Skip to main content

The loss of photosynthesis pathway and genomic locations of the lost plastid genes in a holoparasitic plant Aeginetia indica

Abstract

Background

With three origins of holoparasitism, Orobanchaceae provides an ideal system to study the evolution of holoparasitic lifestyle in plants. The evolution of holoparasitism can be revealed by plastid genome degradation and coordinated changes in the nuclear genome, since holoparasitic plants lost the capability of photosynthesis. Among the three clades with holoparasitic plants in Orobanchaceae, only Clade VI has no available plastid genome sequences for holoparasitic plants. In this study, we sequenced the plastome and transcriptome of Aeginetia indica, a holoparasitic plant in Clade VI of Orobanchaceae, to study its plastome evolution and the corresponding changes in the nuclear genome as a response of the loss of photosynthetic function.

Results

The plastome of A. indica is reduced to 86,212 bp in size, and almost all photosynthesis-related genes were lost. Massive fragments of the lost plastid genes were transferred into the mitochondrial and/or nuclear genomes. These fragments could not be detected in its transcriptomes, suggesting that they were non-functional. Most protein coding genes in the plastome showed the signal of relaxation of purifying selection. Plastome and transcriptome analyses indicated that the photosynthesis pathway is completely lost, and that the porphyrin and chlorophyll metabolism pathway is partially retained, although chlorophyll synthesis is not possible.

Conclusions

Our study suggests the loss of photosynthesis-related functions in A. indica in both the nuclear and plastid genomes. The lost plastid genes are transferred into its nuclear and/or mitochondrial genomes, and exist in very small fragments with no expression and are thus non-functional. The Aeginetia indica plastome also provides a resource for comparative studies on the repeated evolution of holoparasitism in Orobanchaceae.

Background

Chloroplast (plastid) is an organelle of plants that conducts photosynthesis, and the structure and gene content of chloroplast genomes are highly conserved in most flowering plants [1]. Typical chloroplast DNA (cpDNA) is circular, ranging mainly from 110 to 160 kb in length [2], and it contains two inverted repeat (IR) sequences separated by a large single-copy region (LSC) and a small single-copy region (SSC) [3]. Holoparasitic plants offer a good system to study plastid genome evolution due to their loss of photosynthetic capacity. They usually display a reduction of plastid genome including genome size and gene content. The family Orobanchaceae is especially suitable for studying chloroplast (plastid) genome evolution because it contains a full trophic spectrum from autotrophic plants, to hemiparasites and holoparasites. In Orobanchaceae, holoparasites occur in three of six well supported clades, namely, Clade III (Orobancheae, ~ 180 holoparasitic species), Clade V (Rhinantheae, 7 holoparasitic species) and Clade VI (Buchnereae, ~ 70 species) [4,5,6]. So far, plastid genome sequences of holoparasites in Orobanchaceae were mainly from Clade III [7,8,9,10,11,12]. In Clade V, the plastid genome of only one holoparasite, Lathraea squamaria, has been sequenced recently [12]. Although Clade VI includes ~ 70 species from four holoparasitic genera (Hyobanche, Harveya, Aeginetia and Christisonia), no plastid genome sequences of holoparasitic plants from this clade have been characterized.

Plastid genomes of holoparasites in Clade III and Clade V of Orobanchaceae differ markedly in genome size and gene content. Plastid genome sizes of holoparasites in Clade III range from 45,673 (Conopholis americana) to 120,840 bp (Orobanche californica) [9]. However, the plastid genome size of Lathraea squamaria from Clade V is 150,504 bp [12], much larger than those in Clade III. The number of intact genes in the plastid genomes of Conopholis americana and Orobanche species ranges from 21 to 34 [9], and almost all genes related to photosynthesis (pet, psa, psb, and rbcL) were lost or became pseudogenes. Whereas in the plastid genome of Lathraea squamaria, there are 46 intact genes including many genes related to photosynthesis (such as psa, psb and pet). This might be due to holoparasitic lineages in Clade V is younger than those in Clade III [12].

In addition to plastome degradation, the nuclear genomes of holoparasitic plants are also expected to evolve as a response of the loss of photosynthesis capability, since the genes related to photosynthesis in the plastid genome interact with many genes in the nuclear genome [7, 13]. The expressional changes of nuclear genes could be revealed by transcriptome sequencing. For example, the expression of genes in the photosynthesis and chlorophyll synthesis pathways has been examined in some parasitic plants [7, 14, 15].

Aeginetia is a small holoparasitic genus of Orobanchaceae and it consists of about four species distributed in southern and southeastern Asia [16]. According to the phylogenetic analyses of Orobanchaceae, Aeginetia, along with Hyobanche, Harveya and Christisonia, forms a monophyletic holoparasitic lineage in Clade VI [4, 5]. Aeginetia indica is the most widespread species in this genus [17]. It usually parasitizes on the roots of Poaceae plants like Miscanthus and Saccharum [18]. In a recent study, transcriptome data of A. indica have been used to detect horizontally transferred genes from Fabaceae and Poaceae species [19]. So far, plastid genome sequence and the degradation of photosynthesis related pathways have not been studied in this holoparasitic plant.

In this study, we assembled the plastid genome of A. indica using Illumina short reads produced by genome skimming. We also sequenced the transcriptomes from multiple tissues to examine the expressional changes of genes involved in photosynthesis. In addition, we investigated the evolutionary fates of the lost plastid genes in A. indica. The results of this study will contribute to our understanding of the coordinated evolution of plastid and nuclear genomes and also facilitate comparative analysis of convergent evolution of holoparasitism in Orobanchaceae.

Results

Severe shrinkage of the A. indica plastid genome and evolutionary fates of the lost plastid genes

The complete plastid genome of A. indica is 86,212 bp in length, highly reduced relative to the size of most other angiosperms. It has a typical quadripartite structure, with 22,301 bp of the LSC region, 529 bp of the SSC region, and 31,691 bp each of the IR regions (Fig. 1). AT content of this plastid genome was 65.64%. Based on the DOGMA and GeSeq annotation, the plastid genome of A. indica contains 48 putative intact genes and three pseudogenes. These intact genes contain 18 tRNA genes, 4 rRNA genes, 8 rpl genes, 12 rps genes and 6 other genes, namely, ycf1, ycf2, accD, matK, infA and clpP (Table 1). The three pseudogenes are ψatpA, ψatpI and ψndhB. ψatpA and ψatpI genes in the LSC region of A. indica plastome became pseudogenes because of being truncated at the 88th codon and a premature stop codon at the 32nd codon, respectively. ψndhB gene in the IR region became a pseudogene due to an internal stop codon at the 53rd codon.

Fig. 1
figure 1

The plastid genome map of Aeginetia indica. Genes shown outside and inside the outer circle are transcribed counterclockwise and clockwise, respectively. GC and AT contents across the chloroplast genome are shown with the dark and light shading, respectively, inside the inner circle

Table 1 Summary of genes in the Aeginetia indica plastome

The SSC region in plastome of A. indica shows a severe reduction in size and only two genes, rpl15 and trnL-UAG, were found in this region (Fig. 1). The two IR regions have undergone expansions towards both the LSC and SSC regions. In the chloroplast genomes of the autotropic relative Lindenbergia philippensis and other autotropic plants, an intact ycf1 gene usually spans the IR and SSC regions, and rps8, rpl14, rpl16, rps3, rpl22 and rps19 genes were in the LSC region. Whereas, in A. indica plastome, there is an intact ycf1 gene in each of the IR regions, and rps8, rpl14, rpl16, rps3, rpl22 and rps19 genes all shift into the IR regions.

Gene contents in the plastomes of A. indica, four holoparasitic species including Cistanche deserticola, Orobanche austrohispanica and Epifagus virginiana from Clade III, and Lathraea squamaria from Clade V, and the autotrophic relative L. philippensis in Orobanchaceae, were compared (Table S1). Compared with L. philippensis, there is substantial loss of genes in the A. indica plastid genome. Ten ndh (ndhA, ndhC, ndhD, ndhE, ndhF, ndhG, ndhH, ndhI, ndhJ and ndhK) genes were lost, and ndhB gene became a pseudogene, they encode subunits of NADH-dehydrogenase complex. All five psa (psaA, psaB, psaC, psaI and psaJ) and 15 psb (psbA, psbB, psbC, psbD, psbE, psbF, psbH, psbI, psbJ, psbK, psbL, psbM, psbN, psbT and psbZ) genes involved in photosystem I and photosystem II, were lost. Also, all six pet (petA, petB, petE, petF, petH and petI) genes, which encode cytochrome b6/f complex subunits with function in photosynthetic electron transport, were missing. In addition, four atp (atpB, atpE, atpF and atpH) genes encoding F-type ATPase subunits, four genes encoding DNA dependent RNA polymerase (rpoA, rpoB, rpoC1 and rpoC2), and genes encoding envelop membrane protein (cemA), large subunit of Rubisco (rbcL), haem attachment factor (ccsA), and photosystem assembly factors (ycf3 and ycf4) were lost as well. Similar gene loss was found in the plastomes of two holoparasitic species Orobanche austrohispanica and Epifagus virginiana from Clade III. Although most of these genes were not lost in the plastomes of Lathraea squamaria from Clade V and Cistanche deserticola from Clade III, many of them became pseudogenes (Table S1).

To study the evolutionary fates of the lost plastid genes in A. indica, we identified their genomic positions through extracting genome skimming reads matched with the reference plastomes of 64 plant species and other analyses (see Methods). A total of 339 contigs with length ≥ 150 bp were assembled from the extracted reads, 76 of them were annotated as fragments of plastid genes lost in the A. indica plastid genome. These 76 fragments, ranging from 150 to 3086 bp in length, represent 27 lost plastid genes, including four atp genes (atpA, atpB, atpE and atpF), seven ndh genes (ndhA, ndhB, ndhD, ndhE, ndhF, ndhH and ndhJ), four psa genes (psaA, psaB, psaC and psaI), six psb genes (psbA, psbB, psbC, psbD, psbE and psbJ), three rpo genes (rpoB, rpoC1 and rpoC2), and one gene each for petB, rbcL and ycf3 (Table S2). In addition, we found that 15 fragments representing nine genes in the A. indica plastid genome were also transferred to mitochondrial and/or nuclear genomes and exist in small fragments. Among these genes, seven (atpA, ndhA, ndhB, ndhF, petB, rpoB and rpl23) were transferred to both the mitochondrial and nuclear genomes, 11 (ndhD, ndhE, ndhH, ndhJ, rpl14, rpl16, rpl2, rps12, rps14, rps3 and rps4) were transferred to only the mitochondrial genome, and 18 (atpB, atpE, atpF, psaA, psaB, psaC, psaI, psbA, psbB, psbC, psbD, psbE, psbJ, rbcL, rpoC1, rpoC2, ycf3 and ycf2) were transferred to only the nuclear genome based on their sequencing depth. All these fragments were not detected in the transcriptomes of multiple tissues, suggesting that they were non-functional.

Multiple structural rearrangements in the plastid genome of A. indica relative to its autotrophic relative

With Mauve 2.4.0, sequence alignment for the plastomes of A. indica (Clade VI) and five other Orobanchaceae species mentioned above was shown in Fig. 2. We identified nine locally co-linear blocks (LCBs) for these six species, and A. indica plastid genome has undergone two inversions relative to the chloroplast genome of L. philippensis. One inversion is located in the LSC region and contains an intact accD gene, and the other contains the intact SSC region and IRB region. Compared with the L. philippensis chloroplast genome, there were no inversions in the plastomes of Lathraea squamaria, Cistanche deserticola and Epifagus virginiana, while there were two distinct inversions in that of Orobanche austrohispanica (Fig. 2).

Fig. 2
figure 2

Mauve alignment of the plastomes of Aeginetia indica and five other Orobanchaceae species. Gene annotation of the Lindenbergia philippensis plastome is the upper panel. Different colors in the alignment show different locally collinear blocks

Relaxed purifying selection of A. indica plastid genes

A total of 20 protein coding genes shared among the seven species in Orobanchaceae, including 10 rps genes, 7 rpl genes, and accD, infA and matK genes were used for phylogenetic analysis. The maximum likelihood tree was strongly supported, with bootstrap values of all branches being 100 (Figure S1). Three Striga species were clustered into one clade, and Buchnera americana was sister to them. Aeginetia indica was sister to the clade consisting of the former four species. Non-synonymous (dN)/synonymous (dS) substitution rate ratio (ω) can be considered as an indicator for selection pressure. Two-ratio model (M2) was first compared with one-ratio model (M0). ω values of all genes but rpl20 and rps18 in the parasitic plant branch were larger than those of the nonparasitic plant branch (Table S3), and the likelihood ratio test showed that M2 is significantly better than M0 at nine genes, i.e. accD, infA, rpl22, rps11, rps14, rps19, rps2, rps3 and rps7, suggesting that these genes were under relaxed purifying selection in parasitic plants. Using three-ratio branch model (M3), we found that hemiparasitic species had higher or much higher ω than holoparasitic species at 13 of 18 genes (ω values of the remaining two genes are not available), while holoparasitic species had slightly higher ω than hemiparasitic species at only five genes (Table S3). This suggests that protein-coding genes retained in the plastome of A. indica still play important functional roles rather than experiencing more relaxed selective pressure than hemiparasitic species.

Degradation of the photosynthesis pathway in A. indica revealed by transcriptome analysis

We obtained 21.05, 19.04, 18.34 and 18.02 Gb clean reads for four tissues, i.e. flower, sepal, fruit, and stem, respectively. By de novo assembly of read data from the four tissues, we obtained a total of 205,380 transcripts, among which 153,986 were extracted as unigenes. The average length and N50 of these unigenes were 623.18 and 880 bp, respectively. There were 47,480 ORFs (Open Reading Frames) predicted from all unigenes by TransDecoder, and 42,007 of them could be annotated in Swiss-Prot database, among 42,007 Swiss-Prot annotations, 8466 could be assigned to 131 Kyoto Encyclopedia of Genes and Genomes (KEGG) pathways.

To confirm the reliability of gene expression obtained from transcriptome sequencing, the expression of 10 genes were also examined with qRT-PCR. The expression levels of these genes obtained from the two approaches were of relatively high correlation (Pearson correlation coefficient R2 = 0.71; Figure S2), suggesting that gene expression obtained from transcriptome sequencing was reliable.

The photosynthesis pathway (ko00195) from the KEGG pathway database contains 63 genes (30 plastid genes and 33 nuclear genes). In the A. indica plastome, genes that encode proteins involved in photosystem I and II, cytochrome b6f complex, and photosynthetic electron transport are completely lost. The only two F-type ATPase related genes (atpA and atpI) in its plastome are pseudogenes. Based on the transcriptome analysis, only 14 nuclear unigenes enriched in the photosynthesis pathway had expression (Table S4). The 14 genes include one gene encoding PSII 6.1 kDa protein, seven genes encoding proteins implicated in photosynthetic electron transport and six genes encoding components of F-type ATPase (Figure S3). Expression of other genes in this pathway was not detected, indicating that these genes were either lost or non-expressional. The results from plastome and transcriptome analyses indicate that the photosynthesis pathway in A. indica was completely lost.

The porphyrin and chlorophyll metabolism pathway (ko00860) is complicated in plants. Porphyrins are intermediates of heme and chlorophyll, and heme is required for chlorophyll biosynthesis [20]. In the pathway from glutamate to protoporphyrin IX, the expression of eight genes (HemA, HemB, HemC, HemD, HemE, HemF, HemL and HemY) were observed in the transcriptome of A. indica (Figure S4). However, because of the absence of expression of divinyl chlorophyllide a 8-vinyl-reductase [EC:1.3.1.75], which catalyzes divinyl protochlorophyllide to protochlorophyllide [21], the chlorophyll synthesis pathway appears to end at divinyl-proto-chlorophyllide production in A. indica (Figure S4). Obviously, the chlorophyll synthesis pathway is not complete at the later stage and chlorophyll can not be synthesized in A. indica.

Discussion

Features of the A. indica plastid genome

Among the available plastid genomes in Orobanchaceae, the plastid genome of A. indica is much smaller than those of Striga species (S. aspera, S. forbesii and S. hermonthica) from Clade VI of Orobanchaceae, but similar to those of Orobanche species, such as O. pancicii (88,525 bp) and O. crenata (87,529 bp) from Clade III [10]. However, their plastid genome structures are quite different: the sizes of the LSC (22,301 bp) and SSC (529 bp) regions in A. indica are much smaller than those of O. pancicii (42,763 and 7608 bp) and O. crenata (43,054 and 7925 bp), while the size of IR (31,691 bp) is much larger than those of O. pancicii (19,077 bp) and O. crenata (18,275 bp). The IR and SSC regions in plastome of A. indica have undergone expansions and shrinkage, respectively. These phenomena were also observed in four hemiparasites (Striga) from the same clade VI in Orobanchaceae [1]. The IR region in chloroplast genome is usually highly conserved in size and gene content, and it plays an important role in stabilizing chloroplast genome structure [22]. In the present study, IR region expansion occurred in A. indica suggests that dramatic changes in the plastome accompanied by the loss of photosynthesis. Moreover, our study shows that holoparasitic plants in Orobanchaceae exhibit different arrangements of plastid genomes, for example, compared with the L. philippensis chloroplast genome, there were no inversions in the plastomes of Lathraea squamaria, Cistanche deserticola and Epifagus virginiana, while there were two distinct inversions in each of the A. indica and Orobanche austrohispanica plastomes.

All 48 genes in the plastome of A. indica had sequence length similar to the autotrophic Lindenbergia philippensis. The protein-coding genes in A. indica play fundamental roles in plastid function, including large and small ribosome protein subunits (rpl and rps genes), acetyl-CoA carboxylase beta subunit involved in lipid biosynthesis (accD), intron splicing (matK), translational initiation factor (infA), ATP-dependent protease subunit P (clpP), and protein import and turnover (ycf1) [10, 23]. The function of ycf2 gene, which has a conserved open reading frame, is still unknown. Four genes, atpA, clpP, rpl2 and rpl23, contain introns, which is consistent with the retention of matK’s function as intron splicing.

The loss of photosynthesis related genes is a common phenomenon in holoparasitic plants, such as Aphyllon and Orobanche [10, 11]. Loss of housekeeping genes was also observed in other holoparasitic plants, for example, the plastid genome of Balanophora laxiflora is only 15,505 bp in size, with most genes being lost [24], and Rafflesia lagascae has even lost its whole plastid genome [25]. Some housekeeping genes have been transferred to the nuclear genome and their proteins can move back to the plastid to perform their functions [26]. In this study, 27 lost plastid genes in A. indica were transferred to nuclear and/or mitochondrial genomes, however, no expressions of these fragments could be detected from our transcriptome data, suggesting that they are nonfunctional.

Previous studies proposed models of plastome evolution in parasites and the order of gene losses [27,28,29]. The five stages in these models include “Photosynthetic”, “Degradation I”, “Stationary”, “Degradation II” and “Absent” stages. The order of gene losses starts with ndh genes, followed by psa/psb genes and rpo genes, then atp genes, rbcL gene, nonessential housekeeping genes and other metabolic genes like accD, clpP, ycf1 and ycf2, ends with the remaining housekeeping genes like rpl and rps genes. According to their models, plastome of A. indica is in the “Stationary” stage.

The loss of photosynthesis pathway in A. indica

Aeginetia indica has no photosynthetic activity and obtains all carbon through connection with its host [19]. In the present study, the loss of the photosynthesis pathway in A. indica was confirmed based on the loss of photosynthesis genes in its plastome and no detected expression of many genes in the photosynthesis pathway from its transcriptome. Chlorophyll is the primary pigment which absorbs light energy in photosynthesis. Chlorophyll synthesis is impossible in A. indica because some key genes in the later stage of the porphyrin and chlorophyll metabolism pathway were not detected with expression. Accumulation of protochlorophyllide in the light may increase the oxidative risk [30], so the degraded chlorophyll synthesis pathway in A. indica may benefit for its survival and evolution. In contrast, an intact chlorophyll synthesis pathway was ever found in a holoparasitic plant Phelipanche aegyptiaca, suggesting that the expression of the chlorophyll synthesis pathway has other functions (like metabolic signaling, previous studies have proved Mg-protoporphyrin IX is a candidate signaling molecule) other than photosynthesis [31].

Conclusions

The plastid genome of Aeginetia indica, a holoparasitic plant from Clade VI of Orobanchaceae, was de novo assembled in this study. Its plastid genome shows a reduction in size, accompanied with loss or pseudogenization of almost all photosynthesis related genes and some structural rearrangements. The lost plastid genes were transferred into its nuclear and/or mitochondrial genomes, and most of them exist in very small fragments which have no expression and are thus non-functional. Transcriptome analysis from multiple tissues indicates that the photosynthesis pathway of A. indica was completely lost, while the porphyrin and chlorophyll metabolism pathway was partially retained, although chlorophyll synthesis is not possible. Our results suggest coordinated loss of photosynthesis related functions in the plastid and nuclear genomes of a holoparasitic plant. The plastome and transcriptome data of A. indica in the present study provides genetic resources for future studies and will facilitate comparative analysis of convergent evolution of holoparasitism in Orobanchaceae.

Methods

Plant materials, DNA isolation and Illumina sequencing

All the tissues of Aeginetia indica used in this study were collected from Shimentai Forest Park, Yingde, Guangdong, China. This plant was identified by Dr. Renchao Zhou based on its morphological characteristics described in the Flora Republicae Popularis Sinicae [16]. The voucher specimen (Zhou20190721) was deposited at the herbarium of Sun Yat-sen University (SYS). There were no specific permits required for collecting tissue samples of this species for research purpose. The plants were sampled and frozen immediately in liquid nitrogen, then kept at − 80 °C for further analysis. Total DNA was isolated from the flower stalk with a HiPure Plant DNA Mini Kit (Magen Company, Guangzhou, China) following the manufacturer’s instructions. The quality and quantity of DNA were detected with 1% agarose gel electrophoresis and Qubit 3.0 Fluorometer (Invitrogen Corporation, USA), respectively. Shotgun genome sequencing with paired-end reads of 150 bp was performed on an Illumina Hiseq X Ten platform (IGE Biotechnology Ltd., Guangzhou, China). The sequencing data were deposited in NCBI Sequence Read Archive under accession number SRR9878563.

Assembly, annotation and alignment of plastid genome

The plastid genome of A. indica was de novo assembled from Illumina sequencing data using NOVOPlasty [32], with parameters of insert size (300 bp), K-mer (37) and coverage cut off (1500). The rps16 gene sequence (EU572719.1) of A. indica was served as a seed. Annotation of plastid genome was performed by combining the DOGMA program [33] and GeSeq in MPI-MP CHLOROBOX (https://chlorobox.mpimp-golm.mpg.de/) with default parameters. Genes which contain one or more premature stop codons or frameshift mutations were considered as potential pseudogenes. The annotated plastid genome sequence of A. indica was deposited in GenBank under accession number MN529629. The circular map of plastid genome was drawn with OGDRAW [34]. To compare gene loss in the holoparasitic plants from the three clades in Orobanchaceae and potential genomic rearrangements, we downloaded the plastid genome sequences of four holoparasitic species, namely, Cistanche deserticola (NC021111.1), Orobanche austrohispanica (NC031441.1), Epifagus virginiana (NC001568.1) from Clade III, and Lathraea squamaria (NC027838.1) from Clade V, and one autotrophic relative Lindenbergia philippensis (NC022859.1) from Clade I in Orobanchaceae from GenBank. These sequences were aligned using progressiveMauve [35] with default parameters.

Identification of genomic locations of the lost plastid genes in A. indica

The lost plastid genes in A. indica might be transferred to its nuclear and/or mitochondrial genomes. To identify their genomic locations, a total of 64 plastomes from 21 families in four orders (Table S5) were downloaded from GenBank and taken as reference bait sequences. The clean paired-end reads were aligned to all reference bait sequences using Magic-BLAST v1.5.0 [36] with the default parameters. All plastid and plastid-like reads of A. indica were extracted by Samtools v1.10 [37] with the command: samtools view -b -S -F 4. All the extracted reads were merged with Samtools v1.10 and then de novo assembled with Velvet v1.2.10 [38] by setting insert length 350, insert length standard deviation 100 and minimal contig length 150. The identity between these contigs and the A. indica plastome were obtained using Magic-BLAST v1.5.0. Contigs with identity ≤90% or alignment length shorter than contig length were extracted by Seqtk v1.2-r94 (https://github.com/lh3/seqtk) with the default parameters. All protein sequences of the Lindenbergia philippensis chloroplast genome were used to construct a protein database, and the extracted contigs were annotated against this database using BLASTX v2.5.0+ [39] with an e-value ≤10− 5. The lost plastid genes were searched from the extracted contigs. Contigs that were annotated as plastid genes or gene fragments with flanking mitochondrial or nuclear sequences were used to determine their genomic positions (mitochondrial or nuclear). However, for most contigs, the sequences are too short to determine their genomic positions, and we therefore calculated the average depth of these contigs by BWA-MEM [40] mapping to infer the genomic locations of the lost plastid genes. This is based on that mitochondrial DNA is usually one to two orders of magnitude in copy number higher than nuclear DNA. Note that for contigs which have hits to the plastid genes of A. indica, we calculated the depth of no-hit parts to avoid mapping of plastid gene reads. Sequences of these fragments were searched in the assembled transcriptomes (as described below) to assess whether they are expressional or not.

Testing signatures of relaxed purifying selection for plastid genes in A. indica

We downloaded plastid genomes of six species of Orobanchaceae from GenBank, including four hemiparastic plants Striga aspera (MF780872.1), Striga forbesii (MF780873.1), Striga hermonthica (MF780874.1), and Buchnera americana (MF780871.1) from Clade VI, and two autotrophic plants Lindenbergia philippensis (NC_022859.1) and Rehmannia glutinosa (NC_034308.1). Phylogeny of the six species and A. indica was performed using 20 protein-coding genes (7 rpl genes, 10 rps genes, accD, infA and matK) shared in their plastid genomes. After sequence alignment using MAFFT [41], phylogeny was reconstructed using the maximum likelihood algorithm in RAxML [42] with 1000 bootstrap replicates, with Rehmannia glutinosa served as an outgroup. The maximum likelihood tree was shown with FigTree v1.4.3 (http://tree.bio.ed.ac.uk/software/figtree/).

We then used the phylogenetic tree as the input tree to test relaxed purifying selection for plastid genes in A. indica. The coding regions of each gene shared among the seven species were aligned by ClustalW (Codons) with default settings in MEGA-X version 10.0.5 [43]. The ratios (ω) of non-synonymous (dN) to synonymous (dS) substitution rate for 20 shared genes were estimated using codon-based analysis (codeml) in the PAML v.4.8a package [44]. Different branch models were used to analyze selective pressures among these species. The null one-ratio model (M0, it hypothesizes that all branches have one ω) was firstly performed, and then the likelihood of a two-ratio model (M2), with a foreground ω1 for parasitic species and a background ω2 for autotrophic species, was compared with that of M0. Moreover, a branch model with three ratios (M3) which assumes three different ω values for holoparasitic, hemiparasitic and autotrophic species, respectively, was compared with M2. The likelihood ratio test for M0 vs M2, and M2 vs M3, was conducted with the Chi-square distribution, with the degree of freedom equal to the difference in the number of parameters for the models, to evaluate the fit of the data to alternative branch models.

Transcriptome sequencing

Total RNA was isolated from fresh flower, sepal, fruit, and stem tissues of A. indica, respectively. The quality and concentration of RNA were determined using 1% agarose gel electrophoresis and a Qubit spectrophotometer, respectively. mRNAs of these four tissues were purified with Oligo dT, and then used to construct cDNA libraries. Four cDNA libraries were sequenced on an Illumina HiSeq2000 platform (IGE Biotechnology Ltd., Guangzhou, China) with PE150 (paired-end 150 bp) strategy. The raw RNA sequencing data of flower, sepal, fruit, and stem tissues of A. indica were deposited in NCBI Sequence Read Archive under accession number SRR9959046, SRR9959049, SRR9959048 and SRR9959047, respectively.

Transcriptome analysis

Raw sequencing data were filtered by removing the adaptors and low quality reads. Transcripts were assembled using clean reads from four tissues through Trinity v2.8.4 [45], and then unigenes were extracted from these transcripts using a perl script get_longest_isoform_seq_per_trinity_gene.pl in Trinity. Expression level of all unigenes (Transcripts Per Million, TPM) was determined using RSEM [46]. For annotation, all unigenes were aligned to NCBI non-redundant database (NR, http://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/genbank/) using DIAMOND (http://ab.inf.uni-tuebingen.de/software/diamond/). Open reading frames (ORFs) were predicted and putative protein sequences were obtained using TransDecoder tools (http://transdecoder.sourceforge.net/). Putative protein sequences were annotated in Swiss-Prot (http://www.uniprot.org/) and Kyoto Encyclopedia of Genes and Genomes (KEGG) [47] pathway (http://www.genome.jp/kegg/pathway.html) databases with the criterion of E value <1e-5 using Blastx and GhostKOALA (https://www.kegg.jp/ghostkoala/), respectively. Within the KEGG pathways, we focused only on the photosynthesis pathway (ko00195) and porphyrin and chlorophyll metabolism pathway (ko00860), and expression of genes in these two pathways were recorded.

Quantitative real-time PCR experiment

To test the reliability of gene expression obtained from transcriptome sequencing, quantitative real-time PCR (qPCR) for 10 genes was performed using CFX Manager (Bio-Rad, USA) with TB Green® Premix Ex Taq™ II (Tli RNaseH Plus) (Takara Bio Inc., Japan). The protocol of the qPCR was as follows: an initiation step at 95 °C for 2 min, followed by 40 cycles of 5 s at 95 °C, 30 s at 60 °C, and a final melting curve analysis. Each reaction was performed with three biological replicates and three technical replicates, respectively. The actin gene was used as the internal control. Primers used for qPCR were listed in Table S6. The expression level for each gene was calculated using the 2−ΔΔCt method [48].

Availability of data and materials

The shotgun genome sequencing data of Aeginetia indica were deposited in NCBI Sequence Read Archive under accession number SRR9878563; the annotated plastid genome sequence of A. indica was deposited in GenBank under accession number MN529629; and the raw RNA sequencing data of flower, sepal, fruit, and stem tissues of A. indica were deposited in NCBI Sequence Read Archive under accession number SRR9959046, SRR9959049, SRR9959048 and SRR9959047, respectively. The data sets supporting the results of this study are included in this manuscript and its additional files.

Abbreviations

cpDNA:

Chloroplast DNA

IR:

Inverted Repeat

LSC:

Large single-copy region

SSC:

Small single-copy region

DOGMA:

Dual Organellar GenoMe Annotator

LCB:

Locally co-linear block

ORF:

Open Reading Frame

KEGG:

Kyoto Encyclopedia of Genes and Genomes

PSII:

Photosystem II

References

  1. Frailey DC, Chaluvadi SR, Vaughn JN, Coatney CG, Bennetzen JL. Gene loss and genome rearrangement in the plastids of five Hemiparasites in the family Orobanchaceae. BMC Plant Biol. 2018;18(1):30.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  2. Wicke S, Schneeweiss GM, dePamphilis CW, Müller KF, Quandt D. The evolution of the plastid chromosome in land plants: gene content, gene order, gene function. Plant Mol Biol. 2011;76(3–5):273–97.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  3. Shaw J, Lickey EB, Schilling EE, Small RL. Comparison of whole chloroplast genome sequences to choose noncoding regions for phylogenetic studies in angiosperms: the tortoise and the hare III. Am J Bot. 2007;94(3):275–88.

    Article  PubMed  CAS  Google Scholar 

  4. Bennett JR, Mathews S. Phylogeny of the parasitic plant family Orobanchaceae inferred from phytochrome a. Am J Bot. 2006;93(7):1039–51.

    Article  PubMed  CAS  Google Scholar 

  5. McNeal JR, Bennett JR, Wolfe AD, Mathews S. Phylogeny and origins of holoparasitism in Orobanchaceae. Am J Bot. 2013;100(5):971–83.

    Article  PubMed  Google Scholar 

  6. Fu W, Liu X, Zhang N, Song Z, Zhang W, Yang J, Wang Y. Testing the hypothesis of multiple origins of holoparasitism in orobanchaceae: phylogenetic evidence from the last two unplaced holoparasitic genera, Gleadovia and phacellanthus. Front Plant Sci. 2017;8:1380.

    Article  PubMed  PubMed Central  Google Scholar 

  7. Wolfe KH, Morden CW, Palmer JD. Function and evolution of a minimal plastid genome from a nonphotosynthetic parasitic plant. Proc Natl Acad Sci U S A. 1992;89(22):10648–52.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  8. Li X, Zhang TC, Qiao Q, Ren Z, Zhao J, Yonezawa T, Hasegawa M, Crabbe MJC, Li J, Zhong Y. Complete chloroplast genome sequence of Holoparasite Cistanche deserticola (Orobanchaceae) reveals gene loss and horizontal gene transfer from its host Haloxylon ammodendron (Chenopodiaceae). PLoS One. 2013;8(3):3.

    Article  Google Scholar 

  9. Wicke S, Müller KF, de Pamphilis CW, Quandt D, Wickett NJ, Zhang Y, Renner SS, Schneeweiss GM. Mechanisms of functional and physical genome reduction in photosynthetic and nonphotosynthetic parasitic plants of the broomrape family. Plant Cell. 2013;25(10):3711–25.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Cusimano N, Wicke S. Massive intracellular gene transfer during plastid genome reduction in nongreen Orobanchaceae. New Phytol. 2016;210(2):680–93.

    Article  PubMed  CAS  Google Scholar 

  11. Schneider AC, Chun H, Tefanovic S, Baldwin BG. Punctuated plastome reduction and host-parasite horizontal gene transfer in the holoparasitic plant genus Aphyllon. P Roy Soc Lond B: Bio. 2018;285(1887):20181535.

  12. Samigullin TH, Logacheva MD, Penin AA, Vallejo-Roman CM. Complete plastid genome of the recent holoparasite lathraea squamaria reveals earliest stages of plastome reduction in orobanchaceae. PLoS One. 2016;11(3):3.

    Article  CAS  Google Scholar 

  13. Barkan A, Goldschmidt-Clermont M. Participation of nuclear genes in chloroplast gene expression. Biochimie. 2000;82(6–7):559–72.

    Article  PubMed  CAS  Google Scholar 

  14. Ng SM, Lee XW, Mat-Isa MN, Aizat-Juhari MA, Adam JH, Mohamed R, Wan KL, Firdaus-Raih M. Comparative analysis of nucleus-encoded plastid-targeting proteins in Rafflesia cantleyi against photosynthetic and non-photosynthetic representatives reveals orthologous systems with potentially divergent functions. Sci Rep-UK. 2018;8(1):17258.

    Article  CAS  Google Scholar 

  15. Ravin NV, Gruzdev EV, Beletsky AV, Mazur AM, Prokhortchouk EB, Filyushin MA, Kochieva EZ, Kadnikov VV, Mardanov AV, Skryabin KG. The loss of photosynthetic pathways in the plastid and nuclear genomes of the non-photosynthetic mycoheterotrophic eudicot Monotropa hypopitys. BMC Plant Biol. 2016;16:238.

    Article  PubMed  CAS  Google Scholar 

  16. Zhang Z. Orobanchaceae. In: Wang W, editor. Flora Republicae Popularis Sinicae. Beijing: Science Press; 1990. p. 69–124.

    Google Scholar 

  17. Morawetz JJ, Randle CP, Wolfe AD. Phylogenetic relationships within the tropical clade of Orobanchaceae. Taxon. 2010;59(2):416–26.

    Article  Google Scholar 

  18. Kato Y, Hisano K. In vitro culture of a root parasite, Aeginetia indica L. - I. general survey. Bot Mag Tokyo. 1983;96(3):203–9.

    Article  Google Scholar 

  19. Kado T, Innan H. Horizontal gene transfer in five parasite plant species in Orobanchaceae. Genome Biol Evol. 2018;10(12):3196–210.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  20. Oborník M, Green BR. Mosaic origin of the heme biosynthesis pathway in photosynthetic eukaryotes. Mol Biol Evol. 2005;22(12):2343–53.

    Article  PubMed  CAS  Google Scholar 

  21. Schoefs B, Bertrand M. Chlorophyll biosynthesis - a review. In: Pessarakli M, editor. Handbook of photosynthesis. Boca Raton, London, New York: Taylor & Francis; 2005. p. 37–54.

    Google Scholar 

  22. Perry AS, Wolfe KH. Nucleotide substitution rates in legume chloroplast DNA depend on the presence of the inverted repeat. J Mol Evol. 2002;55(5):501–8.

    Article  PubMed  CAS  Google Scholar 

  23. Cho WB, Choi BH, Kim JH, Lee DH, Lee JH. Complete Plastome sequencing reveals an extremely diminished SSC region in Hemiparasitic Pedicularis ishidoyana (Orobanchaceae). Ann Bot Fenn. 2018;55(1–3):171–83.

    Article  Google Scholar 

  24. Su HJ, Barkman TJ, Hao W, Jones SS, Naumann J, Skippington E, Wafula EK, Hu JM, Palmer JD, DePamphilis CW. Novel genetic code and record-setting AT-richness in the highly reduced plastid genome of the holoparasitic plant Balanophora. Proc Natl Acad Sci U S A. 2019;116(3):934–43.

    Article  PubMed  CAS  Google Scholar 

  25. Molina J, Hazzouri KM, Nickrent D, Geisler M, Meyer RS, Pentony MM, Flowers JM, Pelser P, Barcelona J, Inovejas SA, et al. Possible loss of the chloroplast genome in the parasitic flowering plant Rafflesia lagascae (Rafflesiaceae). Mol Biol Evol. 2014;31(4):793–803.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  26. Börner T, Zhelyazkova P, Legen J, Schmitz-Linneweber C. Chloroplast gene expression—RNA synthesis and processing. In: Theg SM, Wollman F-A, editors. Plastid Biology. New York: Springer New York; 2014. p. 3–47.

    Chapter  Google Scholar 

  27. Wicke S, Müller KF, DePamphilis CW, Quandt D, Bellot S, Schneeweiss GM. Mechanistic model of evolutionary rate variation en route to a nonphotosynthetic lifestyle in plants. Proc Natl Acad Sci U S A. 2016;113(32):9045–50.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  28. Wicke S, Naumann J. Chapter Eleven - Molecular Evolution of Plastid Genomes in Parasitic Flowering Plants. In: Chaw S-M, Jansen RK, editors. Advances in Botanical Research. Massachusetts: Academic Press; 2018. p. 315–47.

  29. Naumann J, Der JP, Wafula EK, Jones SS, Wagner ST, Honaas LA, Ralph PE, Bolin JF, Maass E, Neinhuis C, et al. Detecting and characterizing the highly divergent plastid genome of the nonphotosynthetic parasitic plant Hydnora visseri (Hydnoraceae). Genome Biol Evol. 2016;8(2):345–63.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Schoefs B, Franck F. Protochlorophyllide reduction: mechanisms and evolution. Photochem Photobiol. 2003;78(6):543–57.

  31. Kleine T, Voigt C, Leister D. Plastid signalling to the nucleus: messengers still lost in the mists? Trends Genet. 2009;25(4):185–92.

    Article  PubMed  CAS  Google Scholar 

  32. Dierckxsens N, Mardulyn P, Smits G. NOVOPlasty: De novo assembly of organelle genomes from whole genome data. Nucleic Acids Res. 2017;45(4):4.

    Google Scholar 

  33. Wyman SK, Jansen RK, Boore JL. Automatic annotation of organellar genomes with DOGMA. Bioinformatics. 2004;20(17):3252–5.

    Article  PubMed  CAS  Google Scholar 

  34. Lohse M, Drechsel O, Kahlau S, Bock R. OrganellarGenomeDRAW--a suite of tools for generating physical maps of plastid and mitochondrial genomes and visualizing expression data sets. Nucleic Acids Res. 2013;41(Web Server issue):W575–81.

    Article  PubMed  PubMed Central  Google Scholar 

  35. Darling AE, Mau B, Perna NT. Progressivemauve: multiple genome alignment with gene gain, loss and rearrangement. PLoS One. 2010;5(6):e11147.

  36. Boratyn GM, Thierry-Mieg J, Thierry-Mieg D, Busby B, Madden TL. Magic-BLAST, an accurate RNA-seq aligner for long and short reads. BMC Bioinformatics. 2019;20(1):405.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, Marth G, Abecasis G, Durbin R. The sequence alignment/map format and SAMtools. Bioinformatics. 2009;25(16):2078–9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Zerbino DR, Birney E. Velvet: Algorithms for de novo short read assembly using de Bruijn graphs. Genome Res. 2008;18(5):821–9.

  39. Camacho C, Coulouris G, Avagyan V, Ma N, Papadopoulos J, Bealer K, Madden TL. BLAST+: architecture and applications. BMC bioinformatics. 2009;10(1):421.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Li H. Aligning sequence reads, clone sequences and assembly contigs with BWA-MEM. arXiv preprint. 2013;00(2013):1–3.

    Google Scholar 

  41. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30(4):772–80.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics. 2014;30(9):1312–3.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  43. Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol. 2018;35(6):1547–9.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  44. Yang Z. PAML 4: phylogenetic analysis by maximum likelihood. Mol Biol Evol. 2007;24(8):1586–91.

    Article  PubMed  CAS  Google Scholar 

  45. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, 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(7):644–52.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Li B, Dewey CN. RSEM: accurate transcript quantification from RNA-Seq data with or without a reference genome. BMC Bioinformatics. 2011;12:323.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Kanehisa M, Goto S. KEGG: Kyoto encyclopedia of genes and genomes. Nucleic Acids Res. 2000;28(1):27–30.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2−ΔΔCT method. Methods. 2001;25(4):402–8.

    Article  PubMed  CAS  Google Scholar 

Download references

Acknowledgments

Not applicable.

Funding

This study was supported by the National Natural Science Foundation of China (31811530297), and Guangdong Provincial Key Laboratory of Applied Botany, South China Botanical Garden, Chinese Academy of Sciences (AB2018014), Science and Technology Program of Guangzhou (201707020035) and Chang Hungta Science Foundation of Sun Yat-sen University. The funding body played no role in the design of the study and collection, analysis, and interpretation of data and in writing the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

RZ conceived the study and designed the experiments; JC, RY and JD performed the experiments and data analysis; JC and RZ wrote the manuscript; RZ and YL revised the manuscript. All authors read and approved the final manuscript.

Corresponding author

Correspondence to Renchao Zhou.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

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

Supplementary information

Additional file 1: Figure S1.

Maximum likelihood tree of seven species in Orobanchaceae based on sequences of 20 plastid genes shared among them. Numbers in the nodes are bootstrap values. Scale in substitutions per site.

Additional file 2: Figure S2.

The correlation between the transcriptome analysis and the qRT-PCR measurements. Each qRT-PCR reaction was performed with three biological replicates and three technical replicates.

Additional file 3: Figure S3.

The expression of genes in the photosynthesis pathway observed in the Aeginetia indica transcriptome. Genes with detected expression were in the red boxes. With courtesy of© www.genome.jp/kegg/kegg1.html.

Additional file 4: Figure S4.

The expression of genes in the porphyrin and chlorophyll metabolism pathway detected in the Aeginetia indica transcriptome. Genes with detected expression were in the red boxes. With courtesy of© www.genome.jp/kegg/kegg1.html.

Additional file 5: Table S1.

Gene contents of the plastomes of five holoparasitic plants and one autotrophic relative Lindenbergia philippensisin. Black boxes indicate the absence of genes, white boxes indicate the presence of genes, and gray boxes indicate pseudogenes.

Additional file 6: Table S2.

Genomic locations of the lost and transferred plastid genes in Aeginetia indica. The annotation, sequence alignment, alignment position and sequence identity are all based on the chloroplast genes of Lindenbergia philippensis. * indicates gene also present in the plastome of Aeginetia indica.

Additional file 7: Table S3

. Relaxation of purifying selection in parasitic plants of Orobanchaceae based on branch model analysis of 20 protein coding genes shared by seven species of Orobanchaceae. The likelihood ratio test was used to compare the three models (M0: one ratio model; M2: two ratio model; M3: three ratio model). P-values are in bold when they are less than 0.05.

Additional file 8: Table S4.

Expression level of unigenes of Aeginetia indica in the photosynthesis pathway based on transcriptome analysis.

Additional file 9: Table S5.

Data sources of 64 reference plastid genomes used in this study.

Additional file 10: Table S6.

Primers used for quantitative real-time PCR.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. 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 in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chen, J., Yu, R., Dai, J. et al. The loss of photosynthesis pathway and genomic locations of the lost plastid genes in a holoparasitic plant Aeginetia indica. BMC Plant Biol 20, 199 (2020). https://0-doi-org.brum.beds.ac.uk/10.1186/s12870-020-02415-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12870-020-02415-2

Keywords