Skip to main content

Complete chloroplast genome of the desert date (Balanites aegyptiaca (L.) Del. comparative analysis, and phylogenetic relationships among the members of Zygophyllaceae

Abstract

Background

Balanites aegyptiaca (L.) Delile, commonly known as desert date, is a thorny evergreen tree belonging to the family Zygophyllaceae and subfamily Tribuloideae that is widespread in arid and semiarid regions. This plant is an important source of food and medicines and plays an important role in conservation strategies for restoring degraded desert ecosystems.

Results

In the present study, we sequenced the complete plastome of B. aegyptiaca. The chloroplast genome was 155,800 bp, with a typical four-region structure: a large single copy (LSC) region of 86,562 bp, a small single copy (SSC) region of 18,102 bp, and inverted repeat regions (IRa and IRb) of 25,568 bp each. The GC content was 35.5%. The chloroplast genome of B. aegyptiaca contains 107 genes, 75 of which coding proteins, 28 coding tRNA, and 4 coding rRNA.

We did not observe a large loss in plastid genes or a reduction in the genome size in B. aegyptiaca, as found previously in some species belonging to the family Zygophyllaceae. However, we noticed a divergence in the location of certain genes at the IR-LSC and IR-SSC boundaries and loss of ndh genes relative to other species. Furthermore, the phylogenetic tree constructed from the complete chloroplast genome data broadly supported the taxonomic classification of B. aegyptiaca as belonging to the Zygophyllaceae family. The plastome of B. aegyptiaca was found to be rich in single sequence repeats (SSRs), with a total of 240 SSRs.

Conclusions

The genomic data available from this study could be useful for developing molecular markers to evaluate population structure, investigate genetic variation, and improve production programs for B. aegyptiaca. Furthermore, the current data will support future investigation of the evolution of the family Zygophyllaceae.

Peer Review reports

Introduction

The Zygophyllaceae family includes approximately 22 genera and 230–240 species [1, 2]. Balanites aegyptiaca (L.) Delile, desert date or heglig, belonging to the Zygophyllaceae family, is native to dry and semiarid regions in Asia and Africa. It is a thorny evergreen shrub or small tree with multiple branches, grey bark [3, 4], compound leaves, hermaphrodite flowers that are usually greenish-yellow, and brown fruit with a hardstone seed [4].

B. aegyptiaca has been the subject of various classic studies because of its importance as a source of food and medicines. In traditional medicine, the fruit of this plant is used to treat diabetes, asthma, epilepsy, and malaria [5]. The seeds of B. aegyptiaca fruit have a high concentration of oils (46.0–54.7%), especially unsaturated fatty acids (up to 75% of total fatty acids), and proteins (26.1–34.3%). The oil extracted from B. aegyptiaca can potentially be used as biodiesel, an alternative to chemical diesel [5]. Furthermore, B. aegyptiaca plays a key role in conservation strategies owing to its resistance to drought, and it has been used as a native plant in the restoration of a degraded ecosystem in Africa [4,5,6]. In the Great Green Wall (GGW) project, B. aegyptiaca was selected as one of the native plants that was convenient for restoring degraded Sahelian ecosystems. The GGW project aimed to plant a green belt of trees extending south of the Sahara Desert across 11 countries [5].

Despite the ability of plants in the Zygophyllaceae family to adapt to harsh conditions, the stability and natural diversity of B. aegyptiaca currently face threats from anthropogenic pressure (wood is used as a fuel), animal overgrazing, and environmental pressure due to the increasing occurrence of drought episodes [5]. Civil wars and consequent instability also represent a worrisome factor, as the migration of people negatively impacts certain areas, which become vulnerable to the depletion of natural resources [7]. In addition, B. aegyptiaca faces deterioration when large areas of virgin land are urbanised [8].

Historically, the genus Balanites has undergone numerous changes in name and taxonomic position. The species B. aegyptiaca was first described by Alpino in 1592 under the name Agihalid [9]. In 1753, Linnaeus described it as Ximenia aegyptiaca, while in 1813, Delile replaced the name Agihalid with Balanites, from a Greek word meaning “the fruit” [10]. Harms in1904 [11] proposed keeping the name Balanites, and Balanites aegyptiaca was formally adopted at the Vienna Botanical Congress in 1905 [9]. Initial classifications in the genus Balanites were based on morphological characteristics and were vulnerable to conflict, which led to the movement of Balanites between plant families. Initially, Bentham in 1862 [12] placed it within the family Simaroubaceae, while Engler 1896 [13] moved it to the family Zygophyllaceae. Cronquist in 1968 [14] returned it again to Simaroubaceae. Hegnauer in 1973 [15] provided evidence that Balanites did not contain quassia-like alkaloids, which are the main chemical characteristic of the family Simaroubaceae. Hegnauer 1973, Scholz 1964 and Cronquist 1981 [15,16,17] supported the return of Balanites to the Zygophyllaceae family again. Maksoud 1988 [18] mentioned that similarities between the flavonoids of Balanites and Zygophyllaceae did not support the treatment of Balanites as a separate family. Sheahan 1993 [19] studied 37 species in 19 genera within Zygophyllaceae. Based on anatomy and c4 activity in 27 species, the results of their study supported the separation of Balanites into an independent family named Balanitaceae. Boesewinkel 1994 [20] also supported the separation of Balanites in a special family based on distinguishing features of ovule and seed characters. Next, the molecular and anatomical characteristics of flowers and embryos, in addition to pollen characteristics, supported the classification of Balanites within the family Zygophyllaceae [21, 22] . However, he last comprehensive review performed by [23] supported the separation of Balanites and its placement in the family Balanitaceae.

While a few molecular phylogenetic studies have been conducted for the Zygophyllaceae family, no study has specifically addressed B. aegyptiaca. The molecular aspects of Zygophyllaceae were studied by [24] using the plastid gene rbcL, in combination with anatomical and morphological data from 20 Zygophyllaceae species, including B. maughamii. Based on morphological and phylogenetic data, the redistribution of Zygophyllaceae into five subfamilies, namely, Morkillioideae, Tribuloideae, Seetzenioideae, Larreoideae, and Zygophylloideae, was proposed [24]. Recently, [1] presented a phylogenetic tree for Zygophyllaceae based on the Bayesian analysis tree of combined sequence data (rbcL, trnL-F, and ITS) for these five subfamilies. The results were consistent with previous studies [21, 22, 24, 25], and the Balanites genus was affiliated with the Tribuloideae subfamily within the Zygophyllaceae family. On the other hand, [26] reported that B. aegyptiaca includes five varieties; however, this number was later reduced to two by [26]. In a study by [27], three ecotypes were reported within the B. aegyptiaca population in Egypt based on RAPD markers. This discrepancy in the number of varieties belonging to Balanites aegyptiaca may be due to variation in morphological characteristics affected by environmental conditions.

Chloroplasts, which have an essential role in photosynthesis, are organelles in plant cells that contain their own genome, the plastome [28]. Typically, chloroplast genome sizes range between 120 and 170 kilobase pairs (kb) [29]. A chloroplast genome presents a four-region structure comprising a large single copy (LSC), a small single copy (SSC), and two inverted repeats (IRa and IRb). During the evolutionary history of plant families, plastomes have been subjected to strong selective pressure [30]. Thus, chloroplast genomes include useful phylogenetic information that can be used to study evolutionary relationships at different taxonomic levels and resolve difficult problems in plant phylogenetics. In addition, chloroplast genomes represent a database for identifying and developing efficient polymorphic molecular markers for studying genetic diversity and population structure, and for DNA barcoding, which is a tool for identification [31, 32].

Zygophyllaceae is an angiosperm family that utilizes the C4 pathway [33], which helps plants adapt to harsh, dry environments. Molecular phylogenetic analyses suggest that the evolutionary history of Zygophyllaceae is related to an arid period that began in the Oligocene in Asia and Africa [34, 35]. The most recent study of three chloroplast genomes of Zygophyllum species belonging to the Zygophyllaceae family reported a significant reduction in the genome size of these species [36]. In this study, we validated the hypothesis that dry environments will significantly reduce the genome size of B. aegyptiaca in the family Zygophyllaceae. The aims of the present study are to provide baseline molecular information on the B. aegyptiaca plastome (B. aegyptiaca is considered typical of plants in dry and semiarid environments), to verify the plastome structure of B. aegyptiaca, and to perform comparative analyses and investigate phylogenetic relationships and variations between B. aegyptiaca and related Zygopyllaceae species, based on an available complete plastome dataset in GenBank.

Results

Characteristics of the B. aegyptiaca chloroplast Genome

The cp genome of B. aegyptiaca, shown in Fig. 1, is a circular molecule with a length of 155,800 bp. It has a four-region structure comprising a large single copy, a small single copy, and two inverted repeats. The LSC and SSC regions were 86,562 bp and 18,102 bp long, respectively, while IRa and IRb regions were 25,568 bp each (Table 1). The length of the coding region is 75,890 bp and represents 49% of the whole genome, while the noncoding region length is 79,910 bp (51%). The percentage of AT in the whole genome was 64.5%, whereas the percentage of GC was 35.5%. The genome of B. aegyptiaca consists of A = 31.8%, T(U) = 32.7%, C = 18.1% and G = 17.4%, as shown in Table 1.

Fig. 1
figure 1

Gene map of Balanites aegyptiaca chloroplast genome. Genes outside the circles are transcribed counter clockwise, while those inside are transcribed clockwise. Known functional genes are indicated in the coloured bar

Table 1 Chloroplast genome features of Balanites aegyptiaca

The total number of unique genes was 107, and the number of duplicated genes in the inverted region was 18. The LSC region was composed of 60 protein-coding genes and 21 tRNA genes, the SSC region contains 11 protein-coding genes and one tRNA, and 10 protein-coding genes and 22 tRNAs are located in the IR regions. Most of the protein-coding genes start with a methionine codon (AUG).

Sixteen of the 107 genes in B. aegyptiaca contain introns, 11 are protein coding genes, and five were tRNA genes, as illustrated in Table 2. The clpP and ycf3 genes present two introns, while the remaining genes present only one intron. Ten introns were included in the LSC region, one intron is included in the SSC region, and five introns were specifically located within the IRa and IRb regions.

Table 2 Genes with introns in the chloroplast genome of Balanites aegyptiaca

Relative synonymous codon usage (RSCU)

The nucleotides of protein coding and tRNA genes were used to determine the codon usage bias of the plastome. The results obtained from the analysis of protein-coding genes and tRNA genes (78,624 bp) of the B. aegyptiaca plastome are shown in Table S2. Genes are encoded by 22,415 codons. Leucine is the most frequent amino acid (11.1%), as shown in Fig. 2, whereas cysteine is the least frequent (1.2%). The RSCU values in Table S2 show that half (30) of the codons are > 1, all with an A/T ending. It can be seen from the data that tryptophan and methionine with no codon usage bias have an RSCU value of 1.

Fig. 2
figure 2

Amino acid frequencies of the protein-coding sequences of Balanites aegyptiaca chloroplast genome

RNA editing sites

The PREP suite was used to predict the RNA editing sites in the B. aegyptiaca plastome, and the first codon position of the first nucleotide was used in the analysis.

The RNA editing sites are presented in Table S3. Overall, there were 41 editing sites in the genome distributed among protein-coding genes. Most of the codon position exchanges involved serine (S) and leucine (L) amino acids (S to L). The results obtained show the highest number of editing sites in the ndhB and ndhD genes (12 and 6 sites, respectively), followed by rpoB and ndhF genes with three sites. Moreover, the accD, atpA, matK, and ndhA genes have 2 sites, while 1 site is present for each of the remaining genes. The results of RNA editing show that certain genes, namely, atpI, ccsA, petB, petD, petG, petL, psaB, psaI, psbB, psbE, psbF, rpl2, rpl23, rpoA, ycf3, and atpB, do not possess a predicted site in the first codon of the first nucleotide.

Repeat Analysis

Long repeats

There are four types of repeats in the cp genome of B. aegyptiaca: palindromic (21), forward (12), reverse (15), and complement (1), as evident from Table S4. Overall, there were 49 repeats in the B. aegyptiaca plastome. The majority (75.5%) of repeats were found in the intergenic spacer (IGS) region. The sizes of most of the repeats range from 20 to 29 bp (71.4%), followed by 10–19 bp (14.3%), 30–39 bp (10.2%), and 40–49 bp (4.1%). The tRNAs include 4 repeats (8.2%), while the remaining 8 repeats (16.3%) were in the protein-coding genes ndhC, ycf2, clpP, ycf1, and ndhA. We noticed that the protein-coding gene ycf2 had the most repeat locations: 2 palindromic repeats and 2 forward repeats.

A comparison of the number of repeats in six Zygophyllaceae species (Balanites aegyptiaca (L.), Guaiacum angustifolium Engelm., Larrea tridentata (DC.) Coville, Tetraena mongolica Maxim., Tribulus terrestris L., and Zygophyllum xanthoxylon (Bunge) Maxim.) is provided in Fig. 3 (species description is available in Table S5); the genome of L. tridentata has the highest frequency of palindromic repeats (26), while that of Z. xanthoxylon has the lowest (15). L. tridentata and T. mongolica have the same number of forward repeats (17), while the same number of reverse repeats occur in the cp genome of T. mongolica and T. terrestris (13). We can also see that complement repeats are the least abundant in the three genomes; B. aegyptiaca, L. tridentata and G. angustifolium have only one complement repeat, and complement repeats are absent in the other three species.

Fig. 3
figure 3

Number of different repeats among six Zygophyllaceae plastomes. F, forward repeats; P, palindromic repeats; R, reverse repeats; and C, complement repeats

Simple sequence repeats (SSRs)

A total of 240 SSRs are present in the plastid genome of B. aegyptiaca, which is a larger number than in the five other species of Zygophyllaceae shown in Table 3. The majority of SSRs in the cp genome are mononucleotides (86. 3%), mostly poly T (53.6%) and A (44%) (Figs. 4 and 5); poly C and G represent 1.5 and 1% of repeats, respectively. The results obtained from the analysis of SSRs frequency in the genomes of six species of Zygophyllaceae are presented in Table 3 and Fig. 4. The dinucleotide AT/AT was found in the genome of all species, while the dinucleotide AG/CT was found in four species but was absent from L. tridentata and G. angustifolium. Furthermore, the existence of three trinucleotides (AAT/ATT, AAG/CTT, and AAC/GTT), as well as nine tetra-repeats (AAAC/GTTT, AAAG/CTTT, AAAT/ATTT, ACAT/ATGT, AATC/ATTG, AATC/ATTG, AATG/ATTC, ACCT/AGGT, and ACTG/AGTC), and three pentanucleotides (AATAT/ATATT, AATCG/ATTCG, and AAATAT/ATATTT) was observed only in T. mongolica, Z. xanthoxylon, and L. tridentata (Fig. 4).

Table 3 cpSSRs detected in six Zygophyllaceae chloroplast genomes
Fig. 4
figure 4

Frequency of different microsatellite motifs in different repeat types of six Zygophyllaceae plastome genomes

Fig. 5
figure 5

Number of different SSR types in the six Zygophyllaceae chloroplast genomes

A comparison between the frequency of SSRs in the plastomes of the six species is presented in Fig. 5; it is clear that mononucleotides are the most frequent in all genomes. B. aegyptiaca had the highest number of mononucleotides, trinucleotides, and tetranucleotides (207, 6, and 12, respectively). However, pentanucleotides were not present in the cp genomes of B. aegyptiaca, G. angustifolium, or T. terrestris. In addition, hexa-repeats were not present in any of the six species.

Overall, the IGS region included most of the SSRs repeats (83.3%), shown in Fig. 6, followed by the coding regions (16.7%).

Fig. 6
figure 6

Number of SSR types in the complete chloroplast genome, protein-coding regions, and non-coding genes of Balanites aegyptiaca

Sequence divergence

To investigate the degree of genome sequence divergence, the program mVISTA was used to align the cp genome sequence of B. aegyptiaca with five Zygophyllaceae chloroplast genomes available in GenBank: T. mongolica, Z. xanthoxylon, L. tridentata, G. angustifolium, and T. terrestris. The alignment showed that genes were less conserved in the genomes of T. mongolica and Z. xanthoxylon, especially the ndh genes rps16 and ycf2. In general, the protein-coding genes were more conserved than the noncoding regions (Fig. 7). The noncoding regions presented high divergence in the following genes: psabA-trnK-UUU, trnK-UUU-rps16, psbK-psbI, trnG-UCC-trnR-UCU, atpF-atpH, atpH-atpI, rps2-rpoC2, rpoC2-rpoC1, rpoC1-rpoB, rpoB-trnC-GCA, psbD-trnS-UGA, psaA-ycf3, ycf3-trnS-GGA, trnS-GGA-rps4, trnT-UGU-trnL-UAA, ndhC-trnM-CAU, ntrnM-CAU-atpE, atpB-rbcL, rbcL-accD, cemA-petA, petA-psbJ, psbJ-psbF, psbE-petL, psaJ-rpl33, rps18-rpl20, rps12-psbB, psbB-psbT, psbH-petB, petB-petD, petD-rpoA, rps8-rpl14, rpl14-rps3, rpl2-rpl23, ycf2-trnL-CAA, trnL-CAA-ndhB, ndhF-trnL-UAG, psaC-ndhE, trnN-GUU-trnR-ACG, trnA-UGC-trnI-GAU, trnI-GAU-rrn16S, rrn16S-trnV-GAC, trnL-CAA-ycf2, trnI-CAU-rpl23, and rpl23-rpI2. However, the following protein-coding genes showed divergence in fewer regions: matK, rpoC2, psaA, accD, cemA, clpP, rpl23, ndhF, ccsA, and rpl23.

Fig. 7
figure 7

Whole chloroplast genome alignments for six Zygophyllaceae species via the mVISTA program, using the annotation of Balanites aegyptiaca as reference. The x-axis represents the coordinates in the cp genome, while the y-axis indicates percentage identity from 50 to 100%. The top grey arrows indicate the position and direction of each gene. Pink indicates non-coding sequences (NCS), blue indicates protein-coding genes, and light green indicates tRNAs and rRNAs

Boundary between LSC/SSC and IR regions

The comparisons between IR-SC boundaries for the six species of Zygophyllaceae are shown in Fig. 8. In general, the variation in length of the two LSC/SSC regions was lower than that of the IRa/IRb regions. The shortest LSC and SSC regions were 80,458 bp and 13,767 bp, respectively, in T. mongolica and G. angustifolium, while the longest LSC and SSC regions were 88,878 bp and 18,102 bp, in T. terrestris and B. aegyptiaca, respectively. The IRa/IRb regions of B. aegyptiaca and T. terrestris were the longest (25,568/25842), and contraction was noticed in the IR region in both Z. monogolica (4315 bp) and Z. xanthoxylon (5084 bp) compared to the other species (Table S6).

Fig. 8
figure 8

Comparison of the LSC, SSC, and IR region borders among the chloroplast genomes of six Zygophyllaceae species

The ycf1 gene was located within the SSC/IR border in B. aegyptiaca, T. terrestris, and G. angustifolium but was located within the IRa region in L. tridentata (701 bp). The pseudogene ycf1 was present in the IRb/SSC border in B. aegyptiaca, G. angustifolium and T. terrestris. In contrast, the ycf1 gene was not present in Z. xanthoxylon or T. mongolica; instead, the rpl32 gene was located in the SSC region, close to the SSC/IRa border (45 bp) in T. mongolica, and the rps7 gene was present in the SSC region in Z. xanthoxylon. Both T. mongolica and Z. xanthoxylon possessed two copies of the trnL gene (in the IRa and IRb regions).

The ndhF gene was in the IRb/SSC border of L. tridentata (1 bp), whereas gene ndhF was in the SSC (92 bp) of IRb/SSC border in B. aegyptiaca and T. terrestris (147 bp), while G. angustifolium having ndhF gene in (1701 bp) of the IRb/SSC border. However, the ndhF gene was missing from the SSC region in both Z. xanthoxylon and T. mongolica as well.

The trnH gene showed variation in its location in the IR/LSC border. B. aegyptiaca had the trnH gene in the IRa/LSC border, while it was located in the LSC region and far from the IRb/LSC border in Z. xanthoxylon (540 bp), L. tridentata (108 bp), G. angustifolium (83 bp), and T. terrestris (202 bp). T. mongolica was different; the trnH gene was located in the IRa region far from the IRa/LSC border (128 bp). T. mongolica and Z. xanthoxylon had another copy of the trnH gene; it was present in the IRb region and far from the IRb/SSC border (128 bp) in T. mongolica, and it was in the LSC region far from the IRb/SSC border (656 bp) in Z. xanthoxylon.

Variations in the location of the rps19 gene in the IR/LSC border also occurred, with it being located at different sites in the cp genomes (Fig. 8). The rps19 gene spanned the border of LSC/IR in B. aegyptiaca and Z. xanthoxylon. Z. xanthoxylon duplicated the rps19 gene in the IRa region. rps19 was located in the LSC region, distant to the LSC/IR border in L. tridentata (223 bp) and T. terrestris (16 bp). Both species possessed duplicated rpl2 genes in the IRa and IRb regions. T. mongolica had two copies of the rps19 gene located in the inverted repeat region (IRa and IRb), whereas the rpl22 gene was in the LSC region. The rpl22 and rpl2 genes were located in the LSC and IRa regions in G. angustifolium, respectively.

Characterisation of substitution rate

The value of synonymous (Ks) and nonsynonymous (Ka) substitutions and the Ka/Ks ratio were calculated among the 70 protein-coding genes that represent the common genes in the species that were compared within Zygophyllaceae (T. mongolica, Z. xanthoxylon, L. tridentata, G. angustifolium, and T. terrestris). Several genes were under positive selection with Ka/Ks values > 1, as shown in Fig. 9: atpF, ndhG, petB, petD, psaI, psbH, psbT, rpl2, rps14, rps4, rps7, ycf4, rpl23, and matK. In addition, most Ks values were < 1 in all genes (Fig. 9), except for atpA, atpB, psaC, rpl14, and ycf4.

Fig. 9
figure 9

The synonymous (Ks) and Ka/Ks ratio values of 70 protein-coding genes of the six Zygophyllaceae plastomes

Phylogenetic Analysis

Phylogenetic relationships were assessed among all currently available cp genome sequences in GenBank for the order Zygophyllales (six species from the family Zygophyllaceae), two species from Krameriaceae, and 11 plastome sequences from the orders Sapindales and Santalales, where B. aegyptiaca was classified formerly. Phylogenetic analysis was performed via MP and BI analyses. The two phylogenetic trees were topologically similar, with the majority of nodes having 100% bootstrap (BP) values and 1.00 Bayesian posterior probabilities (PP) (Fig. 10). The monophyly of the Zygophyllales order is strongly supported (PP = 1.00; BS = 100%) by this study.

Fig. 10
figure 10

Phylogenetic tree construction inferred from the coding sequence (CDS) of 20 taxa, using Bayesian Inference (BI) and Maximum Parsimony (MP) methods. The tree shows the relationships between Zygophyllales (Krameriaceae & Zygophyllaceae), Sapindales, and Santalales. The numbers in the branch nodes represent bootstrap (BP) values of Bayesian posterior probabilities (PP)

Zygophyllales was divided into two highly supported clades (PP = 1.00; BS = 100%): one clade contained species from the Krameriaceae family, and the other comprised Zygophyllaceae species. Zygophyllaceae was further divided into three separate clades representing three subfamilies: Tribuloideae (PP = 0.84; BS = 73%), Larreoideae PP = 1.00; BS = 100%), and Zygophylloideae (PP = 1.00; BS = 100%). In this study, B. aegyptiaca was nested in the Zygophyllaceae family and formed a moderately supported clade with T. terrestris, representing the Tribuloideae subfamily.

Discussion

This study presents the first chloroplast genome from B. aegyptiaca, a plant within the Tribuloideae (Zygophyllaceae) subfamily. The length of the cp genome in B. aegyptiaca (155 kb) was similar to that seen in the cp genome of Tribulus terrestris (158 kb), which is also within the Tribuloideae subfamily. Previous studies have [36] indicated a great reduction in plastid genome size in three species of the subfamily Zygophylloideae (Zygophyllaceae), namely, T. mongolica, Z. xanthoxylon, and Z. fabago. The sizes of these genomes ranged between 104 and 106 kb. In the current study, a large decrease in the size of the B. aegyptiaca genome, which has mentioned in regard to some Zygophyllaceae species, was not observed [36]. Size, structure, gene content, and organization are usually conserved in the chloroplast genomes of angiosperms [30]. However, the chloroplast genomes in some species are smaller than those in most other angiosperms. The cp genome in Saguaro (Carnegiea giganteais) considered to the smallest known plastid genome in autotrophic angiosperms having size (113 kb), where lost whole of IR region and ndh genes [37]. As well in Astragalus membranaceus of the Fabaceae whose chloroplast genome size is 124 kb [38]. In addition, to decrease in genome size is a common feature of parasitic plants and is caused by their conversion from an autotrophic to parasitic strategy. This decrease in genome size is accompanied by several other changes, such as pseudogenization, gene loss, structural rearrangement, and size reduction [39, 40]. For example, the plastid genome size in parasitic species of the Loranthaceae family ranges from 116 to 139 kb and the reductions in the plastid sizes of some holoparasites may be even greater. Indeed, the root parasite species in the family Cynomoriaceae has a total plastid genome length of 45,519 bp [41].

The organization, size, and structure of the B. aegyptiaca chloroplast genome is similar to those of other angiosperms; the size of the LSC region is 86 kb, the SSC region is 18 kb, and the two IR regions are 25 kb. The LSC region in angiosperms ranges from 80 to 90 kb, SSC regions are approximately 16–27 kb, and the sizes of the two IRs range from 20 to 28 kb [36].

A typical angiosperm chloroplast genome consists of 113 genes, including 79 protein coding genes, 30 tRNA genes and four rRNA genes [30]. The B. aegyptiaca chloroplast genome had a similar number of genes (107 genes), including 75 protein-coding genes, 28 tRNA genes and 4 rRNA genes. This was similar to other species in the Zygophyllum genus of the family Zygophyllaceae [36], which contained 107 genes, including 75 protein-coding genes, 33 tRNA genes and 4 rRNA genes.

We found that a GC content of 35% in the B. aegyptiaca cp genome was akin to GC content in Zygophyllum spp. [36], which ranged between 33 and 36%. Codon usage bias plays an important role in chloroplast genome evolution and occurs as a result of natural selection and mutations [42, 43]. Codons encoding the amino acid leucine were the most frequently observed in the plastome of B. aegyptiaca, whereas those encoding cysteine were the least frequently observed. This finding has also been reported in the chloroplast genomes of Populus species [44]. While isoleucine is the most encoded amino acid among Zygophyllum spp. (Zygophyllaceae), methionine is a less abundant amino acid [36].

Introns play a significant role in gene expression and regulation [45, 46]. The clpP and ycf3 genes in the B. aegyptiaca plastome had two introns, while the remaining genes contained only one intron; these findings are consistent with previously published reports [47, 48]. The chloroplast genomes of three species of Zygophyllum also contained two introns in the ycf3 gene [36].

RNA editing includes the processes of inserting, deleting, or modifying nucleotides, which leads to changes in the DNA coding sequence during RNA transcription [49]; in turn, this allows the creation of different protein transcripts [50]. Only C-U RNA editing has been described in plastomes [51]. The results obtained from the present study show that in the B. aegyptiaca plastome, most of the transformation in the codons was from the amino acid exchange of serine to leucine and that the ndhB gene had the highest number of editing sites, which could be attributed to high preservation of RNA editing [52, 53].

Prior studies have noted the importance of repeat sequences in the cp genome, which play a significant role in genomic recombination [54, 55]. In the B. aegyptiaca plastome, most repeats are in the ycf2 gene, which is consistent with data obtained from other angiosperms [56, 57].

Several reports [58,59,60,61] have shown the importance of chloroplast SSRs (cpSSRs) as reliable molecular markers to discriminate between specimens at lower taxonomic levels and in studying population structure. In the B. aegyptiaca cp genome, there are 240 SSRs repeats, the largest number compared to five other species of Zygophyllaceae. Mononucleotide A/T SSRs were the most frequent, and most SSRs were marked in noncoding regions (83.3%). Thus, we recommended used plastid genome data in this study for developing cpSSR loci and for studying the levels of genetic variation in the B. aegyptiaca population.

Although the plastid genome is conserved in angiosperm plants as previously reported [30], several studies have reported variations in size and boundaries among IR/LSC and IR/SSC regions and variation in gene location [62, 63]. In the present study, comparisons between the IR-LSC and IR-SSC boundaries in the six complete Zygophyllaceae chloroplast genomes showed clear variation in the inverted repeat region in Zygophyllaceae chloroplast genomes and significant contraction in the IR region in the Z. xanthoxylon and T. mongolica genomes (synonym of Zygophyllum mongolicum). Our results are consistent with recent results published by [36], who mention that the most conspicuous change occurs in two IRs reduced by approximately 16–24 kb in size in the plastid genomes of T. mongolica, Z. xanthoxylon, and Z. fabago.

There is variation in the border of the IR-SC region among the six species of Zygophyllaceae (in this study based on the position of the rps19, rpl2 and trnH genes) at the LSC/IRa junction, which could occur as a result of contraction and expansion in the inverted repeat region. A similar observation of variation in the location of the trnH, rpl2, and rps19 genes at the border was reported for species in the subfamily Acanthoideae by [64]. Furthermore, two copies of the trnL gene appeared in the IRs of both T. mongolica and Z. xanthoxylon. This could be related to IR border changes due to either gene loss, such as ndhF, or expansion/reduction. The trnL genes at the IR borders of Tetraena mongolica and Zygophyllum xanthoxylon are not the same, trnL-CAA and trnL-UAG respectively. The trnL- CAA is usually located in the IR regions and trnL-UAG in the SSC region. We noticed that the ycf1 gene was located in the IRb/SSC border in B. aegyptiaca, G. angustifolium and T. terrestris, and studies [65] have indicated that the ycf1 gene in the IRb region is a pseudogene in the angiosperm chloroplast genome.

In the current comparison, we observed that all the ndh genes that were usually located in the SSC and IR regions were lost in T. mongolica and Z. xanthoxylon. This was similar to the mVISTA results, which also showed a loss of ndh genes (specifically in the sequences from T. mongolica and Z. xanthoxylon). A similar loss of the ndh group (11 genes) from the chloroplast genomes of T. mongolica, Z. xanthoxylon, and Z. fabago was also reported by [36], and it was reported that rRNA genes, which are usually present in IRs, were in the SSC region. In our study, although we used plastome sequences from GenBabk with accession numbers that were different than those used by [36] for the species T. mongolica and Z. xanthoxylon, we observed a similar loss of all ndh genes and the presence of ribosomal proteins codes (rpl32 & rps7) genes in the SSC region in Zygophyllum spp., which could be a result of contraction of the IR regionas. The ndh genes represent a complex group consisting of approximately 30 subunits, and 11 out of 30 subunits are used in encoding the NADH dehydrogenase complex in plant plastids and are involved in photosynthesis [66]. The partial or complete loss of genes associated with photosynthesis (ndh) has been reported in some species of the Cactacaeae [37], Pinaceae [67], and Orchidaceae families [68]. It is also a common phenomenon in hemiparasites and holoparasites of the Santalales and Orobanchaceae [69,70,71,72,73,74].

On the other hand, in a study by [36] on three species of the subfamily Zygophylloideae, the authors indicated a significant loss in chloroplast genes (11 genes), and it was not certain if this loss was because of adaptation by plants to living in arid and semiarid environments or as a result of gene transmission to the nuclear genome. In our study, we present the cp genome of B. aegyptiaca (subfamily Tribuloideae and family Zygophyllaceae), which is adapted to arid and semiarid lands. However, we did not observe a large loss in plastid genes and all 11 ndh genes were present in the Balanites plastome. Another possible explanation could be that gene loss observed in Zygophyllum spp. may be related to the evolutionary history of the subfamily (Zygophylloideae); this point is interesting, and we recommend future study of different species in the family Zygophyllaceae. We also recommend that future studies include species from subfamilies Morkillioideae, Seetzenioideae and Larreoideae in Zygophyllaceae.

The Ka/Ks value is usually used for evaluating sequence variations in different species or taxonomical species with unknown evolutionary status and to detect substitution, selection and beneficial mutations in genes under selective pressure [75]. The values of synonymous (Ks) and nonsynonymous (Ka) substitutions and the Ka/Ks ratio showed that 14 protein-coding genes (atpF, ndhG, petB, petD, psaI, psbH, psbT, rpl2, rps14, rps4, rps7, ycf4, rpl23, and matK) were under positive selection and may have a faster evolutionary rate [57]. Most of these genes play a role in maintaining the efficiency of photosynthesis. In the current study, the ycf4 gene was under positive selection. ycf4 is located in the thylakoid membrane and involved in the assembly of the photosystem I complex [76], it is possible that ycf4 has high substitution rates in arid plant species. Further research is required to investigate the use of these regions in detecting phylogenetic relationships among Zygophyllaceae species.

The plastome consists of many highly efficient genes that may resolve phylogenetic questions at different levels of angiosperm taxonomy [77, 78]. In this study, we found that B. aegyptiaca was distantly related to the family Simaroubaceae. This provides additional evidence to confirm the current position of B. aegyptiaca in the family Zygophyllaceae, as suggested by previous studies [21, 22, 25]. The chloroplast phylogenetic tree showed a strong relationship among Zygophyllaceae species. B. aegyptiaca and T. terrestris formed a group representing the Tribuloideae subfamily, L. tridentata and G. angustifolium (within the same branch) represent the Larreoideae subfamily, while Z. mongolica, Z. fabago, and Z. xanthoxylon were part of the same group and represent the Zygophylloideae subfamily. These findings broadly support the previous results of [1, 21].

Materials and methods

Sample collection and DNA extraction

Fresh leaves of B. aegyptiaca were collected from the Wadi Fatima (Al-Jamoum) Makkah district (21° 38′ 49.6“ N, 39° 41’ 49.3” E) in Saudi Arabia. The plants were identified by Dr. Widad S. Al-Juhani, assistant professor of taxonomy and supervisor of the herbarium in the Biology Department of Umm AlQura University, based on herbarium specimens and morphologies in relevant literature. A sample specimen was prepared and deposited in the herbarium of Umm Al-Qura University, Makkah, with accession number UQU072021. Samples of fresh leaves were dried in silica gel for DNA extraction. DNA was extracted from the silica gel-dried leaves of B. aegyptiaca using the CTAB Plant DNA extraction protocol [79].

Library construction and De novo Genome sequencing

Library construction and sequencing using Illumina sequencing and read length 151 bp paired-ends were carried out by Macrogen (https://dna.macrogen.com/, Seoul, South Korea), with a final yield raw data of 3.5 Gb.

Genome assembly and annotation

The FastQC tool was used to check raw read quality and remove adaptors. A Phred score above 30 was used. Clean reads were processed for genome assembly using NOVOPlasty 4.3.0 version [80] with kmer (K-mer = 33). The contig N50 value was high, and the plastome was assembled using the whole genomic sequence of B. aegyptiaca. Tribulus terrestris (NC_046758.1) was used as a reference in the assembly. Single contigs containing the plastome were generated. Gene prediction and annotation of the B. aegyptiaca chloroplast genome were carried out using the GeSeq tool [81], with default parameters and percent identity cut-off for protein-coding genes and RNAs set at ≥60 and ≤ 85, respectively. tRNA genes were identified with trnAscan-SE version 2.0 [82]. The annotated (gb) format sequence files were used to draw the circular chloroplast genome maps with the OGDRAW tool (Organellar Genome DRAW), version 1.3.1 [83]. The sequence of the chloroplast genome of B. aegyptiaca was deposited in the GenBank database with accession number (OL703321).

Sequence Analysis

The relative synonymous codon usage (RSCU) values, base composition, and codon usage were analysed using MEGA software [84] version 11.0. Potential RNA editing sites present in the protein-coding genes were predicted by the PREP suite [85] with a cut-off value of 0.8.

Repeat Analysis in the chloroplast Genome

The online software MIcroSAtellite (MISA) v2.1 [86] was used to identify simple sequence repeats (SSRs) in the chloroplast genome of B. aegyptiaca and five other species from the Zygophyllaceae family, namely, Guaiacum angustifolium, Larrea tridentata, Tetraena mongolica, Tribulus terrestris, and Zygophyllum xanthoxylon. Parameters eight, five, four and three repeats units were used: eight repeats for mononucleotide, five for dinucleotides, four for trinucleotides, three for each tetranucleotides, pentanucleotides, and hexanucleotides respectively.

In addition, REPuter [85] software was used with default settings to detect the size and location of long palindromic, forward, reverse, and complementary repeats in the B. aegyptiaca cp genome and the genomes of five species from Zygopyllaceae.

Sequence divergence and boundary

Comparison of the genome of B. aegyptiaca with five chloroplast genome sequences from Zygophyllaceae (G. angustifolium, L. tridentata, T. mongolica, T. terrestris, and Z. xanthoxylon; GenBank accession numbers are shown in Table 4) was performed using the mVISTA program [87]; the annotation of B. aegyptiaca was used as a reference in the Shuffle-LAGAN mode. Furthermore, comparisons between the borders of the IR, SSC, and LSC regions were generated using IRSCOPE [88].

Table 4 Accession numbers of chloroplast genome analysed in the study

Characterisation of the substitution rate

The methods for estimating nonsynonymous and synonymous substitution rates (Ka and Ks), selection, and beneficial mutations among protein-coding sequences followed. Nonsynonymous (Ka) substitution, synonymous (Ks) substitution, and Ka/Ks ratios were calculated to detect variable mutation rates across chloroplast genom sequence, that contain an important information related to evolutionary history in B. aegyptiaca compared with those in the five aforementioned Zygophyllaceae species. We employed Ka/Ks Calculator version 2.0 [75] with default parameters, and the Nei and Gojobori substitution model was used.

Phylogenetic Analysis

Phylogenetic analysis was conducted based on the cp genome sequences of the members of the Zygophyllales order, including cp genome sequences available in GenBank of species from the families Zygophyllaceae and Krameriaceae.

Because an older taxonomy placed Balanites aegyptiaca in the orders Sapindales and Santalales, the current phylogenetic analysis included sequesters of the cp genome of some species belonging to these orders. Information on species names, families, and GenBank accessions is available in Table 4. Two species from the Malvaceae family (Malva parviflora and Malva wigandii) were used as an outgroup.

All the common genes from the organisms were retrieved and the coding sequence CDS were joined in order. The sequences were further aligned using MAFFT version 7.475 [89]. Then phylogenetic trees were reconstructed based on the maximum parsimony (MP) method using 1000 bootstrap values and Mega software [84] version 11.0. The MP search method consisted of subtree-pruning-regrafting, with the number of initial trees (random addition) set at 10 and the number of threads set at 5.

The optimal evolutionary model was the GTR + I + G model, as calculated by ModelFinder using [90] Akaike’s information criterion (AIC) [91]. Bayesian inference analysis (BI) was conducted using MrBayes v. 3.2.6 [92, 93] in CIPRES Science Gateway 3.3 [94]. BI analysis included two separate runs; each of four Markov chain Monte Carlo chains was run for 10 million generations with sampling every 10,000 generations. Trees from the first 25% of the sampled generations were discarded as burn-in. The convergence of the runs was tested by using the effective sample size (ESS), calculated with Tracer v1.7.1 [95], with ESS values greater than 200 for all parameters considered good evidence. The majority rule (> 50%) consensus tree of BI was visualised using FigTree 1.4.3 [96].

Conclusions

The aim of the present research was to provide the complete chloroplast genome of B. aegyptiaca, a plant in subfamily Tribuloideae and family Zygophyllaceae, which has medical and nutritional importance and plays a key role in ecosystem conservation in arid lands. We compared the cp genomes of available genera and species in the Zygopyllaceae family to assess the systematic relationships within the family and between related families as well as the genome conservation state.

This study confirmed the taxonomic status of the species B. aegyptiaca as a member of the Zygophyllaceae family. This study did not record loss in the chloroplast genome of B. aegyptiaca, as mentioned for species in subfamily Zygophylloideae. However, plastomes vary mainly at the SC/IR boundary among members of Zygophyllaceae, and there is clear genome reduction and gene loss in some species of Zygopyllaceae. Thus, we recommended further study to investigate changes that could have occurred in the structure of the genome during the evolutionary history of the family. It is necessary that future studies also include samples from subfamilies Morkillioideae, Seetzenioideae and Larreoideae in Zygophyllaceae.

Availability of data and materials

The data presented in this study are available in this article and Supplementary Material. The complete chloroplast genome sequence of Balanites aegyptiaca was deposited in GenBank at https://www.ncbi.nlm.nih.gov, (accession numbers: OL703321).

Abbreviations

Cp:

Chloroplast

LSC:

Large single copy region

SSC:

Small single copy region

IR:

Inverted repeat

RSCU:

Relative synonymous codon usage

SSR:

Simple sequence repeats

IGS:

Intergenic spacer

CNS:

Conserved non coding sequence

cpSSRs:

Chloroplast simple sequence repeats

RAPD:

Random Amplified Polymorphic DNA

MP:

Maximum Parsimony

BP:

Bootstrap percentage

BI:

Bayesian Inference

PP:

Posterior probability

Ks:

Synonymous

Ka:

Non-synonymous

CTAB:

Cetrimonium bromide

PCR:

Polymerase chain reaction

References

  1. Godoy-Bürki AC, Acosta JM, Aagesen L. Phylogenetic relationships within the New World subfamily Larreoideae (Zygophyllaceae) confirm polyphyly of the disjunct genus Bulnesia. Syst Biodivers. 2018;16(5):453–68 https://www.tandfonline.com/doi/abs/10.1080/14772000.2018.1451406.

    Article  Google Scholar 

  2. Al-Thobaiti SA, Abu ZI. Medicinal properties of desert date plants (Balanites aegyptiaca)-an overview. Glob J Pharmacol. 2018;12(1):01–12.

    CAS  Google Scholar 

  3. Orwa C, Mutua A, Kindt R, Jamnadass R, Anthony S. Agroforestree database: a tree reference and selection guide; version 4.0. World Agroforestry Centre, Kenya; 2009.

  4. Chothani DL, Vaghasiya H. A review on Balanites aegyptiaca Del (desert date): phytochemical constituents, traditional uses, and pharmacological activity. Pharmacogn Rev. 2011;5(9):55–62. https://doi.org/10.4103/0973-7847.79100.

    Article  PubMed  PubMed Central  Google Scholar 

  5. Abdelaziz SM, Medraoui L, Alami M, Pakhrou O, Makkaoui M, Mohamed AO, et al. Inter simple sequence repeat markers to assess genetic diversity of the desert date (Balanites aegyptiaca Del.) for Sahelian ecosystem restoration. Sci Rep. 2020;10(1):1–8 https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7486294/.

    Google Scholar 

  6. Gardette J-L, Baba M. FTIR and DSC studies of the thermal and photochemical stability of Balanites aegyptiaca oil (Toogga oil). Chem Phys Lipids. 2013;170(171):1–7. https://doi.org/10.1016/j.chemphyslip.2013.02.008.

    Article  CAS  PubMed  Google Scholar 

  7. Abdoulaye B, Béchir AB, Mapongmetsem PM. Morphological variability of Balanites aegyptiaca (L.) Del. In the region of Ouaddaï in Chad Int. J. Biol. Chem. Sci. 2016;10(4):1733–46 http://ajol.info/index.php/ijbcs.

    Google Scholar 

  8. Thomas J. Plant Diversity in Saudi Arabia. http://www.plantdiversityofsaudiarabia.info/index.htm. 2011. Accessed 10 Sept 2021.

  9. Sands MJS. The desert date and its relatives: a revision of the genus Balanites. Kew Bull. 2001;56(1):1–128 http://www.jstor.org/stable/4119431.

    Article  Google Scholar 

  10. Hall JB, Walker DH. Balanites aegyptiaca: a monograph. Bangor: School of Agricultural and Forest Sciences, University of Wales; 1991.

    Google Scholar 

  11. Harms HAT. Vorschlag zur Erginzung der "Lois de la nomenclature botanique de 1867". Notizbl K6nigl Bot Gart Berlin-Dahlem, App. 1904;13:1–37.

    Google Scholar 

  12. Bentham G, Hooker JD. Simarubeae. In: Genera Plantarum. London; Lovell Reeve & Co., London, 1862, 1(1).

  13. Engler HGA. Zygophyllaceae. In: Engler HGA, Prantl, K, Die Natfirlichen Pflanzenfamilien, Leipzig; W Engelmann 1896: , 3(4).

  14. Cronquist A. Evolution and classification of flowring plants. London & Edinburgh. 1968.

  15. Hegnauer R. Chemotaxonomy der pflanzen. Basel und Stuttgart: Birkhäuser Verlag; 1973. p. 6.

    Book  Google Scholar 

  16. Scholz H. Zygophyllaceae. In: Melchior, H, 12th edition (ed). A. Vol. 2. Berlin: Engler's Syllabus der Pfannzenfamilien;1964: 251-252c.

  17. Cronquist AJ. An integrated system of classification of flowering plants. New York, Columbia university press, 1981.

  18. Maksoud SA, El-Hadidi MN. The flavonoids of Balanites aegyptiaca from Egypt. Plant Syst Evol. 1988;160:153–8.

    Article  CAS  Google Scholar 

  19. Sheahan MC, Culter DF. Contribution of vegetation anatomy to systematics of the Zygophyllaceae r.br. Bot J Linn Soc. 1993;113:227–62.

    Article  Google Scholar 

  20. Boesewinkel FD. Ovule and seed characters of Balanites aegyptiaca and the classification of the Linales-Geraniales-Polygalales assembly. Acta Bot Neerl. 1994;43(1):15–25. https://doi.org/10.1111/j.1438-8677.1994.tb00730.x.

    Article  Google Scholar 

  21. Sheahan MC, Chase MW. Phylogenetic relationships within Zygophyllaceae based on DNA sequences of three plastid regions, with special emphasis on Zygophylloideae. Syst Bot. 2000;25(2):371–84. https://doi.org/10.2307/2666648.

    Article  Google Scholar 

  22. Singh KK, Samanta AK, Kundu SS, Sharma D. Evaluation of certain feed resources for carbohydrate and protein fractions and in situ digestion characteristics. Indian J Anim Sci. 2002;72(9):794–7.

    Google Scholar 

  23. Sands MJS. Flora of tropical East Africa: Balanitaceae. In: Beentje HJ, editor. and Ghazanfar S.a. (subed.). Royal Botanic Gardens, Kew: Flora of tropical East Africa; 2013. p. 1–17.

    Google Scholar 

  24. Sheahan MC, Chase MW. A phylogenetic analysis of Zygophyllaceae R. Br. Based on morphological, anatomical and rbc L DNA sequence data. Bot J Linn Soc. 1996;122(4):279–300. https://doi.org/10.1111/j.1095-8339.1996.tb02077.x.

    Article  Google Scholar 

  25. Boulos L. Fora of Egypt, checklist. Cairo: Al-Hadara Publishing; 2000.

    Google Scholar 

  26. Sands MJS. Balanitaceae, in Flora of Ethiopia and Eritrea S. Edwards, M. Tadesse, and I. Hedberg, editors. National Herbarium, Addis Ababa; Department of Systematic Botany: Uppsala University; 1989.

    Google Scholar 

  27. Amer WM, Soliman MM, Sheded MM. Biosystematics studies for Balanaties aegyptica (Balanitaceae) populations in Egypt. Flora Mediterr. 2002;12:353–67.

    Google Scholar 

  28. Palmer JD. Comparative organization of chloroplast genomes. Annu Rev Genet. 1985;19:325–54. https://doi.org/10.1146/annurev.ge.19.120185.001545.

    Article  CAS  PubMed  Google Scholar 

  29. 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. https://doi.org/10.3732/ajb.94.3.275 PMID 21636401.

    Article  CAS  PubMed  Google Scholar 

  30. 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;76:273–97. https://doi.org/10.1007/s11103-011-9762-4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Straub SCK, Parks M, Weitemier K, Fishbein M, Cronn RC, Liston A. Navigating the tip of the genomic iceberg: next-generation sequencing for plant systematics. Am J Bot. 2012;99:349–64.

    Article  CAS  PubMed  Google Scholar 

  32. Liu HJ, Ding CH, He J, Cheng J, Pei LY, Xie L. Complete chloroplast genomes of Archiclematis, Naravelia and Clematis (Ranunculaceae), and their phylogenetic implications. Phytotaxa. 2018;343:214–26.

    Article  Google Scholar 

  33. Sage RF, Christin P-A, Edwards EJ. The C4 plant lineages of planet earth. J Exp Bot. 2011;62(9):3155–69 https://academic.oup.com/jxb/article/62/9/3155/474202.

    Article  CAS  PubMed  Google Scholar 

  34. Bellstedt DU, Galley C, Pirie MD, Linder HP. The migration of the palaeotropical arid flora: Zygophylloideae as an example. Syst Bot. 2012;37:951–9.

    Article  Google Scholar 

  35. Wu SD, Lin L, Li HL, Yu SX, Zhang LJ, Wang W. Evolution of Asian interior arid-zone biota: evidence from the diversification of Asian Zygophyllum (Zygophyllaceae). PLoS One. 2015;10(9):e0138697. https://doi.org/10.1371/journal.pone.0138697.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Wang X, Dorjee T, Chen Y, Gao F, Zhou Y. The complete chloroplast genome sequencing analysis revealed an unusual IRs reduction in three species of subfamily Zygophylloideae. Plos One. 2022;17(2):e0263253. https://doi.org/10.1371/journal.pone.0263253.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Sanderson MJ, Copetti D, Búrquez A, Bustamante E, Charboneau JLM, Eguiarte LE, et al. Exceptional reduction of the plastid genome of saguaro cactus ( Carnegiea gigantea ): loss of the ndh gene suite and inverted repeat. Am J Bot. 2015;102(7):1115–27. https://doi.org/10.3732/ajb.1500184.

    Article  CAS  PubMed  Google Scholar 

  38. Lei WJ, Ni DP, Wang YJ, Shao JJ, Wang XC, Yang D, et al. Intraspecific and heteroplasmic variations, gene losses and inversions in the chloroplast genome of Astragalus membranaceus. Sci Rep. 2016;6:21669. https://doi.org/10.1038/srep21669 PMID: 26899134.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Wicke S, Müller KF, Claude WD, 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;113:9045–50. https://doi.org/10.1073/pnas.1607576113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Wicke S, Naumann J. Molecular evolution of plastid genomes in parasitic flowering plants. Adv Bot Res. 2018;85:315–47. https://doi.org/10.1016/bs.abr.2017.11.014.

    Article  CAS  Google Scholar 

  41. Bellot S, Cusimano N, Luo S, Sun G, Zarre S, Groger A, et al. Assembled plastid and mitochondrial genomes, as well as nuclear genes, place the parasite family Cynomoriaceae in the Saxifragales. Genome Biol. Evol. 2016;8(7):2214–30. https://doi.org/10.1093/gbe/evw147.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Staden R, McLachlan AD. Codon preference and its use in identifying protein coding regions in long DNA sequences. Nucleic Acids Res. 1982;10:141–56. https://doi.org/10.1093/nar/10.1.141 PMID: 7063399.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Li B, Lin F, Huang P, Guo W, Zheng Y. Complete chloroplast genome sequence of Decaisnea insignis: Genome organization, genomic resources and comparative analysis. Sci Rep. 2017;7(1):1–10. https://doi.org/10.1038/s41598-017-10409-8.

    Article  CAS  Google Scholar 

  44. Zong D, Zhang Y, Zou X, Li D, Duan A, He C. Characterization of the complete chloroplast genomes of five Populus species from the western Sichuan plateau, Southwest China: comparative and phylogenetic analyses. PeerJ. 2019;7:e6386. https://doi.org/10.7717/peerj.6386.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Mattick JS, Gagen MJ. The evolution of controlled multitasked gene networks: the role of introns and other noncoding RNAs in the development of complex organisms. Mol Biol Evol. 2001;18:1611–30. https://doi.org/10.1093/oxfordjournals.molbev.a003951 PMID: 11504843.

    Article  CAS  PubMed  Google Scholar 

  46. Taberlet P, Coissac E, Pompanon F, Gielly L, Miquel C, Valentini A, et al. Power and limitations of the chloroplast trnL (UAA) intron for plant DNA barcoding. Nucleic Acids Res. 2007, 35(3):e14. doi: https://doi.org/10.1093/nar/gkl938. Epub 2006 Dec 14. PMID: 17169982; PMCID: PMC1807943.

  47. Raman G, Park S. The complete chloroplast genome sequence of Ampelopsis: gene organization, comparative analysis, and phylogenetic relationships to other angiosperms. Front Plant Sci. 2016;7:341. https://doi.org/10.3389/fpls.2016.00341.

    Article  PubMed  PubMed Central  Google Scholar 

  48. Alzahrani D, Albokhari E, Yaradua S, Abba A. Complete chloroplast genome sequences of Dipterygium glaucum and Cleome chrysantha and other Cleomaceae species, comparative analysis and phylogenetic relationships. Saudi J Biol Sci. 2021;28(4):2476–90. https://doi.org/10.1016/j.sjbs.2021.01.049.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Mower JP. The PREP suite: predictive RNA editors for plant mitochondrial genes, chloroplast genes and user-defined alignments. Nucleic Acids Res. 2009;37:253–9. https://doi.org/10.1093/nar/gkp337.

    Article  CAS  Google Scholar 

  50. Bundschuh R, Altmuller J, Becker C, Nurnberg P, Gott JM. Complete characterization of the edited transcriptome of the mitochondrion of Physarum polycephalum using deep sequencing of RNA. Nucleic Acids Res. 2011;39:6044–55.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Smith DR. Unparalleled variation in RNA editing among Selaginella Plastomes. Plant Physiol. 2020;182(1):12–4. https://doi.org/10.1104/pp.19.00904.

    Article  CAS  PubMed  Google Scholar 

  52. Wang W, Yu H, Wang J, Lei W, Gao J, Qiu X, et al. The complete chloroplast genome sequences of the medicinal plant Forsythia suspensa (Oleaceae). Int J Mol Sci. 2017;18(11):2288. https://doi.org/10.3390/ijms18112288.

    Article  CAS  PubMed Central  Google Scholar 

  53. Park M, Park H, Lee H, Lee B, Lee J. The complete plastome sequence of an Antarctic bryophyte Sanionia uncinata (Hedw.) Loeske. Int J Mol Sci. 2018;19:709. https://doi.org/10.3390/ijms19030709.

    Article  CAS  PubMed Central  Google Scholar 

  54. Asano T, Tsudzuki T, Takahashi S, Shimada H, Ki K. Complete nucleotide sequence of the sugarcane (Saccharum officinarum) chloroplast genome: a comparative analysis of four monocot chloroplast genomes. DNA Res. 2004;11(2):93–9. https://doi.org/10.1093/dnares/11.2.93.

    Article  CAS  PubMed  Google Scholar 

  55. Weng ML, Ruhlman TA, Jansen RK. Expansion of inverted repeat does not decrease substitution rates in Pelargonium plastid genomes. New Phytol. 2017;214(2):842–51. https://doi.org/10.1111/nph.14375.

    Article  CAS  PubMed  Google Scholar 

  56. Yang Y, Zhou T, Duan D, Yang J, Feng L, Zhao G. Comparative Analysis of the complete chloroplast genomes of five Quercus species. Front Plant Sci. 2016;7(959) https://www.frontiersin.org/article/10.3389/fpls.2016.00959.

  57. Zhou T, Ruhsam M, Wang J, Zhu H, Li W, Zhang X, et al. The complete chloroplast Genome of Euphrasia regelii, Pseudogenization of ndh genes and the Phylogenetic Relationships within Orobanchaceae. Front Genet. 2019;10(444):1–15 https://www.frontiersin.org/article/10.3389/fgene.2019.00444.

    Google Scholar 

  58. Provan J, Powell W, Hollingsworth PM. Chloroplast microsatellites: new tools for studies in plant ecology and evolution. Trends Ecol Evol. 2001;16(3):142–7. https://doi.org/10.1016/s0169-5347(00)02097-8.

    Article  CAS  PubMed  Google Scholar 

  59. Yang A-H, Zhang J-J, Yao X-H, Huang H-W. Chloroplast microsatellite markers in Liriodendron tulipifera (Magnoliaceae) and cross-species amplification in L. chinense. Am J Bot. 2011;98(5):e123–6 https://bsapubs.onlinelibrary.wiley.com/doi/abs/10.3732/ajb.1000532.

    Article  CAS  PubMed  Google Scholar 

  60. Xue J, Wang S, Zhou S-L. Polymorphic chloroplast microsatellite loci in Nelumbo (Nelumbonaceae). Am J Bot. 2012;99(6):e240–4 https://bsapubs.onlinelibrary.wiley.com/doi/abs/10.3732/ajb.1100547.

    Article  PubMed  Google Scholar 

  61. Hu Y, Woeste KE, Zhao P. Completion of the chloroplast genomes of five Chinese Juglans and their contribution to chloroplast phylogeny. Front Plant Sci. 2016;7(1955) https://www.frontiersin.org/article/10.3389/fpls.2016.01955.

  62. Ruhsam M, Clark A, Finger A, Wulff AS, Mill RR, Thomas PI, et al. Hidden in plain view: cryptic diversity in the emblematic Araucaria of New Caledonia. Am J Bot. 2016;103(5):888–98. https://doi.org/10.3732/ajb.1500487.

    Article  CAS  PubMed  Google Scholar 

  63. Wang RJ, Cheng CL, Chang CC, Wu C-L, Su T-M, Chaw S-M. Dynamics and evolution of the inverted repeat-large single copy junctions in the chloroplast genomes of monocots. BMC Evol Biol. 2008;8(1):3650. https://doi.org/10.1186/1471-2148-8-36.

    Article  CAS  Google Scholar 

  64. Alzahrani DA, Yaradua SS, Albokhari EJ, Abba A. Complete chloroplast genome sequence of Barleria prionitis, comparative chloroplast genomics and phylogenetic relationships among Acanthoideae. BMC Genomics. 2020;21:393. https://doi.org/10.1186/s12864-020-06798-2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Yao X, Tan Y-H, Liu Y-Y, Song Y, Yang J-B, Corlett RT. Chloroplast genome structure in Ilex (Aquifoliaceae). Sci Rep. 2016;6:28559.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Raubeson LA, Peery R, Chumley TW, Dziubek C, Fourcade HM, Boore JL. Comparative chloroplast genomics: analyses including new sequences from the angiosperms Nuphar advena and Ranunculus macranthus. BMC Genomics. 2007;8(1):174. https://doi.org/10.1186/1471-2164-8-174.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Braukmann TWA, Kuzmina M, Stefanović S. Loss of all plastid ndh genes in Gnetales and conifers: extent and evolutionary significance for the seed plant phylogeny. Curr Genet. 2009;55:323–37.

    Article  CAS  PubMed  Google Scholar 

  68. Dong WL, Wang RN, Zhang NY, Fan WB, Fang MF, Li Z-H. Molecular evolution of chloroplast genomes of orchid species: insights into Phylogenetic relationship and adaptive evolution. Int J Mol Sci. 2018;19:716. https://doi.org/10.3390/ijms19030716 PMID: 29498674.

    Article  CAS  PubMed Central  Google Scholar 

  69. Friedrich T, Steinmüller K, Weiss H. The proton–pumping respiratory complex I of bacteria and mitochondria and its homologue in chloroplasts. FEBS Lett. 1995;367:107–11. https://doi.org/10.1016/0014-5793(95)00548-n.

    Article  CAS  PubMed  Google Scholar 

  70. Petersen G, Cuenca A, Seberg O. Plastome evolution in Hemiparasitic mistletoes. Genome Biol Evol. 2015;7:2520–32. https://doi.org/10.1093/gbe/evv165.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Shin HW, Lee NS. Understanding plastome evolution in Hemiparasitic Santalales: complete chloroplast genomes of three species, Dendrotrophe varians, Helixanthera parasitica, and Macrosolen cochinchinensis. PLoS One. 2018;13. https://doi.org/10.1371/journal.pone.0200293.

  72. Zhu ZX, Wang JH, Cai YC, Zhao KK, Moore MJ, Wang HF. Complete plastome sequence of Erythropalum scandens (Erythropalaceae), an edible and medicinally important liana in China. Mitochondr DNA Part B. 2018;3:139–40. https://doi.org/10.1080/23802359.2017.1413435.

    Article  Google Scholar 

  73. Su HJ, Hu JM. The complete chloroplast genome of hemiparasitic flowering plant Schoepfia jasminodora. Mitochondr DNA Part B. 2016;1:767–9. https://doi.org/10.1080/23802359.2016.1238753.

    Article  Google Scholar 

  74. Frailey DC, Chaluvadi SR, Vaughn JN, Goatney CG, Bennetzen JL. Gene loss and genome rearrangement in the plastids of five Hemiparasites in the family Orobanchaceae. BMC Plant Biol. 2018;18:30. https://doi.org/10.1186/s12870-018-1249-x.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Wang D, Zhang Y, Zhang Z, Zhu J, Yu J. KaKs_Calculator 2.0: a toolkit incorporating gamma-series methods and sliding window strategies. GPB. 2010;8(1):77–80. https://doi.org/10.1016/S1672-0229(10)60008-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Boudreau E, Takahashi Y, Lemieux C, Turmel M, Rochaix JD. The chloroplast ycf3 and ycf4 open reading frames of Chlamydomonas reinhardtii are required for the accumulation of the photosystem I complex. EMBO J 1997,16 (20): 6095–6104. doi:https://doi.org/10.1093/emboj/16.20.6095. PMC 1326293. PMID 9321389.

  77. Yaradua SS, Alzahrani DA, Albokhary EJ, Abba A, Bello A. Complete Chloroplast Genome Sequence of Justicia flava: Genome Comparative Analysis and Phylogenetic Relationships among Acanthaceae. Biomed Res. Int. 2019: 17 pages, Article ID 4370258. https://doi.org/10.1155/2019/4370258.

  78. Dong W, Xu C, Li W, Xie X, Lu Y, Liu Y, et al. Phylogenetic resolution in Juglans based on complete chloroplast genomes and nuclear DNA sequences. Front Plant Sci. 2017;8:1148. https://doi.org/10.3389/fpls.2017.01148.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Doyle J, Doyle J. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem Bull. 1987;19(1):11–5.

    Google Scholar 

  80. Dierckxsens N, Mardulyn P, Smits G. NOVOPlasty: de novo assembly of organelle genomes from whole genome data. Nucleic Acids Res. 2016. https://doi.org/10.1093/nar/gkw955.

  81. Tillich M, Lehwark P, Pellizzer T, Ulbricht-Jones ES, Fischer A, Bock R, et al. GeSeq – versatile and accurate annotation of organelle genomes. Nucleic Acids Res. 2017;45(W1):W6–W11 https://chlorobox.mpimp-golm.mpg.de/geseq.html.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Chan PP, Lowe TM. tRNAscan-SE: searching for tRNA genes in genomic sequences. In: Kollmar M, editor. Gene prediction: methods and protocols. New York: Springer; 2019. p. 1–14.

    Google Scholar 

  83. Greiner S, Lehwark P, Bock R. OrganellarGenomeDRAW (OGDRAW) version 1.3.1: expanded toolkit for the graphical visualization of organellar genomes. Nucleic Acids Res. 2019;47(W1):W59–64. https://doi.org/10.1093/nar/gkz238.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Kumar S, Stecher G, Tamura K. MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol Biol Evol. 2016;33(7):1870–4. https://doi.org/10.1093/molbev/msw054.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Kurtz S, Choudhuri JV, Ohlebusch E, Schleiermacher C, Stoye J, Giegerich R. REPuter: the manifold applications of repeat analysis on a genomic scale. Nucleic Acids Res. 2001;29:4633–42. https://doi.org/10.1093/nar/29.22.4633.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Thiel T, Michalek W, Varshney R. Exploiting EST databases for the development and characterization of gene-derived SSR-markers in barley (Hordeum vulgare L.). Theor Appl Genet. 2003;106:411–22.

    Article  CAS  PubMed  Google Scholar 

  87. Mayor C, Brudno M, Schwartz JR, Poliakov A, Rubin EM, Frazer KA, et al. VISTA: visualizing global DNA sequence alignments of arbitrary length. Bioinformatics. 2000;16:1046. https://doi.org/10.1093/bioinformatics/16.11.1046.

    Article  CAS  PubMed  Google Scholar 

  88. Amiryousefi A, Hyvönen J, Poczai P. IRscope: an online program to visualize the junction sites of chloroplast genomes. Bioinformatics. 2018;34(17):3030–1. https://doi.org/10.1093/bioinformatics/bty220.

    Article  CAS  PubMed  Google Scholar 

  89. Madeira F, Pearce M, Tivey ARN, Basutkar P, Lee J, Edbali O, et al. Search and sequence analysis tools services from EMBL-EBI in 2022. Nucleic Acids Res. 2022;50(W1):W276–9. https://doi.org/10.1093/nar/gkac240 Online ahead of print.

    Article  PubMed Central  Google Scholar 

  90. Kalyaanamoorthy S, Minh BQ, Wong TKF, Haeseler AV, Jermiin LS. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat Methods. 2017;14(6):587–9. https://doi.org/10.1038/nmeth.4285https://doi.org/10.1038/nmeth.4285.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Akaike H. A new look at the statistical model identification. IEEE. 1974;19(6):716–23. https://doi.org/10.1109/TAC.1974.1100705.

    Article  Google Scholar 

  92. Huelsenbeck JP, Ronquist F, Nielsen R, Bollback JP. MRBAYES: Bayesian inference of phylogeny. Bioinformatics. 2001;17:754–5.

    Article  CAS  PubMed  Google Scholar 

  93. Ronquist F, Huelsenbeck JP. MRBAYES 3: Bayesian phylogenetic inference under mixed models. Bioinformatics. 2003;19:1572–4. https://doi.org/10.1093/bioinformatics/btg180.

    Article  CAS  PubMed  Google Scholar 

  94. Miller MA, Pfeiffer W, Schwartz T. Creating the CIPRES science gateway for inference of large phylogenetic trees. In: 2010 gateway computing environments workshop (GCE): Ieee; 2010.

    Google Scholar 

  95. Rambaut A, Drummond AJ, Xie D, Baele G, Suchard MA. Posterior summarization in Bayesian Phylogenetics using tracer 1.7. Syst Biol. 2018;67(5):901–4. https://doi.org/10.1093/sysbio/syy032.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Rambaut A, Drummond A. FigTree v. 1.4.0; 2012.

    Google Scholar 

Download references

Acknowledgments

Not applicable.

Funding

This research received no external funding.

Author information

Authors and Affiliations

Authors

Contributions

WAJ and SAA collected the data, designed and performed the experiment, WAJ, SAA, NMA, and AYA analyzed the data, investigation and drafted the manuscript, WAJ supervised the project, all the authors edited and agreed to the published version of the manuscript.

Corresponding author

Correspondence to Widad S. AL-Juhani.

Ethics declarations

Ethics approval and consent to participate

The experiment was conducted with relevant institutional, national, and international guidelines and legislation. Permission was obtained to collect the samples.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interest.

Additional information

Publisher’s Note

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

Supplementary Information

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

AL-Juhani, W.S., Alharbi, S.A., Al Aboud, N.M. et al. Complete chloroplast genome of the desert date (Balanites aegyptiaca (L.) Del. comparative analysis, and phylogenetic relationships among the members of Zygophyllaceae. BMC Genomics 23, 626 (2022). https://doi.org/10.1186/s12864-022-08850-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-022-08850-9

Keywords