ARTICLE   Open Access    

Characterization and comparative analysis of the first mitochondrial genome of Michelia (Magnoliaceae)

  • # Authors contributed equally: Suyan Wang, Jing Qiu

More Information
  • Genus Michelia has various functions and is valuable in medicine, food, and agriculture. Many plastid genomes (plastomes) of Michelia have been released, but no mitochondrial genomes (mitogenomes) have been reported. In this study, using third-generation HIFI sequencing techniques, Michelia figo (M. figo) mitogenome was de novo assembled into a circular chromosome spanning 773,377 bp with a total GC content of 46.83%. Sixty six genes in total were annotated, including 41 protein-coding genes, 21 tRNA genes, and three rRNA genes. The mitogenome contains 1,514 dispersed repeats (> 30 bp), 39 tandem repeats, and 262 simple sequence repeats. Eighty one fragments originating from the M. figo plastome were detected in its mitogenome and three tRNA genes (trnD-GUC, trnW-CCA, and trnV-GAC) completely transferred from the plastome to the mitogenome. Repeats and collinearity analyses of four Magnoliaceae mitogenomes reveal substantial structural variations, a relatively low degree of collinearity, and significant genetic diversity of this genus. Phylogenetic analysis showed that two phylogenetic trees constructed separately based on mitogenomes and plastomes accurately depict the phylogenetic relationship of M. figo. This study offers the first comprehensive comparative genomic and phylogenetic analysis of the M. figo mitogenome, facilitating the development of genetic markers, taxonomic classification, and resource exploration within the Michelia genus.
  • Plants are continuously subjected to unpredictable environmental conditions and encounter a multitude of stressors throughout their growth and development, posing a significant challenge to global crop production and food security[1]. Heat and drought are undoubtedly the two most important stresses that have a huge impact on crops. Both elicit a wide array of biochemical, molecular, and physiological alterations and responses, impacting diverse cellular processes and ultimately influencing crop yield and quality[2].

    A primary physiological consequence of both stresses is the diminished photosynthetic capacity, partially resulting from the degradation of chlorophyll due to leaf senescence under stress conditions. Chlorophyll accumulation was diminished in numerous plants subjected to drought or heat stress conditions[3,4]. Various environmental stresses prompt excessive generation of reactive oxygen species (ROS), initiating oxidative damage that compromises lipids, and proteins, and poses a serious threat to cellular functions[2]. To mitigate oxidative stress and minimize damage, plants have developed various protective mechanisms to neutralize ROS. Several antioxidant enzymes, such as SOD, POD, and CAT, are integral to cellular antioxidative defense mechanisms. Additionally, antioxidants such as anthocyanins and proline serve as crucial ROS scavengers[5,6]. The elevation in temperature typically induces the transient synthesis of heat shock proteins (Hsps), which function as molecular chaperones in protecting proteins from denaturation and aggregation, with their activity primarily regulated at the transcriptional level by heat shock factors (Hsfs)[7]. The significance of Hsps and Hsfs in all organisms, including plants, has been assessed in various stress conditions that could disrupt cellular homeostasis and result in protein dysfunction[7]. Drought stress can also trigger the transcription of a suite of marker genes, including RD29A, RD29B, NCED3, AREB1, Rab18, etc., which assist plants in mitigating cellular damage during dehydration and bolstering their resilience to stress[810].

    Previous research efforts focusing on the regulatory control of stress-related genes have largely centered around protein-coding genes. In recent years, non-protein-coding transcripts have emerged as important regulatory factors in gene expression. Among them, long non-coding RNAs (lncRNAs) lncRNAs have been identified as implicated in various abiotic stresses[11,12]. LncRNAs are a class of non-coding RNAs (ncRNAs) exceeding 200 nucleotides in length. They possess minimal or no protein-coding potential[13]. In plants, lncRNAs are specifically transcribed by RNA polymerases Pol IV, Pol V, Pol II, and Pol III[14,15]. LncRNAs exhibit low abundance and display strong tissue and cellular expression specificity relative to mRNAs. Moreover, sequence conservation of lncRNAs is was very poor across different plant species[13,16,17]. The widespread adoption of high-throughput RNA sequencing technology has revealed lncRNAs as potential regulators of plant development and environmental responses. In cucumber, RNA-seq analysis has predicted 2,085 lncRNAs to be heat-responsive, with some potentially acting as competitive endogenous RNAs (ceRNAs) to execute their functions[18]. In radish, a strand-specific RNA-seq (ssRNA-seq) technique identified 169 lncRNAs that were differentially expressed following heat treatment[19]. In Arabidopsis, asHSFB2a, the natural antisense transcript of HSFB2a was massively induced upon heat stress and exhibited a counteracted expression trend relative to HSFB2a. Overexpression of asHSFB2a entirely suppressed the expression of HSFB2a and impacted the plant's response to heat stress[20]. For drought stress resistance, 244 lncRNAs were predicted in tomatoes to be drought responsive probably by interacting with miRNAs and mRNAs[21]. Under drought stress and rehydration, 477 and 706 lncRNAs were differentially expressed in drought-tolerant Brassica napus Q2 compared to drought-sensitive B. napus, respectively[22]. In foxtail millet and maize, 19 and 644 lncRNAs, respectively, were identified as drought-responsive[23,24]. Despite the identification of numerous lncRNAs by high-throughput sequencing, which suggests their potential involvement in various abiotic stress processes, only a minority have been experimentally validated for function.

    In our previous study, we characterized 1,229 differentially expressed (DE) lncRNAs in Chinese cabbage as heat-responsive, and subsequent bioinformatics analysis reduced this number to 81, which are more likely associated with heat resistance[25]. lnc000283 and lnc012465 were selected from among them for further functional investigation. The findings indicated that both lnc000283 and lnc012465 could be promptly induced by heat shock (HS). Overexpression of either lnc000283 or lnc012465 in Arabidopsis plants enhanced their capacity to tolerate heat stress. Additionally, both lnc000283 and lnc012465 conferred drought tolerance to transgenic Arabidopsis.

    The lncRNA sequences examined in this study were from Chiifu-401-42 Chinese cabbage and all Arabidopsis plants were of the Col-0 background. Transgenic plants expressing lnc000283 and lnc012465 were generated using the Agrobacterium tumefaciens-mediated floral dip method[26]. Single-copy and homozygous T3 plants were identified through genetic segregation on an agar medium supplemented with kanamycin. The T3 generation plants, or their homozygous progeny, were utilized in the experiments.

    For phenotypic assessment, Arabidopsis seeds were initially sown on filter paper moistened with ddH2O and placed in a 4 °C freezer for 2 d. Subsequently, they were evenly planted in nutrient-rich soil and transferred to a growth chamber operating a 16-h day/8-h night cycle, with day/night temperatures of 22 °C/18 °C and a light intensity of 250 μmol·m−2·s−1. After 10 d of growth, Arabidopsis plants with uniform growth were transferred to 50-hole plates. Arabidopsis plants grown in Petri dishes were firstly seed-sterilized and then sown on 1/2 MS medium supplemented with 10 g·L−1 sucrose. The seeds were then placed in a 4 °C refrigerator for 2 d in the dark before transferring them to a light incubator. The day/night duration was set to 16 h/8 h, the day/night temperature to 21 °C/18 °C, and the light intensity to 100 μmol·m−2·s−1.

    For heat treatment, 3-week-old seedlings were subjected to 38 °C for 4 d within a light incubator, subsequently transferred to their original growth conditions under the same light/dark cycles. For drought treatment, 3-week-old Arabidopsis seedlings were deprived of water for 10 d, followed by rehydration to facilitate a 2-d recovery period. Plants were photographed and surveyed both before and after treatment.

    The lncRNA sequences (lnc000283 and lnc012465) were chemically synthesized based on RNA-seq data, with restriction sites for BamH1 and Kpn1 engineered upstream and downstream. The resultant lncRNA constructs were subcloned into the pCambia2301 binary vector, incorporating a cauliflower mosaic virus (CaMV) 35S promoter. The recombinant vectors were transformed into Escherichia coli TOP10 competent cells (Clontech), incubated at 37 °C overnight, after which single clones were selected for PCR verification, and the confirmed positive colonies were submitted for sequencing. Following verification, the correct plasmids were introduced into A. tumefaciens strain GV3101 using the freeze-thaw method and subsequently transformed into Arabidopsis wild-type (Col) plants.

    To quantify the chlorophyll content, the aerial portions of wild-type and transgenic Arabidopsis plants, grown in Petri dishes were weighed, minced, and then subjected to boiling in 95% ethanol until fully decolorized. Aliquots of 200 μL from the extract were transferred to a 96-well plate and the absorbance at 663 nm and 645 nm was measured via spectrophotometry by a microplate reader (Multiskan GO, Thermo Scientific, Waltham, MA, USA). Three biological replicates were analyzed for WT and each transgenic line. Chlorophyll content was determined according to the formula of the Arnon method[27]: Chlorophyll a = (12.72A663 − 2.59A645) v/w, Chlorophyll b = (22.88A645 − 4.67A663) v/w, Total chlorophyll = (20.29A645 + 8.05A663) v/w.

    The quantification of anthocyanin was performed as follows: aerial parts of wild-type and transgenic Arabidopsis plants, cultivated in Petri dishes, were weighed and ground to powder in liquid nitrogen. Subsequently, the samples were incubated in 600 μL of acidified methanol (containing 1% HCl) at 70 °C for 1 h. Following this, 1 mL of chloroform was added, and the mixture was vigorously shaken to remove chlorophyll. The mixture was then centrifuged at 12,000 rpm for 5 min, after which the absorbance of the aqueous phase was determined at 535 nm using a spectrophotometer (Shimadzu, Kyoto, Japan). Three biological replicates were analyzed for WT and each transgenic line. The relative anthocyanin content was calculated according to anthocyanin concentration and extraction solution volume. One anthocyanin unit is defined as an absorption unit at a wavelength of 535 nm in 1 mL of extract solution. In the end, the quantity was normalized to the fresh weight of each sample.

    Three-week-old transgenic and WT A. thaliana plants, subjected to normal conditions or varying durations of heat or drought stress, were utilized for subsequent physiological assessments. All assays were performed in accordance with the method described by Chen & Zhang[28]. In brief, 0.1 g of fresh leaf tissue was homogenized in 500 μL of 100 mM PBS (pH 7.8) while chilled on ice. The homogenate was then centrifuged at 4 °C, and the resultant supernatant was employed for further analysis. For the determination of MDA content, 100 μL of the supernatant was combined with 500 μL of a 0.25% thiobarbituric acid (TBA) solution (which was prepared by dissolving 0.125 g of TBA in 5 mL of 1 mol·L−1 NaOH before being added to 45 mL of 10% TCA) and boiled for 15 min. Following a 5 min cooling period on ice, the absorbance was measured at 532 nm and 600 nm. The activity of POD was determined as follows: initially, 28 μL of 0.2% guaiacol and 19 μL of 30% H2O2 were sequentially added to 50 mL of 10mM PBS (pH 7.0), after thorough heating and mixing, 1 mL was transferred into a cuvette, then 50 μL of the supernatant was added to the cuvette and the absorbance at 470 nm was monitored every 15 s for 1 min. To determine the proline content, a reaction solution was prepared by mixing 3% sulfosalicylic acid, acetic acid, and 2.5% acidic ninhydrin in a ratio of 1:1:2, then 50 μL of the supernatant was added to 1 mL of the reaction solution, which was then subjected to a boiling water bath for 15 min (the solution turned red after the boiling water bath). Following cooling on ice, the absorbance at 520 nm was recorded. For the quantification of proline, an L-proline standard curve was prepared by dissolving 0, 5, 10, 15, 20, 25, and 30 μg of L-proline in 0.5 mL of ddH2O, followed by the addition of 1 mL of the reaction solution and measuring the absorbance at 520 nm. The proline content in the samples was then determined based on the L-proline standard curve.

    Total RNA was isolated from the aerial parts of Arabidopsis using the TaKaRa MiniBEST Plant RNA Extraction Kit, followed by purification and reverse transcription using the PrimeScript RT reagent Kit with gDNA Eraser (Takara). The cDNA product was diluted 10 times and real-time PCR was conducted in triplicate for each biological replicate using SYBR PCR Master Mix (Applied Biosystems) on the ABI 7500 system under the following conditions: 98 °C for 3 min, followed by 40 cycles of 98 °C for 2 s and 60 °C for 30 s. The relative expression levels of each gene were normalized against the transcript abundance of the endogenous control UBC30 (At5g56150) and calculated using the 2−ΔCᴛ method. The specific primers employed for qRT-PCR are detailed in Supplemental Table S1.

    In our prior investigation, dozens of lncRNAs associated with the heat stress response in Chinese cabbage were identified through informatics analysis. Two lncRNAs (lnc000283 and lnc012465) were chosen for genetic transformation in Arabidopsis to elucidate their functions comprehensively. Transcriptome data analysis indicated that the expression of lnc000283 and lnc012465 in Chinese cabbage were both induced by HS. To verify the accuracy, the expression patterns of lnc000283 and lnc012465 were confirmed through quantitative real-time PCR (qRT-PCR), and the results from qRT-PCR were consistent with those obtained from RNA-seq (Fig. 1a). The corresponding homologous genes in Arabidopsis were identified as CNT2088434 and CNT2088742, exhibiting sequence similarities of 88% and 87%, respectively (Supplemental Fig. S1). Subcellular localization predictions using the lnclocator database (www.csbio.sjtu.edu.cn/bioinf/lncLocator) suggested that both lncRNAs are localized within the nucleus (Supplemental Table S2). Bioinformatics analysis was conducted using the CPC tool (http://cpc.cbi.pku.edu.cn/) indicated that lnc000283 and lnc012465 are noncoding sequences, with coding probabilities of 0.0466805 and 0.0432148, respectively comparable to the well-characterized lncRNAs COLDAIR and Xist, but significantly lower than those of the protein-coding genes UBC10 and ACT2 (Fig. 1b).

    Figure 1.  Characteristics of lnc000283 and lnc012465. (a) Expression level of lnc000283 and lnc012465 in Chinese cabbage leaves treated at 38 °C at different time points, as determined by qRT-PCR and RNA-seq. CK is a representative plant before heating, and T1, T4, T8, and T12 denote plants that were subjected to 38 °C for 1, 4, 8, and 12 h, respectively. The expression levels were normalized to the expression level of Actin. (b) Analysis of coding potential for lnc000283 and lnc012465. The coding potential scores were calculated using the CPC program. UBC10 (At5g53300) and ACT2 (At3g18780) are positive controls that encode proteins. COLDAIR (HG975388) and Xist (L04961) serve as negative controls, exhibiting minimal protein-coding potential.

    To elucidate the role of lnc000283 and lnc012465 in response to abiotic stress, overexpression vectors were constructed for these lncRNAs, driven by the CaMV 35S promoter, and they were introduced into Arabidopsis thaliana (Col-0 ecotype). Through PCR identification and generational antibiotic screening, two homozygous positive lines for lnc012465 and lnc000283 were obtained. The relative expression levels of these lncRNAs were assessed using qRT-PCR (Fig. 2a). When plants were grown in 1/2 MS medium, with the consumption of nutrients, and reduction of water, the leaves of WT began to turn yellow, but the lnc000283 and lnc012465 overexpression lines developed a deep purple color of leaf veins (Fig. 2b). Examination of chlorophyll and anthocyanin contents in the plants revealed that both overexpression lines had higher levels of chlorophyll and anthocyanin compared to the WT, suggesting that the transgenic plants might possess enhanced resistance to nutritional or water stress (Fig. 2c, d).

    Figure 2.  Arabidopsis plants overexpressing lnc000283 and lnc012465 had higher anthocyanins and chlorophyll content. (a) The relative expression level of lnc000283 and lnc012465 in WT and different transgenic lines. UBC10 (At5g53300) was used as an internal control. Each value is mean ± sd (n = 3). (b) The phenotype of WT and Arabidopsis overexpressing lnc000283 or lnc012465 grown on 1/2 MS medium 50 d after sowing. The (c) anthocyanin and (d) chlorophyll content of WT and transgenic Arabidopsis overexpressing lnc000283 or lnc012465. The asterisks above the bars indicate statistical significance using Student's t-test (*, p < 0.05; **, p < 0.01).

    Given that lnc000283 and lnc012465 were highly induced by heat, the thermotolerance of the overexpressing (OE) plants were compared to that of the wild type. Arabidopsis plants were initially exposed to a an HS treatment at 38 °C for 4 d, followed by recovery at room temperature. The death caused by HS was processive. Post-severe HS challenge for 4 d, OE plants initially appeared similar to WT, but upon recovery, their leaves started to fold or curl, followed by a transition to yellow, white, and eventually drying out (Fig. 3a). OE lnc000283 and OE lnc012465 plants exhibited enhanced thermotolerance compared to WT, with lnc012465 showing particularly strong tolerance (Fig. 3a; Supplemental Fig. S2a). After 5 d of recovery, leaf coloration indicated that transgenic plants maintained a significantly higher percentage of green leaves and a lower percentage of bleached leaves compared to WT (Fig. 3b; Supplemental Fig. S2b). Under non-heat-stress conditions, WT and OE plants possessed comparable water content. However, following heat stress, the fresh-to-dry weight ratio of OE lnc000283 and lnc012465 lines was significantly greater than that of WT (Fig. 3c; Supplemental Fig. S2c). Abiotic stresses frequently trigger the production of excessive reactive oxygen species (ROS), which are believed to cause lipid peroxidation of membrane lipids, leading to damage to macromolecules. Leaf MDA content is commonly used as an indicator of lipid peroxidation under stress conditions; therefore, the MDA content in both transgenic and WT plants was assessed. Figure 3d shows that the MDA content in WT plants progressively increased after heat treatment, whereas in the two lines overexpressing lnc012465, the MDA content increased only slightly and remained significantly lower than that in WT at all time points. In plants overexpressing lnc000283, the MDA content did not significantly differ from that of WT before heat stress. However, after 4 d of heat treatment, the MDA content was significantly lower compared to WT (Supplemental Fig. S2d). The results suggested that the expression of both lnc012465 and lnc000283 can mitigate injury caused by membrane lipid peroxidation under heat-stress conditions. Peroxidase (POD) is a crucial antioxidant enzyme involved in ROS scavenging. Figure 3e and Supplemental Fig. S2e demonstrate that POD activity increased in both transgenic and WT plants after heat treatment. However, the increase in WT plants was modest, whereas OE lnc000283 and OE lnc012465 plants exhibited consistently higher POD activity. As anticipated, proline levels were induced in response to stress in all studied plants (Fig. 3f; Supplemental Fig. S2f). However, under normal conditions and 2 d post-heat stress treatment, the proline content in OE lnc000283 and OE lnc012465 plants did not exhibit significant changes compared to WT (Fig. 3f; Supplemental Fig. S2f). Moreover, after 4 d of heat stress, the proline content in OE lnc012465 lines was significantly lower than in WT, and the OE lnc000283 transgenic line 12-6 also showed a marked decrease in proline content compared to WT (Fig. 3f; Supplemental Fig. S2f). The results indicated that the thermotolerance of plants overexpressing either lnc000283 or lnc012465 was independent of proline accumulation.

    Figure 3.  Overexpressing lnc012465 lines are more tolerant to heat stress. (a) Phenotypes of WT and OE lnc012465 plants were assessed before and after exposure to heat stress. The heat treatment was applied to 25-day-old Arabidopsis plants. (b) The percentage of leaves with different colors in Arabidopsis after heat treatment and recovery for 5 d. (c) The fresh-to-dry weight ratio of Arabidopsis leaves was measured before and after 38 °C heat treatment. (d)−(f) depict the MDA content, POD activity, and proline content in Arabidopsis leaves at varying durations of heat stress. The asterisks above the bars indicate statistical significance using Student's t-test (*, p < 0.05; **, p < 0.01; ***, p < 0.001).

    To elucidate the molecular mechanisms by which lncRNAs enhance thermotolerance in Arabidopsis, the expression of the Hsf gene HsfA7a and three Hsps (Hsp25.3, Hsa32, and Hsp18.1-CI) in OE lnc000283, OE lnc012465, and WT Arabidopsis plants were investigated at various time points following heat treatment. As shown in Fig. 4 and Supplemental Fig. S3, both Hsf and Hsps exhibited a rapid response to heat stress with strong induction. Notably, the transcripts of HsfA7a and Hsp25.3 were significantly upregulated at 1 h after heat exposure, then experienced a sharp decrease. Hsa32 and Hsp18.1-CI were highly induced at 1 h and, unlike the other proteins, sustained high expression levels at 3 h (Fig. 4; Supplemental Fig. S3). At 1 h post-heat treatment, the transcript levels of Hsa32 and HsfA7a in OE lnc000283 did not significantly differ from those in WT. However, by 3 h, Hsa32 expression was roughly 50% of the WT level, while HsfA7a expression was approximately double that of WT (Supplemental Fig. S3). The overexpression of lnc000283 did not significantly affect the transcript level of Hsp25.3 at any of the tested time points. Notably, Hsp18.1-CI expression in both lines overexpressing lnc000283 was significantly induced at all three detection points post-heat treatment, reaching approximately 4-9-fold higher levels than in the WT (Supplemental Fig. S3). In Arabidopsis plants with elevated expression of lnc012465, the expression patterns of all Hsp and Hsf genes were similar to those in plants overexpressing lnc000283, with the notable exception of Hsa32. Unlike the WT, Hsa32 did not show a trend of down-regulation at 3 h post-heat treatment (Fig. 4). The findings suggest that the substantial induction of Hsp18.1-CI may play a role in enhancing the thermotolerance of Arabidopsis plants overexpressing lnc000283 and lnc012465.

    Figure 4.  The expression of HSF and HSP genes in lnc012465 overexpressing lines before and after different heat treatment times. Gene expression levels were quantified using RT-qPCR and normalized to UBC10 (At5g53300). Each value represents the mean ± standard deviation (n = 3). The asterisks above the bars indicate statistical significance using Student's t-test (*, p < 0.05; **, p < 0.01; ***, p < 0.001).

    Prior research has implicated a significant proportion of genes in conferring resistance to various abiotic stresses. To elucidate the functions of lnc000283 and lnc012465 more thoroughly, WT and transgenic plants were subjected to drought stress by depriving them of water for 9 d. It was noted that the majority of leaves in WT plants withered and dried, whereas the OE lnc000283 and OE lnc012465 plants exhibited reduced withering, with only a minority displaying dryness (Fig. 5a; Supplemental Fig. S4a). Eight days post-rewatering, a negligible fraction of WT seedlings exhibited recovery, whereas the overwhelming majority of transgenic plants regained vigorous growth (Fig. 5a; Supplemental Fig. S4a). The transgenic plants demonstrated a significantly higher survival rate compared to the WT plants. Following 9 d of water deficit treatment, less than 40% of the WT plants survived, whereas the OE 012465 lines 8-7 and 9-1 exhibited survival rates of 100% and 95%, respectively, and the OE 000283 lines 11-10 and 12-6 had survival rates of 87% each. (Fig. 5b; Supplemental Fig. S4b). Water loss serves as a critical metric for assessing plant drought tolerance, hence the fresh-to-dry weight ratio of excised leaves was assessed via desiccation analysis. Following 4 d of drought treatment, the fresh-to-dry weight ratio for WT plants was reduced to 43%, whereas for OE lnc000283 lines 11-10 and 12-6, it was reduced to 73% and 75%, respectively. For OE 012465 lines 8-7 and 9-1, the ratios were reduced to 67% and 62%, respectively (Fig. 5c; Supplemental Fig. S4c). The findings indicated that lnc000283 and lnc012465 endow the transgenic plants with drought tolerance.

    Figure 5.  Overexpressing lnc012465 lines are more tolerant to drought stress. (a) Phenotype of WT and OE lnc012465 plants before and after subjecting to drought stress. Drought treatment was carried out on 20-day-old Arabidopsis plants. (b) The percentage of leaves with different colors in Arabidopsis after heat treatment and recovery for 5 d. (c) The fresh weight to dry weight ratio of Arabidopsis leaves before and after undergoing 38 °C heat treatment. (d)−(f) MDA content, POD activity, and proline content in Arabidopsis leaves under different times of heat stress. The asterisks above the bars indicate statistical significance using Student's t-test (*, p < 0.05; **, p < 0.01; ***, p < 0.001)

    MDA content in leaves is a standard biomarker for assessing the extent of drought stress-induced damage. Prior to drought stress exposure, MDA levels in WT and transgenic plants were comparable. However, following 7 and 9 d of water deficit, the MDA content in the transgenic plants was markedly reduced compared to the WT, suggesting a less severe degree of membrane lipid peroxidation in the transgenic plants (Fig. 5d; Supplemental Fig. S4d). Oxidative stress frequently coincides with drought stress, hence the activity of POD was assessed to evaluate the ROS scavenging ability. The findings indicated that as the duration of drought treatment increased, POD activity progressively rose. Before drought exposure, the POD activity in lines 11-10 and 12-6 of OE 000283 was 2.4-fold and 2.2-fold higher than that of the WT, respectively (refer to Supplemental Fig. S4e). Following drought treatment, the POD activity in the transgenic lines remained significantly elevated compared to the wild type, although the enhancement was less pronounced than before the treatment (Supplemental Fig. S4e). In the OE 012465 plants, the POD activity in lines 8-7 and 9-1 significantly surpassed that of the wild type, with the discrepancy being more pronounced during drought stress (Fig. 5e). The proline content in WT and OE 000283 plants exhibited no significant differences before and after 7 d of treatment. However, after 9 d of drought, the proline content in OE 000283 plants was significantly lower compared to that in the WT (Supplemental Fig. S4f). OE 000465 plants showed no significant difference from the wild type before and after drought treatment (Fig. 5f). The findings were consistent with those under heat stress, indicating that the enhanced stress resistance due to the overexpression of lnc000283 and lnc012465 in Arabidopsis is not reliant on proline accumulation.

    Following drought stress treatment, the expression levels of drought-related genes such as RD29A, RD29B, NCED3, AREB1, and Rab18 were significantly elevated in plants overexpressing lnc000283 and lnc012465 compared to WT plants. These findings suggest that lnc000283 and lnc012465 modulate Arabidopsis drought tolerance by regulating the expression of genes associated with the drought stress response (Fig. 6; Supplemental Fig. S5).

    Figure 6.  The expression of drought-responsive genes in lnc012465 overexpressing lines before and after different drought treatment time. Gene expression levels were determined by qRT-PCR normalized against UBC10 (At5g53300). Each value is mean ± sd (n = 3). The asterisks above the bars indicate statistical significance using Student's t-test (*, p < 0.05; **, p < 0.01; ***, p < 0.001).

    The integrity of global food security is under threat due to the confluence of rapid population expansion and profound climatic shifts[29]. Amidst the shifting climatic landscape, heat and drought stress have emerged as primary limitations to crop yield and global food security. Understanding how plants detect stress cues and acclimate to challenging conditions is a pivotal biological inquiry. Moreover, enhancing plant resilience to stress is essential for maintaining agricultural productivity and fostering environmental sustainability[2]. Concurrently, the advancement of next-generation sequencing (NGS) technology has led to the identification of a substantial number of lncRNAs that participate in diverse stress responses, with functional analyses having been conducted on several of these molecules.[30] For instance, in the case of potatoes, the lncRNA StFLORE has been identified to modulate water loss through its interaction with the homologous gene StCDF1[31]. LncRNA TCONS_00021861 can activate the IAA biosynthetic pathway, thereby endowing rice with resistance to drought stress[32]. In wheat, the expression of TalnRNA27 and TalnRNA5 was upregulated in response to heat stress[33]. Our prior investigation identified a total of 81 lncRNAs in Chinese cabbage that engage in intricate interactions with their respective mRNA targets across various phases of heat treatment[25]. Two lncRNAs, lnc000283 and lnc012465, were chosen for subsequent functional analysis. Findings confirmed that these lncRNAs endow transgenic Arabidopsis plants with enhanced tolerance to both heat and drought, thereby offering novel resources for enhancing stress resistance through genetic engineering.

    Abiotic stresses frequently trigger the synthesis of anthocyanins, serving as natural antioxidants that mitigate oxidative damage by neutralizing surplus reactive oxygen species (ROS), thereby protecting plants from growth inhibition and cell death, allowing plants to adapt to abiotic stresses[34,35]. For instance, during chilling stress, the accumulation of anthocyanins within leaves can mitigate oxidative damage, thereby enhancing the photosynthetic rate[36]. Consequently, the level of abiotic stress tolerance can be inferred from the concentration of anthocyanins. The reduction of photosynthetic ability is one of the key physiological phenomena of stresses, which is partly due to the degradation of chlorophyll caused by leaf senescence during stress. The reduced accumulation of chlorophyll in the plants was seen in many plants when exposed to drought or heat stress conditions. The current investigation revealed that lncRNA-overexpressing plants cultivated in Petri dishes exhibited increased accumulation of both chlorophyll and anthocyanins in advanced growth phases, indicating that these transgenic plants, overexpressing lnc000283 and lnc012465, demonstrated enhanced stress tolerance and superior growth performance relative to WT (Fig. 2c, d).

    Upon exposure to heat stress, there is a marked induction of transcription for numerous genes that encode molecular chaperones in plants, with the vast majority of these genes contributing to the prevention of protein denaturation-related damage and the augmentation of thermotolerance[3739]. The present investigation identified multiple heat-inducible genes in plants overexpressing lnc000283 and lnc012465, as well as in WT (Fig. 4; Supplemental Fig. S3). The findings indicated that of the four HSP or HSF genes examined, Hsp18.1-CI exhibited a significantly greater abundance in both OE lnc000283 and OE lnc012465 plants compared to the WT following heat treatment for several days. Hsp18.1-CI, formerly referred to as Hsp18.2 has been the subject of investigation since 1989.[40] Following the fusion of the 5' region of Hsp18.2 in frame with the uidA gene of Escherichia coli, the activity of GUS, serving as the driver gene was observed to increase upon exposure to HS[40]. The Arabidopsis hsfA2 mutant exhibited diminished thermotolerance after heat acclimation, with the transcript levels of Hsp18.1-CI being substantially reduced compared to those in wild-type plants following a 4-h recovery period[41]. The findings revealed that the upregulation of Hsp18.1-CI protein is a critical mechanism by which plants achieve enhanced protection against heat stress in adverse environmental conditions, thereby bolstering their thermotolerance.

    Plants cultivated in natural settings are often subjected to concurrent multiple abiotic stresses, which can exacerbate threats to their routine physiological functions, growth, and developmental processes[42,43]. Elucidating the molecular mechanisms underlying plant responses to abiotic stress is crucial for the development of new crop varieties with enhanced tolerance to multiple abiotic stresses. Previous research has indicated that the overexpression of certain protein-coding genes can endow plants with resistance to a variety of abiotic stresses. For instance, tomatoes with robust expression of ShCML44 demonstrated significantly enhanced tolerance to drought, cold, and salinity stresses[44]. Overexpression of PeCBF4a in poplar plants confers enhanced tolerance to a range of abiotic stresses[45]. With respect to lncRNAs, transgenic Arabidopsis plants that overexpress lncRNA-DRIR displayed marked increased tolerance to salt and drought stresses compared to the wild-type[46]. In the present study, both overexpression lines of lnc000283 and lnc012465 exhibited resistance to heat and drought stresses, thereby contributing to the enhancement of plant resilience against multiple stresses (Figs 3, 5; Supplemental Figs S2, S4).

    The number of genes implicated in plant drought resistance is regulated by both ABA-dependent and ABA-independent pathways[47,48]. It is well established that the expression of RD29A exhibits a high level of responsiveness to drought stress, operating through both ABA-dependent and ABA-independent mechanisms[49]. RD29B, AREB1, and RAB18 are governed by an ABA-dependent regulatory pathway[10,49,50]. NCED3 is involved in ABA biosynthesis[51]. In the present study, the transcript levels of RD29A, RD29B, NCED3, AREB1, and RAB18 were significantly elevated in OE lnc000283 and OE lnc012465 plants compared to those in the WT plants (Fig. 6; Supplemental Fig. S5). The findings indicated that the drought tolerance imparted by OE lnc000283 and OE lnc012465 plants is contingent upon an ABA-dependent mechanism.

    Prior research has indicated that certain long non-coding RNAs (lncRNAs) can assume analogous roles across diverse biological contexts. For example, the lncRNA bra-miR156HG has been shown to modulate leaf morphology and flowering time in both B. campestris and Arabidopsis[52]. Heterogeneous expression of MSL-lncRNAs in Arabidopsis has been associated with the promotion of maleness, and similarly, it is implicated in the sexual lability observed in female poplars[53]. In the present study, lnc000283 and lnc012465 were induced by heat in Chinese cabbage, and their heterologous expression was found to confer heat tolerance in Arabidopsis. Additionally, sequences homologous to lnc000283 and lnc012465 were identified in Arabidopsis (Supplemental Fig. S1). The data suggest that these sequences may share a comparable function to that of heat-inducible sequences, potentially accounting for the conservation of lnc000283 and lnc012465'os functionality across various species.

    In conclusion, the functions of two heat-inducible lncRNAs, lnc000283 and lnc012465 have been elucidated. Transgenic Arabidopsis lines overexpressing these lncRNAs accumulated higher levels of anthocyanins and chlorophyll at a later stage of growth compared to the WT when grown on Petri dishes. Furthermore, under heat and drought stress conditions, these OE plants exhibited enhanced stress tolerance, with several genes related to the stress resistance pathway being significantly upregulated. Collectively, these findings offer novel insights for the development of new varieties with tolerance to multiple stresses.

    The authors confirm contribution to the paper as follows: study conception and supervision: Li N, Song X; experiment performing: Wang Y, Sun S; manuscript preparation and revision: Wang Y, Feng X, Li N. All authors reviewed the results and approved the final version of the manuscript.

    All data generated or analyzed during this study are included in this published article and its supplementary information files.

    This work was supported by the National Natural Science Foundation of China (32172583), the Natural Science Foundation of Hebei (C2021209019), the Natural Science Foundation for Distinguished Young Scholars of Hebei (C2022209010), and the Basic Research Program of Tangshan (22130231H).

  • The authors declare that they have no conflict of interest.

  • Supplementary Table S1 The relative synonymous codon usage of amino acids in the mitogenome of Michelia figo, Magnolia biondii, Magnolia officinalis, and Liriodendron tulipifera.
    Supplementary Table S2 The frequency of codon usage in the mitogenome of Michelia figo, Magnolia biondii, Magnolia officinalis, and Liriodendron tulipifera.
    Supplementary Table S3 Dispersed repeat sequences identified in the Michelia figo mitogenome.
    Supplementary Table S4 Tandem repeat sequences identified in the Michelia figo mitogenome.
    Supplementary Table S5 The simple sequence repeats identified in the Michelia figo mitogenome.
    Supplementary Table S6 Dispersed repeat sequences identified in the Magnolia biondii mitogenome.
    Supplementary Table S7 Dispersed repeat sequences identified in the Magnolia officinalis mitogenome.
    Supplementary Table S8 Tandem repeat sequences identified in the Magnolia biondii mitogenome.
    Supplementary Table S9 Tandem repeat sequences identified in the Magnolia officinalis mitogenome.
    Supplementary Table S10 The simple sequence repeats identified in the Magnolia biondii mitogenome.
    Supplementary Table S11 The simple sequence repeats identified in the Liriodendron tulipifera mitogenome.
    Supplementary Table S12 The homologous DNA fragment between mitogenome and cpgenome of Michelia figo.
    Supplementary Table S13 The collinear blocks between mitogenomes of Michelia figo and Magnolia officinalis.
    Supplementary Table S14 The collinear blocks between mitogenomes of Michelia figo and Magnolia biondii.
    Supplementary Table S15 The collinear blocks between mitogenomes of Magnolia biondii and Liriodendron tulipifera.
    Supplementary Table S16 Genes used for phylogenetic analysis.
    Supplementary Table S17 The genomic information of the species used in this study.
    Supplementary Fig. S1 The map of genes containing introns. This diagram illustrates the distribution of cis- and trans-introns.
    Supplementary Fig. S2 Heat maps of PCG and intron contents among 15 mitogenomes (a) Comparison of PCG contents among 15 mitogenomes. The gene numbers are shown on the top right. (b) Comparison of intron contents among 15 mitogenomes. The intron numbers are shown on the top right.
  • [1]

    Roger AJ, Muñoz-Gómez SA, Kamikawa R. 2017. The origin and diversification of mitochondria. Current Biology 27(21):R1177−R1192

    doi: 10.1016/j.cub.2017.09.015

    CrossRef   Google Scholar

    [2]

    Dyall SD, Brown MT, Johnson PJ. 2004. Ancient invasions: from endosymbionts to organelles. Science 304:253−57

    doi: 10.1126/science.1094884

    CrossRef   Google Scholar

    [3]

    Yu SB, Pekkurnaz G. 2018. Mechanisms orchestrating mitochondrial dynamics for energy homeostasis. Journal of Molecular Biology 430:3922−41

    doi: 10.1016/j.jmb.2018.07.027

    CrossRef   Google Scholar

    [4]

    van Loo G, Saelens X, van Gurp M, MacFarlane M, Martin SJ, et al. 2002. The role of mitochondrial factors in apoptosis: a Russian roulette with more than one bullet. Cell Death & Differentiation 9:1031−42

    doi: 10.1038/sj.cdd.4401088

    CrossRef   Google Scholar

    [5]

    Bi C, Sun N, Han F, Xu K, Yang Y, et al. 2024. The first mitogenome of Lauraceae (Cinnamomum chekiangense). Plant Diversity 46:144−48

    doi: 10.1016/j.pld.2023.11.001

    CrossRef   Google Scholar

    [6]

    Ma Q, Wang Y, Li S, Wen J, Zhu L, et al. 2022. Assembly and comparative analysis of the first complete mitochondrial genome of Acer truncatum Bunge: a woody oil-tree species producing nervonic acid. BMC Plant Biology 22:29

    doi: 10.1186/s12870-021-03416-5

    CrossRef   Google Scholar

    [7]

    Han F, Qu Y, Chen Y, Xu LA, Bi C. 2022. Assembly and comparative analysis of the complete mitochondrial genome of Salix wilsonii using PacBio HiFi sequencing. Frontiers in Plant Science 13:1031769

    doi: 10.3389/fpls.2022.1031769

    CrossRef   Google Scholar

    [8]

    Wang XD, Xu CY, Zheng YJ, Wu YF, Zhang YT, et al. 2022. Chromosome-level genome assembly and resequencing of camphor tree (Cinnamomum camphora) provides insight into phylogeny and diversification of terpenoid and triglyceride biosynthesis of Cinnamomum. Horticulture Research 9:uhac216

    doi: 10.1093/hr/uhac216

    CrossRef   Google Scholar

    [9]

    Morley SA, Nielsen BL. 2017. Plant mitochondrial DNA. Frontiers in Bioscience-Landmark (FBL) 22:1023−32

    doi: 10.2741/4531

    CrossRef   Google Scholar

    [10]

    Wynn EL, Christensen AC. 2019. Repeats of unusual size in plant mitochondrial genomes: identification, incidence and evolution. G3 Genes| Genomes| Genetics 9:549−59

    doi: 10.1534/g3.118.200948

    CrossRef   Google Scholar

    [11]

    Møller IM, Rasmusson AG, Van Aken O. 2021. Plant mitochondria – past, present and future. The Plant Journal 108:912−59

    doi: 10.1111/tpj.15495

    CrossRef   Google Scholar

    [12]

    Skippington E, Barkman TJ, Rice DW, Palmer JD. 2015. Miniaturized mitogenome of the parasitic plant Viscum scurruloideum is extremely divergent and dynamic and has lost all nad genes. Proceedings of the National Academy of Sciences of the United States of America 112:E3515−E24

    doi: 10.1073/pnas.1504491112

    CrossRef   Google Scholar

    [13]

    Putintseva YA, Bondar EI, Simonov EP, Sharov VV, Oreshkova NV, et al. 2020. Siberian larch (Larix sibirica Ledeb.) mitochondrial genome assembled using both short and long nucleotide sequence reads is currently the largest known mitogenome. BMC Genomics 21:654

    doi: 10.1186/s12864-020-07061-4

    CrossRef   Google Scholar

    [14]

    Bi C, Qu Y, Hou J, Wu K, Ye N, et al. 2022. Deciphering the Multi-Chromosomal Mitochondrial Genome of Populus simonii. Frontiers in Plant Science 13:914635

    doi: 10.3389/fpls.2022.914635

    CrossRef   Google Scholar

    [15]

    Logacheva MD, Schelkunov MI, Fesenko AN, Kasianov AS, Penin AA. 2020. Mitochondrial genome of Fagopyrum esculentum and the genetic diversity of extranuclear genomes in buckwheat. Plants 9:618

    doi: 10.3390/plants9050618

    CrossRef   Google Scholar

    [16]

    Adams KL, Qiu YL, Stoutemyer M, Palmer JD. 2002. Punctuated evolution of mitochondrial gene content: High and variable rates of mitochondrial gene loss and transfer to the nucleus during angiosperm evolution. Proceedings of the National Academy of Sciences of the United States of America 99:9905−12

    doi: 10.1073/pnas.042694899

    CrossRef   Google Scholar

    [17]

    Filip E, Skuza L. 2021. Horizontal gene transfer involving chloroplasts. International Journal of Molecular Sciences 22:4484

    doi: 10.3390/ijms22094484

    CrossRef   Google Scholar

    [18]

    Rodríguez-Moreno L, González VM, Benjak A, Carmen Martí M, Puigdomènech P, et al. 2011. Determination of the melon chloroplast and mitochondrial genome sequences reveals that the largest reported mitochondrial genome in plants contains a significant amount of DNA having a nuclear origin. BMC Genomics 12:424

    doi: 10.1186/1471-2164-12-424

    CrossRef   Google Scholar

    [19]

    Veltjen E, Testé E, Palmarola Bejerano A, et al. 2022. The evolutionary history of the Caribbean magnolias (Magnoliaceae): Testing species delimitations and biogeographical hypotheses using molecular data. Molecular Phylogenetics and Evolution 167:107359

    doi: 10.1016/j.ympev.2021.107359

    CrossRef   Google Scholar

    [20]

    Law YW. 1984. A preliminary study on the taxonomy of the family magnoliaceae. Journal of Systematics and Evolution 22(2):89−109

    Google Scholar

    [21]

    Taprial S. 2015. A review on phytochemical and pharmacological properties of Michelia champaca Linn. family: Magnoliaceae. International Journal Of Pharmaceutical Sciences And Research 2:430−36

    Google Scholar

    [22]

    Shang C, Hu Y, Deng C, Hu K. 2002. Rapid determination of volatile constituents of Michelia alba flowers by gas chromatography-mass spectrometry with solid-phase microextraction. Journal of chromatography. A 942:283−8

    doi: 10.1016/S0021-9673(01)01382-6

    CrossRef   Google Scholar

    [23]

    Khan MR, Kihara M, Omoloso AD. 2002. Antimicrobial activity of Michelia champaca. Fitoterapia 73:744−48

    doi: 10.1016/S0367-326X(02)00248-4

    CrossRef   Google Scholar

    [24]

    Cheng KK, Nadri MH, Othman NZ, Rashid SNAA, Lim YC, et al. 2022. Phytochemistry, bioactivities and traditional uses of Michelia × alba. Molecules 27:3450

    doi: 10.3390/molecules27113450

    CrossRef   Google Scholar

    [25]

    Chericoni S, Testai L, Campeol E, Calderone V, Morelli I, et al. 2004. Vasodilator activity of Michelia figo Spreng (Magnoliaceae) by in vitro functional study. Journal of Ethnopharmacology 91:263−66

    doi: 10.1016/j.jep.2003.12.021

    CrossRef   Google Scholar

    [26]

    Zhai M. 2020. The complete chloroplast genome sequence of Michelia figo based on landscape design, and a comparative analysis with other Michelia species. Mitochondrial DNA Part B 5:2723−24

    doi: 10.1080/23802359.2020.1788446

    CrossRef   Google Scholar

    [27]

    Hinsinger DD, Strijk JS. 2017. The chloroplast genome sequence of Michelia alba (Magnoliaceae), an ornamental tree species. Mitochondrial DNA Part B 2:9−10

    doi: 10.1080/23802359.2016.1275850

    CrossRef   Google Scholar

    [28]

    Li Y, Zhou M, Wang L, Wang J. 2021. The characteristics of the chloroplast genome of the Michelia chartacea (Magnoliaceae). Mitochondrial DNA Part B 6:493−95

    doi: 10.1080/23802359.2020.1871432

    CrossRef   Google Scholar

    [29]

    Sima Y, Li Y, Yuan X, Wang Y. 2020. The complete chloroplast genome sequence of Michelia chapensis Dandy: an endangered species in China. Mitochondrial DNA Part B 5:1594−95

    doi: 10.1080/23802359.2020.1742619

    CrossRef   Google Scholar

    [30]

    Arseneau JR, Steeves R, Laflamme M. 2017. Modified low-salt CTAB extraction of high-quality DNA from contaminant-rich tissues. Molecular Ecology Resources 17:686−93

    doi: 10.1111/1755-0998.12616

    CrossRef   Google Scholar

    [31]

    Shi Q, Tian D, Wang J, Chen A, Miao Y, et al. 2023. Overexpression of miR390b promotes stem elongation and height growth in Populus. Horticulture Research 10:uhac258

    doi: 10.1093/hr/uhac258

    CrossRef   Google Scholar

    [32]

    Bi C, Shen F, Han F, Qu Y, Hou J, et al. 2024. PMAT: an efficient plant mitogenome assembly toolkit using low coverage HiFi sequencing data. Horticulture Research 11:uhae023

    doi: 10.1093/hr/uhae023

    CrossRef   Google Scholar

    [33]

    Dong S, Chen L, Liu Y, Wang Y, Zhang S, et al. 2020. The draft mitochondrial genome of Magnolia biondii and mitochondrial phylogenomics of angiosperms. Plos One 15:e0231020

    doi: 10.1371/journal.pone.0231020

    CrossRef   Google Scholar

    [34]

    Wick RR, Schultz MB, Zobel J, Holt KE. 2015. Bandage: interactive visualization of de novo genome assemblies. Bioinformatics 31:3350−52

    doi: 10.1093/bioinformatics/btv383

    CrossRef   Google Scholar

    [35]

    Li J, Ni Y, Lu Q, Chen H, Liu C. 2024. PMGA: A plant mitochondrial genome annotator. Plant Communications 00:101191

    doi: 10.1016/j.xplc.2024.101191

    CrossRef   Google Scholar

    [36]

    Chen Y, Ye W, Zhang Y, Xu Y. 2015. High speed BLASTN: an accelerated MegaBLAST search tool. Nucleic Acids Research 43:7762−68

    doi: 10.1093/nar/gkv784

    CrossRef   Google Scholar

    [37]

    Lowe TM, Eddy SR. 1997. tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Research 25:955−64

    doi: 10.1093/nar/25.5.955

    CrossRef   Google Scholar

    [38]

    Zhang X, Chen H, Ni Y, Wu B, Li J, et al. 2024. Plant mitochondrial genome map (PMGmap): a software tool for the comprehensive visualization of coding, noncoding and genome features of plant mitochondrial genomes. Molecular Ecology Resources 24:e13952

    doi: 10.1111/1755-0998.13952

    CrossRef   Google Scholar

    [39]

    Beier S, Thiel T, Münch T, Scholz U, Mascher M. 2017. MISA-web: a web server for microsatellite prediction. Bioinformatics 33:2583−85

    doi: 10.1093/bioinformatics/btx198

    CrossRef   Google Scholar

    [40]

    Benson G. 1999. Tandem repeats finder: a program to analyze DNA sequences. Nucleic Acids Research 27:573−80

    doi: 10.1093/nar/27.2.573

    CrossRef   Google Scholar

    [41]

    Kurtz S, Choudhuri JV, Ohlebusch E, Schleiermacher C, Stoye J, et a l. 2001. REPuter: the manifold applications of repeat analysis on a genomic scale. Nucleic Acids Research 29:4633−42

    doi: 10.1093/nar/29.22.4633

    CrossRef   Google Scholar

    [42]

    Mao J, Wei S, Chen Y, Yang Y, Yin T. 2023. The proposed role of MSL-lncRNAs in causing sex lability of female poplars. Horticulture Research 10:uhad042

    doi: 10.1093/hr/uhad042

    CrossRef   Google Scholar

    [43]

    Chen C, Wu Y, Li J, Wang X, Zeng Z, et al. 2023. TBtools-II: a “one for all, all for one” bioinformatics platform for biological big-data mining. Molecular Plant 16:1733−42

    doi: 10.1016/j.molp.2023.09.010

    CrossRef   Google Scholar

    [44]

    Marçais G, Delcher AL, Phillippy AM, Coston R, Salzberg SL, et al. 2018. MUMmer4: A fast and versatile genome alignment system. PLOS Computational Biology 14:e1005944

    doi: 10.1371/journal.pcbi.1005944

    CrossRef   Google Scholar

    [45]

    He W, Yang J, Jing Y, Xu L, Yu K, et al. 2023. NGenomeSyn: an easy-to-use and flexible tool for publication-ready visualization of syntenic relationships across multiple genomes. Bioinformatics 39:btad121

    doi: 10.1093/bioinformatics/btad121

    CrossRef   Google Scholar

    [46]

    Katoh K, Rozewicki J, Yamada KD. 2019. MAFFT online service: multiple sequence alignment, interactive sequence choice and visualization. Briefings in Bioinformatics 20:1160−66

    doi: 10.1093/bib/bbx108

    CrossRef   Google Scholar

    [47]

    Capella-Gutiérrez S, Silla-Martínez JM, Gabaldón T. 2009. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25:1972−73

    doi: 10.1093/bioinformatics/btp348

    CrossRef   Google Scholar

    [48]

    Nguyen LT, Schmidt HA, von Haeseler A, Minh BQ. 2015. IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Molecular Biology and Evolution 32:268−74

    doi: 10.1093/molbev/msu300

    CrossRef   Google Scholar

    [49]

    Stamatakis A. 2006. RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22:2688−90

    doi: 10.1093/bioinformatics/btl446

    CrossRef   Google Scholar

    [50]

    Letunic I, Bork P. 2021. Interactive tree of life (iTOL) v5: an online tool for phylogenetic tree display and annotation. Nucleic Acids Research 49:W293−W296

    doi: 10.1093/nar/gkab301

    CrossRef   Google Scholar

    [51]

    Wu ZQ, Liao XZ, Zhang XN, Tembrock LR, Broz A. 2022. Genomic architectural variation of plant mitochondria—A review of multichromosomal structuring. Journal of Systematics and Evolution 60:160−68

    doi: 10.1111/jse.12655

    CrossRef   Google Scholar

    [52]

    Sloan DB, Alverson AJ, Chuckalovcak JP, Wu M, McCauley DE, et al. 2012. Rapid evolution of enormous, multichromosomal genomes in flowering plant mitochondria with exceptionally high mutation rates. PLoS Biology 10:e1001241

    doi: 10.1371/journal.pbio.1001241

    CrossRef   Google Scholar

    [53]

    Alverson AJ, Wei X, Rice DW, Stern DB, Barry K, et al. 2010. Insights into the evolution of mitochondrial genome size from complete sequences of Citrullus lanatus and Cucurbita pepo (Cucurbitaceae). Molecular Biology and Evolution 27:1436−48

    doi: 10.1093/molbev/msq029

    CrossRef   Google Scholar

    [54]

    Gualberto JM, Newton KJ. 2017. Plant mitochondrial genomes: dynamics and mechanisms of mutation. Annual Review of Plant Biology 68:225−52

    doi: 10.1146/annurev-arplant-043015-112232

    CrossRef   Google Scholar

    [55]

    Gualberto JM, Mileshina D, Wallet C, Niazi AK, Weber-Lotfi F, et al. 2014. The plant mitochondrial genome: dynamics and maintenance. Biochimie 100:107−20

    doi: 10.1016/j.biochi.2013.09.016

    CrossRef   Google Scholar

    [56]

    Rice DW, Alverson AJ, Richardson AO, Young GJ, Sanchez-Puerta MV, et al. 2013. Horizontal transfer of entire genomes via mitochondrial fusion in the angiosperm Amborella. Science 342:1468−73

    doi: 10.1126/science.1246275

    CrossRef   Google Scholar

    [57]

    Yu R, Sun C, Zhong Y, Liu Y, Sanchez-Puerta MV, et al. 2022. The minicircular and extremely heteroplasmic mitogenome of the holoparasitic plant Rhopalocnemis phalloides. Current Biology 32:470−479.e5

    doi: 10.1016/j.cub.2021.11.053

    CrossRef   Google Scholar

    [58]

    Yang H, Ni Y, Zhang X, Li J, Chen H, et al. 2023. The mitochondrial genomes of Panax notoginseng reveal recombination mediated by repeats associated with DNA replication. International Journal of Biological Macromolecules 252:126359

    doi: 10.1016/j.ijbiomac.2023.126359

    CrossRef   Google Scholar

    [59]

    Cole LW, Guo W, Mower JP, Palmer JD. 2018. High and variable rates of repeat-mediated mitochondrial genome rearrangement in a genus of plants. Molecular Biology and Evolution 35:2773−85

    Google Scholar

    [60]

    Sun M, Zhang M, Chen X, Liu Y, Liu B, et al. 2022. Rearrangement and domestication as drivers of Rosaceae mitogenome plasticity. BMC Biology 20:181

    doi: 10.1186/s12915-022-01383-3

    CrossRef   Google Scholar

    [61]

    Zardoya R. 2020. Recent advances in understanding mitochondrial genome diversity. F1000Research 9:270

    doi: 10.12688/f1000research.21490.1

    CrossRef   Google Scholar

    [62]

    Richardson AO, Rice DW, Young GJ, Alverson AJ, Palmer JD. 2013. The “fossilized” mitochondrial genome of Liriodendron tulipifera: ancestral gene content and order, ancestral editing sites, and extraordinarily low mutation rate. BMC Biology 11:29

    doi: 10.1186/1741-7007-11-29

    CrossRef   Google Scholar

    [63]

    Mower JP, Sloan DB, Alverson AJ. 2012. Plant mitochondrial genome diversity: The genomics revolution. In Plant Genome Diversity Volume 1: Plant Genomes, their Residents, and their Evolutionary Dynamics, ed. JF Wendel, J Greilhuber, J Dolezel, IJ Leitch: 123-44. Vienna: Springer Vienna. Number of 123-44 pp

    [64]

    Sloan DB, Alverson AJ, Štorchová H, Palmer JD, Taylor DR. 2010. Extensive loss of translational genes in the structurally dynamic mitochondrial genome of the angiosperm Silene latifolia. BMC Evolutionary Biology 10:274

    doi: 10.1186/1471-2148-10-274

    CrossRef   Google Scholar

    [65]

    Hecht J, Grewe F, Knoop V. 2011. Extreme RNA editing in coding islands and abundant microsatellites in repeat sequences of Selaginella moellendorffii mitochondria: the root of frequent plant mtDNA recombination in early tracheophytes. Genome Biology and Evolution 3:344−58

    doi: 10.1093/gbe/evr027

    CrossRef   Google Scholar

    [66]

    Regina TMR, Quagliariello C. 2010. Lineage-specific group II intron gains and losses of the mitochondrial rps3 gene in gymnosperms. Plant Physiology and Biochemistry 48:646−54

    doi: 10.1016/j.plaphy.2010.05.003

    CrossRef   Google Scholar

    [67]

    Timmis JN, Ayliffe MA, Huang CY, Martin W. 2004. Endosymbiotic gene transfer: organelle genomes forge eukaryotic chromosomes. Nature Reviews Genetics 5:123−35

    doi: 10.1038/nrg1271

    CrossRef   Google Scholar

    [68]

    Cheng Y, He X, Priyadarshani SVGN, Wang Y, Ye L, et al. 2021. Assembly and comparative analysis of the complete mitochondrial genome of Suaeda glauca. BMC Genomics 22:167

    doi: 10.1186/s12864-021-07490-9

    CrossRef   Google Scholar

    [69]

    Sloan DB, Wu Z. 2014. History of plastid DNA insertions reveals weak deletion and AT mutation biases in angiosperm mitochondrial genomes. Genome Biology and Evolution 6:3210−21

    doi: 10.1093/gbe/evu253

    CrossRef   Google Scholar

    [70]

    Zhang T, Fang Y, Wang X, Deng X, Zhang X, et al. 2012. The complete chloroplast and mitochondrial genome sequences of Boea hygrometrica: insights into the evolution of plant organellar genomes. PLOS ONE 7:e30531

    doi: 10.1371/journal.pone.0030531

    CrossRef   Google Scholar

    [71]

    Wang D, Rousseau-Gueutin M, Timmis JN. 2012. Plastid sequences contribute to some plant mitochondrial genes. Molecular Biology and Evolution 29:1707−11

    doi: 10.1093/molbev/mss016

    CrossRef   Google Scholar

    [72]

    Wang D, Wu YW, Shih AC, Wu CS, Wang YN, et al. 2007. Transfer of Chloroplast Genomic DNA to Mitochondrial Genome Occurred At Least 300 MYA. Molecular Biology and Evolution 24:2040−48

    doi: 10.1093/molbev/msm133

    CrossRef   Google Scholar

    [73]

    Notsu Y, Masood S, Nishikawa T, Kubo N, Akiduki G, et al. 2002. The complete sequence of the rice (Oryza sativa L.) mitochondrial genome: frequent DNA sequence acquisition and loss during the evolution of flowering plants. Molecular Genetics and Genomics 268:434−45

    doi: 10.1007/s00438-002-0767-1

    CrossRef   Google Scholar

    [74]

    Clifton SW, Minx P, Fauron CMR, Gibson M, Allen JO, et al. 2004. Sequence and Comparative Analysis of the Maize NB Mitochondrial Genome. Plant Physiology 136:3486−503

    doi: 10.1104/pp.104.044602

    CrossRef   Google Scholar

    [75]

    Nie ZL, Wen J, Azuma H, Qiu YL, Sun H, et al. 2008. Phylogenetic and biogeographic complexity of Magnoliaceae in the Northern Hemisphere inferred from three nuclear data sets. Molecular Phylogenetics and Evolution 48:1027−40

    doi: 10.1016/j.ympev.2008.06.004

    CrossRef   Google Scholar

    [76]

    Group TAP, Chase MW, Christenhusz MJM, Fay MF, Byng JW, et al. 2016. An update of the Angiosperm Phylogeny Group classification for the orders and families of flowering plants: APG IV. Botanical Journal of the Linnean Society 181:1−20

    doi: 10.1111/boj.12385

    CrossRef   Google Scholar

  • Cite this article

    Wang S, Qiu J, Sun N, Han F, Wang Z, et al. 2025. Characterization and comparative analysis of the first mitochondrial genome of Michelia (Magnoliaceae). Genomics Communications 2: e001 doi: 10.48130/gcomm-0025-0001
    Wang S, Qiu J, Sun N, Han F, Wang Z, et al. 2025. Characterization and comparative analysis of the first mitochondrial genome of Michelia (Magnoliaceae). Genomics Communications 2: e001 doi: 10.48130/gcomm-0025-0001

Figures(7)  /  Tables(1)

Article Metrics

Article views(763) PDF downloads(157)

ARTICLE   Open Access    

Characterization and comparative analysis of the first mitochondrial genome of Michelia (Magnoliaceae)

Genomics Communications  2 Article number: e001  (2025)  |  Cite this article

Abstract: Genus Michelia has various functions and is valuable in medicine, food, and agriculture. Many plastid genomes (plastomes) of Michelia have been released, but no mitochondrial genomes (mitogenomes) have been reported. In this study, using third-generation HIFI sequencing techniques, Michelia figo (M. figo) mitogenome was de novo assembled into a circular chromosome spanning 773,377 bp with a total GC content of 46.83%. Sixty six genes in total were annotated, including 41 protein-coding genes, 21 tRNA genes, and three rRNA genes. The mitogenome contains 1,514 dispersed repeats (> 30 bp), 39 tandem repeats, and 262 simple sequence repeats. Eighty one fragments originating from the M. figo plastome were detected in its mitogenome and three tRNA genes (trnD-GUC, trnW-CCA, and trnV-GAC) completely transferred from the plastome to the mitogenome. Repeats and collinearity analyses of four Magnoliaceae mitogenomes reveal substantial structural variations, a relatively low degree of collinearity, and significant genetic diversity of this genus. Phylogenetic analysis showed that two phylogenetic trees constructed separately based on mitogenomes and plastomes accurately depict the phylogenetic relationship of M. figo. This study offers the first comprehensive comparative genomic and phylogenetic analysis of the M. figo mitogenome, facilitating the development of genetic markers, taxonomic classification, and resource exploration within the Michelia genus.

    • Mitochondria, as semi-autonomous organelles with their unique genetic material and systems, play crucial roles in plant energy metabolism by generating ATP through oxidative phosphorylation[1,2]. In plants, mitochondria not only participate in energy production, but also collaborate with other organelles to maintain cellular homeostasis[3]. In addition, mitochondria also participate in regulating processes such as apoptosis, playing an important regulatory role in plant life activities[4]. However, despite the undisputed significance of mitochondria, previous research on plant mitochondrial genomes (mitogenomes) has been scant. Initially, sequencing and assembly techniques which were primarily developed for nuclear genomes, encountered significant challenges in handling mitogenomes due to their complicated structures, leading to fragmented or incomplete mitogenome assemblies. Moreover, early bioinformatic tools were not optimized for the unique characteristics of mitogenomes. This limitation further hindered the accurate assembly and analysis of plant mitogenomes. Nevertheless, with the continuous advancement of sequencing technology, especially the widespread application of long-read sequencing, research on plant mitogenomes has gradually increased in recent years[58].

      The mitogenomes of most higher plants exhibit substantial variability in structure and size[9]. Plant mitogenomes contain numerous repetitive sequences (repeats), leading to significant variations in size and structure through frequent recombination events[10]. The size of plant mitogenomes is 100–1,000 times larger than that of animals (15-18 kb)[11]. Plant mitogenomes stand out not just for their exceptional size, but also for the notable variation in size they display across diverse species. For example, Viscum scurruloideum[12] has a mitogenome of only 66 kb, while the mitogenome of Larix sibirica is 11.7 Mb[13]. Additionally, the intricate structure of plant mitogenomes further adds to their complexity, with most of them existing as a single circular molecule, while a minority exist as linear or branched molecules. It has been reported that the mitogenomes of Populus simonii[14] and Fagopyrum esculentum[15] are composed of three and ten circular molecules, respectively. Although plant mitogenomes differ significantly in terms of size and structure, the number of genes remains comparably stable and conserved, with a similar core set of PCGs, rRNAs, and tRNAs which are essential for respiratory function, and translation processes[16]. Intracellular gene transfer (IGT) can further complicate the mitogenome, as sequences from both the plastid and nuclear genomes coexist in plant mitogenomes[17]. For instance, the sequences of the nuclear and plastid genomes account for 46.5% and 1.4% of the Cucumis melo mitogenome, respectively[18]. Therefore, the plant mitogenome, due to its complex characteristics, is an ideal system for exploring genome complexity.

      Michelia figo (M. figo) belongs to the genus Michelia of the Magnoliaceae family, which is the second-largest genus and a relatively evolved group in the Magnoliaceae family. There are about 80 Michelia species in the world, predominantly distributed in tropical, subtropical, and temperate regions of Asia, of which approximately 70 species are distributed in China[19,20]. The broad spectrum of physiological activities exhibited by the genus Michelia underscores its potential applications in medicine, food, agriculture, and other domains[21]. The flowers, leaves, branches, and other parts of these species contain abundant aromatic oil that has been traditionally used in China, India, and other regions for treating fever, leprosy, inflammation, and other ailments[22]. Michelia species usually serve as valuable sources of bioactive compounds, exhibiting antibacterial[23], and antioxidant properties[24]. Furthermore, the methanolic extract from the leaves of M. figo has a concentration-dependent vasodilatory effect, having widespread applications in medicine[25].

      A previous study has reported the complete plastome of M. figo and analyzed its phylogenetic relationship with other Michelia species based on (plastid genomes) plastomes[26]. However, the mitogenome of M. figo and its phylogenetic status based on mitogenomes remain unexplored. Additionally, many plastomes of Michelia have been released[2729], but no mitogenomes have been reported for this genus. Consequently, to further explore the evolution and genetics of M. figo, this study has successfully assembled the complete mitogenome of M. figo. Comparative genomic and phylogenetic analyses were undertaken to elucidate the characteristics of the mitogenomes of M. figo and other Magnoliaceae species. These analyses will offer crucial theoretical and data-driven supports for genomic research, biological functions, and mitogenome evolution in M. figo and other Michelia species.

    • In this study, we collected fresh leaves of M. figo at Nanjing Forestry University, Nanjing, Jiangsu Province, China (118.81° E, 32.07° S). Before DNA extraction, fresh leaves were immediately frozen in liquid nitrogen to preserve their integrity and subsequently stored in a laboratory freezer maintained at −80 °C. The total genomic DNA was extracted using the CTAB method[30]. The quality of the DNA sample was evaluated using 1% agarose gel electrophoresis, while its concentration was accurately determined using a NanoDrop ND 2000 (ThermoFisher Scientific, Waltham, MA, USA)[31]. The size of the genomic insert fragments is 15−18 kb. Then the sequencing libraries were constructed using the high-integrity genomic DNA through SMRTbell Express Template Prep Kit 2.0 (PacBio Biosciences, Menlo Park, CA, USA). We ultimately obtained the HiFi sequencing data from the PacBio Revio platform.

    • The HiFi sequencing data was fed into PMAT v1.31[32] to assemble the mitogenome of M. figo. The parameters were 'autoMito -st hifi -g 2.2G -CPU 50'. The nuclear genome size of M. figo was estimated using the genome of Magnolia biondii as a reference[33]. After using PMAT, the raw assembly graph of M. figo mitogenome was composed of 12 contigs, containing four pairs of repeats. Using Bandage[34], we obtained the circular mitogenome of M. figo by decoding the raw assembly graph, taking into account the copy number of each contig. The mitogenome of M. figo was annotated using the online program PMGA[35]. The rRNA and tRNA genes were then verified by BLASTN[36] and tRNAscan-SE v2.0[37], respectively. Finally, an online tool PMGmap[38] was used to draw the mitogenome map.

    • The online tool MISA[39] was used to detect simple sequence repeats (SSRs) of the M. figo, M. biondii, and M. officinalis mitogenomes. We set the repetition thresholds at 10, 5, 4, 3, 3, and 3 for mononucleotides, dinucleotides, trinucleotides, tetranucleotides, pentanucleotides, and hexanucleotides, respectively. The minimal distance between two SSRs was established as 100 bp. Meanwhile, the online tool TRF[40] was utilized to detect tandem repeats with default parameters. REPuter[41] was used to detect dispersed repeats and the parameters were set as follows: hamming distance of three, maximum computed repeats of 5000, and a minimal repeat size of 30 bp[42]. Codon composition and usage of the M. figo mitogenome were analyzed using CondonW v1.4.4 (https://codonw.sourceforge.net/) with default parameters.

    • We obtained the plastid genome (plastome) of M. figo from the NCBI with the accession number of NC_053861.1. Then, we used BLASTN[36] to identify the homologous fragments between the mitogenome and plastome, and utilized TBtools to visualize the results[43]. We selected three mitogenomes of Magnoliaceae (L. tulipifera, M. biondii, and M. officinalis) for the collinearity analysis with M. figo. The collinear blocks were identified using MUMmer v4.0[44] with default parameters. We chose collinear blocks that exceeded 5,000 bp for subsequent analysis. NGenomeSyn v1.0[45] was finally used to visualize the results.

    • To further clarify the phylogenetic location of M. figo, two phylogenetic trees were constructed using 15 plant mitogenomes and plastomes respectively, including two species of Gymnosperm (Cycas taitungensis, and Ginkgo biloba), three species of ANA clade (Amborella trichopoda, Nymphaea colorata, and Schisandra sphenanthera), four species of Magnoliidae (L. tulipifera, M. figo, M. officinalis, and M. biondii), three species of monocots (Apostasia shenzhenica, Cocos nucifera, and Sorghum bicolor), and three species of core eudicots (Ilex pubescens, Sapindis mukorossi, and Ficus carica). Among these plant species, C. taitungensis and G. biloba were selected as outgroups. We used in-house Python scripts to select shared genes and used MAFFT v7.407[46] to compare the shared genes. After trimming the results using trimAl v1.4[47], IQ-TREE v2.0.3[48] was utilized to construct the phylogenetic trees based on the maximum likelihood (ML) method with 1,000 bootstraps.[49]. Both plastid and mitochondrial trees were found to be best fit by the GTR + F + I + G4 model. Finally, the online tool iTOL[50] was used to visualize and optimize the results.

    • Using the Revio sequencing platform, we obtained a total of 410,107 HiFi sequencing reads with 5.83 Gb in length and the N50 value of 14,355 bp. After using PMAT v1.31 to generate the raw assembly graph of the M. figo mitogenome (Fig. 1a), we utilized Bandage to disentangle the mitogenome graph resulting in a circular molecule with 773,377 bp in length (Fig. 1b). The total GC content is 46.83%, with 26.56%, 26.21%, 23.37%, and 23.46% for bases A, T, G, and C, respectively. The M. figo mitogenome was annotated with 66 genes, comprising 41 protein-coding genes (PCGs), 21tRNA, and three rRNA, as detailed in Table 1. Figure 1c provides a visual representation of the functional classification and specific positions of the annotated genes. The majority of genes are present in a single-copy format, with the exception of three genes (rps7, trnM-CAU, and trnP-UGG), possessing multiple copies. Moreover, we found that a total of 10 genes harbor introns (ccmFc, rpl2, rps3, rps10, cox2, nad1, nad2, nad4, nad5, and nad7) (Supplementary Fig. S1). Most of these introns are cis-spliced, with nad1, nad2, and nad5 containing a few trans-spliced introns.

      Figure 1. 

      Structural and functional features of the M. figo mitogenome. (a) The raw assembly graph of the M. figo mitogenome. (b) The disentangled graph of the M. figo mitogenome. (c) Circular mitogenome map of M. figo. Genes depicted outside the outer circle undergo clockwise transcription, while those positioned within the inner circle undergo counter-clockwise transcription. The legends of different colors positioned in the bottom left corner serve to distinguish genes based on their specific functionalities.

      Table 1.  Gene compositions of the Michelia figo mitogenome.

      Group of genes Name Start codon Stop codon Length Amino acids
      ATP synthase atp1 ATG TGA 1,530 509
      atp4 ATG TAA 582 193
      atp6 ATG TAG 891 296
      atp8 ATG TAA 480 159
      atp9 ATG TAA 261 86
      Cytochrome c biogenesis ccmB ATG TGA 621 206
      ccmC ATG TAA 960 319
      ccmFc* ATG TAA 1,359 452
      ccmFn ATG TAG 1,806 602
      Ubichinol cytochrome c reductase cob ATG TGA 1,182 393
      Cytochrome c oxidase cox1 ACG TAA 1,584 527
      cox2** ATG TAA 759 252
      cox3 ATG TGA 798 265
      Maturases matR ATG TAG 1,959 652
      Transport membrane protein mttB ACG TGA 768 255
      NADH dehydrogenase nad1***+ ACG TAA 978 325
      nad2**** ATG TAA 1,467 488
      nad3 ATG TAA 357 118
      nad4*** ATG TGA 1,488 495
      nad4L ACG TAA 303 100
      nad5**++ ATG TAA 2,013 670
      nad6 ATG TGA 735 244
      nad7**** ATG TAG 1,185 394
      nad9 ATG TAA 573 190
      Large subunit of ribosome (LSU) rpl10 ATG TAA 471 156
      rpl16 GTG TAA 435 144
      rpl2* ATG TAG 1,665 697
      rpl5 ATG TAA 561 186
      Small subunit of ribosome (SSU) rps1 ATG TAA 606 201
      rps10* ATG TGA 420 139
      rps11 ATG TGA 552 183
      rps12 ATG TGA 378 125
      rps13 ATG TGA 351 116
      rps14 ATG TAG 303 100
      rps19 ATG TAA 282 93
      rps2 ATG TAA 657 218
      rps3* ATG TAA 1,572 523
      rps4 ACG TAA 1,071 356
      rps7 (2) ATG/ATG TAA/TAA 450 149
      Succinate dehydrogenase sdh3 ATG TAA 330 109
      sdh4 ATG TGA 447 148
      Ribosomal RNAs rrn5 117
      rrnL 3,560
      rrnS 2,087
      Transfer RNAs trnC-GCA 71
      trnD-GUC 74
      trnE-UUC 72
      trnF-GAA 74
      trnfM-CAU 74
      trnG-GCC 73
      trnH-GUG 74
      trnI-CAU 81
      trnK-UUU 73
      trnM-CAU (2) 73/73
      trnN-GUU 72
      trnP-UGG (3) 75/74/75
      trnQ-UUG 72
      trnS-GCU 88
      trnS-UGA 87
      trnV-UAC 73
      trnW-CCA 74
      trnY-GUA 83
      * Indicates the cis-spliced introns, and + indicates the trans-spliced introns. The number of * and + represents the number of introns. The number in parentheses represents the number of genes.
    • The relative synonymous codon usage (RSCU) value is equal to 1 when there is no synonymous codon usage preference. In the M. figo mitogenome, the RSCU values of AUG (Met), UGG (Trp), and AGC (Ser) are 1 (Supplementary Table S1). Twenty nine codons exhibit RSCU values above 1, among which the codon AGA (Arg) possesses the highest RSCU value, especially 1.57. Additionally, the RSCU values of 32 codons are lower than 1, with CGU(Arg) exhibiting the lowest RSCU value of 0.64. We also compared the RSCU of M. figo mitogenome with the other three Magnoliaceae mitogenomes. The result shows that the relative synonymous codon usage is highly consistent (Fig. 2a), with the codon of AGA (Arg) exhibiting the highest RSCU value in these mitogenomes (Supplementary Table S1). We further calculated the frequency of codon usage, revealing a remarkable similarity across the mitogenomes of Magnoliaceae (Fig. 2b & Supplementary Table S2).

      Figure 2. 

      Codon usage of four Magnoliaceae mitogenomes. (a) Stacked column plots of the relative synonymous codon usage. (b) Heatmap of the codon usage frequencies.

    • The M. figo mitogenome contains abundant repeats (Fig. 3). Using the online tool REPuter, we detected 1,514 pairs of dispersed repeats (≥ 30 bp), including 758 pairs of forward repeats and 756 pairs of palindromic repeats (Supplementary Table S3). However, there are no complementary and reverse repeats. Additionally, the M. figo mitogenome was found to harbor 39 tandem repeats, with lengths varying from 14 to 52 bp (Supplementary Table S4), with matching identity greater than 64%. Altogether, 262 SSRs were identified in the M. figo mitogenome (Supplementary Table S5), most of which are tetranucleotides (96), followed by mononucleotides (55), and dinucleotides (55).

      Figure 3. 

      The distribution of repeats in the M. figo mitogenome. From the center outward, the first circle shows the mitogenome of M. figo, the second and third circle shows tandem repeats and simple sequence repeats, respectively. The inner lines represent the dispersed repeats. The legends of different colors positioned in the bottom left corner represent the dispersed repeats of different lengths.

      To further investigate the repeats in the mitogenomes of Magnoliaceae, we detected and compared the tandem, SSRs and dispersed repeats in the mitogenomes of three Magnoliaceae species (M. biondii, M. officinalis, and M. figo). The results show that only M. officinalis contains two pairs of reverse repeats and one pair of complementary repeats (Fig. 4a), while the number of tandem repeats does not exhibit significant variation (Supplementary Tables S6 & S7). The M. officinalis mitogenome exhibits the highest number of dispersed repeats (3,609), followed by M. biondii (2,800) and M. figo (1,514) (Fig. 4a, Supplementary Tables S8 & S9). The distribution of dispersed repeat lengths across the three mitogenomes is also similar (Fig. 4b), with most repeats ranging from 30 to 49 bp, and only a few exceeding 500 bp. Additionally, comparative results of SSRs reveal that all three mitogenomes of Magnoliaceae contain six SSR types (Fig. 4c, Supplementary Tables S10 & S11), with M. officinalis exhibiting the highest number of SSRs (327). The diversity of SSR types in Magnoliaceae mitogenomes does not vary significantly, with the exception of a notable difference in the case of mononucleotides.

      Figure 4. 

      (a) Type and number of simple sequence repeats in the mitogenomes of two Magnolia species and M. figo. (b) Length and number of dispersed repeats in the mitogenomes of two Magnolia species and M. figo. (c) The different colored legends indicate different species.

    • We identified 81 fragments transferred from the plastome to the mitogenome of M. figo (Fig. 5 & Supplementary Table S12), ranging from 50 to 4,665 bp. The entire length of MTPTs measures 42,791 bp, constituting 5.53% of the whole mitogenome. Most of these MTPTs range from 50 to 500 bp in length, and only 11 fragments exceeding 1 kb, with the longest fragment reaching 4,666 bp. A total of 15 plastid genes were found to be located on MTPTs, including nine PCGs (psbL, psbF, psbE, petL, petG, rps8, rpl14, rps7, and ndhB) and six tRNA genes (trnD-GUC, trnY-GUA, trnE-UUC, trnW-CCA, trnP-UGG, and trnV-GAC). Notably, trnD-GUC, trnW-CCA, and trnV-GAC are completely transferred from the plastome to the mitogenome. Additionally, we found that MTPT22, MTPT65, MTPT68, and MTPT69 are located in repeat regions.

      Figure 5. 

      (a) Homologous sequences between mitogenome and plastome. The plastome is represented by the green circular segment and the mitogenome by the gray circular segment, and two different kinds of yellow lines represent the homologous fragments. The legends of different colors positioned in the bottom right corner represent fragments of different lengths. (b) Lengths and numbers of these homologous fragments in the M. figo mitogenome.

    • We conducted a collinearity analysis by comparing the mitogenome of M. figo with three other Magnoliaceae mitogenomes (L. tulipifera, M. officinalis, and M. biondii). As illustrated in Fig. 6a, a total of 40 locally collinear blocks (LCBs) were identified between the mitogenomes of M. figo and M. officinalis (Supplementary Table S13). The cumulative length of these colinear blocks amounts to 416,577 bp, comprising approximately 53.86% of the M. figo mitogenome. Among these colinear blocks, the longest is 34,737 bp, and the average length is 10,413 bp. Between the mitogenomes of M. biondii and M. figo, we detected 42 LCBs, accounting for 59.04% (456,622 bp) of the M. figo mitogenome (Supplementary Table S14). The longest colinear block is 44,187 bp, and the average length is 10,871 bp. Between the mitogenomes of M. biondii and L. tulipifera, a total of 27 LCBs were identified (Supplementary Table S15), accounting for 52.14% (287,725 bp) of the L. tulipifera mitogenome. The longest colinear block is 30,420 bp, and the average length is 10,656 bp. The average colinear lengths of the four mitogenomes are highly consistent (Fig. 6b).

      Figure 6. 

      Schematic representation of the collinearity among four Magnoliaceae mitogenomes. (a) Collinearity plots of the four Magnoliaceae mitogenomes. The mitogenomes are shown by the bars in each row, and collinear regions are indicated by the connecting lines in the center. (b) Lengths and numbers of collinear blocks. The different colored legends indicate homologous fragments between different species.

    • To further elucidate the phylogenetic position of M. figo, we constructed two phylogenetic trees based on 18 mitochondrial and 61 plastid PCGs from 15 species, respectively (Supplementary Tables S16 & S17). As illustrated in Fig. 7, 91.67% of the total nodes possess bootstrap support values exceeding 80%, including 20 nodes that achieve the maximum support of 100%. From the basal group downward, the bootstrap value for the separation of Magnoliidae from the clade consisting of monocots and core eudicots is 100%. In the Magnoliidae, the bootstrap value for the separation of Magnoliales and Laurales is 100%. Furthermore, we found that M. biondii and M. officinalis firstly grouped, this clade subsequently grouped with M. figo with a 100% bootstrap value, indicating that Michelia is closely related to Magnolia. The phylogenetic trees constructed based on mitogenomes and plastomes exhibit remarkable consistency, supporting that Michelia is closely related to Magnolia.

      Figure 7. 

      The phylogenetic trees constructed based on M. figo and other 14 plant mitogenomes and plastomes. The bootstrap values are clearly displayed within each node. The utilization of distinct colors serves to show the various groups to which the specific species belong. (a) The tree was constructed based on 18 shared mitochondrial genes. (b) The tree was constructed based on 61 shared plastid genes.

    • Plant mitogenomes frequently undergo recombination events mediated by repeats, resulting in great differences in their size[51]. Despite the closely related species, notable variations in mitogenome size can still be observed. For example, the size of the Silene latifolia mitogenome (253 kb) differs by 45 times compared to that of S. conica (11.3 Mb)[52]. The mitogenomes of Cucumis melo (2.9 Mb) and Citrullus lanatus (379 kb) differ by more than seven times[53]. The frequent recombination events of plant mitogenomes may integrate a large amount of foreign DNA during evolution, potentially contributing to the great differences in plant mitogenomes size. In this study, the mitogenome length of M. figo (773,377 bp) is relatively short in Magnoliaceae, with the longest in M. biondii (967,100 bp), followed by M. officinalis (930,306 bp) and M. liliiflora (865,191 bp). The shortest mitogenome is L. tulipifera (551,806 bp), accounting for only 60% of the mitogenomes of M. biondii and M. officinalis.

      Frequent recombination events not only lead to great differences in mitogenome size, but also contribute to complex and diverse structures of plant mitogenomes[54], ranging from single circular and linear structures to more complex branched linear, branched circular, and other complex structures[55]. It has been reported that the mitogenomes of Amborella trichopoda[56], Rhopalocnemis phalloi[57], and Panax notoginseng[58] are complex dynamic structures resulting from recombination. The mitogenome structures of Magnoliaceae are relatively conserved, with the majority being assembled into a single circular chromosome (M. biondii, M. officinalis, M. figo, and L. tulipifera). However, the mitogenome of M. liliiflora exhibits a linear chromosome. Additionally, the mitogenomes of angiosperms exhibit rapid structural differentiation and loss of collinearity, even those of closely related species[59,60]. In this study, using the nucmer program of MUMmer, numerous colinear regions and genomic rearrangements were identified among four Magnoliaceae mitogenomes. The lengths of these colinear blocks account for more than half of each mitogenome. The results of collinearity analysis reveal there may have been significant genomic rearrangements in the mitogenomes of Magnoliaceae species during their evolutionary history.

      The mitogenomes of angiosperms generally encode a core set of 24 PCGs: nad1-7, 9, and 4L; cob; cox1-3; ccmB, C, Fc, and Fn; atp1, 4, 6, 8, and 9; mttB/tatC; and matR. Although these core genes are present in most mitogenomes, there are significant variations in their quantity, position, and arrangement, even within mutants of the same species. In addition to the 24 conserved PCGs, plant mitogenomes also possess 19 standard variable genes, consisting of five large subunits of ribosome proteins (rpl2, 5, 6, 10, and 16), 12 small subunits of ribosome proteins (rps1-4, 7, 8, 10-14, and 19), and two respiratory genes (sdh3-4). Among these variable genes, the large and small subunits of ribosome genes are missing relatively frequently[61]. The mitogenome of M. figo harbors all 24 core PCGs, with only two variable PCGs (rpl6 and rps18) being lost. Similarly, the mitogenomes of M. biondii[33] and L. tulipifera[62] have retained nearly all ancestral PCGs. However, the Silene vulgaris mitogenome has nearly lost all variable PCGs with the exception of rps13. Moreover, the Viscum scurruloideum mitogenome has lost the entirety of 11 of the 24 core PCGs, including ccmB, matR, and all NADH dehydrogenase genes[12]. The gene content in the mitogenomes of Magnoliaceae is relatively abundant[63], suggesting that they may have undergone less gene loss during the mitogenome evolution.

      Plant mitogenomes vary significantly in the number of introns. The Silene latifolia mitogenome has only 19 introns[64], while the Selaginella moellendorffii mitogenome contains the largest number of 37 introns[65]. The M. figo mitogenome contains 25 introns in 10 PCGs (ccmFc, rpl2, rps3, rps10, cox2, nad1, nad2, nad4, nad5, and nad7), consisting of 22 cis-splicing and three trans-splicing introns. Cis-splicing is prevalent in most introns of angiosperm mitogenomes, whereas nad1, nad2, and nad5 evolved a split structure that requires trans-splicing[63]. Similar to the majority of angiosperm mitogenomes, the intron rps3i257 in the M. figo mitogenome is completely lost during differentiation[66]. These results indicate that introns are frequently gained or lost during the evolution of plant mitogenomes (Supplementary Fig. S2)[63].

      Plant mitogenomes are characterized by the abundance of repeats, contributing to the complexity and diversity of mitogenome sizes and structures through frequent recombination events[10]. The intense recombination events mediated by long repeats (> 500 bp) facilitate reversible recombination, regulate the molecular conformation of the mitogenome, and ultimately contribute to the expansion and complexity of plant mitogenomes[54]. In this study, the M. figo mitogenome exhibits the lowest abundance of SSRs and long repeats (> 500 bp), whereas M. biondii mitogenome displays the highest abundance in Magnoliaceae. It can be inferred that it is likely to undergo less recombination events during the evolution of M. figo mitogenome, while the M. biondii mitogenome may experience more recombination events. Simultaneously, variations in the quantity of repeats may result in significant differences in the size of mitogenomes. For example, the mitogenome sizes of bryophytes remain relatively stable at approximately 110 kb, probably due to the scarcity of repeats within their mitogenomes. This scarcity contributes to the conserved and stable structure of the bryophyte mitogenomes. By contrast, the mitogenomes of ferns exhibit a significant number of repeats, accounting for their relatively large sizes[10]. In this study, the mitogenomes of M. biondii and M. officinalis exhibit a significantly higher numbers of repeats compared to M. figo, potentially explaining the differences in their mitogenome sizes.

      DNA fragment transfer events between the plastomes and mitogenome, as well as among different species, are recurrent phenomena that occur during the evolution of the plant mitogenome[67]. The lengths and similarities of these cp-derived fragments vary among different species[68]. The total length of MTPTs in the M. figo mitogenome is 42,791 bp, constituting 5.53% of the whole mitogenome. This proportion is significantly higher than that observed in numerous other mitogenomes, such as Arabidopsis thaliana (0.8%) , Glycine max (0.6%), Silene conica (0.2%), and Vigna angularis (0.1%)[69]. At the other extreme, the length of MTPTs accounts for 10.5% of the Boea hygrometrica mitogenome[70]. The MTPTs in the M. figo mitogenome are notably abundant, with the longest fragment spanning 4,666 bp, and the majority of MTPTs ranging from 50 to 500 bp. These sizable MTPTs are presumed to have significant impacts on plant mitogenome evolution, thereby contributing to genetic diversity[17,71]. Additionally, it is frequently observed that these transferred fragments contain PCGs. The number of PCGs in MTPTs exhibits significant variation in plant mitogenomes, ranging from seven in Brassica to 22 in Nicotiana[72]. In the mitogenome of M. figo, nine PCGs (psbL, psbF, psbE, petL, petG, rps8, rpl14, rps7, and ndhB) are located in MTPTs. However, PCGs in MTPTs turned out to degenerate as a result of sequence alterations and the absence of RNA editing[73,74]. Consequently, PCGs in MTPTs may have limited functional significance in mitogenomes, potentially acting as non-essential sequences[72].

      In this study, we reconstructed two phylogenetic trees based on 18 mitochondrial and 61 plastid PCGs from 15 species, respectively. Both trees exhibit remarkable consistency, supporting that Michelia is closely related to Magnolia, which is consistent with previous studies[26,75]. Additionally, the topological structure of the two phylogenetic trees is also highly consistent with the Angiosperm Phylogeny Group IV (APG IV) system[76]. However, due to the scarcity of mitogenomes in Michelia, we are unable to expand our discussion on the phylogenetic relationships in this genus.

    • In this study, we have successfully sequenced and assembled the mitogenome of M. figo for the first time. The circular mitogenome of M. figo is 773,377 bp in length, encoding 41 PCGs, 21 tRNA genes and three rRNA genes. A total of 22 cis- and three trans-splicing introns were identified in the M. figo mitogenome. The M. figo mitogenome contains abundant repeats, with 1,514 pairs of dispersed repeats, 39 tandem repeats, and 262 SSRs. Additionally, we identified 81 fragments (42,791 bp) that were transferred from the plastome to the mitogenome of M. figo, constituting 5.53% of the whole mitogenome. The M. figo mitogenome is characterized by the abundance of repeats and MTPTs, contributing to the complexity and diversity of mitogenome size and structure. Furthermore, comparative analyses of four Magnoliaceae mitogenomes reveal significant genetic diversity of this genus. Two phylogenetic trees, constructed independently based on the mitogenomes and plastomes of 15 species, depicted the phylogenetic relationship of M. figo. This study presents the first comprehensive genomic and phylogenetic analyses of the M. figo mitogenome, providing crucial theoretical insights and data support for the development of genetic markers, classification, and resource utilization within the Michelia genus.

      • The work is supported by the Natural Science Foundation of Jiangsu Province (BK20220414) and Jiangsu Students' Innovation and Entrepreneurship Training Program (202210298119Y). We thank Assoc. Prof. Kewang Xu from Nanjing Forestry University for collecting the sample of M. figo.

      • This study has rigorously adhered to relevant institutional, national, and international guidelines and regulations. Moreover, the study did not involve the use of any endangered or protected species. The M. figo plant leaves utilized in this experiment were collected at Nanjing Forestry University.

      • The authors confirm contribution to the paper as follows: study conception and design: Bi C, Yang Y; analysis and interpretation of results: Wang S, Sun N, Qiu J, Han F; materials collection and experiments conduct: Bi C, Qiu J; draft manuscript preparation:Wang S; manuscript revision presentation of comments: Bi C, Wang Z, Yang Y. All authors reviewed the results and approved the final version of the manuscript.

      • The mitochondrial genome supporting this study is available at GenBank with accession number: NC_082234.1. The HiFi sequencing data of M. figo is deposited in the SRA repository under SRR28267342.

      • The authors declare that they have no conflict of interest.

      • # Authors contributed equally: Suyan Wang, Jing Qiu

      • Supplementary Table S1 The relative synonymous codon usage of amino acids in the mitogenome of Michelia figo, Magnolia biondii, Magnolia officinalis, and Liriodendron tulipifera.
      • Supplementary Table S2 The frequency of codon usage in the mitogenome of Michelia figo, Magnolia biondii, Magnolia officinalis, and Liriodendron tulipifera.
      • Supplementary Table S3 Dispersed repeat sequences identified in the Michelia figo mitogenome.
      • Supplementary Table S4 Tandem repeat sequences identified in the Michelia figo mitogenome.
      • Supplementary Table S5 The simple sequence repeats identified in the Michelia figo mitogenome.
      • Supplementary Table S6 Dispersed repeat sequences identified in the Magnolia biondii mitogenome.
      • Supplementary Table S7 Dispersed repeat sequences identified in the Magnolia officinalis mitogenome.
      • Supplementary Table S8 Tandem repeat sequences identified in the Magnolia biondii mitogenome.
      • Supplementary Table S9 Tandem repeat sequences identified in the Magnolia officinalis mitogenome.
      • Supplementary Table S10 The simple sequence repeats identified in the Magnolia biondii mitogenome.
      • Supplementary Table S11 The simple sequence repeats identified in the Liriodendron tulipifera mitogenome.
      • Supplementary Table S12 The homologous DNA fragment between mitogenome and cpgenome of Michelia figo.
      • Supplementary Table S13 The collinear blocks between mitogenomes of Michelia figo and Magnolia officinalis.
      • Supplementary Table S14 The collinear blocks between mitogenomes of Michelia figo and Magnolia biondii.
      • Supplementary Table S15 The collinear blocks between mitogenomes of Magnolia biondii and Liriodendron tulipifera.
      • Supplementary Table S16 Genes used for phylogenetic analysis.
      • Supplementary Table S17 The genomic information of the species used in this study.
      • Supplementary Fig. S1 The map of genes containing introns. This diagram illustrates the distribution of cis- and trans-introns.
      • Supplementary Fig. S2 Heat maps of PCG and intron contents among 15 mitogenomes (a) Comparison of PCG contents among 15 mitogenomes. The gene numbers are shown on the top right. (b) Comparison of intron contents among 15 mitogenomes. The intron numbers are shown on the top right.
      • Copyright: © 2025 by the author(s). Published by Maximum Academic Press, Fayetteville, GA. This article is an open access article distributed under Creative Commons Attribution License (CC BY 4.0), visit https://creativecommons.org/licenses/by/4.0/.
    Figure (7)  Table (1) References (76)
  • About this article
    Cite this article
    Wang S, Qiu J, Sun N, Han F, Wang Z, et al. 2025. Characterization and comparative analysis of the first mitochondrial genome of Michelia (Magnoliaceae). Genomics Communications 2: e001 doi: 10.48130/gcomm-0025-0001
    Wang S, Qiu J, Sun N, Han F, Wang Z, et al. 2025. Characterization and comparative analysis of the first mitochondrial genome of Michelia (Magnoliaceae). Genomics Communications 2: e001 doi: 10.48130/gcomm-0025-0001

Catalog

  • About this article

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return