Skip to main content
  • Research article
  • Open access
  • Published:

Study on differentially expressed genes related to defoliation traits in two alfalfa varieties based on RNA-Seq

Abstract

Background

Alfalfa (Medicago sativa) is a widely cultivated, essential commercial forage crop. The protein content in its leaves is the critical factor in determining the quality of alfalfa. Thus far, the understanding of the molecular mechanism of alfalfa defoliation traits remains unclear. The transcriptome database created by RNA-Seq is used to identify critical genes related to defoliation traits.

Results

In this study, we sequenced the transcriptomes of the Zhungeer variety (with easy leaf abscission) and WL319HQ variety (without easy leaf abscission). Among the identified 66,734 unigenes, 706 differentially expressed genes (DEGs) upregulated, and 392 unigenes downregulated in the Zhungeer vs WL319HQ leaf. KEGG pathway annotations showed that 8,414 unigenes were annotated to 87 pathways and contained 281 DEGs. Six DEGs belonging to the “Carotenoid biosynthesis”, “Plant hormone signal transduction” and “Circadian rhythm-plant” pathways involved in defoliation traits were identified and validated by RT-qPCR analyses.

Conclusions

This study used RNA-Seq to discover genes associated with defoliation traits between two alfalfa varieties. Our transcriptome data dramatically enriches alfalfa functional genomic studies. In addition, these data provide theoretical guidance for field production practice and genetic breeding, as well as references for future study of defoliation traits in alfalfa.

Background

Alfalfa (Medicago sativa L.) is a member of the plant family Fabaceae. Alfalfa is widely distributed across the world and cultivated as a commercial forage crop due to its high protein, vitamin and mineral content [1, 2]. Mature alfalfa leaves are the primary crude protein storage organs (approximately 260 g kg− 1 dry basis), and protein content is a crucial indicator of the quality of alfalfa hay [3]. However, the leaves fall off very easily due to external pressure during harvest [4], and the resulting low-quality hay cannot satisfy livestock nutritional requirements.

The main factors in leaf senescence are endogenous abscisic acid (ABA) and ethylene (ETH) [5, 6]. ABA is a natural plant hormone identified in the 1960s, with roles in promoting plant organ shedding and seed germination [7, 8]. ABA is present in higher plant organs and tissues and especially abundant in mature and aging tissues or dormant organs. When plants are under biotic or abiotic stress, such as pathogen infection, drought or salinity, leaf senescence is induced by increasing ABA content [9,10,11,12]. ETH is another early-discovered plant hormone with function in the regulation of plant growth, such as the promotion of fruit ripening, acceleration of organ aging and shedding [13]. ETH is synthesized by conversion from methionine with enough oxygen, which occurs in various plant tissues and organs [14]. ETH synthesis is regulated by a variety of growth signals with tissue specificity at different development stages, and the accumulation of ETH can also stimulate further ETH production [15]. The content of ETH is significantly higher in the processes of fruit ripening, seed germination, and leaf and flower aging off [16, 17]. In addition, the synthesis of ETH is affected by a variety of external factors, such as mechanical damage, pathogen infection, low temperature, drought and salinity [18].

Previous studies showed that both ABA and ETH have the function of accelerating plant organ shedding. However, the physiological mechanisms are still unclear and remain controversial [19]. A study by Addicott [20] showed that endogenous ABA had positive effects on plant organ shedding, whereas field experiments failed, possibly because indoleacetic acid (IAA), gibberellin (GA) and cytokinin (CTK) in leaves have an adverse effect on ABA. Furthermore, the effect of introducing exogenous ABA on plant organ abscission was weaker than the effect of exogenous ETH. Osborne [21] concluded that ABA caused the premature senescence of organ cells, but that this may be due to increasing ETH content in plant organs that leads to organ shedding.

Transcriptome refers to a collection of all the transcription products under a specific physiological condition, including messenger RNA, ribosomal RNA, transport RNA and noncoding RNA [22]. Transcriptome sequencing research is essential for novel gene discovery, functional gene annotation, gene differential expression and the development of molecular markers [23,24,25]. Next Generation Sequencing (NGS) technology enables complete and rapid access to almost all transcript sequence information of a particular tissue or organ of a species and has been widely used in modal and nonmodel plant research [26,27,28]. To our knowledge, there have been no previous reports of research on DEGs related to defoliation traits in alfalfa varieties. This study aimed to construct a transcriptome database of two alfalfa leaf varieties in the early blooming stage. Based on DEGs, we aimed to identify the critical genes related to defoliation traits in the Zhungeer variety (characterized by easy leaf abscission) and the WL319HQ variety (without easy leaf abscission). The results of this study provide a theoretical basis for future alfalfa molecular breeding and provide a reference for subsequent study of the transcription of alfalfa leaves.

Results

De novo assembly transcriptome

Six libraries of the total RNA extracted from the leaves of Zhungeer and WL319HQ varieties at the early flowering stage were constructed for high-throughput sequencing. A total of 261,775,004 clean reads with a total of 39,202,954,927 nucleotides (nt) were obtained from the six sequencing libraries. Trinity method assembled a total of 66,734 unigenes with an N50 of 1496 nt. The maximum, minimum and average lengths were 143,789 nt, 201 nt and 869 nt, respectively (Fig. 1).

Fig. 1
figure 1

The size distribution of M. sativa unigenes. The abscissa is the length of the assembled unigenes from 200 nt to ≥3000 nt, and the ordinate is the number of unigenes of the corresponding length

Gene annotation

A total of 45,657 unigenes (68.42% of total 66,734 unigenes) were annotated against Nr, Swiss-Prot, KOG and KEGG databases using BLASTx (E-value < 1 × 10− 5). Among them, 42,888 (67.26%), 29,190 (43.74%), 24,844 (37.23%) and 15,647 (23.45%) unigenes were annotated to the Nr, Swiss-Prot, KOG and KEGG databases, respectively (Fig. 2a). According to the Nr database, 19.26% of unigenes showed homology (1 x E− 20 < E-value ≤1 x E− 5), 42.48% of unigenes showed strong homology (1 x E− 100 < E-value ≤1E− 20), and the remaining of 38.26% of unigenes had very strong homology (E-value ≤1E− 100) (Fig. 2b). For the species distribution of the top BLAST hits, 3,441 unigenes matched to the homologous sequences of Medicago truncatula, while 2,679, 1,058 and 551 unigenes matched to the homologous sequences of Cicer arietinum, Cajanus cajan and Glycine max, respectively (Fig. 2c).

Fig. 2
figure 2

Homology search of M. sativa unigenes. a Venn diagram of unigene numbers annotated by BLASTx with a cut-off E-value of 1 × 10− 5 against protein databases. The numbers in the circles indicate the number of unigenes annotated by single or multiple databases. b E-value distribution of the top BLASTx hits against the Nr database. c Ten top BLASTx hits of homologous sequences for the species distribution

Identification and analysis of DEGs

Using the Nr annotation results, Blast2GO software was used to analyze GO functional annotations of unigenes, and WEGO software was used to perform the functional classification of unigenes. A total of 1,098 significant DEGs were assigned to one or more ontologies by the standard of |log2FC| > 1 and FDR < 0.05 (Additional file 1: Table S1; Additional file 2: Table S2; Fig. 3), 706 unigenes were upregulated, and the other 392 unigenes were downregulated (Zhungeer vs WL319HQ). The significant DEGs were used for the subsequent analysis.

Fig. 3
figure 3

GO classification of assembled unigenes. A total of 17,156 unigenes were categorized into three main categories: biological process, cellular component and molecular function

In this group, “biological process”, “metabolic process”, “cellular process”, “response to stimulus” and “single-organism process” were the most frequent terms and contained 288, 260, 174 and 159 unigenes, respectively. For “metabolic process”, there were 195 upregulation unigenes and 93 downregulation unigenes. For “cellular process”, there were 170 upregulation unigenes and 90 downregulation unigenes. For “response to stimulus”, there were 131 upregulation unigenes and 43 downregulation unigenes. For “single-organism process”, there were 109 upregulation unigenes and 50 downregulation unigenes, whereas “immune system process” (two unigenes) and “rhythmic process” (two unigenes) were infrequent. Under the “cellular component” group, most of those were classified into “cell” (210 upregulation unigenes, 46 downregulation unigenes), “cell part” (210 upregulation unigenes, 46 downregulation unigenes), and “organelle” (156 upregulation unigenes, 30 downregulation unigenes). For the “molecular function” group, “binding” (169 upregulation unigenes, 97 downregulation unigenes) and “catalytic activity” (124 upregulation unigenes, 99 downregulation unigenes) were the most abundant subcategories.

KEGG pathway enrichment analysis

A total of 8414 unigenes were assigned to 87 KEGG pathways (Additional file 3: Table S3). The 15 top KEGG pathways with the highest representation of DEGs were: “Ribosome” (ko03010), “Spliceosome” (ko03040), “Protein processing in endoplasmic reticulum” (ko0414), “Carbon metabolism” (ko01200), “Endocytosis” (ko04144), “Biosynthesis of amino acids” (ko01230), “RNA transport” (ko03013), “Plant-pathogen interaction” (ko04626), “Starch and sucrose metabolism” (ko00500), “Oxidative phosphorylation” (ko00190), “Nucleotide excision repair” (ko03420), “Plant hormone signal transduction” (ko04075), “DNA replication” (ko0303), “Ubiquitin-mediated proteolysis” (ko04120), and “Homologous recombination” (ko03440). 622 (14.58%) unigenes and 64 DEGs accounting for 22.78% of 281 DEGs were annotated to the “Ribosome” pathway.

In the 87 KEGG pathways, the DEGs related to direct or indirect effects on the content of ABA and ETH were predicted. In the “Plant hormone signal transduction” pathway (ko04075), 294 unigenes with 6 DEGs were annotated. In the “Circadian rhythm-plant” pathway (ko04712), 61 unigenes with 4 DEGs were annotated. In addition, the “Carotenoid biosynthesis” pathway (ko00906) contained 44 unigenes with 1 DEG crucial for ABA synthesis, which was annotated.

Real-time quantitative PCR analysis of DEGs related to defoliation

To test the reliability of transcriptome sequencing data in the three pathways, Unigene0044746 (Auxin response factor, ARF), Unigene0002039 (Phytochrome interacting factor 3, PIF3), Unigene0027311 (Ethylene receptor protein, ETR), Unigene0053251 (Phytochrome B, PHYB), Unigene0053032 (Cryptochrome, CRY) and Unigene0014585 (9-cis-epoxycarotenoid dioxygenase 3, NCED3) (Additional file 4: Table S4) were selected for One-Step RT-qPCR analysis. In addition to DEGs, Unigene0030464 (Tubulin alpha, TUBA) and Unigene0032887 (indole-3-acetic acid-amido synthetase, GH) were selected in the RT-qPCR analysis as unaffected genes. The alfalfa GAPDH gene was used as the endogenous reference. The significant difference of RT-qPCR data between two varieties was analyzed by t-test, and the expression patterns revealed by RT-qPCR analysis were similar to those revealed by transcriptome sequencing for the same genes (Fig. 4).

Fig. 4
figure 4

The relative expression level of DEGs involved in “Plant hormone signal transduction”, “Circadian rhythm-plant” and “Carotenoid biosynthesis” pathways revealed by RT-qPCR analysis. TUBA: Tubulin alpha, no significant difference (P = 0.165); GH: indole-3-acetic acid-amido synthetase, no significant difference (P = 0.246); ARF: Auxin response factor, significant difference (P = 0.033); PIF3: Phytochrome interacting factor 3, significant difference (P = 0.043); ETR: Ethylene receptor protein, significant difference (P = 0.014); PHYB: Phytochrome B, significant difference (P = 0.047); CRY: Cryptochrome, significant difference (P = 0.049); NCED3: 9-cis-epoxycarotenoid dioxygenase 3, significant difference (P = 0.04)

Discussion

ABA, ETH, jasmonic acid, salicylic acid, CTK and auxin are the primary phytohormones that regulate plant senescence [29]. When plants are under environmental stress, such as drought, salinity or low temperature, the leaves will accelerate senescence and abscission due to the elevation of ABA concentration [11, 30]. In addition, ETH stimulates the synthesis of cellulase and controls the release of cellulase from protoplasts into the cell wall, which promotes cell wall degradation and leaf abscission [5]. Many studies reported that ETH levels would rise if plant tissues were damaged by mechanical force [4, 31, 32]. Therefore, the abscission of alfalfa leaves may be due to mechanical damage to plant tissues during harvesting, resulting in elevation of ETH concentration in leaf tissues. The mechanism by which phytohormone interactions promote the abscission of leaves is not well understood and is difficult to study [33].

In higher plants, xanthoxin is not only the synthesis precursor of ABA but also a metabolic intermediate of carotenoid catabolism [34]. The RNA-Seq data revealed that the alfalfa 9-cis-epoxycarotenoid dioxygenase 3 (NCED3) gene differentially expressed between WL319HQ and Zhungeer varieties. The RT-qPCR analysis showed that the transcription of the Zhungeer variety was approximately 2.3 times higher than the WL319HQ variety. NCED is the enzyme that catalyzes the cleavage reaction of 9-cis-epoxycarotenoids to produce xanthoxin and is the rate-limiting enzyme of ABA biosynthesis in higher plants [35,36,37]. NCED has been identified in maize, kidney bean, tomato and Arabidopsis thaliana [4, 35, 36, 38, 39]. A. thaliana NCED and its role in ABA biosynthesis are the most well-studied [37]. Overexpression of A. thaliana NCED3 resulted in a significantly higher level of endogenous ABA in the plant [40], and similar results were obtained in transgenic tobacco to overexpress the tomato NCED1 gene [38].

The “Plant hormone signal transduction” (ko04075) pathway plays a vital role in the metabolism of phytohormones in higher plants. In the RNA-Seq data, the expression level of the ethylene receptor protein (ETR) gene in Zhungeer variety was approximately 1.4 times higher than the WL319HQ variety. ETR is a protein that initiates the ETH signal transduction pathway by binding the ETH and relaying the ETH signal to constitutive triple response 1 (CTR1) Raf-like protein kinase [41]. The high-affinity binding activity has been confirmed in A. thaliana, as well as the tomato [42,43,44]. In the present study, the ETR gene was downregulated in WL319HQ compared to the Zhungeer variety, indicating that Zhungeer is more likely to accumulate ETH than the WL319HQ variety.

Auxin response factors (ARF) are transcription factors that bind specifically to the DNA sequence 5’-TGTCTC-3′ found in auxin-responsive promoter elements (AuxREs). Although ARF genes are not involved in leaf senescence directly, they act as transcriptional activators or repressors and therefore promote flowering, stamen development, floral organ abscission and fruit dehiscence [45]. Study of the A. thaliana ARF1 and ARF2 mutant has shown that ARF2 may affect plant tissue shedding [46]. Nagpal et al. [47] found that the A. thaliana ARF6 gene promotes flower organ maturation and senescence. In addition, NPH4/ARF7 and ARF19 not only enhance the ARF2 phenotype but also induce leaf senescence [46]. The ARF gene was significantly downregulated in alfalfa leaves in the WL319HQ variety.

Plant circadian rhythm is a hereditary physiological regulatory mechanism that synchronizes with circadian alternation in the plant. The response of plants to environmental factors mainly depends on the hormone regulatory network, including CTK, IAA, ABA and GA [48]. Circadian rhythm plays an essential role in the plant hormonal regulation pathway [49, 50]. At the same time, the plant hormone signaling network also acts on circadian rhythm to transmit different metabolic signals and environmental signals to the endogenous circadian rhythm system, thus forming a complex regulatory network [51, 52]. In this study, three DEGs involved in the plant circadian rhythm pathway were identified that may affect the regulation of ABA and ETH. Compared with Zhungeer, the PHYB (− 1.775-fold), PIF3 (− 3.715-fold) and CRY (− 1.952) gene of WL319HQ were downregulated. PIF3 plays a critical role not only in light and temperature-mediated environmental signaling but also in the signaling of ABA and ETH. Light activates the phytochrome pathway to induce the degradation of PIF3 and promotes the accumulation of carotenoids, which may lead to an increase in ABA content.

In summary, it is likely that the Zhungeer alfalfa variety is more susceptible to abscission than the WL319HQ variety due to higher levels of plant hormone accumulation. Higher expression of NCED3 and PIF3 genes caused increasing ABA content, which may be a reason why the Zhungeer variety is more likely to defoliate. In addition, upregulation of ETR and ARF expression affects ETH; auxin accumulation is another indirect evidence as well.

Conclusion

This study is the first to use RNA-Seq to identify the defoliation traits between alfalfa varieties. It is essential to understand the physiological effects of various hormones on plants, the interaction between hormones and their relationship with the agricultural production environment. Our transcriptome data dramatically enriches alfalfa genome research. This research will provide theoretical guidance for field production practice and genetic breeding and provides a reference for the future study of defoliation traits in alfalfa.

Methods

Sample collection

We selected two varieties of the tetraploid M. sativa: the Zhungeer variety that has easy leaf abscission and the WL319HQ variety that is not characterized by easy leaf abscission. Both varieties were grown in experimental fields located in Baotou City, Inner Mongolia, China (110°37′E, 40°5’N). Leaves of the Zhungeer and WL319HQ varieties were collected from two-year-old plants at the early flowering stage in June. All samples were immediately frozen in liquid nitrogen and stored at − 80 °C for future RNA extraction.

RNA extraction and RNA-Seq library construction

Total RNA was extracted from three biological replicates of the leaves of each M. sativa variety using the RNeasy Plant Mini Kit (Qiagen, Germany), following the manufacturer’s protocol. The integrity of RNA samples was measured by Agilent 2100 Bioanalyzer (Agilent Technologies, USA). The Poly (A) mRNA was enriched by NEBNext Oligo(dT)25 beads (NEB, USA) from 50 μl total RNA. Then, the enriched mRNA was constructed to a cDNA library by NEBNext Ultra RNA Library Prep Kit for Illumina (NEB, USA), following the manufacturer’s protocol.

Sequencing and raw reads filtering

The transcriptome sequencing was performed using an Illumina HiSeq 4000 sequencing platform with paired-end 150 bp (PE150) sequencing strategy at Gene Denovo Biotechnology Co. (Guangzhou, China). Raw reads obtained from the sequencing that may affect the following assembly and analysis were removed, including reads with adapters, reads containing more than 10% unknown nucleotides and low-quality reads containing more than 50% low-quality bases (Q-values ≤10).

De novo transcriptome assembly and unigenes detection

Since M. sativa genome information was not previously available, full-length transcriptome de novo assembly was carried out using the reference genome independent Trinity method [53]. Trinity is a software package consisting of three modules: Inchworm, Chrysalis and Butterfly. First, Inchworm assembled clean reads to a collection of linear contigs by the greedy k-mer method. Second, the related minimal overlapping contigs created by Inchworm were clustered into connected components by Chrysalis, and a de Bruijn graph was constructed for each cluster of related contigs. Finally, Butterfly analyzed the paths from the de Bruijn graphs that were taken by reads and pairs of reads, and the complete transcripts and unigenes sets were obtained from the outputs [54].

Unigenes annotation

The unigenes were annotated using the BLASTx program (http://www.ncbi.nlm.nih.gov/BLAST/) with an E-value threshold of 1 × 10− 5 to Nr (NCBI nonredundant protein) database (http://www.ncbi.nlm.nih.gov), the Swiss-Prot (A manually annotated and reviewed protein sequence database) protein database (http://www.expasy.ch/sprot), the KEGG (Kyoto Encyclopedia of Genes and Genomes) database (http://www.genome.jp/kegg) and the KOG (Clusters of Orthologous Groups of proteins) database (http://www.genome.jp/kegg/ko.html). Protein functional annotations were obtained according to the best alignment results.

Analysis of DEGs

Gene expression analysis of unigenes was performed and normalized by RPKM (reads per kb per million reads). The edgeR package (http://www.r-project.org/) was used to identify DEGs across samples or groups. Genes with a fold change ≥2 and a false discovery rate (FDR) < 0.05 were considered significant DEGs. DEGs were further subjected to GO functional enrichment analysis and KEGG pathway analysis.

Real-time quantitative PCR analysis

One-Step Real-time quantitative PCR (RT-qPCR) was carried out to confirm the RNA-Seq data from the extracted M. sativa leaves’ total RNA. The experiments were performed using One-Step TB Green PrimeScript RT-PCR Kit II (Takara, Japan) on an ABI 7500 system (Applied Biosystems, USA). Primers for One-Step RT-qPCR were designed using Primer Premier 5.0 software (Premier, Canada). The alfalfa GAPDH gene was selected as the endogenous reference [55]. All the primers were synthesized by Invitrogen (Beijing, China) and are listed in Additional file 5: Table S5. A 20 μl reaction mix was set up containing 50 ng total RNA and 0.4 μM of each primer. The One-Step RT-qPCR analysis was performed on the extracted total RNA in three biological replicates of the leaves of each M. sativa variety, and each RNA sample was carried out in technical triplicate. The program was set at 42 °C for 5 min then 95 °C for 10 s, followed by 40 cycles of 95 °C for 5 s then 60 °C for 34 s. The final melting curve analysis was set at 95 °C for 15 s, 60 °C for 1 min and 95 °C for 15 s. The relative expression changes of the endogenous reference and tested genes were analyzed by the 2 - CT method [56].

Statistical analysis

Statistical analyses were performed using the software SigmaPlot 12.5 (Systat Software, Inc. USA). The Shapiro-Wilk test was used to assess the data for normality. Then, the t-test was used to test the expression levels of target genes between the Zhungeer and WL319HQ varieties. Results were considered significant at 5%.

Abbreviations

ABA:

Abscisic acid

ARF :

Auxin response factor

AuxREs:

Auxin-responsive promoter elements

BLAST:

Basic Local Alignment Search Tool

CRY :

Cryptochrome

CTK:

Cytokinin

CTR1 :

Constitutive triple response 1

DEGs:

Differentially expressed genes

ETH:

Ethylene

ETR :

Ethylene receptor protein

GA:

Gibberellin

GH :

Indole-3-acetic acid-amido synthetase

GO:

Gene Ontology

IAA:

Indoleacetic acid

KEGG:

Kyoto Encyclopedia of Genes and Genomes

KOG:

Clusters of Orthologous Groups of proteins

NCED3 :

9-cis-epoxycarotenoid dioxygenase 3

Nr:

NCBI nonredundant protein

nt:

Nucleotides

PHYB :

Phytochrome B

PIF3 :

Phytochrome interacting factor 3

RT-qPCR:

Real-time quantitative PCR

Swiss-Prot:

A manually annotated and reviewed protein sequence database

TUBA :

Tubulin alpha

References

  1. Yacoubi R, Job C, Belghazi M, Chaibi W, Job D. Toward characterizing seed vigor in alfalfa through proteomic analysis of germination and priming. J Proteome Res. 2011;10(9):3891–903.

    Article  CAS  Google Scholar 

  2. Jia C, Wu X, Chen M, Wang Y, Liu X, Gong P, Xu Q, Wang X, Gao H, Wang Z. Identification of genetic loci associated with crude protein and mineral concentrations in alfalfa (Medicago sativa) using association mapping. BMC Plant Biol. 2017;17:97.

    Article  Google Scholar 

  3. Hojilla-Evangelista MP, Selling GW, Hatfield R, Digman M. Extraction, composition, and functional properties of dried alfalfa (Medicago sativa L.) leaf protein. J Sci Food Agr. 2017;97(3):882–8.

    Article  CAS  Google Scholar 

  4. Lieberman M. Biosynthesis and action of ethylene. Annu Rev Plant Biol. 1979;30:533–91.

    Article  CAS  Google Scholar 

  5. Reid PD, Strong HG, Lew F, Lewis LN. Cellulase and abscission in the red kidney bean (Phaseolus vulgaris). Plant Physiol. 1974;53:732–7.

    Article  CAS  Google Scholar 

  6. Bleecker AB, Patterson SE. Last Exit: Senescence, abscission, and meristem arrest in Arabidopsis. Plant Cell. 1997;9:1169–79.

    Article  CAS  Google Scholar 

  7. Walker-Simmons M. ABA levels and sensitivity in developing wheat embryos of sprouting resistant and susceptible cultivars. Plant Physiol. 1987;84(1):61–6.

    Article  CAS  Google Scholar 

  8. Finkelstein RR, Gampala SSL, Rock CD. Abscisic acid signaling in seeds and seedlings. Plant Cell. 2002;14(1):15–45.

    Article  Google Scholar 

  9. Adie BAT, Pérez-Pérez J, Pérez-Pérez MM, Godoy M, Sánchez-Serrano J, Schmelz EA, Solano R. ABA is an essential signal for plant resistance to pathogens affecting JA biosynthesis and the activation of defenses in Arabidopsis. Plant Cell. 2007;19(5):1665–81.

    Article  CAS  Google Scholar 

  10. Figueiredo MVB, Burity HA, Martínez CR, Chanway CP. Alleviation of drought stress in the common bean (Phaseolus vulgaris L.) by co-inoculation with Paenibacillus polymyxa and rhizobium tropici. Appl Soil Ecol. 2008;40(1):182–8.

    Article  Google Scholar 

  11. Zhang J, Liang S, Duan J, Wang J, Chen S, Cheng Z, Zhang Q, Liang X, Li Y. De novo assembly and characterisation of the transcriptome during seed development, and generation of genic-SSR markers in Peanut (Arachis hypogaea L.). BMC Genomics. 2012;13:90.

    Article  CAS  Google Scholar 

  12. Lovelli S, Sofo A, Perniola M, Scopa A. Abscisic acid and biomass partitioning in tomato under salinity. In: Ahmad P, Azooz M, Prasad M, editors. Ecophysiology and responses of plants under salt stress. New York: Spinger; 2013. p. 267–82.

    Chapter  Google Scholar 

  13. Burg SP. Ethylene in plant growth. PNAS. 1973;70(2):591–7.

    Article  CAS  Google Scholar 

  14. Davies PJ. Ethylene in plant biology. Cell. 1993;72(1):11–2.

    Article  Google Scholar 

  15. Reig C, Martínez-Fuentes A, Mesejo C, Rodrigo MJ, Zacarías L, Agustí M. Loquat fruit lacks a ripening-associated autocatalytic rise in ethylene production. J Plant Growth Regul. 2016;35(1):232–44.

    Article  CAS  Google Scholar 

  16. McMurchie EJ, McGlasson WB, Eaks IL. Treatment of fruit with propylene gives information about the biogenesis of ethylene. Nature. 1972;237:235–6.

    Article  CAS  Google Scholar 

  17. Machabé ES, Saini HS. Differences in the requirement for endogenous ethylene during germination of dormant and non-dormant seeds of Chenopodium album L. J Plant Physiol. 1991;138(1):97–101.

    Article  Google Scholar 

  18. Fluhr R, Mattoo AK, Dilley DR. Ethylene-biosynthesis and perception. Crit Rev Plant Sci. 1996;15(5&6):479–523.

    CAS  Google Scholar 

  19. Kitsaki CK, Vemmos SN, Tzoutzoukou CG. Changes of respiration rate, ethylene evolution, and abscisic acid content in developing inflorescence and young fruit of olive (Olea europaea L. cv. Konservolia). J Plant Growth Regul. 1999;18:1–7.

    Article  CAS  Google Scholar 

  20. Addicott FT. Abscisic acid. New York: Praeger; 1983.

    Google Scholar 

  21. Osborne DJ, Morgan PW. Abscission. Crit Rev Plant Sci. 1989;8(2):103–29.

    Article  CAS  Google Scholar 

  22. Costa V, Angelini C, de Feis I, Ciccodicola A. Uncovering the complexity of transcriptomes with RNA-seq. Biomed Res Int. 2010. https://doi.org/10.1155/2010/853916.

    Article  Google Scholar 

  23. Lu T, Lu G, Fan D, Zhu C, Li W, Zhao Q, Feng Q, Zhao Y, Guo Y, Li W, Huang X, Han B. Function annotation of the rice transcriptome at single-nucleotide resolution by RNA-seq. Genome Res. 2010;20(9):1238–49.

    Article  CAS  Google Scholar 

  24. Wang Z, Fang B, Chen J, Zhang X, Luo Z, Huang L, Chen X, Li Y. De novo assembly and characterization of root transcriptome using Illumina paired-end sequencing and development of cSSR markers in sweetpotato (Ipomoea batatas). BMC Genomics. 2010;11:726.

    Article  CAS  Google Scholar 

  25. Zhang K, Xia X, Zhang Y, Gan S. An ABA-regulated and Golgi-localized protein phosphatase controls water loss during leaf senescence in Arabidopsis. Plant J. 2012;69:667–78.

    Article  CAS  Google Scholar 

  26. Kang YJ, Lee J, Kim YH, Lee S. Identification of tissue-specific gene clusters and orthologues of nodulation-related genes in Vigna angularis. Plant Genet Resour. 2014;12(S1):S21–6.

    Article  CAS  Google Scholar 

  27. O'Rourke JA, Bolon Y, Bucciarelli B, Vance CP. Legume genomics: understanding biology through DNA and RNA sequencing. Annuls of Bot. 2014;113(7):1107–20.

    Article  CAS  Google Scholar 

  28. Gailing O, Staton ME, Lane T, Schlarbaum SE, Nipper R, Owusu SA, Carlson JE. Construction of a framework genetic linkage map in Gleditsia triacanthos L. Plant Mol Biol Rep. 2017;35(2):177–87.

    Article  CAS  Google Scholar 

  29. Lim PO, Kim HJ, Nam HG. Leaf senescence. Annu Rev Plant Biol. 2007;58:115–36.

    Article  CAS  Google Scholar 

  30. Guo Y, Gan S. Leaf senescence: signals, execution, and regulation. In: Schatten GP, editor. Current topics in developmental biology. New York: Academic Press; 2005. p. 83–112.

    Google Scholar 

  31. Yuan R. Effects of temperature on fruit thinning with ethephon in ‘Golden delicious’ apples. Sci Hortic-Amsterdam. 2007;113:8–12.

    Article  CAS  Google Scholar 

  32. Zhu H, Dardick CD, Beers EP, Callanhan AM, Xia R, Yuan R. Transcriptomics of shading-induced and NAA-induced abscission in apple (Malus domestica) reveals a shared pathway involving reduced photosynthesis, alterations in carbohydrate transport and signaling and hormone crosstalk. BMC Plant Biol. 2011;11:138.

    Article  Google Scholar 

  33. De Smet I, Signora L, Beeckman T, Inzé D, Foyer CH, Zhang H. An abscisic acid-sensitive checkpoint in lateral root development of Arabidopsis. Plant J. 2003;33:543–55.

    Article  CAS  Google Scholar 

  34. Lefebvre V, North H, Frey A, Sotta B, Seo M, Okamoto M, Nambara E, Marion-Poll A. Functional analysis of Arabidopsis NCED6 and NCED9 genes indicates that ABA synthesized in the endosperm is involved in the induction of seed dormancy. Plant J. 2006;45:309–19.

    Article  CAS  Google Scholar 

  35. Tan BC, Schwartz SH, Zeevaart JAD, McCarty DR. Genetic control of abscisic acid biosynthesis in maize. PNAS. 1997;94:12235–40.

    Article  CAS  Google Scholar 

  36. Qin X, Zeevaart JAD. The 9-cis-epoxycarotenoid cleavage reaction is the key regulatory step of abscisic acid biosynthesis in water-stressed bean. PNAS. 1999;96:15354–61.

    Article  CAS  Google Scholar 

  37. Tan B, Joseph LM, Deng W, Liu L, Li Q, Cline K, McCarty DR. Molecular characterization of the Arabidopsis 9-cis epoxycarotenoid dioxygenase gene family. Plant J. 2003;35:44–56.

    Article  CAS  Google Scholar 

  38. Thompson AJ, Jackson AC, Symonds RC, Mulholland BJ, Dadswell AR, Blake PS, Burbidge A, Taylor IB. Ectopic expression of a tomato 9-cis-epoxycarotenoid dioxygenase gene causes over-production of abscisic acid. Plant J. 2000;23(3):363–74.

    Article  CAS  Google Scholar 

  39. Schwartz SH, Qin X, Zeevaart JAD. Characterization of a novel carotenoid cleavage dioxygenase from plants. J Biol Chem. 2001;276(27):25208–11.

    Article  CAS  Google Scholar 

  40. Iuchi S, Kobayashi M, Taji T, Naramoto M, Seki M, Kato T, Tabata S, Kakubari Y, Yamaguchi-Shinozaki K, Shinozaki K. Regulation of drought tolerance by gene manipulation of 9-cis-epoxycarotenoid dioxygenase, a key enzyme in abscisic acid biosynthesis in Arabidopsis. Plant J. 2001;24(4):325–33.

    Article  Google Scholar 

  41. Chang C, Stadler R. Ethylene hormone receptor action in Arabidopsis. BioEssays. 2001;23:619–27.

    Article  CAS  Google Scholar 

  42. Chang C, Kwok SF, Bleecker AB, Meyerowitz EM. Arabidopsis ethylene-response gene ETR1: similarity of product to two-component regulators. Science. 1993;262(22):539–44.

    Article  CAS  Google Scholar 

  43. Sakai H, Hua J, Chen QG, Chang C, Medrano LJ, Bleecker AB, Meyerowitz EM. ETR2 is an ETR1-like gene involved in ethylene signaling in Arabidopsis. PNAS. 1998;95:5812–7.

    Article  CAS  Google Scholar 

  44. O'Malley RC, Rodriguez FI, Esch JJ, Binder BM, O'Donnell P, Klee HJ, Bleecker AB. Ethylene-binding activity, gene expression levels, and receptor system output for ethylene receptor family members from Arabidopsis and tomato. Plant J. 2005;41:651–9.

    Article  CAS  Google Scholar 

  45. Ulmasov T, Hagen G, Guilfoyle TJ. Activation and repression of transcription by auxin-response factors. PNAS. 1999;96:5844–9.

    Article  CAS  Google Scholar 

  46. Ellis CM, Nagpal P, Young JC, Hagen G, Guilfoyle T, Reed JW. AUXIN RESPONSW FACTOR1 and AUXIN RESPONSE FACTOR2 regulate senescence and floral organ abscission in Arabidopsis thaliana. Development. 2005;132:4563–74.

    Article  CAS  Google Scholar 

  47. Nagpal P, Ellis CM, Weber H, Ploense SE, Barkawi LS, Guilfoyle TJ, Hagen G, Alonso JM, Cohen JD, Farmer EE, Ecker JR, Reed JW. Auxin response factors ARF6 and ARF8 promote jasmonic acid production and flower maturation. Development. 2005;132:4107–18.

    Article  CAS  Google Scholar 

  48. Fedoroff NV. Cross-talk in abscisic acid signaling. Sci Signal. 2002;140:10.

    Article  Google Scholar 

  49. Mizuno T, Yamashino T. Comparative transcriptome of diurnally oscillating genes and hormone-responsive genes in Arabidopsis thaliana: insight into circadian clock-controlled daily responses to common ambient stresses in plants. Plant Cell Physiol. 2008;49(3):481–7.

    Article  CAS  Google Scholar 

  50. Robertson FC, Skeffington AW, Gardner MJ, Webb AA. Interactions between circadian and hormonal signalling in plants. Plant Mol Biol. 2009;69:419–27.

    Article  CAS  Google Scholar 

  51. Edwards KD, Akman OE, Knox K, Lumsden PJ, Thomson AW, Brown PE, Pokhilko A, Kozma-Bognar L, Nagy F, Rand DA, Millar AJ. Quantitative analysis of regulatory flexibility under changing environmental conditions. Mol Syst Biol. 2010;6:424.

    Article  Google Scholar 

  52. Umezawa T. Systems biology approaches to abscisic acid signaling. J Plant Res. 2011;124:539–48.

    Article  CAS  Google Scholar 

  53. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, Adiconis X, Fan L, Raychowdhury R, Zeng Q, Chen Z, Mauceli E, Hacohen N, Gnirke A, Rhind N, di Palma F, Birren BW, Nusbaum C, Lindblad-Toh K, Friedman N, Regev A. Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nat Biotechnol. 2011;29(7):644–52.

    Article  CAS  Google Scholar 

  54. Haas BJ, Papanicolaou A, Yassour M, Grabherr M, Blood PD, Bowden J, Couger MB, Eccles D, Li B, Lieber M, MacManes MD, Ott M, Orvis J, Pochet N, Strozzi F, Weeks N, Westerman R, Wiliam T, Dewey CN, Henschel R, LeDuc RD, Friedman N, Friedman N, Regev A. De novo transcript sequence reconstruction from RNA-seq: reference generation and analysis with trinity. Nat Protoc. 2013;8(8):1494–521.

    Article  CAS  Google Scholar 

  55. Guerriero G, Legay S, Hausman J. Alfalfa cellulose synthase gene expression under abiotic stress: a Hitchhiker’s guide to RT-qPCR normalization. PLoS One. 2014;9(8):e103808.

    Article  Google Scholar 

  56. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2 - CT method. Methods. 2001;25:402–8.

    Article  CAS  Google Scholar 

Download references

Acknowledgments

We thank Gene Denovo Biotechnology Co. (Guangzhou, China) for providing technical support.

Funding

This study was supported by the National Natural Science Foundation of China (31572461). The funding body had no role in the design of the study, collection, analysis, or interpretation of data or in the writing of the manuscript.

Availability of data and materials

Our transcriptome datasets are available at NCBI project PRJNA472434 with accession number SRP148624, and SRA with accession number SRR7193615, SRR7193614, SRR7193613, SRR7193612, SRR7193611, SRR7193610.

Author information

Authors and Affiliations

Authors

Contributions

YSJ and SQB designed the experiments and revised the manuscript. QMC, LYL performed the experiments. QMC, CDZ wrote the manuscript. GTG and PL carried out the data analysis. All authors reviewed, considered and approved the manuscript.

Corresponding authors

Correspondence to Chengdong Zhang or Yushan Jia.

Ethics declarations

Ethics approval and consent to participate

The two alfalfa cultivars in this study were collected from Institute of Grassland Research of CAAS, located in Hohhot, Inner Mongolia, North China (latitude 40°48’N, longitude 111°42′E, elevation 1070 m). No specific permit was required for the field collection of alfalfa cultivars and the location is not privately-owned or protected in any way.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Additional files

Additional file 1:

Table S1. DEGs generated from two varieties of alfalfa leaves. (DOCX 144 kb)

Additional file 2:

Table S2. GO functional annotations and the number of DEGs statistics. (DOCX 18 kb)

Additional file 3:

Table S3. KEGG pathway annotation of DEGs between two varieties of alfalfa leaves. (DOCX 26 kb)

Additional file 4:

Table S4. Genes for RT-qPCR analysis and the list of GO ID. (DOCX 14 kb)

Additional file 5:

Table S5. Primers used for RT-qPCR analysis. (DOCX 12 kb)

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cheng, Q., Bai, S., Ge, G. et al. Study on differentially expressed genes related to defoliation traits in two alfalfa varieties based on RNA-Seq. BMC Genomics 19, 807 (2018). https://doi.org/10.1186/s12864-018-5180-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-018-5180-1

Keywords