be an important factor potentially underlying differences in carbohydrate synthesis and carbon allocation. Knockdown/knockout experiments could shed more light on the specific role expression of Glyma.10G141200 on starch synthesis and carbon allocation.
Glyma.18G112500 was upregulated in high 11S soybeans and was identified to be closely related to the Tetratricopeptide repeat (TPR)-like superfamily protein in Arabidopsis according to our TAIR10 results (Table 3 and Table S2). TPR-like proteins are important determents during signal transduction, mediated by plant hormones [46]. Increased expression of Glyma.18G112500 in high 11S soybeans compared to low 11S soybeans suggests it may have an underlying role in signal transduction and/or hormone signaling which influence 11S content. It would be of interest to conduct protein–protein interaction (PPI) prediction analysis on Glyma.18G112500 to determine interacting partners.
Glyma.01G127800 is downregulated in the high 11S soybean lines (Table 3). While yet uncharacterized in G. max, the BLASTP and TAIR10 results identify an Arabidopsis transducin family protein/WD-40 repeat family protein as the most closely related known sequence (Table S2). WD40 domains (sometimes called WD-repeat proteins) are prominent features within proteins spanning ~40–60 amino acids, typically terminated by a WD motif. WD40 domains are sites of PPIs, including multicomplex interactions and sometimes act as transient regulators for PPIs [47]. WD40 proteins conserved in Arabidopsis have a spectrum of known functions: auxin response, light signaling, meristem maintenance, time of flowering, flowering and seed development. Evidently the transducin family/WD-40 repeat family proteins influence a wide range of processes; the exact function of the protein encoded by Glyma.01G127800 is not currently understood. Downregulation of Glyma.01G127800 across high 11S soybeans suggests there is a decrease in WD40-related PPIs, signaling, or regulation of one of these pathways. PPI prediction for Glyma.01G127800 would be in the interest of directing further investigation into the putative role of this gene on seed content.
Intriguingly, Glyma.13G077600 and Glyma.03G054100 both made an appearance on the high TP and the high 11S downregulated shortlists. Glyma.13G077600 is downregulated in 7/10 high vs. low TP DE analyses and also in 7/10 high vs. low 11S DE analyses (Table 2 and Table 3). Glyma.13G077600 is uncharacterized in G. max, but the top BLASTP hit identified a protein of unknown function in Arabidopsis of the DUF538 protein superfamily as the most closely related protein on record (Table S2). Proteins of the DUF538 superfamily are widely distributed in monocots and dicots, with a conserved recognizable ß-sheet-rich domain called the DUF358 domain [48]. DUF538 proteins have been predicted to play regulatory roles in plants under different stress conditions and may be chlorophyll hydrolyzing enzymes induced by stress response stimuli [48]. The ambiguity of DUF358 proteins in literature makes pinpointing the exact relationship between Glyma.13G077600 and seed protein difficult. PPI prediction for the protein encoded by Glyma.13G077600 could be useful in investigating this connection.
Glyma.03G054100 is downregulated in 5/10 high vs. low TP datasets and also downregulated in 5/10 high 11S datasets (Table 2 and Table 3). Glyma.03G054100 is predicted in G. max to encode a TMV resistance protein N-like isoform X3 (Table S2). The most closely related protein according to BLASTP results is the TIR-NBS-LRR RCT1-like resistance protein in Medicago sativa (Table S2). Resistance (R) proteins are an important part of the defense system. A specific elicitor such as an avirulence (Avr) protein requires a specific R protein from the host to recognize it to signal the effector triggered immunity (ETI) pathway. Recognizing the Avr protein by specific R proteins leads to a cascade of responses resulting in an immune response by the plant in the form of a localized programmed cell death, through a hypersensitive response [49]. The downregulation of Glyma.03G054100 in both high TP and high 11S samples may be indicative of reduced ETI pathway signaling in these soybeans. The low 11S and low TP soybeans may induce signaling cascades in response to an environmental stress, which does not affect the high TP and high 11S soybeans in the same way.
Revigo is a tool used to reduce the number of GO terms in a given set to those scored least dispensable and to visualize the ontologies relative to one another in semantic space. Revigo scores GO terms from a given set (GO enrichment analysis) based on frequency, uniqueness and dispensability. Based on these parameters, a relative size for each group of terms is calculated, giving an indication of the relative number of daughter terms which fall under a parent term. GO enrichment for a given set of genes can be evaluated semantically and the GO term population structure and term relationships can be assessed. Using Revigo, we were able to assess GO based on the most indispensable terms across the shortlisted genes. After redundancy reduction, embryo development ending in seed dormancy (GO:0009793) was among the top indispensable downregulated BP GOs in high TP soybeans (Figure 6B), indicating high TP soybeans (lines 8, 9 and 10) are downregulating genes with roles in embryo development and seed dormancy. Upregulated genes in high 11S soybeans (lines 2, 4 and 8) had two key indispensable BP GO terms, lipid metabolic process (GO:0006629) and carbohydrate metabolic process (GO:0005975), signifying a significant evidence of upregulation in expression of genes involved in these two processes in high 11S soybeans (Figure 7A). Among the most important terms upregulated in high 11S soybeans as determined by Revigo include glucose metabolic process (GO:0006006), starch metabolic process (GO:0005982), maltose catabolic process (GO:0000023) and oxylipin metabolic process (GO:0031407) (Figure 7A, Table S4). These results indicate clear differences in expression of genes related to carbohydrate metabolism and lipid metabolism, particularly in high 11S soybeans.
When comparing the shortlisted data between East and West, the average DE of the majority of the genes is similar, with the exception of a few genes of interest. Glyma.03G057800 (Rhodanese-like family protein-like protein in M. truncatula) and Glyma.10G092400 (With no lysine (WNK) kinase 3, top Arabidopsis homolog) are more highly upregulated in high TP soybeans grown in the East than in the West. Eastern-grown high TP soybeans upregulated Glyma.03G057800 and Glyma.10G092400 by an average of log2FC 25.1 and log2FC 24.2, respectively; western-grown high TP soybeans upregulated these genes by an average of log2FC 11.9 and log2FC 10.5, respectively (Table S5). Rhodaneses are versatile sulfur carrying proteins, which have reactive roles in metabolism and regulatory pathways, including cell cycle control [35]. The observed difference between East and West high TP soybean data could be in part due to the products of Glyma.03G057800 (regulatory protein) and Glyma.10G092400 (a kinase) potentially having a relationship. It could be suggested that the high TP soybeans in the East are more tightly regulating cell cycle control and have enhanced intracellular signaling by increasing expression of Glyma.03G057800 and Glyma.10G092400, respectively. However DE analysis between genotypically identical high TP cultivars across East and West growing locations would need to be carried out to confirm this suggestion.
In comparing East vs. West high 11S average DE, the two locations show similar values with the exception of one gene, Glyma.03G054100, which is significantly more downregulated in western-grown high 11S soybeans (log2FC −38.9) than eastern-grown high 11S soybeans (log2FC −7.29) (Table S5). In high TP soybeans Glyma.03G054100 was downregulated in the East by an average log2FC of −18.58 and in the West by −25.25 (standard deviation of 4.71). Glyma.03G054100 is predicted in G. max to encode a TMV resistance protein N-like isoform X3 and TIR-NBS-LRR RCT1-like resistance protein in M. sativa is the most closely related known protein according to BLASTP (Table S2). Significant downregulation of a resistance gene in high 11S soybeans in the West compared to eastern-grown high 11S soybeans was observed without any differences seen in pathogen pressure. This gene was shortlisted for downregulation in both the high TP and high 11S soybeans, thus, there may be some attribution to a potential role in protein quality.

4. Materials and Methods

4.1. Lines, Locations, and Planting

The ten selected soybean lines, ranging in seed protein content, were grown in replicated trials at four locations across Canada: Ottawa Ontario (latitude 45.39, longitude −75.72), Morden Manitoba (49.18, X98.08), Brandon Manitoba (49.86, −99.98) and Saskatoon Saskatchewan (52.15, −106.57). Listed from lowest to highest average seed protein content, the lines under investigation are as follows; X5583-1-041-5-5 (line 1); AC Harmony [50] (line 2), AAC Halli (line 3), 90A01 [51] (line 4), Maple Amber (line 5), OT13-08 (line 6), OT14-03 (line 7), AAC Springfield (line 8), Jari (line 9) and AC Proteus [52] (line 10). These lines were selected as a representation of the spectrum of seed protein and oil in Canadian soybean agriculture. This spectrum of phenotypic information offered multi-cultivar information on high and low TP and/or 11S soybeans to dilute any genotypic bias. Planting was done in mid-end of May each year. Trial arrangement was carried out in 4 × 5 rectangular lattices and each had four replicates. Plot planting density was at a rate of 50 seeds m2 and best management practices were taken by each site. For more details on lines, see [53].

4.2. Sampling and RNA Extraction

Triplicate leaf tissue samples from R5 stage [54] soybeans for each of the 10 lines in each location were collected annually since 2018 and subjected to RNA-seq creating a large, high-quality data set. Tissue samples were crushed in liquid nitrogen using RNase-treated mortar and pestle. RNA extractions using SPLIT Total mRNA Extraction Kit (Lexogen, Vienna, Austria) was performed on approximately 200 mg of crushed leaf tissue from each sample according to the manufacturer’s instructions. RNA quality was initially tested using a NanoDrop™ 2000 Spectrophotometer (Thermo Fisher Scientific, Waltham, MA, USA), followed by a number of quality analysis checkpoints including agarose gel electrophoresis, TapeStation 4200 RNA ScreenTape (Agilent, Santa Clara, CA, USA) and 2100 Bioanalyzer (Agilent, Santa Clara, CA, USA) at Génome Québec (Montréal, QC, Canada) and the Ottawa Research and Development Centre (Ottawa, ON, Canada). RNA with RIN values of at least 4.5 and a Q30 score of at least 36 were used for library preparation. Spike-in RNA variants (SIRVs) (Lexogen, Vienna, Austria) were integrated within the RNA samples as controls to monitor and compare key parameters (such as sensitivity and quantification). For this work the E0 SIRV mix is used, which contains 69 different isoform variants with known sequences added at the same molar concentration.

4.3. RNA-seq Library Preparation, Alignment, Read Mapping, Read Counting, and DE

cDNA library preparation with paired-end sequencing was carried out using the Illumina HiSeq 4000 platform (Illumina, San Diego, CA, USA). RNA-seq analysis was carried out by Génome Québec and the preliminary DE analysis was performed by the department of bioinformatics at McGill University (Montréal, QC, Canada). Each read was sequenced and exported as a FASTA file for downstream RNA-seq quality control (QC) and analysis. dupRadar [55] (v3.16, Biberach an der Riß, Germany; Bioconductor, R) was used for duplication rate QC for the RNA-seq data.
Read normalization at the individual-sample level was performed using edgeR [56] (v3.16, Parkville, Victoria, Australia) and normalized data was subjected to exploratory data analysis (EDA) using R. PCA outputs for RNA-seq normalized read data were used to construct Figure 2 and Figure 3 in order to compare variance between transcript data. Each sample on the PCA plots is identified by its line number. The PCA plots do not directly represent the seed content variability data, but in some instances TP/11S content is responsible for the first PC (e.g., Ottawa 2018 and Saskatoon 2019 for the TP data).
QualiMap [57] (v2.2.1, Berlin, Germany) is a program which runs on an independent platform and was used as a QC step for the alignment of the sequencing data and features (genes or transcripts). Preseq [58] (v3.1.1, Los Angeles, CA, USA; Bioconductor, R) was used to predict the number of distinct reads from a sequencing library (in this case, RNA-seq). RSeQC [59] (v4.0.0, Nanjing, China; Bioconductor, R) was used to comprehensively evaluate the RNA-seq read data. RSeQC calculates the semantic read distribution of a sample, the inner distance between two reads, presence of exact read duplications, junction saturation, etc.
GenPipes is the main in-house framework of the Canadian Centre for Computational Genomics used to perform major processing steps [60]. Genpipes is the pipeline used to automatically run all the pre-processing steps mentioned below. Adaptor sequences and low-quality score containing bases (Phred score < 30) were trimmed from reads using Trimmomatic [61] (v0.36, Jülich, Germany). The resulting reads were aligned to the soybean genome (Glycine_max_v2.1, INSDC Assembly GCA_000004515.4, Jul 2018), using STAR [62] (v2.7.7a, Menlo Park, CA, USA) under the following command—runMode alignReads, after generating out own index files from the aforementioned genome. Read counts are obtained using HTSeq [63] (v0.12.3, Heidelberg, Germany) using the following options: “-m intersection-nonempty”.
The R package DESeq2 [64] (v3.16, Heidelberg, Germany) was used to identify differences in expression levels between the groups using negative Binomial GLM fitting and Wald statistics: nbinomWaldTest, similarly “ashr” [65] was used to shrink log2 fold changes in gene expression data. Genes were considered to be differentially expressed when they had a adjusted p-value < 0.05 as well as a log2 Fold change > |2| for shortlist considerations. For TP DE analyses, “low TP” samples serve as the “control” expression values; the expression of genes in low TP soybeans is being used to compare the expression of genes in high TP soybeans. With this, it is important to note when discussing log2FC differences in expression that the difference is seen in the high TP group. Upregulated genes would therefore be upregulated in high TP soybeans; downregulated genes are downregulated in high TP soybeans. Analyses for 11S protein content and the 11S:7S ratio was analyzed following the same fashion. Location was always controlled for each DE dataset, i.e., high protein soybeans grown in Ottawa were only compared with low protein soybeans grown in Ottawa.

4.4. Determination of Soybean Sample Seed Content

Seeds from individual plots from each soybean line-location-year trial were analyzed for the content of protein, oil, 11S, 7S and 11S:7S ratio at the central grain quality lab at Agriculture and Agri-Food Canada (AAFC) Ottawa Research and Development Centre (ORDC). Protein and oil content were determined using a Foss grain analyzer (Infratec 1241, Foss, Eden Prairie, MN, USA). Protein quality samples were milled in a ball mill before defatting via Accelerated Solvent Extraction Machine (Thermo Scientific Dionex ASE 350, Waltham, MA, USA). Protein separation was carried out using a Bioanalyzer (Agilent 2100 Bioanalyzer, Santa Clara, CA, USA) using a Protein 230 kit for sample preparation. Electropherogram analysis was used to assess 11S and 7S subunits and subsequently used to calculate the 11S:7S ratio. See [53] for a detailed description. Three samples with the highest and three samples with the lowest average values across all years (2018–2020) were designated “high” and “low”, respectively.

4.5. Candidate Gene Selection

As stated above, the DE datasets were trimmed at an adjusted p-value < 0.05 and a log2FC of at least 2. The total number of DE genes following trimming for each dataset is listed in Supplementary Table S6. The average number of remaining genes following trimming was 101 genes, thus the top and bottom 15% of the average number of genes with a log2FC of at least 2 was selected for cross examination (15 genes). From each DE dataset (10 TP datasets or 10 11S datasets), the top 15 upregulated genes (150 genes) and 15 most downregulated genes (150 genes) were extracted for a total of 300 genes differentially expressed in high TP soybeans and 300 genes differentially expressed in high 11S soybeans. From each of the resultant lists, the most frequently occurring genes were selected for candidacy. Genes that were found to be within the top 15 DE genes in at least 5/10 DE datasets were shortlisted. Table 4 lists all shortlisted genes and their corresponding NCBI gene identities.

4.6. Gene Ontology and Pathway Analysis

Shortlisted genes were assessed for GO to better understand the key BPs and MFs of genes in the curated candidate lists. Resultant gene lists, both up- and downregulated, were run independently through the SoyBase GO Term Enrichment Tool [22] (https://www.soybase.org/ (accessed on 17 May 2022)) to curate a list of GO terms representative of each gene list.
To search specifically for protein- and lipid-related genes, GO terms were identified using QuickGO (https://www.ebi.ac.uk/QuickGO/ (accessed on 17 May 2022)) [67] and supported by findings from a recent study on identifying soybean genes related to seed protein content [68]. Search keyword terms related to lipids include those with descriptions including “fat”, “glyce” and “lip” to encompass descriptive terms related to lipid prosses. Terms related to seed storage protein were searched based on the GO term for cupins, vicilins and globulins; “nutrient reservoir”. Because “nutrient reservoir” could include genes other than seed storage proteins, genes associated with this term were confirmed using the Nation Centre for Biotechnology Information (NCBI) for their annotation as a seed storage protein-related gene. Each curated GO list was then run through Revigo [23,24] (http://revigo.irb.hr/ (accessed on 17 May 2022)) at a threshold of 0.5 (50%) using Arabidopsis thaliana as the closest reference species.

4.7. Expression Profile Matrices and Heatmaps

Expression profiles were normalized across each dataset and exported as a matrix by edgeR [56]. edgeR uses a RNA-seq-specific normalization function for expression data for all samples in an expression matrix (produced by salmon). Genes of interest to be explored via heatmaps were selected based on the above criteria for curation of candidate gene shortlists.
Expression profiles can be thought of as a library of all genes and their relative expression in each individual sample. Heatmaps are a convenient way to compare the expression of a set of genes across all individual samples while simultaneously looking at the expression data globally. A matrix of all relative expression data of the genes of interest across all individual samples is read into the data processing tool, Heatmapper [69,70] (http://www.heatmapper.ca/ accessed on 14 June 2022), including all normalized expression data regardless of log2FC DE values for comparative expression purposes. Genes and their corresponding expression data are arranged according to the specified hierarchical gene clustering method (average linkage) and distance measurements (Euclidean). The correlation coefficient (distance measurement) between each pair of variables (columns, or individual samples) is calculated and a matrix of pairwise coefficients is created [69]. The row Z-score is used to scale the normalized expression data for enhanced visualization of trends in heatmaps and is calculated by (gene expression value in sample of interest)—(mean expression across all samples)/(standard deviation) [71].

5. Conclusions

In this work we sought to find DE genes between high and low TP soybean lines and between high and low 11S protein soybean lines from samples grown in four locations across Canada’s growing regions over 3 years. We identified shortlists of upregulated and downregulated genes in high TP and high 11S soybeans which may be of significance to soybean seed protein breeding programs. Ontologies of these genes include those within embryonic development, lipid metabolic pathways and carbohydrate metabolic pathways which may hold the key to the difference in seed quality between our selected lines. Within these shortlisted genes, a handful of key genes were found to be disproportionately DE between East and West growing locations. This suggests that these genes are likely to underlie the molecular mechanism responsible for the long withstanding observation of different seed protein and oil content, influenced by the different environmental factors at play.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24010222/s1, Table S1: The average seed protein, oil, 11S, 7S and 11S:7S ratio overall years + locations.; Table S2: Annotations for all short-listed genes including G. max Wm81.a2.0 ID, top BLASTP hit, top descriptive BLASTP hit and top TAIR10 BLASTP hit. Table S3: Shortlisted differential expressed genes in high TP and high 11S soybeans, including log2fold-change difference in expression across all year-location permutations and gene annotations.; Table S4: Revigo results, indispensable BS GOs up- and downregulated in high TP and high 11S soybeans.; Table S5: The average log2FC DE of soybean samples grown between East (Ottawa) and West (Morden, Brandon, Saskatoon) locations over 2018–2020. Table S6: The total number of remaining DE genes per dataset after trimming at an adjusted p-value below 0.05 and a log2FC DE of at least 2.

Author Contributions

Conceptualization, E.R.C. and B.S.; methodology, E.R.C. and B.S.; software, G.Z.; validation, J.C.H. and G.Z.; formal analysis, B.S., J.C.H., G.Z. and F.L.; investigation, J.C.H., B.S., D.L., A.H., R.M.M., T.D.W., K.A.D. and B.B.; resources, J.C.H., D.L., M.C., G.Z., A.H., R.M.M., T.D.W., K.A.D., B.B., F.L., E.R.C., M.H. and B.S.; data curation, J.C.H. and G.Z..; writing—original draft preparation, J.C.H. and N.N.; writing—review and editing, J.C.H., N.N., D.L., M.C., G.Z., A.H., R.M.M., T.D.W., K.A.D., B.B., F.L., A.G., M.H., E.R.C. and B.S.; visualization, J.C.H. and G.Z.; supervision, A.H., R.M.M., T.D.W., F.L., A.G., E.R.C. and B.S.; project administration, E.R.C. and B.S.; funding acquisition, E.R.C. and B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Agriculture and Agri-Food Canada (Management Driven Genomics Call) and the Canadian Field Crop Research Alliance (CFCRA).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to graciously thank the field crew at the farms in Ottawa, Morden, Brandon and Saskatoon. The authors would like to thank the Molecular Technology Lab at the Ottawa Research and Development Centre for their help with this project and with a special thank you to Kasia Dadej. We would also like to thank Genome Québec (Montréal, QC, Canada) for their contributions in RNA-sequencing. J.C.H. would like to thank S.M.B., C.E.C.H, M.R.W., E.A.H., M.S., J.S, H.H. and M.H.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

11SGlycinin
7Sβ-Conglycinin
ABAAbscisic acid
AcActivator
ACPEnoyl-Acyl Carrier Protein
ArfADP-ribosylation factor
AvrAvirulence
BBrandon
BLASTPBasic Local Alignment Search Tool
BPBiological process
bZipBasic Leucine Zipper domain
CysCysteine
DEDifferential Expression/Differentially Expressed
DinDark Inducible
DPEDisproportionating Enzyme
DUFDomain of Unknown Function
EMBEmbryo defective
ETIEffector Triggered Immunity
GOGene Ontology
hATHistone Acetyltransferases
JAJasmonic Acid
log2FClog2 fold change
LRRLeucine Rich Repeat
MMorden
MetMethionine
MFMolecular Function
NAD(P)Nicotinamide Adenine Dinucleotide Phosphate
NTRNitrate Transporter
OOttawa
OPDAOxylipin 12-Oxophytodienoic Acid
OPR12-Oxophytodienoate Reductase
PCPrinciple Component
PFAMProtein Family database
PPIProtein-Protein Interaction
PPRPentatricopeptide Repeat
PTRPeptide Transporter
QTLQuantitative Trait Loci
RResistance
R5Reproductive stage 5
RNA-seqRNA-sequencing
ROM1Regulator of MAT1
SSaskatoon
SASalicylic Acid
SBTSubtilase
SPLSquamosa Promoter-binding protein-like
TAIR10The Arabidopsis Information Resource
TIRToll-Interleukin-1-Receptor
TPTotal seed storage Protein
TPP2Tripeptidyl-Peptidase II
TPRTetratricopeptide Repeat
TTFTransposes Transcription Factor
WNKWith no lysine

References

  1. Tilman, D.; Balzer, C.; Hill, J.; Befort, B.L. Global Food Demand and the Sustainable Intensification of Agriculture. Proc. Natl. Acad. Sci. USA 2011, 108, 20260–20264. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Sprent, J.I.; Sprent, P. Nitrogen Fixing Organisms, 1st ed.; Sprent, J.I., Ed.; Springer: Dordrecht, The Netherlands, 1990; ISBN 978-0-412-34690-3. [Google Scholar]
  3. Snyder, C.S.; Bruulsema, T.W.; Jensen, T.L.; Fixen, P.E. Review of Greenhouse Gas Emissions from Crop Production Systems and Fertilizer Management Effects. Agric. Ecosyst. Environ. 2009, 133, 247–266. [Google Scholar] [CrossRef]
  4. Ma, Y.; Kan, G.; Zhang, X.; Wang, Y.; Zhang, W.; Du, H.; Yu, D. Quantitative Trait Loci (QTL) Mapping for Glycinin and β-Conglycinin Contents in Soybean (Glycine Max L. Merr.). J. Agric. Food Chem. 2016, 64, 3473–3483. [Google Scholar] [CrossRef] [PubMed]
  5. Yamada, T.; Mori, Y.; Yasue, K.; Maruyama, N.; Kitamura, K.; Abe, J. Knockdown of the 7S Globulin Subunits Shifts Distribution of Nitrogen Sources to the Residual Protein Fraction in Transgenic Soybean Seeds. Plant Cell Rep. 2014, 33, 1963–1976. [Google Scholar] [CrossRef] [Green Version]
  6. Breene, W.M.; Lin, S.; Hardman, L.; Orf, J. Protein and Oil Content of Soybeans from Different Geographic Locations. J. Am. Oil Chem. Soc. 1988, 65, 1927–1931. [Google Scholar] [CrossRef]
  7. Hooker, J.C.; Nissan, N.; Luckert, D.; Zapata, G.; Hou, A.; Mohr, R.M.; Glenn, A.J.; Barlow, B.; Daba, K.A.; Warkentin, T.D.; et al. GmSWEET29 and Paralog GmSWEET34 Are Differentially Expressed between Soybeans Grown in Eastern and Western Canada. Plants 2022, 11, 2337. [Google Scholar] [CrossRef]
  8. Zhang, H.; Goettel, W.; Song, Q.; Jiang, H.; Hu, Z.; Wang, M.L.; Charles An, Y.Q. Selection of GmSWEET39 for Oil and Protein Improvement in Soybean. PLoS Genet. 2020, 16, e1009114. [Google Scholar] [CrossRef]
  9. Borisjuk, L.; Nguyen, T.H.; Neuberger, T.; Rutten, T.; Tschiersch, H.; Claus, B.; Feussner, I.; Webb, A.G.; Jakob, P.; Weber, H.; et al. Gradients of Lipid Storage, Photosynthesis and Plastid Differentiation in Developing Soybean Seeds. New Phytol. 2005, 167, 761–776. [Google Scholar] [CrossRef]
  10. Rawsthorne, S. Carbon Flux and Fatty Acid Synthesis in Plants. Prog. Lipid Res. 2002, 41, 182–196. [Google Scholar] [CrossRef]
  11. Herman, E.M.; Larkins, B.A. Protein Storage Bodies and Vacuoles. Plant Cell 1999, 11, 601–613. [Google Scholar] [CrossRef]
  12. Peng, I.C.; Quass, D.W.; Dayton, W.R.; Allen, C.E. Physico Chemical Properties of Soybean 11S Globulin-A Review. Cereal Chem. 1984, 61, 480–490. [Google Scholar]
  13. Tsukada, Y.; Kitamura, K.; Harada, K.; Kaizuma, N. Genetic Analysis of Subunits of Two Major Storage Proteins (β-Conglycinin and Glycinin) in Soybean Seeds. Jpn. J. Breed. 1986, 36, 390–400. [Google Scholar] [CrossRef] [Green Version]
  14. Krishnan, H.B.; Jez, J.M. Review: The Promise and Limits for Enhancing Sulfur-Containing Amino Acid Content of Soybean Seed. Plant Sci. 2018, 272, 14–21. [Google Scholar] [CrossRef]
  15. Panthee, D.R.; Kwanyuen, P.; Sams, C.E.; West, D.R.; Saxton, A.M.; Pantalone, V.R. Quantitative Trait Loci for β-Conglycinin (7S) and Glycinin (11S) Fractions of Soybean Storage Protein. JAOCS J. Am. Oil Chem. Soc. 2004, 81, 1005–1012. [Google Scholar] [CrossRef]
  16. Ma, L.; Li, B.; Han, F.; Yan, S.; Wang, L.; Sun, J. Evaluation of the Chemical Quality Traits of Soybean Seeds, as Related to Sensory Attributes of Soymilk. Food Chem. 2015, 173, 694–701. [Google Scholar] [CrossRef] [Green Version]
  17. Nielsen, N.C.; Dickinson, C.D.; Cho, T.J.; Thanh, V.H.; Scallon, B.J.; Fischer, R.L.; Sims, T.L.; Drews, G.N.; Goldberg, R.B. Characterization of the Glycinin Gene Family in Soybean. Plant Cell 1989, 1, 313–328. [Google Scholar] [CrossRef]
  18. Li, C.; Zhang, Y.M. Molecular Evolution of Glycinin and β-Conglycinin Gene Families in Soybean (Glycine Max L. Merr.). Heredity 2011, 106, 633–641. [Google Scholar] [CrossRef] [Green Version]
  19. Qiu, H.; Hao, W.; Gao, S.; Ma, X.; Zheng, Y.; Meng, F.; Fan, X.; Wang, Y.; Wang, Y.; Wang, S. Gene mining of sulfur-containing amino acid metabolic enzymes in soybean. Yi Chuan 2014, 36, 934–942. [Google Scholar] [CrossRef]
  20. Ma, B.; Zhang, A.; Zhao, Q.; Li, Z.; Lamboro, A.; He, H.; Li, Y.; Jiao, S.; Guan, S.; Liu, S.; et al. Genome-Wide Identification and Analysis of Long Non-Coding RNAs Involved in Fatty Acid Biosynthesis in Young Soybean Pods. Sci. Rep. 2021, 11, 7603. [Google Scholar] [CrossRef]
  21. Canadian Grain Commission. Quality of Canadian Oilseed-Type Soybeans; Canadian Grain Commission: Winnipeg, MB, Canada, 2021.
  22. Soybase. Available online: https://soybase.org/ (accessed on 17 May 2022).
  23. Supek, F.; Bošnjak, M.; Škunca, N.; Šmuc, T. Revigo Summarizes and Visualizes Long Lists of Gene Ontology Terms. PLoS ONE 2011, 6, e21800. [Google Scholar] [CrossRef] [Green Version]
  24. Revigo. Available online: http://revigo.irb.hr/ (accessed on 17 May 2022).
  25. Schaller, A.; Stintzi, A.; Graff, L. Subtilases–Versatile Tools for Protein Turnover, Plant Development, and Interactions with the Environment. Physiol. Plant. 2012, 145, 52–66. [Google Scholar] [CrossRef] [PubMed]
  26. Lanoix, J.; Ouwendijk, J.; Lin, C.C.; Stark, A.; Love, H.D.; Ostermann, J.; Nilsson, T. GTP Hydrolysis by Arf-1 Mediates Sorting and Concentration of Golgi Resident Enzymes into Functional COP I Vesicles. EMBO J. 1999, 18, 4935–4948. [Google Scholar] [CrossRef] [PubMed]
  27. Dascher, C.; Balch, W.E. Dominant Inhibitory Mutants of ARF1 Block Endoplasmic Reticulum to Golgi Transport and Trigger Disassembly of the Golgi Apparatus. J. Biol. Chem. 1994, 269, 1437–1448. [Google Scholar] [CrossRef] [PubMed]
  28. Hillmer, S.; Movafeghi, A.; Robinson, D.G.; Hinz, G. Vacuolar Storage Proteins Are Sorted in the Cis-Cisternae of the Pea Cotyledon Golgi Apparatus. J. Cell Biol. 2001, 152, 41–50. [Google Scholar] [CrossRef] [PubMed]
  29. Hara-Nishimura, I.; Shimada, T.; Hatano, K.; Takeuchi, Y.; Nishimura, M. Transport of Storage Proteins to Protein Storage Vacuoles Is Mediated by Large Precursor-Accumulating Vesicles. Plant Cell 1998, 10, 825–836. [Google Scholar] [CrossRef] [Green Version]
  30. Guan, X.; Okazaki, Y.; Zhang, R.; Saito, K.; Nikolau, B.J. Dual-Localized Enzymatic Components Constitute the Fatty Acid Synthase Systems in Mitochondria and Plastids. Plant Physiol. 2020, 183, 517–529. [Google Scholar] [CrossRef] [Green Version]
  31. Léran, S.; Varala, K.; Boyer, J.-C.; Chiurazzi, M.; Crawford, N.; Daniel-Vedele, F.; David, L.; Dickstein, R.; Fernandez, E.; Forde, B.; et al. A Unified Nomenclature of Nitrate Transporter 1/Peptide Transporter Family Members in Plants. Trends Plant Sci. 2014, 19, 5–9. [Google Scholar] [CrossRef]
  32. Huang, W.; Bai, G.; Wang, J.; Zhu, W.; Zeng, Q.; Lu, K.; Sun, S.; Fang, Z. Two Splicing Variants of OsNPF7.7 Regulate Shoot Branching and Nitrogen Utilization Efficiency in Rice. Front. Plant Sci. 2018, 9, 300. [Google Scholar] [CrossRef] [Green Version]
  33. Wang, Y.; Suo, H.; Zheng, Y.; Liu, K.; Zhuang, C.; Kahle, K.T.; Ma, H.; Yan, X. The Soybean Root-Specific Protein Kinase GmWNK1 Regulates Stress-Responsive ABA Signaling on the Root System Architecture. Plant J. 2010, 64, 230–242. [Google Scholar] [CrossRef]
  34. Chern, M.S.; Eiben, H.G.; Bustos, M.M. The Developmentally Regulated BZIP Factor ROM1 Modulates Transcription from Lectin and Storage Protein Genes in Bean Embryos. Plant J. 1996, 10, 135–148. [Google Scholar] [CrossRef]
  35. Bordo, D.; Bork, P. The Rhodanese/Cdc25 Phosphatase Superfamily. Sequence-Structure-Function Relations. EMBO Rep. 2002, 3, 741–746. [Google Scholar] [CrossRef]
  36. Jiao, Y.; Wang, Y.; Xue, D.; Wang, J.; Yan, M.; Liu, G.; Dong, G.; Zeng, D.; Lu, Z.; Zhu, X.; et al. Regulation of OsSPL14 by OsmiR156 Defines Ideal Plant Architecture in Rice. Nat. Genet. 2010, 42, 541–544. [Google Scholar] [CrossRef]
  37. Burton, J.W.; Wilcox, C.R. Soybeans: Improvement, Production, and Uses. Agronomy 1987, 16, 211–247. [Google Scholar]
  38. Manna, S. An Overview of Pentatricopeptide Repeat Proteins and Their Applications. Biochimie 2015, 113, 93–99. [Google Scholar] [CrossRef] [Green Version]
  39. Herridge, R.P.; Day, R.C.; Macknight, R.C. The Role of the MCM2-7 Helicase Complex during Arabidopsis Seed Development. Plant Mol. Biol. 2014, 86, 69–84. [Google Scholar] [CrossRef]
  40. Schaller, F.; Biesgen, C.; Müssig, C.; Altmann, T.; Weiler, E.W. 12-Oxophytodienoate Reductase 3 (OPR3) Is the Isoenzyme Involved in Jasmonate Biosynthesis. Planta 2000, 210, 979–984. [Google Scholar] [CrossRef]
  41. Vick, B.A.; Zimmerman, D.C. Biosynthesis of Jasmonic Acid by Several Plant Species 1. Plant Physiol. 1984, 75, 458–461. [Google Scholar] [CrossRef] [Green Version]
  42. Farhangi-Abriz, S.; Ghassemi-Golezani, K. Improving Amino Acid Composition of Soybean under Salt Stress by Salicylic Acid and Jasmonic Acid. J. Appl. Bot. Food Qual. 2016, 89, 243–248. [Google Scholar] [CrossRef]
  43. Markovič, O.; Janeček, Š. Pectin Degrading Glycoside Hydrolases of Family 28: Sequence-Structural Features, Specificities and Evolution. Protein Eng. Des. Sel. 2001, 14, 615–631. [Google Scholar] [CrossRef]
  44. Hadfield, K.A.; Bennett, A.B. Polygalacturonases: Many Genes in Search of a Function 1. Plant Physiol. 1998, 117, 337–343. [Google Scholar] [CrossRef] [Green Version]
  45. Dong, X.; Zhang, D.; Liu, J.; Liu, Q.Q.; Liu, H.; Tian, L.; Jiang, L.; Qu, L.Q. Plastidial Disproportionating Enzyme Participates in Starch Synthesis in Rice Endosperm by Transferring Maltooligosyl Groups from Amylose and Amylopectin to Amylopectin. Plant Physiol. 2015, 169, 2496–2512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Schapire, A.L.; Valpuesta, V.; Botella, M.A. TPR Proteins in Plant Hormone Signaling. Plant Physiol. 2006, 1, 229–230. [Google Scholar] [CrossRef] [PubMed]
  47. Stirnimann, C.U.; Petsalaki, E.; Russell, R.B.; Müller, C.W. WD40 Proteins Propel Cellular Networks. Trends Biochem. Sci. 2010, 35, 565–574. [Google Scholar] [CrossRef] [PubMed]
  48. Gholizadeh, A. DUF538 Protein Superfamily Is Predicted to Be Chlorophyll Hydrolyzing Enzymes in Plants. Physiol. Mol. Biol. Plants 2016, 22, 77–85. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Burch-Smith, T.M.; Schiff, M.; Caplan, J.L.; Tsao, J.; Czymmek, K.; Dinesh-Kumar, S.P. A Novel Role for the TIR Domain in Association with Pathogen-Derived Elicitors. PLOS Biol. 2007, 5, e68. [Google Scholar] [CrossRef] [Green Version]
  50. Voldeng, H.D.; Guillemette, R.J.D.; Leonard, D.A.; Cober, E.R. AC Harmony Soybean. Can. J. Plant Sci. 1996, 76, 477–478. [Google Scholar] [CrossRef]
  51. Cober, E.R.; Bing, D.; Voldeng, H.D.; Soper, J.; Guillemette, R.J.D.; Sloan, A.; Hedges, B.R. 90A01 Soybean. Can. J. Plant Sci. 2006, 86, 481–482. [Google Scholar] [CrossRef]
  52. Voldeng, H.D.; Guillemette, R.J.D.; Leonard, D.A.; Cober, E.R. AC Proteus Soybean. Can. J. Plant Sci. 1996, 76, 153–154. [Google Scholar] [CrossRef]
  53. Cober, E.R.; Daba, K.A.; Warkentin, T.D.; Tomasiewicz, D.J.; Mooleki, P.S.; Karppinen, E.M.; Frey, J.; Mohr, R.M.; Glenn, A.J.; Shaw, L.; et al. Soybean Seed Protein Content Is Lower but Protein Quality Is Higher in Western Canada Compared to Eastern Canada. 2022; unpublished. [Google Scholar]
  54. Soybean Growth and Development. 2014. Available online: https://store.extension.iastate.edu/product/Soybean-Growth-and-Development (accessed on 17 May 2022).
  55. Sayols, S.; Scherzinger, D.; Klein, H. DupRadar: A Bioconductor Package for the Assessment of PCR Artifacts in RNA-Seq Data. BMC Bioinform. 2016, 17, 428. [Google Scholar] [CrossRef] [Green Version]
  56. Robinson, M.D.; McCarthy, D.J.; Smyth, G.K. EdgeR: A Bioconductor Package for Differential Expression Analysis of Digital Gene Expression Data. Bioinformatics 2009, 26, 139–140. [Google Scholar] [CrossRef] [Green Version]
  57. García-Alcalde, F.; Okonechnikov, K.; Carbonell, J.; Cruz, L.M.; Götz, S.; Tarazona, S.; Dopazo, J.; Meyer, T.F.; Conesa, A. Qualimap: Evaluating next-Generation Sequencing Alignment Data. Bioinformatics 2012, 28, 2678–2679. [Google Scholar] [CrossRef] [Green Version]
  58. Daley, T.; Smith, A.D. Modeling Genome Coverage in Single-Cell Sequencing. Bioinformatics 2014, 30, 3159–3165. [Google Scholar] [CrossRef]
  59. Wang, L.; Nie, J.; Sicotte, H.; Li, Y.; Eckel-Passow, J.E.; Dasari, S.; Vedell, P.T.; Barman, P.; Wang, L.; Weinshiboum, R.; et al. Measure Transcript Integrity Using RNA-Seq Data. BMC Bioinform. 2016, 17, 58. [Google Scholar] [CrossRef] [Green Version]
  60. Bourgey, M.; Dali, R.; Eveleigh, R.; Chen, K.C.; Letourneau, L.; Fillon, J.; Michaud, M.; Caron, M.; Sandoval, J.; Lefebvre, F.; et al. GenPipes: An Open-Source Framework for Distributed and Scalable Genomic Analyses. Gigascience 2019, 8, giz037. [Google Scholar] [CrossRef] [Green Version]
  61. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A Flexible Trimmer for Illumina Sequence Data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [Green Version]
  62. Dobin, A.; Davis, C.A.; Schlesinger, F.; Drenkow, J.; Zaleski, C.; Jha, S.; Batut, P.; Chaisson, M.; Gingeras, T.R. STAR: Ultrafast Universal RNA-Seq Aligner. Bioinformatics 2013, 29, 15–21. [Google Scholar] [CrossRef]
  63. Anders, S.; Pyl, P.T.; Huber, W. HTSeq-A Python Framework to Work with High-Throughput Sequencing Data. Bioinformatics 2015, 31, 166–169. [Google Scholar] [CrossRef] [Green Version]
  64. Love, M.I.; Huber, W.; Anders, S. Moderated Estimation of Fold Change and Dispersion for RNA-Seq Data with DESeq2. Genome Biol. 2014, 15, 550. [Google Scholar] [CrossRef] [Green Version]
  65. Stephens, M. False Discovery Rates: A New Deal. Biostatistics 2017, 18, 275–294. [Google Scholar] [CrossRef] [Green Version]
  66. National Center for Biotechnology Information. Available online: https://www.ncbi.nlm.nih.gov/ (accessed on 3 October 2022).
  67. QuickGO. Available online: https://www.ebi.ac.uk/QuickGO/ (accessed on 17 May 2022).
  68. Huang, S.; Yu, J.; Li, Y.; Wang, J.; Wang, X.; Qi, H.; Xu, M.; Qin, H.; Yin, Z.; Mei, H.; et al. Identification of Soybean Genes Related to Soybean Seed Protein Content Based on Quantitative Trait Loci Collinearity Analysis. J. Agric. Food Chem. 2019, 67, 258–274. [Google Scholar] [CrossRef]
  69. Babicki, S.; Arndt, D.; Marcu, A.; Liang, Y.; Grant, J.R.; Maciejewski, A.; Wishart, D.S. Heatmapper: Web-Enabled Heat Mapping for All. Nucleic Acids Res. 2016, 44, W147–W153. [Google Scholar] [CrossRef] [PubMed]
  70. Heatmapper. Available online: http://www.heatmapper.ca/ (accessed on 14 June 2022).
  71. Anders, S.; Huber, W. Differential Expression Analysis for Sequence Count Data. Genome Biol. 2010, 11, R106. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Seed composition across each variety from 2018–2020. (A) Average TP and seed oil content as a percentage of the seed weight. (B) Average 11S and 7S protein content as a percentage of total seed protein (primary y axis) and the average 11S:7S ratio (secondary y axis). Error bars indicate the least squares difference per series.
Figure 1. Seed composition across each variety from 2018–2020. (A) Average TP and seed oil content as a percentage of the seed weight. (B) Average 11S and 7S protein content as a percentage of total seed protein (primary y axis) and the average 11S:7S ratio (secondary y axis). Error bars indicate the least squares difference per series.
Ijms 24 00222 g001
Figure 2. PC analysis of normalized RNA-seq expression data for samples included in each individual DE analysis across the four locations and three years. Each TP year-location dataset is represented: (A) Ottawa 2018, (B) Ottawa 2019, (C) Ottawa 2020, (D) Morden 2018, (E) Morden 2019, (F) Morden 2020, (G) Brandon 2018, (H) Brandon 2019, (I) Saskatoon 2019, (J) Saskatoon 2020. Orange datapoints represent high TP samples, blue datapoints represent low TP samples. No data collected from Saskatoon 2018 and Brandon 2020. The line number (1, 2, 3, 8, 9, 10) for each corresponding datapoint is indicated.
Figure 2. PC analysis of normalized RNA-seq expression data for samples included in each individual DE analysis across the four locations and three years. Each TP year-location dataset is represented: (A) Ottawa 2018, (B) Ottawa 2019, (C) Ottawa 2020, (D) Morden 2018, (E) Morden 2019, (F) Morden 2020, (G) Brandon 2018, (H) Brandon 2019, (I) Saskatoon 2019, (J) Saskatoon 2020. Orange datapoints represent high TP samples, blue datapoints represent low TP samples. No data collected from Saskatoon 2018 and Brandon 2020. The line number (1, 2, 3, 8, 9, 10) for each corresponding datapoint is indicated.
Ijms 24 00222 g002
Figure 3. PC analysis of normalized RNA-seq expression data for samples included in each individual DE analysis across the four locations and three years. Each 11S year-location dataset is represented: (A) Ottawa 2018, (B) Ottawa 2019, (C) Ottawa 2020, (D) Morden 2018, (E) Morden 2019, (F) Morden 2020, (G) Brandon 2018, (H) Brandon 2019, (I) Saskatoon 2019, (J) Saskatoon 2020. Orange datapoints represent high 11S samples, blue datapoints represent low 11S samples. No data for Saskatoon 2018 and Brandon 2020. The line number (1, 2, 4, 5, 8, 9) for each corresponding datapoint is indicated.
Figure 3. PC analysis of normalized RNA-seq expression data for samples included in each individual DE analysis across the four locations and three years. Each 11S year-location dataset is represented: (A) Ottawa 2018, (B) Ottawa 2019, (C) Ottawa 2020, (D) Morden 2018, (E) Morden 2019, (F) Morden 2020, (G) Brandon 2018, (H) Brandon 2019, (I) Saskatoon 2019, (J) Saskatoon 2020. Orange datapoints represent high 11S samples, blue datapoints represent low 11S samples. No data for Saskatoon 2018 and Brandon 2020. The line number (1, 2, 4, 5, 8, 9) for each corresponding datapoint is indicated.
Ijms 24 00222 g003
Figure 4. Expression heatmaps for shortlisted genes differentially expressed in high TP soybeans (lines 8, 9, 10 vs. lines 1, 2, 3) for each year-location analysis. Each TP year-location dataset is represented: (A) Ottawa 2018, (B) Ottawa 2019, (C) Ottawa 2020, (D) Morden 2018, (E) Morden 2019, (F) Morden 2020, (G) Brandon 2018, (H) Brandon 2019, (I) Saskatoon 2019, (J) Saskatoon 2020. Upregulation is indicated by shades of yellow, downregulation is indicated by shades of blue. Grey indicates that a gene was not differentially expressed between high and low TP soybeans at an adjusted p-value < 0.05. The legend at the bottom left provides the gene identities and their correspondence with the numbers on the right y-axis of the heatmaps.
Figure 4. Expression heatmaps for shortlisted genes differentially expressed in high TP soybeans (lines 8, 9, 10 vs. lines 1, 2, 3) for each year-location analysis. Each TP year-location dataset is represented: (A) Ottawa 2018, (B) Ottawa 2019, (C) Ottawa 2020, (D) Morden 2018, (E) Morden 2019, (F) Morden 2020, (G) Brandon 2018, (H) Brandon 2019, (I) Saskatoon 2019, (J) Saskatoon 2020. Upregulation is indicated by shades of yellow, downregulation is indicated by shades of blue. Grey indicates that a gene was not differentially expressed between high and low TP soybeans at an adjusted p-value < 0.05. The legend at the bottom left provides the gene identities and their correspondence with the numbers on the right y-axis of the heatmaps.
Ijms 24 00222 g004
Figure 5. Expression heatmaps for shortlisted genes differentially expressed in high 11S soybeans (lines 2, 4, 8 vs. lines 1, 5, 9) for each year-location analysis. Each 11S year-location dataset is represented: (A) Ottawa 2018, (B) Ottawa 2019, (C) Ottawa 2020, (D) Morden 2018, (E) Morden 2019, (F) Morden 2020, (G) Brandon 2018, (H) Brandon 2019, (I) Saskatoon 2019, (J) Saskatoon 2020. Upregulation is indicated by shades of yellow, downregulation is indicated by shades of blue. Grey indicates that a gene was not differentially expressed between high and low 11S soybeans at an adjusted p-value < 0.05. The legend at the bottom left provides the gene identities and their correspond with the numbers on the right y-axis of the heatmaps.
Figure 5. Expression heatmaps for shortlisted genes differentially expressed in high 11S soybeans (lines 2, 4, 8 vs. lines 1, 5, 9) for each year-location analysis. Each 11S year-location dataset is represented: (A) Ottawa 2018, (B) Ottawa 2019, (C) Ottawa 2020, (D) Morden 2018, (E) Morden 2019, (F) Morden 2020, (G) Brandon 2018, (H) Brandon 2019, (I) Saskatoon 2019, (J) Saskatoon 2020. Upregulation is indicated by shades of yellow, downregulation is indicated by shades of blue. Grey indicates that a gene was not differentially expressed between high and low 11S soybeans at an adjusted p-value < 0.05. The legend at the bottom left provides the gene identities and their correspond with the numbers on the right y-axis of the heatmaps.
Ijms 24 00222 g005
Figure 6. Revigo plots summarizing the relationships between the most indispensable biological process (BP) gene ontologies (GOs) for upregulated (A) and downregulated (B) shortlist genes across the high vs. low TP DE analyses. Circle size represents logSize value, higher logSize values indicate a strong presence of a term and/or its daughter terms; more general terms have larger bubbles. Color represents significance of a term among the query set of GOs.
Figure 6. Revigo plots summarizing the relationships between the most indispensable biological process (BP) gene ontologies (GOs) for upregulated (A) and downregulated (B) shortlist genes across the high vs. low TP DE analyses. Circle size represents logSize value, higher logSize values indicate a strong presence of a term and/or its daughter terms; more general terms have larger bubbles. Color represents significance of a term among the query set of GOs.
Ijms 24 00222 g006
Figure 7. Revigo plots summarizing the relationships between the most indispensable BP GO terms for the upregulated (A) and downregulated (B) shortlist genes across the high vs. low 11S DE analyses. Circle size represents logSize value, higher logSize values indicate a strong presence of a term and/or its daughter terms; more general terms have larger bubbles. Color represents significance of a term among the query set of GOs.
Figure 7. Revigo plots summarizing the relationships between the most indispensable BP GO terms for the upregulated (A) and downregulated (B) shortlist genes across the high vs. low 11S DE analyses. Circle size represents logSize value, higher logSize values indicate a strong presence of a term and/or its daughter terms; more general terms have larger bubbles. Color represents significance of a term among the query set of GOs.
Ijms 24 00222 g007
Table 1. The average seed protein and oil content (as a percentage of the seed weight) for the lines selected for high vs. low TP DE (lines 1, 2, 3 vs. lines 8, 9, 10). The average 11S and 7S content (as a percentage of total seed protein content) and the average ratio of 11S:7S from high 11S lines selected for high vs. low 11S DE analysis (lines 2, 4, 8 vs. lines 1, 5, 9). The p-values represent the statistical difference between high and low soybean groups for each attribute via two-tailed t-tests.
Table 1. The average seed protein and oil content (as a percentage of the seed weight) for the lines selected for high vs. low TP DE (lines 1, 2, 3 vs. lines 8, 9, 10). The average 11S and 7S content (as a percentage of total seed protein content) and the average ratio of 11S:7S from high 11S lines selected for high vs. low 11S DE analysis (lines 2, 4, 8 vs. lines 1, 5, 9). The p-values represent the statistical difference between high and low soybean groups for each attribute via two-tailed t-tests.
High and Low TP Lines for DEHigh and Low 11S Lines for DE
LineTPoil Line11S7S11S:7S
Low TP137.522.1Low 11S168.930.82.38
236.722.4569.030.22.37
338.921.2968.030.82.28
High TP844.618.5High 11S274.124.93.19
943.019.0472.525.82.97
1046.916.5872.427.32.78
p-value 0.0100.018p-value 0.0050.0160.027
Table 2. The log2FC DE between low TP soybeans (lines 1, 2, 3) and high TP soybeans (lines 8, 9, 10) from 2018–2020 over 4 different locations. The top 15 most upregulated and 15 most downregulated genes from each DE analysis were selected. Genes DE in at least 5 of 10 datasets are shortlisted here including the log2FC DE in the high seed protein samples. The 10 DE datasets are identified by their Year-Location shorthand; as an example, 18.B indicates DE of high vs. low 11S soybeans in 2018 from Brandon. O, Ottawa; M, Morden; B, Brandon; S, Saskatoon.
Table 2. The log2FC DE between low TP soybeans (lines 1, 2, 3) and high TP soybeans (lines 8, 9, 10) from 2018–2020 over 4 different locations. The top 15 most upregulated and 15 most downregulated genes from each DE analysis were selected. Genes DE in at least 5 of 10 datasets are shortlisted here including the log2FC DE in the high seed protein samples. The 10 DE datasets are identified by their Year-Location shorthand; as an example, 18.B indicates DE of high vs. low 11S soybeans in 2018 from Brandon. O, Ottawa; M, Morden; B, Brandon; S, Saskatoon.
log2foldchange−25−20−15−10−50510152025
High Total Protein vs. Low Total Protein
Gene_id18.B18.M18.O19.B19.M19.O19.S20.M20.O20.STotal
Glyma.03G05780018.6 22.0 24.327.55.935.1425.75.518
Glyma.10G0924008.625.08 8.835.3622.924.210.825.5 8
Glyma.16G081500 23.828.529.631.029.531.830.830.28
Glyma.01G17910024.7 24.59.37 7.417.055.652.537
Glyma.02G060600 26.727.634.335.735.9 35.134.87
Glyma.10G09230025.64.84 25.0 22.525.012.427.1 7
Glyma.19G14020018.4 21.221.522.822.121.2 21.57
Glyma.09G184300 22.623.623.2 22.823.55
Glyma.16G060600 23.223.623.824.3 23.05
Glyma.16G08220018.9 7.32 6.50 5.065.00 5
Glyma.18G060700 20.821.5 23.5 21.120.4 5
Glyma.13G077600−13.4−35.2−34.7 −35.4−32.6−35.1−31.7 7
Glyma.15G246500 −19.2−20.8 −26.0−26.3 −24.2−22.5 6
Glyma.02G197100−11.3−33.3−36.7−42.7 −38.1 5
Glyma.03G053500 −22.4−9.52 −25.8−25.9 −24.1 5
Glyma.03G054100 −23.4−25.9 −10.2−24.6 −22.2 5
Glyma.03G068900 −23.9 −24.6−24.3−24.0 −23.2 5
Glyma.06G205700 −18.1−20.7 −21.8−23.9 −19.6 5
Table 3. log2FC DE between low 11S soybeans (lines 1, 5, 9) and high 11S soybeans (lines 2, 4, 8) from 2018–2020 over 4 different locations. The top 15 most upregulated and 15 most downregulated genes from each DE analysis were selected. Genes are DE in at least 5 of 10 datasets are listed here including the log2FC DE in the high 11S samples. DE datasets are identified by their year-location shorthand; as an example, 18.B indicates DE of high vs. low 11S soybeans in 2018 from Brandon. O, Ottawa; M, Morden; B, Brandon; S, Saskatoon.
Table 3. log2FC DE between low 11S soybeans (lines 1, 5, 9) and high 11S soybeans (lines 2, 4, 8) from 2018–2020 over 4 different locations. The top 15 most upregulated and 15 most downregulated genes from each DE analysis were selected. Genes are DE in at least 5 of 10 datasets are listed here including the log2FC DE in the high 11S samples. DE datasets are identified by their year-location shorthand; as an example, 18.B indicates DE of high vs. low 11S soybeans in 2018 from Brandon. O, Ottawa; M, Morden; B, Brandon; S, Saskatoon.
log2foldchange−25−20−15−10−50510152025
High 11S vs. Low 11S
Gene_id18.B18.M18.O19.B19.M19.O19.S20.M20.O20.STotal
Glyma.19G084500 24.722.325.322.927.221.527.527.324.79
Glyma.02G07730016.9 22.016.121.819.022.022.222.28
Glyma.17G209900 6.317.446.167.635.898.2410.63.458
Glyma.01G091300 7.43 4.937.259.216.865
Glyma.06G28780018.6 13.8 22.118.7 21.25
Glyma.10G141200 18.9 20.717.0 11.54.505
Glyma.14G204900 23.622.4 22.322.722.2 5
Glyma.18G112500 20.320.823.824.024.8 5
Glyma.13G077600 −33.2−33.5−30.8 −33.5−30.6−32.8 −29.67
Glyma.17G261800−15.0−38.6−40.0 −46.0 −44.9 −42.4−42.07
Glyma.01G127800 −40.2−43.2 −45.2−43.5 −41.65
Glyma.03G054100−18.6−42.7 −46.1 −7.29−48.3 5
Glyma.12G156500 −20.6−20.9−21.3 −22.6 −20.85
Glyma.18G082700 −39.8−42.2−43.9 −41.0−38.75
Table 4. All short-listed G. max gene IDs and corresponding NCBI Gene ID (https://www.ncbi.nlm.nih.gov/ accessed on 3 October 2022) [66].
Table 4. All short-listed G. max gene IDs and corresponding NCBI Gene ID (https://www.ncbi.nlm.nih.gov/ accessed on 3 October 2022) [66].
Gene ID (Wm82.a2)NCBI Gene ID
Glyma.01G091300NA
Glyma.01G127800NA
Glyma.01G179100102669100
Glyma.02G060600100808728
Glyma.02G077300NA
Glyma.02G197100NA
Glyma.03G053500NA
Glyma.03G054100NA
Glyma.03G057800100780425
Glyma.03G068900100527900
Glyma.06G205700NA
Glyma.06G287800NA
Glyma.09G184300106794632
Glyma.10G092300NA
Glyma.10G092400100805392
Glyma.10G141200NA
Glyma.12G156500NA
Glyma.13G077600NA
Glyma.14G204900NA
Glyma.15G246500100812621
Glyma.16G060600NA
Glyma.16G081500NA
Glyma.16G082200100791376
Glyma.17G209900100817099
Glyma.17G261800100794722
Glyma.18G060700NA
Glyma.18G082700NA
Glyma.18G112500100787722
Glyma.19G084500NA
Glyma.19G140200NA
NA = Not applicable; no corresponding NCBI gene ID.
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

Hooker, J.C.; Nissan, N.; Luckert, D.; Charette, M.; Zapata, G.; Lefebvre, F.; Mohr, R.M.; Daba, K.A.; Warkentin, T.D.; Hadinezhad, M.; et al. A Multi-Year, Multi-Cultivar Approach to Differential Expression Analysis of High- and Low-Protein Soybean (Glycine max). Int. J. Mol. Sci. 2023, 24, 222. https://doi.org/10.3390/ijms24010222

AMA Style

Hooker JC, Nissan N, Luckert D, Charette M, Zapata G, Lefebvre F, Mohr RM, Daba KA, Warkentin TD, Hadinezhad M, et al. A Multi-Year, Multi-Cultivar Approach to Differential Expression Analysis of High- and Low-Protein Soybean (Glycine max). International Journal of Molecular Sciences. 2023; 24(1):222. https://doi.org/10.3390/ijms24010222

Chicago/Turabian Style

Hooker, Julia C., Nour Nissan, Doris Luckert, Martin Charette, Gerardo Zapata, François Lefebvre, Ramona M. Mohr, Ketema A. Daba, Thomas D. Warkentin, Mehri Hadinezhad, and et al. 2023. "A Multi-Year, Multi-Cultivar Approach to Differential Expression Analysis of High- and Low-Protein Soybean (Glycine max)" International Journal of Molecular Sciences 24, no. 1: 222. https://doi.org/10.3390/ijms24010222

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