Next Article in Journal
Genetic Structure of Racing Pigeons (Columba livia) Kept in Poland Based on Microsatellite Markers
Previous Article in Journal
Genome-Wide Evolutionary Analysis of Putative Non-Specific Herbicide Resistance Genes and Compilation of Core Promoters between Monocots and Dicots
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Transcriptome Profiling of a Common Mistletoe Species Parasitizing Four Typical Host Species in Urban Southwest China

1
Key Laboratory of Bio-Resource and Eco-Environment of Ministry of Education, College of Life Sciences, Sichuan University, Chengdu 610064, China
2
Institute of Molecular Plant Sciences, The University of Edinburgh, Edinburgh EH9 3JH, UK
3
Institute for Advanced Study, Chengdu University, Chengdu 610064, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Genes 2022, 13(7), 1173; https://doi.org/10.3390/genes13071173
Submission received: 29 March 2022 / Revised: 22 June 2022 / Accepted: 24 June 2022 / Published: 29 June 2022
(This article belongs to the Section Plant Genetics and Genomics)

Abstract

:
Comparing gene expressions among parasitic plants infecting different host species can have significant implications for understanding host–parasite interactions. Taxillus nigrans is a common hemiparasitic species in Southwest China that parasitizes a variety of host species. However, a lack of nucleotide sequence data to date has hindered transcriptome-level research on T. nigrans. In this study, the transcriptomes of T. nigrans individuals parasitizing four typical host species (Broussonetia papyrifera (Bpap), a broad-leaved tree species; Cryptomeria fortunei (Cfor), a coniferous tree species; Cinnamomum septentrionale (Csep), an evergreen tree species; and Ginkgo biloba (Gbil), a deciduous-coniferous tree species) were sequenced, and the expression profiles and metabolic pathways were compared among hosts. A total of greater than 400 million reads were generated in nine cDNA libraries. These were de novo assembled into 293823 transcripts with an N50 value of 1790 bp. A large number of differentially expressed genes (DEGs) were identified when comparing T. nigrans individuals on different host species: Bpap vs. Cfor (1253 DEGs), Bpap vs. Csep (864), Bpap vs. Gbil (517), Cfor vs. Csep (259), Cfor vs. Gbil (95), and Csep vs. Gbil (40). Four hundred and fifteen unigenes were common to all six pairwise comparisons; these were primarily associated with Cytochrome P450 and environmental adaptation, as determined in a KEGG enrichment analysis. Unique unigenes were also identified, specific to Bpap vs. Cfor (808 unigenes), Bpap vs. Csep (329 unigenes), Bpap vs. Gbil (87 unigenes), Cfor vs. Csep (108 unigenes), Cfor vs. Gbil (32 unigenes), and Csep vs. Gbil comparisons (23 unigenes); partial unigenes were associated with the metabolism of terpenoids and polyketides regarding plant hormone signal transduction. Weighted gene co-expression network analysis (WGCNA) revealed four modules that were associated with the hosts. These results provide a foundation for further exploration of the detailed molecular mechanisms involved in plant parasitism.

1. Introduction

Parasitic plants are a diverse group of 4750 species that obtain water, mineral nutrients, and carbon from other plants using a specialized feeding organ called a haustorium; within the angiosperms, parasitism has evolved independently at least twelve times [1,2]. The majority of parasitic plants are hemiparasites, which feed directly on other plants but also maintain green leaves and photosynthesize [3]. Many host characteristics interact to determine parasite performance, including nitrogen content [4], carbon content [5], the presence of secondary compounds [6,7]), host condition [8], defenses and immunity [9], biomass [10], and genotype [11]. This complexity has impeded research into hemiparasite host range evolution, particularly as many of the host variables are confounded (depending on the host species).
The ecological significance of parasitic plants depends on the host plant’s preferences and specificity; most mistletoe species have a wide host range and may attach to a diversity of host plants belonging to co-occurring plant species. For instance, Parentucellia viscosa, Rhinanthus minor, and Triphysaria versicolor are known to parasitize 27, 50, and 25 different plant species, respectively [12,13]. However, among all possible host species, each mistletoe species may have a preferred host(s), for which it shows higher specificity than other host species [14]. Santalum acuminatum performs better (as measured by its biomass, percent cover, and haustoria abundance) when parasitizing nitrogen-fixing woody plants [15]; the haustoria of Thesium chinense are also larger when parasitizing the Fabaceae species as compared to other host species [16]. This enhanced performance on leguminous hosts appears to result from a combination of less effective anti-parasitic defenses and the availability of sufficient amounts of easily absorbed nitrogen-containing compounds within the host, rather than being a direct consequence of nitrogen fixation [17]. In addition, germination stimulants also affect parasite host selectivity [6,18].
The Loranthaceae family comprises approximately 70 genera and 950 species, all of which are hemiparasites [19]. Within the family, Taxillus nigrans is a hemiparasitic species that retains the ability to photosynthesize while obtaining water, carbon, and nutrients from host plants via the haustoria [19,20]. T. nigrans can only be propagated by the seeds. It produces fleshy berries that are eaten by birds, promoting seed dispersal; seeds rapidly germinate when deposited on a suitable host, forming new haustoria and drilling into the host’s cortex [21]. The semiparasitism seen in T. nigrans is similar to that of other Loranthaceae species, such as Agelanthus natalitius and Struthanthus aff. polyanthus [19,22]. T. nigrans can parasitize 51 host species [23], causing the death of branches or the entire plant of the host. However, very little is known about the mechanisms driving host species preference.
With the advent of next-generation sequencing (NGS), the de novo assembly and analysis of parasitic plant transcriptomes have offered important insights into gene expression differences related to host species identity [24]. Here, RNA-seq was used to generate transcriptome profiles for T. nigrans individuals parasitizing four different host species (Broussonetia papyrifera, Cinnamomum septentrionale, Cryptomeria fortune, and Ginkgo biloba) to reveal underlying molecular differences. A fully annotated transcriptome was obtained and used as a reference to examine expression differences associated with each host species. These results will be helpful in understanding gene expression in the mistletoe species colonizing different host species.

2. Materials and Methods

2.1. Sample Collection

In July 2016, fresh T. nigrans were collected from four different host species (Figure 1), namely, Broussonetia papyrifera (Bpap), Cinnamomum septentrionale (Csep), Cryptomeria fortunei (Cfor), and Ginkgo biloba (Gbil), on the Wangjiang Campus of Sichuan University, Chengdu, Sichuan, China. T. nigrans has 41 known host species [21]; the host species selected for this study included two deciduous broad-leaved species (Gbil and Bpap), a deciduous coniferous species (Cfor), and an evergreen broad-leaved species (Csep). For the leaf samples, three biological replicates were collected per host species. Samples were immediately frozen in liquid nitrogen and then stored in a −80 C refrigerator prior to RNA extraction. All samples were used for RNA extraction and RNA-Seq library construction.

2.2. RNA-Seq Library Preparation and Sequencing

Total RNA was extracted from collected leaves using Guanidine thiocyanate (Sigma, 50983, Beijing, China)-Chloroform (Sigma, 472476, Beijing, China) according to the manufacturer’s instructions. For each host species, two hundred milligrams of leaf tissue was ground in liquid nitrogen to extract total RNA. The total RNA samples were then treated with a DNA-free TM DNA Removal Kit (Ambion, AM1906, Shanghai, China) to remove contaminating genomic DNA. RNA purity was checked using a Nano-Photometer spectrophotometer (IMPLEN, Westlake Village, CA, USA). Before cDNA synthesis, RNA concentrations were measured using a Qubit RNA Assay Kit (Life Technologies, Q32852, Shanghai, China), and RNA integrity was assessed using the RNA Nano 6000 Assay Kit for the Agilent Bioanalyzer 2100 system (Agilent Technologies, Santa Clara, CA, USA).
A total of 3 μ g of RNA per sample was used as input material for sample preparation. Twelve sequencing libraries (i.e., three replicates per host species) were generated using the NEBNext ® Ultra TM RNA Library Prep Kit for Illumina (NEB, Beijing, China) following the provided protocol. Clustering of the index-coded samples was then performed with the cBot Cluster Generation System using a TruSeq PE Cluster Kit v3-cBot-HS (Illumina, Shanghai, China) according to the manufacturer’s instructions. After cluster generation, the library preparations were sequenced on an Illumina HiSeq 2000 platform by Novogene Bioinformatics Technology Co., Ltd. (Beijing, China), and 150-bp paired-end reads were generated.
A total of 3 μ g of RNA per sample was used as input material for the sample preparation. Twelve sequencing libraries (i.e., three replicates per host species) were generated using the NEBNext ® Ultra TM RNA Library Prep Kit for Illumina (NEB, Beijing, China) following the provided protocol. Clustering of the index-coded samples was then performed with the cBot Cluster Generation System using a TruSeq PE Cluster Kit v3-cBot-HS (Illumina, Beijing, China) according to the manufacturer’s instructions. After cluster generation, the library preparations were sequenced on an Illumina HiSeq 2000 platform by Novogene Bioinformatics Technology Co., Ltd. (Beijing, China), and 150-bp paired-end reads were generated.

2.3. Preprocessing of Illumina Reads and De Novo Transcriptome Assembly

Raw reads were filtered using Trimmomatic [25] to obtain high-quality, clean reads for assembly: adapter sequences, reads containing poly-N runs (≥10%), and low quality (sQ ≤ 5) reads were removed. The number of sequence duplications, GC content, Q20, and Q30 was then calculated for the clean reads. All downstream analyses were based on clean data. The clean data of all samples were uploaded to the Sequence Read Archive (SRA) with accession number PRJNA851568. De novo assembly of the transcriptome from the RNA-seq data was performed using the Trinity software package [26]), with min_kmer_cov set to two and default values for other parameters. Benchmarking Universal Single-Copy Orthologues (BUSCO) provides measures for quantitative assessment of transcriptome completeness based on evolutionarily informed expectations of gene content from conserved single-copy orthologs [27]. In this study, we evaluated the completeness of the assembly using BUSCO based on the embryophyta_odb10 database.

2.4. Functional Annotation of the Transcriptome

To annotate the assembled T. nigrans transcriptome, transcripts were aligned against the eggNOG Nr, and Swiss-Prot databases using BLASTP with a significance threshold of E ≤ 10 5 . For functional categorization, the Gene Ontology (GO) and Kyoto Encyclopedia of Genes and Genomes (KEGG) pathways were analyzed and annotated using the TBtools [28].

2.5. Differential Expression Analysis and Co-Expression Network Analysis

To estimate the abundance of the de novo assembled transcripts, RSEM [29] was used; this protocol assesses transcript abundance based on the mapping of RNA-seq reads to the assembled transcriptome. The DESeq2 and edgeR in IDEAMEX [30] were used to identify differentially expressed genes (DEGs) among the host species. To minimize the false positive rate, only transcripts with threshold p values < 0.05 and absolute values of log 2 ( f o l d c h a n g e ) > 1 (as screened by IDEAMEX) were considered differentially expressed. The differentially expressed transcripts were divided into six sets, each comparing two host species, as follows: set I (Bpap vs. Cfor), set II (Bpap vs. Csep), set III (Bpap vs. Gbil), set IV (Cfor vs. Csep), set V (Cfor vs. Gbil), and set VI (Csep vs. Gbil). Pearson correlation coefficients were then calculated for all sets using gene expression data. A GO term enrichment analysis and KEGG pathways were performed for the DEGs using TBtools. We performed a co-expression network analysis using the WGCNA package [31] based on log 2 ( F P K M + 1 ) values.

3. Results

3.1. Transcriptome Sequencing and De Novo Assembly of T. nigrans Transcriptome

Fresh leaves of T. nigrans collected from four host species were used to construct cDNA libraries, which were subsequently sequenced on an Illumina HiSeq 4000 platform. For each host species, three biological replicates were analyzed. After filtering the raw sequencing data, a total of 5,553,239,700 (Bpap-001), 6,299,403,300 (Bpap-002), 4,693,123,881 (Bpap-003), 4,183,362,900 (Cfor-001), 5,784,479,700 (Cfor-002), 5,156,214,900 (Cfor-003), 4,763,816,700 (Csep-001), 4,571,633,850 (Csep-002), 5,665,991,700 (Csep-003), 5,116,978,800 (Gbil-001), 4,066,401,900 (Gbil-002), and 4,932,610,200 (Gbil-003) clean bases were obtained (Table 1). The de novo assembly yielded 332,439 transcripts, with an N50 size of 1790 bp (Table 2). The BUSCO analysis showed 92.3% of 1614 single-copy genes in the embryophyta_odb10 database (Table 2). The results of the functional annotations of all transcripts are shown in Table S1.

3.2. Analysis of Differentially Expressed Genes (DEGs) among Different Hosts

To identify DEGs among the host species, six pairwise comparisons were performed (Bpap vs. Cfor, Bpap vs. Csep, Bpap vs. Gbil, Cfor vs. Csep, Cfor vs. Gbil, and Csep vs. Gbil) using DESeq2 and edgeR in IDEAMEX. After combining the two methods (DESeq2 and edgeR), a total of 1526, 1093, 617, 352, 133, and 57 DEGs were detected for each of these comparisons, respectively (Figure 2A). Among the DEGs, a total of 1253, 864, 517, 259, 95, and 40 DEGs were functionally annotated, respectively, for each comparison (Tables S2–S4). The principal components analysis (PCA) analysis of each comparison was shown in Figure 3 based on the IDEAMEX results.
After identifying the DEGs for each pairwise comparison, the subset of DEGs common to all samples was then determined. A total of 415 common unigenes were identified based on DESeq2 and degeR; these are presented in a Venn diagram and volcano map (Figure 2B,D). Partially common unigenes encoding putative transcription factors, including members of the cytochrome P450 family, HSP, ABC transporter family, UDP-glycosyltransferase family, and zinc finger families (Table S1). In addition, unique DEGs were identified for the Bpap vs. Cfor comparison (n = 808), Bpap vs. Csep comparison (n = 329), Bpap vs. Gbil comparison (n = 87), Cfor vs. Csep comparison (n = 108), Cfor vs. Gbil comparison (n = 32), and Csep vs. Gbil comparison (n = 23), respectively (Figure 2A).

3.3. GO and KEGG Enrichment Analysis

To further explore differential gene expression in T. nigrans, the DEGs (from all comparisons) were used for GO and KEGG enrichment analyses. For the shared DEGs (415 across all comparisons), the GO enrichment analysis showed that a majority of DEGs belonged to the following categories: “UDP-glucosyltransferase activity” (GO:0035251), “transcription regulator activity” (GO:0140110), “ubiquitin-protein transferase activity” (GO:0004842), “chloroplast” (GO:0009507), “plant-type vacuole” (GO:0000325), “plasmodesma” (GO:0009506), “response to abiotic stimulus” (GO:0009628), “response to abscisic acid” (GO:0009737), “rhythmic process” (GO:0048511), “terpenoid metabolic process” (GO:0006721) (Figure 2C and Table S5). The top five pathways identified in the KEGG enrichment analysis were: Circadian rhythm, cytochrome P450, organismal systems, environmental adaptation, and metabolism of terpenoids and polyketides (Table 3 and Table S6). The unique DEGs identified in the six comparisons were subjected to GO and KEGG enrichment analyses. The most highly represented GO terms in Bpap vs. Cfor comparisons were: “phenylpropanoid metabolic process” (GO:0009698), “flavonoid metabolic process” (GO:0009812), “response to acid chemical” (GO:0001101), "fruit development" (GO:0010154), “seed development” (GO:0048316), “terpenoid biosynthetic process” (GO:0016114), “response to alcohol” (GO:0097305), “UDP-glycosyltransferase activity” (GO:0008194), “plant-type cell wall” (GO:0009505), and “long-chain fatty acid metabolic process” (GO:0001676) (Table S7). In the KEGG analysis, the top five pathways were: Flavonoid biosynthesis, stilbenoid/diarylheptanoid/gingerol biosynthesis, phenylpropanoid biosynthesis, metabolism of terpenoids and polyketides, and amino sugar and nucleotide sugar metabolism (Table S8). For specific DEGs in the Bpap vs. Csep comparison, significantly enriched GO terms included “transmembrane signaling receptor activity” (GO:0004888), “inorganic anion transmembrane transporter activity” (GO:0015103), “signaling receptor activity” (GO:0038023), “anion transmembrane transporter activity” (GO:0008509), “transmembrane transporter activity” (GO:0022857), “NADH metabolic process” (GO:0006734), “cellular response to jasmonic acid stimulus” (GO:0071395), “ATP metabolic process” (GO:0046034), “response to fatty acid” (GO:0070542), “ribonucleoside triphosphate metabolic process” (GO:0009199), etc. (Table S9). In the KEGG analysis, the top five enriched pathways were glycolysis/gluconeogenesis, brassinosteroid biosynthesis, plant–pathogen interaction, plant hormone signal transduction, and environmental information processing (Table S10). The most highly represented GO terms in Bpap vs. Gbil comparison were “plasma membrane” (GO:0005886), “cellular response to acid chemical” (GO:0071229), “hormone-mediated signaling pathway” (GO:0009755), “response to acid chemical” (GO:0001101), “response to salt stress” (GO:0009651), “cellular response to hormone stimulus” (GO:0032870), “cellular response to lipid” (GO:0071396), “response to osmotic stress” (GO:0006970), “signal transduction” (GO:0007165), and “cellular response to endogenous stimulus” (GO:0071495) (Table S11). For specific DEGs in Csep vs. Cfor comparison, significantly enriched GO terms included “transcription regulator activity” (GO:0140110), “DNA-binding transcription factor activity” (GO:0003700), “chloroplast” (GO:0009507), “cell junction” (GO:0030054), “cell wall” (GO:0005618), “defense response” (GO:0006952), “regulation of RNA metabolic process” (GO:0051252), “regulation of RNA biosynthetic process” (GO:2001141), “regulation of nucleic acid-templated transcription” (GO:1903506), “response to hormone” (GO:0009725) (Table S12). The top five enriched KEGG pathways for unique Csep vs. Cfor comparison were cutin/suberine and wax biosynthesis, diterpenoid biosynthesis, metabolism of terpenoids and polyketides, plant hormone signal transduction, and carbon fixation in photosynthetic organisms (Table S13). There were no GO and KEGG enrichments found in the Cfor vs. Gbil and Csep vs. Gbil comparisons.

3.4. Identification of DEGs Co-Expression Modules

We identified co-expressed gene sets based on DEGs that appeared at least once in six comparisons through a weighted gene co-expression network analysis (WGCNA). The WGCNA analysis revealed several major co-expression modules, namely the brown module, blue module, turquoise module, and grey module, respectively (Figure 4A). There was no significant correlation between modules and hosts, which may be a small amount of samples. Further, we associated each of the co-expression modules with four samples. The brown module was correlated with Bpap. The blue module was correlated with Cfor and Gbil, and the turquoise module was correlated with Gbil and Csep (Figure 4B). Subsequently, to further explore genes in related modules, the GO enrichment analyses were followed. The blue module associated with Cfor and Gbil showed enrichment of GO terms related to “DNA-binding transcription factor activity” (GO:0003700), “hydrolase activity, hydrolyzing O-glycosyl compounds” (GO:0004553), “oxidoreductase activity” (GO:0016705), “monooxygenase activity” (GO:0004497), “plant-type cell wall” (GO:0009505), “plasmodesma” (GO:0009506), “apoplast” (GO:0048046), “plant-type cell wall organization or biogenesis” (GO:0071669), “xylan biosynthetic process” (GO:0045492), “glucuronoxylan biosynthetic process” (GO:0010417), and “phenylpropanoid metabolic process” (GO:0009698) (Table S14). The Bpap-associated brown modules represented GO terms related to “RNA binding” (GO:0003723), “nucleic acid binding” (GO:0003676), “translation factor activity, RNA binding” (GO:0008135), “calmodulin binding” (GO:0005516), “ribosome” (GO:0005840), “cytosol” (GO:0005829), “cytosolic small ribosomal subunit” (GO:0022627), “translation” (GO:0006412), “peptide metabolic process” (GO:0006518), “organonitrogen compound biosynthetic process” (GO:1901566), and “cellular response to xenobiotic stimulus” (GO:0071466) (Table S15). In a similar way, the turquoise module associated with Gbil and Csep exhibited representation of terms, such as “transmembrane transporter activity” (GO:0022857), “glucosyltransferase activity” (GO:0046527), “water transmembrane transporter activity” (GO:0005372), “UDP-glucosyltransferase activity” (GO:0035251), “plastid” (GO:0009536), “chloroplast” (GO:0009507), “plasmodesma” (GO:0009506), “response to abiotic stimulus” (GO:0009628), “response to acid chemical” (GO:0001101), “response to abscisic acid” (GO:0009737), and “response to radiation” (GO:0009314) (Table S16).

4. Discussion

4.1. Transcriptome Assembly and Annotation

Advances in transcriptome sequencing technology and data mining platforms have led to rapid progress in comparative transcriptomics for non-model, non-crop plants, such as parasitic plants [32,33,34,35,36,37,38]. Here, transcriptomes were generated and compared for T. nigrans plants parasitizing four different host species. High-throughput sequencing generated more than 405 million clean reads from all samples, yielding approximately 248,751 unigenes after assembly. The assembled transcriptome comprised mostly short transcripts, as seen for other de novo assembled plant transcriptomes, perhaps as a result of the assembly algorithm used [39,40]. The contig N50 was 1790 bp, indicating that the Trinity assembly was of high quality. The assembly and annotation of the T. nigrans transcriptome provides a valuable resource for investigating the processes and pathways involved in T. nigrans host adaptation. For example, transcriptomes produced for the parasitic plants Arceuthobium sichuanense and Cuscuta campestris were used to identify DEGs among samples [24,40].

4.2. Gene Categories Associated with Plant Development and Host Selection

For some parasitic plants, such as Arceuthobium sichuanense and Cuscuta campestris, the establishment of vascular connections with the host plant allows the acquisition of nutrients and solutes; this nutrient sink can significantly reduce the biomass of the host [40,41]. In this study, based on the results of GO and KEGG enrichment, a large number of common unigenes (shared across T. nigrans individuals) were primarily related to carbohydrate metabolism, including glycolysis/gluconeogenesis; pyruvate metabolism; glyoxylate and dicarboxylate metabolism; and amino sugar and nucleotide sugar metabolism. Although the transcriptomic data were not directly related to the parasitic properties of T. nigrans, studies of other parasitic plants have shown that metabolism is related to parasitism [40,41,42,43]. In addition, unigenes involved in the energy metabolism in photosynthetic organisms (i.e., oxidative phosphorylation and carbon fixation) were also enriched in T. nigrans. We also found that part of the unigenes related to the circadian rhythm, cytochrome P450, metabolism of terpenoids and polyketides, and response to abscisic acid. Cytochrome P450 may contribute to environmental and host adaptation [44,45]. Abscisic acid is thought to be involved in the development [46,47]. The observed results may thus reflect the adaptive adjustment of T. nigrans to host selection and the environment. Furthermore, unigenes were identified specific to the Bpap vs. Cfor comparison (n = 808), Bpap vs. Csep comparison (n = 329), Bpap vs. Gbil comparison (n = 87), Cfor vs. Csep comparison (n = 108), Cfor vs. Gbil comparison (n = 32), and Csep vs. Gbil comparison (n = 23), most of which were annotated using the Gene Ontology and KEGG databases. In all annotated comparisons, most of the genes were related to the biosynthesis and metabolism of secondary metabolites. Unique genes of each comparison showed similarities in functional enrichment, such as Bpap vs. Cfor and Csep vs. Cfor comparisons involving the metabolism of terpenoids and polyketides; in Bpap vs. Csep and Csep vs. Cfor comparisons, we found genes related to plant hormone signal transduction that may be related to host differences. In addition, partial unigenes were associated with the plasma membrane, response to endogenous stimuli, ion binding, and organic hydroxy compound metabolic processes. Previous studies have shown that these factors are associated with plant defenses [48,49,50]. We identified co-expressed gene sets that were expressed in four samples via the weighted gene co-expression network analysis (WGCNA), a method widely used in the comparative transcriptome analysis [51,52,53,54]. These results indicate that each sample was associated with one or two co-expression modules that reflected the gene regulatory processes specific to each sample. We anticipate that the comprehensive information presented here will serve as a crucial resource to understand host differences and commonalities.

5. Conclusions

In summary, RNA-seq analysis revealed both host-specific and common pathways involved in T. nigrans parasitism; these results provide a foundation for further study of the molecular factors and functions underlying adaptation to different host plants. The putative key genes and pathways identified here may also be used to explore the molecular factors and mechanisms explaining host diversity. The analysis represents a critical step in promoting the development of molecular ecology tools for parasite–host systems and sustainable integrated management programs.

Supplementary Materials

The supplementary file can be accessed at https://doi.org/10.6084/m9.figshare.19446917. Supplementary Table S1: functional annotation of transcripts based on eggNOG. Supplementary Table S2: GO enrichment for specific DEGs of Bpap vs. Csep. Supplementary Table S3: GO enrichment for specific DEGs of Bpap vs. Gbil. Supplementary Table S4: GO enrichment for specific DEGs of Cfor vs. Csep. Supplementary Table S5: GO enrichment for common DEGs. Supplementary Table S6: KEGG pathways for common DEGs. Supplementary Table S7: GO enrichment for unique DEGs in Bpap vs. Cfor comparison. Supplementary Table S8: KEGG enrichment analysis for unique DEGs in Bpap vs. Cfor comparison. Supplementary Table S9: GO enrichment for specific DEGs in Bpap vs. Csep comparison. Supplementary Table S10: KEGG enrichment analysis for specific DEGs in Bpap vs. Csep comparison. Supplementary Table S11: GO enrichment for specific DEGs in Bpap vs. Gbil comparison. Supplementary Table S12: GO enrichment for specific DEGs in Csep vs. Cfor comparison. Supplementary Table S13: KEGG enrichment analysis for specific DEGs in Csep vs. Cfor comparison. Supplementary Table S14: GO enrichment for DEGs in blue module based on WGCNA. Supplementary Table S15: GO enrichment for DEGs in brown module based on WGCNA. Supplementary Table S16: GO enrichment for DEGs in turquoise module based on WGCNA.

Author Contributions

Conceptualization, N.M. and K.M.; methodology, J.K. and Y.W.; software, Y.W.; validation, J.K.; data curation, N.M. and K.M.; writing—original draft preparation, J.K. and Y.W.; writing—review and editing, J.K., R.M., N.M., M.W., and Y.W.; visualization, J.K. and Y.W.; supervision, N.M.; project administration, N.M.; funding acquisition, N.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China, grant number 2016YFD0600203.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

We thank Maoping Li for the assistance with the field work and sample collection.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Westwood, J.H.; Yoder, J.I.; Timko, M.P.; Depamphilis, C.W. The evolution of parasitism in plants. Trends Plant Sci. 2010, 15, 227–235. [Google Scholar] [CrossRef] [PubMed]
  2. Nickrent, D.L. Parasitic angiosperms: How often and how many? Taxon 2020, 69, 5–27. [Google Scholar] [CrossRef]
  3. Twyford, A.D. Parasitic plants. Curr. Biol. 2018, 28, R857–R859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Pennings, S.C.; Simpson, J.C. Like herbivores, parasitic plants are limited by host nitrogen content. Plant Ecol. 2008, 196, 245–250. [Google Scholar] [CrossRef]
  5. Těšitel, J.; Těšitelová, T.; Fisher, J.P.; Lepš, J.; Cameron, D.D. Integrating ecology and physiology of root-hemiparasitic interaction: Interactive effects of abiotic resources shape the interplay between parasitism and autotrophy. New Phytol. 2015, 205, 350–360. [Google Scholar] [CrossRef] [PubMed]
  6. Cui, S.; Wakatake, T.; Hashimoto, K.; Saucet, S.B.; Toyooka, K.; Yoshida, S.; Shirasu, K. Haustorial hairs are specialized root hairs that support parasitism in the facultative parasitic plant Phtheirospermum japonicum. Plant Physiol. 2016, 170, 1492–1503. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Muche, M.; Muasya, A.M.; Tsegay, B.A. Biology and resource acquisition of mistletoes, and the defense responses of host plants. Ecol. Process. 2022, 11, 24. [Google Scholar] [CrossRef]
  8. Bickford, C.P.; Kolb, T.E.; Geils, B.W. Host physiological condition regulates parasitic plant performance: Arceuthobium vaginatum subsp. cryptopodum on Pinus ponderosa. Oecologia 2005, 146, 179–189. [Google Scholar] [CrossRef]
  9. Irving, L.J.; Cameron, D.D. You are what you eat: Interactions between root parasitic plants and their hosts. Adv. Bot. Res. 2009, 50, 87–138. [Google Scholar]
  10. Matthies, D. Interactions between a root hemiparasite and 27 different hosts: Growth, biomass allocation and plant architecture. Perspect. Plant Ecol. Evol. Syst. 2017, 24, 118–137. [Google Scholar] [CrossRef] [Green Version]
  11. Rowntree, J.K.; Cameron, D.D.; Preziosi, R.F. Genetic variation changes the interactions between the parasitic plant-ecosystem engineer Rhinanthus and its hosts. Philos. Trans. R. Soc. B Biol. Sci. 2011, 366, 1380–1388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Gibson, C.; Watkinson, A. Host selectivity and the mediation of competition by the root hemiparasite Rhinanthus minor. Oecologia 1991, 86, 81–87. [Google Scholar] [CrossRef] [PubMed]
  13. Matvienko, M.; Torres, M.J.; Yoder, J.I. Transcriptional responses in the hemiparasitic plant Triphysaria versicolor to host plant signals. Plant Physiol. 2001, 127, 272–282. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Norton, D.; De Lange, P. Host specificity in parasitic mistletoes (Loranthaceae) in New Zealand. Funct. Ecol. 1999, 13, 552–559. [Google Scholar] [CrossRef]
  15. Tennakoon, K.; Pate, J.; Arthur, D. Ecophysiological aspects of the woody root hemiparasiteSantalum acuminatum (R. Br.) A. DC and its common hosts in south western Australia. Ann. Bot. 1997, 80, 245–256. [Google Scholar] [CrossRef] [Green Version]
  16. Suetsugu, K.; Kawakita, A.; Kato, M. Host range and selectivity of the hemiparasitic plant Thesium chinense (Santalaceae). Ann. Bot. 2008, 102, 49–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Jiang, F.; Jeschke, W.D.; Hartung, W.; Cameron, D.D. Does legume nitrogen fixation underpin host quality for the hemiparasitic plant Rhinanthus minor? J. Exp. Bot. 2008, 59, 917–925. [Google Scholar] [CrossRef] [Green Version]
  18. Ishida, J.K.; Wakatake, T.; Yoshida, S.; Takebayashi, Y.; Kasahara, H.; Wafula, E.; Depamphilis, C.W.; Namba, S.; Shirasu, K. Local auxin biosynthesis mediated by a YUCCA flavin monooxygenase regulates haustorium development in the parasitic plant Phtheirospermum japonicum. Plant Cell 2016, 28, 1795–1814. [Google Scholar] [CrossRef] [Green Version]
  19. Arruda, R.; Carvalho, L.N.; Del-Claro, K. Host specificity of a Brazilian mistletoe, Struthanthus aff. polyanthus (Loranthaceae), in cerrado tropical savanna. Flora-Morphol. Distrib. Funct. Ecol. Plants 2006, 201, 127–134. [Google Scholar] [CrossRef]
  20. Shen, H.; Ye, W.; Hong, L.; Huang, H.; Wang, Z.; Deng, X.; Yang, Q.; Xu, Z. Progress in parasitic plant biology: Host selection and nutrient transfer. Plant Biol. 2006, 8, 175–185. [Google Scholar] [CrossRef]
  21. Ma, R.; Miao, N. Ingestion Species and Ingestion Intensity of Birds on Taxillus nigrans in the Urban of Chengdu Plains. J. Sichuan For. Sci. Technol. 2018, 39, 59–62. [Google Scholar]
  22. Okubamichael, D.; Rasheed, M.; Griffiths, M.; Ward, D. Avian consumption and seed germination of the hemiparasitic mistletoe Agelanthus natalitius (Loranthaceae). J. Ornithol. 2011, 152, 643–649. [Google Scholar] [CrossRef]
  23. Ma, R.; Miao, N.; Zhang, H.; Tao, W.; Mao, K.; Moermond, T.C. Generalist mistletoes and their hosts and potential hosts in an urban area in southwest China. Urban For. Urban Green. 2020, 53, 126717. [Google Scholar] [CrossRef]
  24. Wang, Q.; Liu, J.; Allen, G.A.; Ma, Y.; Yue, W.; Marr, K.L.; Abbott, R.J. Arctic plant origins and early formation of circumarctic distributions: A case study of the mountain sorrel, Oxyria digyna. New Phytol. 2016, 209, 343–353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. 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]
  26. Grabherr, M.G.; Haas, B.J.; Yassour, M.; Levin, J.Z.; Thompson, D.A.; Amit, I.; Adiconis, X.; Fan, L.; Raychowdhury, R.; Zeng, Q.; et al. Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nat. Biotechnol. 2011, 29, 644–652. [Google Scholar] [CrossRef] [Green Version]
  27. Simão, F.A.; Waterhouse, R.M.; Ioannidis, P.; Kriventseva, E.V.; Zdobnov, E.M. BUSCO: Assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics 2015, 31, 3210–3212. [Google Scholar] [CrossRef] [Green Version]
  28. Chen, C.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.; Xia, R. TBtools: An integrative toolkit developed for interactive analyses of big biological data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef]
  29. Li, B.; Dewey, C.N. RSEM: Accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinform. 2011, 12, 323. [Google Scholar] [CrossRef] [Green Version]
  30. Jiménez-Jacinto, V.; Sanchez-Flores, A.; Vega-Alvarado, L. Integrative differential expression analysis for multiple experiments (IDEAMEX): A web server tool for integrated rna-seq data analysis. Front. Genet. 2019, 10, 279. [Google Scholar] [CrossRef]
  31. Langfelder, P.; Horvath, S. WGCNA: An R package for weighted correlation network analysis. BMC Bioinform. 2008, 9, 559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Honaas, L.A.; Wafula, E.K.; Yang, Z.; Der, J.P.; Wickett, N.J.; Altman, N.S.; Taylor, C.G.; Yoder, J.I.; Timko, M.P.; Westwood, J.H.; et al. Functional genomics of a generalist parasitic plant: Laser microdissection of host-parasite interface reveals host-specific patterns of parasite gene expression. BMC Plant Biol. 2013, 13, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Ko, S.M.; Kwon, Y.K.; Kim, J.H.; Song, I.J.; Lee, H.Y.; Choi, D.W.; Liu, J.R.; Kim, S.W. Transcriptome analysis of mistletoe (Viscum album) haustorium development. Hortic. Environ. Biotechnol. 2014, 55, 352–361. [Google Scholar] [CrossRef]
  34. Westwood, J.H.; Depamphilis, C.W.; Das, M.; Fernández-Aparicio, M.; Honaas, L.A.; Timko, M.P.; Wafula, E.K.; Wickett, N.J.; Yoder, J.I. The parasitic plant genome project: New tools for understanding the biology of Orobanche and Striga. Weed Sci. 2012, 60, 295–306. [Google Scholar] [CrossRef]
  35. Wickett, N.J.; Honaas, L.A.; Wafula, E.K.; Das, M.; Huang, K.; Wu, B.; Landherr, L.; Timko, M.P.; Yoder, J.; Westwood, J.H.; et al. Transcriptomes of the parasitic plant family Orobanchaceae reveal surprising conservation of chlorophyll synthesis. Curr. Biol. 2011, 21, 2098–2104. [Google Scholar] [CrossRef] [Green Version]
  36. Yang, Z.; Wafula, E.K.; Honaas, L.A.; Zhang, H.; Das, M.; Fernandez-Aparicio, M.; Huang, K.; Bandaranayake, P.C.; Wu, B.; Der, J.P.; et al. Comparative transcriptome analyses reveal core parasitism genes and suggest gene duplication and repurposing as sources of structural novelty. Mol. Biol. Evol. 2015, 32, 767–790. [Google Scholar] [CrossRef] [Green Version]
  37. Yoshida, S.; Kee, Y.J. Large-scale sequencing paves the way for genomic and genetic analyses in parasitic plants. Curr. Opin. Biotechnol. 2021, 70, 248–254. [Google Scholar] [CrossRef] [PubMed]
  38. Gatica-Soria, L.M.; Ceriotti, L.F.; Garcia, L.E.; Sanchez-Puerta, M.V. Native and foreign mitochondrial and nuclear encoded proteins conform the OXPHOS complexes of a holoparasitic plant. Gene 2022, 817, 146176. [Google Scholar] [CrossRef]
  39. Gruenheit, N.; Deusch, O.; Esser, C.; Becker, M.; Voelckel, C.; Lockhart, P. Cutoffs and k-mers: Implications from a transcriptome study in allopolyploid plants. BMC Genom. 2012, 13, 92. [Google Scholar] [CrossRef] [Green Version]
  40. Ranjan, A.; Ichihashi, Y.; Farhi, M.; Zumstein, K.; Townsley, B.; David-Schwartz, R.; Sinha, N.R. De novo assembly and characterization of the transcriptome of the parasitic weed dodder identifies genes associated with plant parasitism. Plant Physiol. 2014, 166, 1186–1199. [Google Scholar] [CrossRef] [Green Version]
  41. Zhang, D.; Ren, L.; Yue, J.h.; Shi, Y.b.; Zhuo, L.h.; Wang, L.; Shen, X.h. RNA-Seq-based transcriptome analysis of stem development and dwarfing regulation in Agapanthus praecox ssp. orientalis (Leighton) Leighton. Gene 2015, 565, 252–267. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, Y.; Li, X.; Zhou, W.; Li, T.; Tian, C. De novo assembly and transcriptome characterization of spruce dwarf mistletoe Arceuthobium sichuanense uncovers gene expression profiling associated with plant development. BMC Genom. 2016, 17, 771. [Google Scholar] [CrossRef] [Green Version]
  43. Ichihashi, Y.; Kusano, M.; Kobayashi, M.; Suetsugu, K.; Yoshida, S.; Wakatake, T.; Kumaishi, K.; Shibata, A.; Saito, K.; Shirasu, K. Transcriptomic and metabolomic reprogramming from roots to haustoria in the parasitic plant, Thesium chinense. Plant Cell Physiol. 2018, 59, 729–738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Wakabayashi, T.; Hamana, M.; Mori, A.; Akiyama, R.; Ueno, K.; Osakabe, K.; Osakabe, Y.; Suzuki, H.; Takikawa, H.; Mizutani, M.; et al. Direct conversion of carlactonoic acid to orobanchol by cytochrome P450 CYP722C in strigolactone biosynthesis. Sci. Adv. 2019, 5, eaax9067. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Yoda, A.; Mori, N.; Akiyama, K.; Kikuchi, M.; Xie, X.; Miura, K.; Yoneyama, K.; Sato-Izawa, K.; Yamaguchi, S.; Yoneyama, K.; et al. Strigolactone biosynthesis catalyzed by cytochrome P450 and sulfotransferase in sorghum. New Phytol. 2021, 232, 1999–2010. [Google Scholar] [CrossRef]
  46. Mallu, T.S.; Irafasha, G.; Mutinda, S.; Owuor, E.; Githiri, S.M.; Odeny, D.A.; Runo, S. Mechanisms of pre-attachment Striga resistance in sorghum through genome-wide association studies. Mol. Genet. Genom. 2022, 297, 751–762. [Google Scholar] [CrossRef]
  47. Brun, G.; Thoiron, S.; Braem, L.; Pouvreau, J.B.; Montiel, G.; Lechat, M.M.; Simier, P.; Gevaert, K.; Goormachtig, S.; Delavault, P. CYP707As are effectors of karrikin and strigolactone signalling pathways in Arabidopsis thaliana and parasitic plants. Plant Cell Environ. 2019, 42, 2612–2626. [Google Scholar] [CrossRef]
  48. Dmitriev, A.A.; Krasnov, G.S.; Rozhmina, T.A.; Kishlyan, N.V.; Zyablitsin, A.V.; Sadritdinova, A.F.; Snezhkina, A.V.; Fedorova, M.S.; Yurkevich, O.Y.; Muravenko, O.V.; et al. Glutathione S-transferases and UDP-glycosyltransferases are involved in response to aluminum stress in flax. Front. Plant Sci. 2016, 7, 1920. [Google Scholar] [CrossRef] [Green Version]
  49. Gill, S.S.; Tuteja, N. Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants. Plant Physiol. Biochem. 2010, 48, 909–930. [Google Scholar] [CrossRef]
  50. Poppenberger, B.; Berthiller, F.; Lucyshyn, D.; Sieberer, T.; Schuhmacher, R.; Krska, R.; Kuchler, K.; Glossl, J.; Luschnig, C.; Adam, G. Detoxification of the Fusarium mycotoxin deoxynivalenol by a UDP-glucosyltransferase from Arabidopsis thaliana. J. Biol. Chem. 2003, 278, 47905–47914. [Google Scholar] [CrossRef] [Green Version]
  51. Garg, R.; Singh, V.K.; Rajkumar, M.S.; Kumar, V.; Jain, M. Global transcriptome and coexpression network analyses reveal cultivar-specific molecular signatures associated with seed development and seed size/weight determination in chickpea. Plant J. 2017, 91, 1088–1107. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Ali, M.; Cheng, H.; Soomro, M.; Shuyan, L.; Bilal Tufail, M.; Nazir, M.F.; Feng, X.; Zhang, Y.; Dongyun, Z.; Limin, L.; et al. Comparative transcriptomic analysis to identify the genes related to delayed gland morphogenesis in Gossypium bickii. Genes 2020, 11, 472. [Google Scholar] [CrossRef] [PubMed]
  53. Yuan, Y.; Zhang, B.; Tang, X.; Zhang, J.; Lin, J. Comparative transcriptome analysis of different dendrobium species reveals active ingredients-related genes and pathways. Int. J. Mol. Sci. 2020, 21, 861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Barreto Sánchez, A.L.; Wang, Q.; Thiam, M.; Wang, Z.; Zhang, J.; Zhang, Q.; Zhang, N.; Li, Q.; Wen, J.; Zhao, G. Liver Transcriptome Response to Heat Stress in Beijing You Chickens and Guang Ming Broilers. Genes 2022, 13, 416. [Google Scholar] [CrossRef]
Figure 1. The host plants were ’parasited’ by T. nigrans. (A) Broussonetia papyrifera. (B) Cryptomeria fortunei. (C) Cinnamomum septentrionale. (D) Ginkgo biloba.
Figure 1. The host plants were ’parasited’ by T. nigrans. (A) Broussonetia papyrifera. (B) Cryptomeria fortunei. (C) Cinnamomum septentrionale. (D) Ginkgo biloba.
Genes 13 01173 g001
Figure 2. Summary of differentially expressed genes (DEGs). (A) The number of DEGs for each pairwise comparison. (B) Common unigenes based DESeq2 and edgeR. (C) GO enrichment of common unigenes. (D) The volcano map DEGs.
Figure 2. Summary of differentially expressed genes (DEGs). (A) The number of DEGs for each pairwise comparison. (B) Common unigenes based DESeq2 and edgeR. (C) GO enrichment of common unigenes. (D) The volcano map DEGs.
Genes 13 01173 g002
Figure 3. PCA analysis of each comparison. (A) Represented Bpap vs. Cfor comparison. (B) Represented Bpap vs. Csep comparison. (C) Represented Bpap vs. Gbil comparison. (D) Represented Cfor vs. Csep comparison. (E) Represented Cfor vs. Gbil comparison. (F) Represented Csep vs. Gbil comparison.
Figure 3. PCA analysis of each comparison. (A) Represented Bpap vs. Cfor comparison. (B) Represented Bpap vs. Csep comparison. (C) Represented Bpap vs. Gbil comparison. (D) Represented Cfor vs. Csep comparison. (E) Represented Cfor vs. Gbil comparison. (F) Represented Csep vs. Gbil comparison.
Genes 13 01173 g003
Figure 4. Weighted gene co-expression network analysis (WGCNA) in all samples. (A) Hierarchical clustering tree (dendrogram) of unigenes. (B) Heatmap showing the comparison of modules among all samples.
Figure 4. Weighted gene co-expression network analysis (WGCNA) in all samples. (A) Hierarchical clustering tree (dendrogram) of unigenes. (B) Heatmap showing the comparison of modules among all samples.
Genes 13 01173 g004
Table 1. Summary of T. nigrans RNA-sequencing data analyzed in this study.
Table 1. Summary of T. nigrans RNA-sequencing data analyzed in this study.
Sample IDRaw ReadsClean ReadsBases (bp)
Bpap-00137,814,80637,021,5985,553,239,700
Bpap-00242,783,20041,996,0226,299,403,300
Bpap-00334,569,49631,134,6964,693,123,881
Cfor-00129,838,21827,889,0864,183,362,900
Cfor-00239,435,95238,563,1985,784,479,700
Cfor-00334,963,89834,374,7665,156,214,900
Csep-00132,646,19231,758,7784,763,816,700
Csep-00232,698,23630,735,7194,571,633,850
Csep-00338,673,52837,773,2785,665,991,700
Gbil-00134,908,51834,113,1925,116,978,800
Gbil-00227,899,30827,109,3464,066,401,900
Gbil-00333,457,78032,884,0684,932,610,200
Total419,689,132405,353,74740,058,127,750
Table 2. Overview of the de novo transcriptome assembly.
Table 2. Overview of the de novo transcriptome assembly.
TranscriptsUnigenes
Total number332,439234,510
Median contig length (bp)431348
Minimum length (bp)182201
Maximum length (bp)27,61627,616
N50 (bp)1790729
Number, ≤500 bp187,734165,956
Number, >500 bp144,70568,554
Total nucleotides296,766,287136,260,562
BUSCO (%)92.365.5
Table 3. KEGG enrichment analysis of common DEGs.
Table 3. KEGG enrichment analysis of common DEGs.
ClassKEGG Descriptionp Value
Organismal systemsCircadian rhythm—plant6.34 × 10 9
Brite hierarchiesCytochrome P4501.14 × 10 5
Organismal systemsOrganismal Systems2.35 × 10 5
Organismal SystemsEnvironmental adaptation2.35 × 10 5
Brite hierarchiesTranscription factors9.82 × 10 5
MetabolismCarotenoid biosynthesis2.18 × 10 4
MetabolismMetabolism of terpenoids and polyketides2.41 × 10 4
BRITE hierarchiesTransporters0.003102542
Not included in the pathway or BRITESignaling proteins0.003192151
BRITE hierarchiesProtein families: metabolism0.004501004
BRITE hierarchiesChaperones and folding catalysts0.004860346
Environmental information processingPlant hormone signal transduction0.006261807
BRITE hierarchiesProtein phosphatases and associated proteins0.019453528
Environmental information processingEnvironmental information processing0.021745631
Environmental information processingSignal transduction0.021941008
Not included in pathway or BRITEUnclassified: signaling and cellular processes0.032328609
BRITE hierarchiesUbiquitin system0.039933933
Environmental information processing6 MAPK signaling pathway—plant0.042543616
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kuang, J.; Wang, Y.; Mao, K.; Milne, R.; Wang, M.; Miao, N. Transcriptome Profiling of a Common Mistletoe Species Parasitizing Four Typical Host Species in Urban Southwest China. Genes 2022, 13, 1173. https://doi.org/10.3390/genes13071173

AMA Style

Kuang J, Wang Y, Mao K, Milne R, Wang M, Miao N. Transcriptome Profiling of a Common Mistletoe Species Parasitizing Four Typical Host Species in Urban Southwest China. Genes. 2022; 13(7):1173. https://doi.org/10.3390/genes13071173

Chicago/Turabian Style

Kuang, Jingge, Yufei Wang, Kangshan Mao, Richard Milne, Mingcheng Wang, and Ning Miao. 2022. "Transcriptome Profiling of a Common Mistletoe Species Parasitizing Four Typical Host Species in Urban Southwest China" Genes 13, no. 7: 1173. https://doi.org/10.3390/genes13071173

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