Skip to main content

C4-like Sesuvium sesuvioides (Aizoaceae) exhibits CAM in cotyledons and putative C4-like + CAM metabolism in adult leaves as revealed by transcriptome analysis

Abstract

Background

The co-occurrence of C4 and CAM photosynthesis in a single species seems to be unusual and rare. This is likely due to the difficulty in effectively co-regulating both pathways. Here, we conducted a comparative transcriptomic analysis of leaves and cotyledons of the C4-like species Sesuvium sesuvioides (Aizoaceae) using RNA-seq.

Results

When compared to cotyledons, phosphoenolpyruvate carboxylase 4 (PEPC4) and some key C4 genes were found to be up-regulated in leaves. During the day, the expression of NADP-dependent malic enzyme (NADP-ME) was significantly higher in cotyledons than in leaves. The titratable acidity confirmed higher acidity in the morning than in the previous evening indicating the induction of weak CAM in cotyledons by environmental conditions. Comparison of the leaves of S. sesuvioides (C4-like) and S. portulacastrum (C3) revealed that PEPC1 was significantly higher in S. sesuvioides, while PEPC3 and PEPC4 were up-regulated in S. portulacastrum. Finally, potential key regulatory elements involved in the C4-like and CAM pathways were identified.

Conclusions

These findings provide a new species in which C4-like and CAM co-occur and raise the question if this phenomenon is indeed so rare or just hard to detect and probably more common in succulent C4 lineages.

Peer Review reports

Introduction

In a number of eudicot families, C4 photosynthesis evolved in ancestrally succulent lineages (see [1], for an overview), prominent examples are Chenopodiaceae [2], Aizoaceae-Sesuvioideae [3], Portulacaceae [4] and Zygophyllaceae [5]. Some families like Aizoaceae and Portulacaceae also use crassulacean acid metabolism (CAM) as carbon concentrating mechanism (CCM). Both CCMs share the same core metabolic enzymes, both evolved repeatedly multiple times, however, it seems that they rarely co-occur. So far, the co-occurrence of C4 and CAM was verified for only four genera, Portulaca (Portulacaceae; [6]), Spinifex (Poaceae; [7]), Ottelia (Hydrocharitaceae; [8]) and Trianthema (Aizoaceae; [9]). Recent evidence demonstrated that C4 and CAM are operating in the same cells in Portulaca oleracea under drought conditions [10]. This integration is plausible due to several copies of core C4 genes [i.e. phosphoenolpyruvate carboxylase (PEPC)] that are recruited for C4 and CAM, respectively, sharing a set of biochemical reactions. Detecting the co-occurrence of C4 and CAM is laborious, requires living collections and an experimental approach, which is why this phenomenon has not been documented very often. However, we hypothesize that it might be more common in succulent C4 lineages than currently known.

C4 photosynthesis is an adaptive evolutionary response to the harmful effect of photorespiration under hot and dry growing conditions, by concentrating CO2 around RUBISCO [11]. This allows for a remarkably efficient photosynthesis, as well as water and nitrogen use. The C4 pathway is a complex combination of anatomical and biochemical specialization. In succulent C4 lineages the C4 anatomy is particularly diverse [1, 12]. Often the Kranz cells are not arranged as an inner wreath around the vascular bundles like in plants with typical Kranz anatomy, but form a continuous inner chlorenchyma layer around the central water storage tissue of the leaf (e.g., [3, 13]). The general pathway in which CO2 is converted to bicarbonate (HCO3) by carbonic anhydrase in mesophyll cells (MCs) and then fixed to the 3-carbon molecule phosphoenolpyruvate by the enzyme PEPC to form the 4-carbon molecule oxaloacetate (OAA) unifies all plants with C4 photosynthesis. OAA is then either reduced to malate or transaminated to aspartate. After diffusing to an adjacent Kranz cell, malate or aspartate is primarily decarboxylated by the enzymes either NADP-dependent malic enzyme (NADP-ME) or NAD-dependent malic enzyme (NAD-ME). This decarboxylation releases CO2 in high concentrations around RUBISCO and ensures high photosynthetic efficiency. This carbon concentration mechanism (CCM) is supported and facilitated by a decrease in the ratio of mesophyll to Kranz cells as opposed to the C3 ancestors.

Unlike C4 photosynthesis, the CCM of CAM photosynthesis is temporally asynchronous in a single-cell system. During the night, plants open stomata, and CO2 is fixed and converted to malate, which is stored in the vacuole as malic acid. During the day, stomata are closed and stored malate is transported out the vacuole and decarboxylated to release CO2 that is then fixed by RUBISCO and enters the Calvin cycle for sugar production. This asynchronous carbon fixation system allows plants to keep their stomata closed to avoid water loss through evapotranspiration during the hottest period of the day. Thus, plants with this type of metabolism are able to grow in hot and dry environments. While it is straight forward to detect obligate CAM plants by means of a strong carbon isotope signal and consistent differences between morning and evening acid concentrations, it is laborious to detect facultative or weak CAM plants that only induce CAM under stress [14, 15]. Weak CAM can neither be detected by carbon isotope ratios in C3 species nor in C4 species. In C3 species, the discrimination of RUBISCO towards the heavier C isotope and in C4 species the much higher activity of the PEPC in the C4 pathway hides the low CAM signal.

Aizoaceae comprise annual or perennial herbs, rarely shrubs or trees, growing in tropical and subtropical regions, predominantly in South Africa [16]. Most species of the family are succulent and many, especially from subfamily Mesembryanthemoideae and Ruschioideae are documented CAM plants [14]. In Aizoaceae, C4 photosynthesis is restricted to subfamily Sesuvioideae and likely evolved multiple times [3]. A striking diversity of leaf anatomical types and the occurrence of both biochemical subtypes of C4 (NAD-ME and NADP-ME) can be observed. In addition to this photosynthetic diversity, two species from Sesuvioideae have been reported to activate low CAM under drought, i.e., the C3 species Sesuvium portulacastrum and the C4 species Trianthema portulacastrum ([9, 15]). Yet another species of this subfamily arouses curiosity: Sesuvium sesuvioides, a succulent C4-like species with uncommon C4 features and photosynthetic plasticity during leaf aging. Structural, physiological, and biochemical analysis of Sesuvium sesuvioides indicated a relatively high MC/bundle sheath cell (BSC) ratio and the presence of RUBISCO large subunit together with PEPC in the MCs [3]. Furthermore, a decrease of C4 enzyme activities was observed from young to mature to senescent leaves [17]. Although Bohley et al. [17] did not observe any CAM activity under well-watered conditions, they did not exclude the existence of CAM under dry conditions.

Such species that exhibit photosynthetic variability may contain footprints left from the evolution of CAM and C4 photosynthesis and thus provide useful information to either disentangle or gain new insight into the evolution and regulation of C4 and CAM metabolism. From this perspective, the photosynthetic flexibility in subfamily Sesuvioideae represents an excellent potential study group. We need such models as with climate change, many agricultural regions approaching their potential peak of productivity [18], and with an estimated population of 10 billion people by 2050, C4 and CAM represent a promising way to increase productivity and hence yield to meet global demands for food owing to their intrinsic ability to thrive in hot and dry environments. Thus, both CCMs are targets for genetic engineering into C3 species. Substantial efforts taken in the past to introduce C4 and CAM features into C3 plants failed to reach the envisioned goals due to lack of knowledge of C4 and CAM photosynthesis at the system level [19]. Therefore, mechanisms underlying C4 and CAM anatomical structure, gene-specific expression, and regulation network in C4 must be clarified further [19], and each new mosaic stone will help to solve the conundrum.

To test our hypothesis that S. sesuvioides operates combined C4 and CAM photosynthesis, we (1) performed a comparative transcriptome analysis between cotyledons and young leaves of S. sesuvioides (C4-like species) and young leaves of Sesuvium portulacastrum (C3 species) under stressful conditions: high light intensity and drought. We then (2) investigated the integration of C4-like + CAM in S. sesuvioides via the identification of candidate genes linked to C4 and CAM previously identified in Portulaca [20]. Finally, we explored the regulatory elements controlling C4-like and C3 pathways. Our analyses revealed that S. sesuvioides is operating weak CAM in cotyledons and C4-like + CAM in leaves as proved by gene expression analysis and supported by acid titration. Moreover, C4-like + CAM candidate genes were found up-regulated during the day suggesting the integration of C4-like + CAM metabolism in S. sesuvioides.

Materials and methods

Plant materials

Plants of S. sesuvioides (C4-like) and S. portulacastrum (C3) were grown from seeds and cuttings respectively, in the experimental greenhouse of the Munich-Nymphenburg Botanical Garden, Germany. Sesuvium sesuvioides seeds were collected from a location situated ~ 80 km east of Sendelingsdrif, Karas, Namibia [~ 28.20946°S, 17.28936°E, 208 m altitude, voucher: Klak 2431 (BOL)] and S. portulacastrum plant materials were collected in Texas, USA (MSB Serial number 0394523; year collected: 2007). For simplicity, we will sometimes use C4 instead of C4-like. Germinated seedlings (about 1 cm) and one-year-old plants of S. sesuvioides and one-year-old plants of S. portulacastrum were transferred to climate chambers with the following parameters: photoperiod light/dark 14 h/10 h, 60% humidity, [CO2] = 400 ppm, maximum light intensity = 785 μmol/m2/s, day/night temperature of 25/22 °C. Plants were watered every 2 days. This created stressful conditions specifically drought with changes in leaves colors as shown in the pictures (Additional file 1). Samples for transcriptome were harvested two weeks after transferring the plants to the climate chamber. Three young adult leaves of three plants of each species were collected during the day (1 pm) and for S. sesuvioides cotyledons three replicates of four plants were harvested during day (at 1 pm) and night (9 pm, one hour after the light was off). The plant S. sesuvioides was first identified by Cornelia Klak and later confirmed by Katharina Bohley. Sesuvium portulacastrum was also identified by Katharina Bohley. Vouchers of S. sesuvioides (LS 257) and S. portulacastrum (LS 221) have been deposited at the Herbarium MSB with serial numbers M-0356568 and M-0356569 respectively.

Titratable acidity

Possibility of CAM activity under stressful conditions was investigated in cotyledons and leaves of S. sesuvioides and leaves of S. portulacastrum via comparative titratable acidity between 30 min before the end and 30 min before the beginning of the light period (19.30 h and 5.30 h, respectively). Cotyledons and leaves were harvested and snap-frozen in liquid nitrogen and stored at -20 °C. Since cotyledons were small, eight plants of S. sesuvioides for each harvesting time were collected. Stored cotyledons and leaves were chopped and weighed. About 50 mg of the cotyledons and leaves material were incubated at 60 °C in 20% ethanol for 60 min. The extract obtained was aliquoted into three replicates of the same volume (1 mL). The extracted acid was neutralized by adding 0.01 M NaOH in 1ul increments [21].

RNA extraction and sequencing

Total RNA was extracted from leaves and cotyledons as described by Siadjeu et al. [22] using innuPREP Plant RNA Kit (Analytik Jena AG, Jena, Germany). Total RNA quality control was performed using the 2100 Bioanalyzer (Agilent Technologies) and Agarose gel electrophoresis. Messenger RNA was purified from total RNA using poly-T oligo-attached magnetic beads. After quality control and fragmentation, the first-strand cDNA was synthesized using random hexamer primers followed by the second-strand cDNA synthesis. The library was ready after end repair, A-tailing, adapter ligation, size selection, amplification, and purification.

Transcriptome analysis

An overview of the bioinformatics pipeline presenting software and respective versions used is presented in Fig. 1. Sequence read quality control was assessed using FastQC (https://www.bioinformatics.babraham.ac.uk/projects/fastqc/) and summarized with MultiQC [23]. Random sequencing errors in reads were corrected with a k-mer-based method implemented in Rcorrector [24] and uncorrectable reads were removed from the reads using TranscriptomeAssemblyTools (Fig. 1) (https://github.com/harvardinformatics/TranscriptomeAssemblyTools). Low-quality reads and adapters were filtered using TrimGalore v0.6.7 (https://github.com/FelixKrueger/TrimGalore/releases). The rRNA reads were removed by aligning trimmed reads against the SILVA v-138 rRNA database using Bowtie2 v.2.4.5 [25]. De novo transcriptome assembly was performed using Trinity v.2.14.0 [26] with the following parameters (Trinity –seqType fq –SS_lib_type RF –max_memory 200G –min_contig_length 300 –CPU 16). The de novo transcriptome assembly quality was first confirmed by aligning cleaned reads back to the corresponding de novo transcriptome assembly using Bowtie2 v.2.4.5. Secondly, the BUSCO score (odb10), which evaluates the completeness of the transcriptome, was determined using BUSCO v4 [27]. For the downstream analysis, the initial assembly of each species was processed as follows. We reduced the transcriptome data by clustering transcripts with 98% similarity with CD-HIT v4.8.1 [28]. Then, we selected only transcripts harboring coding sequences with TransDecoder v.5.5.0 (https://github.com/sghignone/TransDecoder). TransDecoder performs a precomputed blastX alignment to the Uniprot protein sequence database to improve the prediction of open reading frames. Finally, we used the indexes of transcripts containing coding sequences to subset the initial assemblies. The reduced transcriptomes were used for differential expression analysis. For differential expression analysis, we exclusively focused on S. sesuvioides when comparing leaves collected during the day to cotyledons collected both during the day and at night. For the comparison between C3 and C4-like species, we used young adult leaves collected during the day.

Fig. 1
figure 1

Workflow of transcriptome data analysis. The blue color represents the software used. L: leaves, C: cotelydons

Transcript quantification and differential expression analysis

We quantified transcript abundance with Salmon by aligning the reads of each species to its reduced transcriptome (Fig. 1). The software tximport v. 1.24.0 [29] was used to import transcript level abundances, estimated counts, and effective lengths for differential expression analysis. For comparison between the leaves of C3 and C4-like species, we searched for orthologs between their reduced transcriptomes using OrthoFinder v2.5.4 [30]. Based on the orthology between C3 and C4-like species, we obtained unique indexes by blasting C3 and C4-like species transcripts against each other. The best blast hits with the lowest e-value and high bit scores were selected and considered homologous. Transcripts of the C4-like species were used as a reference for the unique ID. If a C4-like species transcript was hit multiple times with transcripts of the C3 species, only one randomly selected transcript was kept to get a similar number of transcripts for differential expression analysis. Finally, the unique C4-like ID was changed in the Salmon output of the C3 species before importing the data with tximport (Script is available at https://github.com/Siadjeu/Sesuvioideae_C4-CAM). We assessed differential expression with the program Deseq2 [31]. Transcripts with p-value and p-adjusted as false discovery rate < 0.05 were considered significantly expressed andLog2FC was set > 1.

Pathway and gene ontology (GO) annotation

Metabolic pathways and annotations of differentially expressed (DE) transcripts were assigned via the tool Mercator4 [32]. Swiss-prot protein sequences database and prot-scriber were included to improve the annotations. Unassigned transcripts were manually assigned based on the knowledge of the molecular functions in C4/CAM, photorespiration, and starch metabolism. We assigned GO terms to DETs using Blast2GO through OmicsBox with cutoff = 55, GO weight = 5, e-value = 1.e-5, HSP-hit coverage cutoff = 80 and hit filter = 500. We enriched the GO terms using Fisher’s exact test via OmicsBox.

Phylogenetic analysis of PEPC isoforms

To investigate whether various isoforms of PEPCs are involved in CAM or C4 photosynthesis, we conducted a phylogenetic analysis of PEPC derived from our study alongside three distinct genes (PPC-1E1, PPC-1E2 and PPC-2) that encode PEPC from 35 different species. It has been found that PPC-1E1 is consistently involved in both CAM and C4 photosynthesis [20]. We obtained the protein sequences of these genes from the UniProt database (https://www.uniprot.org/, accessed on 06.03.2024). To generate proteins from the transcript sequence of PEPC discovered in our study, we employed TransDecoder v5.7.1 (https://github.com/TransDecoder/TransDecoder), and extracted the longest open reading frames of these transcripts. We constructed a multiple sequence alignment using MAFFT v.7.520 [33], and subsequently built an unrooted maximum likelihood tree using RAxML v8.2.13 [34].

Co-expression network analysis

The co-expression analysis was performed using the unsupervised machine learning algorithm k-means in R. We used the three most popular methods for determining the optimal cluster: the Elbow and silhouette [35] methods and gap statistic [36]. The normalized read counts of DETs were used. The maximum number of clusters (k) was set to 10. If transcription factors (TFs) and phytohormones were clustered with C4-like or CAM genes, they were considered candidate TFs and phytohormones controlling C4-like or CAM photosynthesis. The k-means clustering script is available underhttps://github.com/Siadjeu/Sesuvioideae_C4-CAM.

Motif detection and enrichment

Transcript sequences of k-means clusters were analyzed for motif identification and enrichment. The program MEME suite v.5.5.0 [37] was employed to detect de novo motifs with the following parameters: E-value threshold = 0.05, minimum motif size = 6 bp. We checked for motif redundancy with TOMTOM [38] using the motif database JASPAR nonredundant core 2022. We enriched the detected motifs using AME [39] with the following parameters:ame --verbose 1 --oc. --scoring avg --method fisher --hit-lo-fraction 0.25 --evalue-report-threshold 0.05 --control --shuffle-- --kmer 2 MemeUpC3vsC4photoStach.fasta motif_db/JASPAR.

Results

Titratable acidity

To investigate CAM induction in leaves and cotyledons of S. sesuvioides and leaves of S. portulacastrum, we conducted titratable acidity tests. Interestingly, the tests revealed significant overnight acid accumulation in leaves (t-test, adjusted p-value = 0.001) and cotyledons (t-test, adjusted p-value = 0.018) of S. sesuvioides and in leaves of S. portulacastrum (t-test, adjusted p-value = 0.018) (Fig. 2). Variations in nocturnal acidification were about 7 μmol/g fresh weight of free acids (FA) in cotyledons of S. sesuvioides, and 22 and 20 μmol/g FA in leaves of S. sesuvioides and S. portulacastrum, respectively. Similar overnight acid accumulation has been reported in the leaves of S. portulacastrum under drought stress [40]. Moreover, acid accumulations of 15 and 18 ∆H+ have been observed in drought stress in Portulaca amilis and P. oleracea, respectively, which induce CAM photosynthesis under stress conditions [41]. These differences in acidity between the morning and night indicate a weak CAM induction. Therefore, we analyzed the transcriptome profiles of S. sesuvioides and S. portulacastrum in relation to C4 and CAM photosynthesis.

Fig. 2
figure 2

Boxplots of morning-night differences in titratable acidity of cotyledons (C) and leaves (SS) of S. sesuvioides and leaves of S. portulacastrum (SP). Values in the graph indicate the adjusted p values of significant differences between morning and night. FA: refers to free acidity, which is measured based on a normalization against the fresh weight of a sample

Transcriptome assembly and quality assessment

The initial transcriptome assemblies of S. sesuvioides and S. portulacastrum contained 313,669 and 248,314 transcripts that were highly complete and little fragmented (C:96.2%, F:2.1) and (C:95.9 F:2.4), respectively (Table 1). Only 27 % of transcripts (85,170) of S. sesuvioides and 31 % of S. portulacastrum (76,264) were predicted to possess coding sequences. However, no significant changes were observed in the BUSCO scores (S. sesuvioides: C: 95.3 %, F: 2.5 %; S. portulacastrum: C: 95.1 %, F: 2.8 %). Transcripts with coding sequences were used for differential expression analysis between cotyledons and leaves of S. sesuvioides. For cross-species differential expression analysis, 43,769 and 40,429 orthologous transcripts were identified in S. sesuvioides (C: 84.9 %, F: 3.5 %) and S. portulacastrum (C: 82.8 %, F: 4.6 %) with an overall alignment of 73 % and 76 %, respectively.

Table 1 Statistics of assembly and alignment for transcriptomes of Sesuvium sesuvioides and S. portulacastrum

Differential expression analysis across Sesuvium species

Two comparative analyses were carried out: (1) between leaves and cotyledons of S. sesuvioides, and (2) between leaves of S. sesuvioides and leaves of S. portulacastrum. Leaves and cotyledons of S. sesuvioides were clearly separated based on their expression profile (Fig. 3A). Likewise, C3 and C4-like species were clustered according to their photosynthetic type mainly along PC1 (Fig. 3B). A total of 6,063 transcripts were found to be significantly DE between leaves and cotyledons of S. sesuvioides during the day (L and CD), of which 2,492 were up-regulated in leaves and 3,571 in cotyledons (Fig. 3C). When comparing leaves during the day (L) and cotyledons of S. sesuvioides during the night (CN), 1242 were up-regulated in leaves and 2,941 in cotyledons. Comparison of cotyledons between night (CN) and day (CD) revealed that 1,706 and 821 transcripts were found to be up-regulated at night and day, respectively. Between the C3 (S. portulacastrum) and C4 (S. sesuvioides) species, we found 20,867 orthologous transcripts in a 1:1 relationship. Out of these orthologs, 3,860 transcripts were significantly up-regulated in the C3 species and 2,433 in the C4-like species (Fig. 3C).

Fig. 3
figure 3

Transcriptome expression patterns of Sesuvium. A Principal component analysis (PCA) of leaves vs cotyledons of S. sesuvioides, CD: cotyledons day, CN cotyledons night, L: leaves. B PCA of S. sesuvioides vs S. portulacastrum, SP251-253: S. portulacastrum replicates 1:3, SS251: S. sesuvioides replicates 1:3. C Number of differential expressed transcripts: C4_vs_C3, between leaves of S. sesuvioides (C4-like) and S. portulacastrum (C3); CN_vs_CD, between cotelydons of S. sesuvioides collected during the night and during the day; L_vs_CD, between leaves and cotelydons of S. sesuvioides both collected during the day; L_vs_CD, between leaves and cotyledons S. sesuvioides both collected during the night. The dash (-) sign indicates that transcripts are down-regulated

Functional annotation of DETs between S. sesuvioides and S. portulacastrum

Functional annotations of DETs to land plant protein sequences were assigned using Mercator4. To explore the difference between ancestral C3 photosynthesis to C4-like photosynthesis in S. sesuvioides, we compared the expression profile of S. portulacastrum (C3) and S. sesuvioides (C4-like) with respect to CCM. According to photosynthetic sub-pathways, DETs were clustered and the number of DETs associated with each pathway was plotted (Fig. 4). We found a significant accumulation of genes involved in C4-related pathways (except PEP regeneration that was only found in the C4-like species) in both species (Fig. 4A, Additional file 2). The number of genes involved in carboxylation, proton pump, transfer acid generation, and transporter were higher in S. sesuvioides (C4-like; Fig. 4A, Additional file 2). Conversely, genes related to decarboxylation and photorespiration were abundant in the C3 species as compared to C4-like species.

Fig. 4
figure 4

Functionally annotated DETS between C3 and C4 species. A Stacked bar charts showing the number of functional annotated DETS involved in CCM between C3 and C4 species. B Abundance of selected DETS involved in CCM between C3 and C4 species. C Stacked bar charts showing the number of functional annotated DETS involved in starch metabolism between C3 and C4 species. D Abundance of selected DETS involved in starch metabolism between C3 and C4 species. The stacked bar charts display all transcripts that were differentially expressed

We then investigated genes related to carboxylation and decarboxylation. Surprisingly, while PEPC1 (TRINITY_DN0_c3_g1_i12) was up-regulated in the C4-like species, PEPC3 (TRINITY_DN9611_c0_g1_i3) and PEPC4 (TRINITY_DN14827_c0_g1_i3) were significantly expressed in the C3 species (Fig. 4B). PPDK and PPCK1 were up-regulated in the C4-like species (Additional file 3). The decarboxylation enzymes chloroplastic NADP-ME4, NADP-ME, and NAD-ME were significantly accumulated in the C3 species. However, another NADP-ME copy was up-regulated in the C4-like species. These findings suggest that S. sesuvioides as C4-like species employs NADP-ME as a decarboxylation enzyme but can additionally use NAD-ME (Additional file 3).

Gene ontology (GO) enrichment showed that response to stress was among the top 20 categories that were enriched in both species (Additional file 4). Moreover, we found that ALMT9 and ALMT4 were significantly up-regulated in S. sesuvioides and S. portulacastrum, respectively (Fig. 4B, Additional file 3). A putative photosynthetic cycle was designed for S. sesuvioides and S. portulacastrum (Fig. 5). Several copies of transcripts of genes that control the tonoplast potential were significantly up-regulated in S. sesuvioides (VHA-A, VHA-C, VHA-E1) and in S. portulacastrum (VHA-G1, VHA-B1) (Fig. 5, Additional file 3). Moreover, genes involved in starch phosphorolytic degradation were significantly abundant in S. portulacastrum (Fig. 4C). These genes included AMY2 and AMY3 (Fig. 4D).

Fig. 5
figure 5

Putative photosynthetic cycle in S. sesuvioides A and S. portulacastrum B. Red colour stands for genes that were up-regulated

Functional annotation of DETs between leaves and cotyledons of S. sesuvioides

We examined the process of photosynthesis in the leaves and cotyledons of S. sesuvioides. We identified DE transcripts (p < 0.05, log2FC > 1) related to CCM according to annotation (Fig. 6, Additional file 2). The transcripts related to CCM were subsequently clustered based on their role in decarboxylation, Calvin cycle, carboxylation, citrate generation, PEP regeneration and regulation, photorespiration, transfer acid generation, and transporter. We then counted the number of related transcripts of each cluster to identify the photosynthetic mode of cotyledons and leaves (Fig. 6A, Additional file 2). We found that transcripts involved in Calvin cycle, photorespiration, carboxylation, and decarboxylation were abundant in cotyledons as compared to leaves, whereas transporters involved in CCMs (C4 and CAM), transfer acid generation and PEP regeneration were higher in leaves than in cotyledons (especially those collected during the day). This suggests that different CCMs are acting in leaves and cotyledons of S. sesuvioides. To investigate further, we plotted the abundance of transcripts associated with the functional categories. CAM differs from C4 by a nocturnal CO2 fixation and accumulation of malate or citrate in the vacuole, and a diurnal decarboxylation of accumulated malate by the malic enzymes (e.g. NADP-ME). Our results showed that transcripts encoding carboxylation enzymes (PEPC4 and PPCK1) were upregulated in leaves as compared to cotyledon collected during the day (CD), while transcripts encoding decarboxylation enzymes NADP-ME and chloroplastic NADP-ME4 were significantly abundant in cotyledons collected during the night (CN) (Fig. 6B). Moreover, phylogenetic analysis showed that PEPC4 (TRINITY_DN14827_c0_g1_i3) grouped with a PPC-1E1 gene, along with several PPC-1E2 and PPC-2 genes (Additional file 5). These results indicate a higher decarboxylation rate in CD and CN. Thus, we suspected the possibility of CAM and C4-like photosynthesis in cotyledons and leaves of Sesuvium sesuvioides, respectively.

Fig. 6
figure 6

Functional annotated DETS between leaves (L) and cotelydons (C) of S. sesuvioides. A Stacked bar charts showing the number of functional annotated DETS involved in CCM between L and C of S. sesuvioides. B Abundance of selected DETS involved in CCM between L and C of S. sesuvioides. C Stacked bar charts showing the number of functional annotated DETS involved in starch metabolism L and C of S. sesuvioides. D Abundance of selected DETS involved in starch metabolism L and C of S. sesuvioides. The stacked bar charts display all transcripts that were differentially expressed

To determine whether CAM is occurring in the cotyledons, we compared the transcript profiles of CD and CN (Fig. 6A). In CAM plants, malate generated at night is supposedly transported to the vacuole by a malate transporter aluminium-activated malate transporter (ALMT) [42]. We observed a significant increase in transcripts of ALMT2 in CN (Fig. 6B). The vacuolar malate influx is driven by the difference in membrane potential established by vacuolar-type proton adenosine triphosphatase ATPase (VHA) and the pyrophosphate-energized membrane proton pump (AVP) [43]. However, as in Portulaca [10], no significant expression of these genes was observed in the CN. Although no significant expression of decarboxylation enzymes NADP-ME in CN when compared to CD, a mitochondrial NAD-ME was highly abundant in CN (Fig. 6B). However, we found that RUBISCO small unit (RBCS1) (Additional file 3) was up-regulated in CD. This suggested that the decarboxylation activity of NAD-ME at night is aimed primarily at malate respiration [44]. Additionally, ATP-citrate synthase alpha chain protein 3 (ACL-3) which is involved in citrate synthesis [45] was found up-regulated in CN. Cotyledons showed lower citrate levels at 11 am compared to leaves of S. sesuvioides (Siadjeu, unpublished data). Citrate is produced when acetyl CoA reacts with OAA suggesting that citrate contributes as well to the acidification in S. sesuvioides at night. Facultative CAM species such as Talinum triangulare exhibit increased citrate levels during the night [46].

Another feature of CAM photosynthesis is the nightly regeneration of PEP via phosphorolytic starch degradation [47, 48]. Moreover, Moreno-Villena et al. [10] hypothesized that CAM induction is associated with the increase in sugar transporter. The number of transcripts related to starch/sugar transporter and starch phosphorolytic degradation was high in CN (Fig. 6C, Additional file 2). We found that genes encoding the probable alpha-amylase 2 (AMY2), alpha-amylase (AMY) and alpha-1,4 glucan phosphorylase L-2 isozyme (PHO2) involved in starch phosphorolytic degradation were significantly abundant at night (Fig. 6D, Additional file 2). The sugar transporters, SWEET14, SWEET17, ERD6-like 5 were up-regulated at night (Fig. 6D). It is worth mentioning that we did not observe a significant expression of PEPC at night. However, pyruvate phosphate dikinase (PPDK) was significantly abundant in the cotyledons at night (Fig. 3D). The enzyme PPDK catalyzes the regeneration of the CO2 acceptor PEP via pyruvate [49].

Regulation and hormonal signaling in Sesuvium

To identify potential regulation and signaling elements, we performed an unsupervised k-means clustering. The three methods used showed the best k was two (Additional file 6). We found that almost all transcripts involved in photosynthesis, starch metabolism, transcription factor, and phytohormone signaling were grouped together (Cluster with the highest number of transcripts) for all comparisons (Additional files 7–8). To identify possible candidates involved in the regulation of the CCMs studies, we clustered TFs found in groups including photosynthesis, starch metabolism, and phytohormone signaling according to their families (Fig. 7, Additional file 2).

Fig. 7
figure 7

Functional annotated expressed genes related to TFs. A Stacked bar charts showing the number of functional annotated DET related to TFs between C3 and C4 species. B Abundance of selected DET related to TFs between C3 and C4 species. C Stacked bar charts showing the number of functional annotated DET related to TFs between (L) and cotelydons (C) of S. sesuvioides. D Abundance of selected DET related to TFs between L and C of S. sesuvioides. BBX: B-BOX, bZIP: basic LEUCINE ZIPPER, CEPR: C-TERMINALLY ENCODED PEPTIDE RECEPTOR, COL: CONSTANS-LIKE, HB: HOMEOBOX, NAC (NAM, ATAF and CUC), NFYA: NUCLEAR TRANSCRIPTION FACTOR Y SUBUNIT ALPHA, SCL: SCARECROW-LIKE. The stacked bar charts display all transcripts that were differentially expressed

For comparison between S. sesuvioides and S. portulacastrum, TF families C2C2, C2H2, GRAS, HOMEOBOX, MYB-related, NAC, TEOSINTE BRANCHED1/CYCLOIDEA/PROLIFERATING CELL FACTOR (TCP), WRKY, basic HELIX-LOOP-HELIX (bHLH), and basic LEUCINE ZIPPER (bZIP) were found on top of the list (Fig. 7A, Additional file 2). Except for TFs from the GRAS family present only in C4-like, all others were found in C4-like and C3 species (Additional file 3). We selected based on the literature candidate TFs that are potentially involved in the regulation of C4 and CAM (Fig. 7B). Different isoforms are recruited in the regulation of C3 and C4-like species. Our data showed that REVEILLE 8, bHLH143, SCL6, COL5, and DOF1.2 were significantly up-regulated in C4-like species, whereas REVEILLE 6, COL15, NFYA7, and HB6 were up-regulated in the C3 species (Fig. 7B, Additional file 3). We identified 63 nonredundant motifs (p < 0.01, E-value < 0.05) in sequences of genes upregulated in the C4-like species while 36 motifs were found in the C3 species (Additional file 9). The top five of the most enriched motifs were annotated to C2C2-DOF families in C4-like and the C3 species with the element CTTTTT (Table 2). Although the most enriched motifs were similar between the C4-like species and the C3 species, the second most enriched motifs were elements (GAGA, BBR/BPC family) and (CACCAACM, MYB family) in the C4-like and C3 species, respectively. We found several motifs often present in the same transcript sequences. This suggests a coordinated and regulatory network of TFs controlling photosynthesis in which motif CTTTTT is dominant.

Table 2 Top five motif enrichments of differentially expressed genes related to CCM (Carbon Concentrating Mechanism) and starch metabolism in Sesuvium species

Our results indicated that the same TF families are recruited to a variable degree for controlling cotyledons and leaves of S. sesuvioides. The top 10 TF families abundant in cotyledons and leaves during the day and in CN were bHLH), C2C2, APETALA2/ETHYLENE-RESPONSIVE FACTOR (AP2/ERF), WRKY, MADS/AGL, bZIP, NAC, MYOLOBLASTOSIS (MYB)-related, GARP (GOLDEN2, ARR-B, PSR1), and HOMEOBOX (Fig. 7C, Additional file 2). While during the day, the number of TFs related to C2C2, bHLH, and bZIP families were higher in leaves than in cotyledons, no TFs related to HOMEOBOX and AP2/ERF were enriched in leaves. When comparing CD and CN, we found nearly all TF families were abundant at night. We then specifically looked at TFs that are frequently expressed in these families. Genes were selected by their potential involvement in the regulation of CAM and C4-like photosynthesis. Our data showed the TFs bHLH87, REVEILLE1, WRKY40, NAC83, and SCL15 were up-regulated in the CN, whereas TFs BBX24, and SCL14 were significantly abundant in the leaves (Fig. 7D). To confirm whether these TFs regulate C4-like and CAM photosynthesis, we explored TF binding sites in sequences of transcripts related to C4-like photosynthesis and starch metabolism (Table 2). Motif enrichment analysis revealed elements AAAAAG and CTTTTT from the C2C2-DOF family were the most enriched in leaves during the day, while element CTTTTT was the most enriched in cotyledons (Table 2). When comparing CD and CN, element CGCCGCC from the AP2/ERF family was enriched in CD whereas element (CTTTTT) was enriched in all comparisons.

Phytohormones play critical roles in photosynthesis regulation and developmental processes ranging from organ initiation to senescence [50]. Moreover, phytohormones have been shown to mediate TF action in C4 and CAM plants [51]. When comparing S. sesuvioides to S. portulacastrum, we found that signaling peptides, auxin, cytokinin, abscisic acid, and brassinosteroid were the top five endogenous signaling hormones (Fig. 8A, Additional file 2). There were many genes related to hormonal signals that were specific to either C3 species or C4-like species (Additional file 3). For instance, genes PHP2 (cytokinin), RALF4 (signaling peptide), and IAA9 (auxin) were significantly accumulated in the C3 species while EIR3 (auxin), ERF118 (cytokinin) and CEPR2 (signaling peptide) were significantly accumulated in the C4-like species (Fig. 8B). When comparing leaves to cotyledons, signaling peptides, auxin, cytokinin, jasmonic, and abscisic acids were the most abundant signaling hormones (Fig. 8C, Additional file 2). We found that the expression of CYP707A4 (abscisic acid), LHW (cytokinin), and JAR6 (jasmonic acid) increased significantly in abundance in cotyledons at night (Fig. 8D). In CD, MS17 (Auxin) and GAST1 (signaling peptide) were up-regulated, while IAA14 (auxin) and GASA1 (signaling peptide) were significantly accumulated in the leaves. Many other signaling protein genes were found specific to leaves and cotyledons and are listed in Additional file 3.

Fig. 8
figure 8

Functional annotated differentially expressed genes related to phytohormones. A Stacked bar charts showing the number of functional annotated DET related to phytohormones between C3 and C4 species. B Abundance of selected DET related to phytohormones between C3 and C4 species. C Stacked bar charts showing the number of functional annotated DET related to phytohormones between (L) and cotelydons (C) of S. sesuvioides. D Abundance of selected DET related to phytohormones between (L) and cotelydons (C) of S. sesuvioides. CEPR: C-TERMINALLY ENCODED PEPTIDE RECEPTOR, CYP707A4: CYTOCHROME P450, EIR: ETHYLENE INSENSITIVE ROOT, ERF: ETHYLENE-RESPONSIVE ELEMENT BINDING

Discussion

Photosynthetic mode in Sesuvium species

Our findings confirmed that adult leaves of S. sesuvioides perform C4-like photosynthesis with core C4 enzymes up-regulated. These up-regulated enzymes were involved in carboxylation (βCA, PEPC1), acid regeneration (ALAAT2, ASPAT, NADP-MDH), decarboyxlation (NADP-ME, NAD-ME), transporters (BASS2, DIT1, DIT2, PPT2) and PEP regeneration (PPDK). PEPC1 (TRINITY_DN0_c3_g1_i12) was clustered in the group containing only PPC-1E1 (Additional file 5) which was repeatedly used for both C4 and CAM photosynthesis [20]. This result is in accordance with anatomical, biochemical, and physiological observations [17]. The C4-like status of S. sesuvioides becomes evident in the still relatively high expression of photorespiratory genes (Fig. 3C, Additional file 2).

Intriguingly, when comparing leaves of the adult plants with cotyledons of S. sesuvioides, the main carboxylation enzyme was PEPC isoform 4 (PEPC4) in leaves. There are several plant PEPC copies that are classified into photosynthetic (C4 and CAM) and non-photosynthetic isoforms [52]. In eudicots, there are three distinct lineages that encode PEPC, which are called PPC-1E1, PPC-1E2, and PPC-2 [20]. According to the Uniprot database, these genes are involved in CO2 fixation and the tricarboxylic acid cycle. Phylogenetic analysis of PEPC4 with these three genes from 35 species (Additional file 5) showed that PEPC4 (TRINITY_DN14827_c0_g1_i3) clustered together with a PPC-1E1, as well as several PPC-1E2 and PPC-2 genes. However, the presence of PPC-1E1 in this group suggests that PPC4 may be involved in C4 photosynthesis. Indeed, PEPC4 isoform identified and up-regulated in the leaves is homologous to Arabidopsis AtPEPC4 which is involved in photosynthesis (https://www.uniprot.org/uniprotkb/Q8GVE8/entry) and may indicate a similar role in S. sesuvioides. In line with this result, PEPC3 and PEPC4 were significantly expressed in adult C3 leaves of S. portulacastrum when compared to the leaves of C4-like S. sesuvioides. PEPC3 (TRINITY_DN9611_c0_g1_i3) was also found clustered in a similar group with PEPC4. This could also be explained by the fact that the C3 species S. portulacastrum induces weak CAM photosynthesis under drought stress [15]. Heat and drought stresses repress nitrogen metabolism enzymes [53]. Thus, regulation of nitrogen metabolism is crucial to maintain plant growth under stress conditions. PEPC plays a crucial role in C4 photosynthesis and in modulating the balance of carbon and nitrogen metabolism in Arabidopsis [54]. This indicates that the up-regulation of these PEPC copies may be important to maintain the photosynthetic system and plant growth of Sesuvium species under adverse conditions. The biological role and localization of these PEPC copies in Sesuvium need to be further investigated, however, our results indicate that different PEPC copies are optimized for different photosynthetic functions in leaves and cotyledons of Sesuvium.

Generally, there are three subtypes of C4 photosynthesis depending on the decarboxylation enzymes. The main decarboxylation enzyme in S. sesuvioides was NADP-ME. The up-regulation of NADP-ME indicates that S. sesuvioides is employing NADP-ME in its decarboxylation mechanism. This result is consistent with biochemical and physiological observations in S. sesuvioides [17]. Moreover, in the C3 species, multiple copies of the chloroplastic decarboxylation enzymes NADP-ME and NAD-ME were significantly accumulated. These findings may indicate a possibility of CAM induction in S. portulacastrum.

It was shown indeed that S. portulacastrum is capable of inducing CAM in stressful conditions [15,55]. ALMT9 and ALMT4 were significantly up-regulated in S. sesuvioides and S. portulacastrum, respectively. ALMT was originally responsible for nocturnal malate accumulation caused by an inward-rectifying anion-selective channel that forces only malate influx to the vacuole [56]. However, Meyer et al. [57] demonstrated that AtALMT6 functions as a malate influx or efflux channel depending on the tonoplast potential. This indicates diurnal vacuolar malate efflux in both species, hence the possibility of weak CAM being induced in S. sesuvioides and S. portulacastrum.

Photosynthetic plasticity in Sesuvium

Plants exhibit plasticity for a wide variety of ecologically important traits to adjust to environmental changes [58]. Photosynthetic plasticity underpins the ability of plants to acclimate and grow in adverse environments and may depend on plant ontogeny. Our data provide evidence of photosynthetic plasticity in S. sesuvioides with C4-like and CAM photosynthesis in leaves and cotyledons, respectively. During the day, decarboxylating enzymes were more strongly expressed in cotyledons compared to leaves while carboxylating enzymes were strongly expressed in leaves. This result was further confirmed by titratable acidity, which showed a significant accumulation of acids overnight in cotyledons. The co-occurrence of C4 and CAM photosynthesis has already been reported in other Aizoaceae, namely for Trianthema portulacastrum [9], but this is the first time to report ontogenetic variability with respect to photosynthesis in the Aizoaceae family with CAM in cotyledons and C4-like in leaves. In Amaranthaceae (incl. Chenopodiaceae), Lauterbach et al. [59] based on RNA expression profiles showed the transition from C3 photosynthesis in cotyledons to C4 photosynthesis in adult leaves of Salsola soda. This phenomenon seems to occur in several species of Salsoleae according to C3-like features such as lower carbon isotope ratios and lack of Kranz anatomy in cotyledons (e.g., [60, 61]). However, these species have never been tested for CAM metabolism. The presence of CAM in cotyledons may be induced by environmental cues. Indeed, no CAM was observed in the cotyledons and leaves of S. sesuvioides under well-watered conditions [17]. In the climate chamber, stressful conditions were mainly created by the maximum light intensity. CAM induction has been linked to a photoprotective role in Portulaca oleracea [51]. This suggests a photoprotective role of CAM induction in cotyledons of S. sesuvioides. It is worth mentioning that a significant expression of PEPC (Phosphoenolpyruvate carboxylase) was not observed in CN. This is likely due to the sampling time (one hour after the light was turned off). However, a significant abundance of PPDK was observed in CN. This enzyme catalyzes the regeneration of PEP via pyruvate, which serves as a CO2 acceptor.

Integration of C4-like and CAM photosynthesis

Our data suggested a possible co-occurrence of C4-like and CAM photosynthesis in a single leaf of S. sesuvioides under adverse conditions (Fig. 5, Fig. 6A). In CAM photosynthesis, nocturnally accumulated malate is translocated out of the vacuole by a malate channel for subsequent decarboxylation during the light period. In P. oleracea, a C4 species that performs CAM when drought-stressed [10], AtALMT9 that has been associated with CAM function [62] is a vacuolar malate channel [63]. Interestingly, we found ALMT9 was significantly abundant during the light period in S. sesuvioides. Thus, the up-regulation of ALMT9 in leaves of S. sesuvioides, suggests that ALMT9 may function as a vacuolar malate efflux channel in S. sesuvioides and is therefore linked to CAM function.

Taking all results together, this may imply the integration of the hybrid system C4-like + CAM in S. sesuvioides under stress conditions. However, the modularity of this integration needs to be investigated. This co-occurrence of C4-like and CAM in a single leaf in S. sesuvioides is probably facilitated by the particular C4-like phenotype of S. sesuvioides leaves where Rubisco is present in the MCs. The MCs of S. sesuvioides are succulent and outnumber the Kranz cells by two-fold. When the leaves grow older, the M portion becomes even larger and the carbon isotope ratios drop [17]. This might indicate a photosynthetic plasticity towards a higher proportion of CAM and or C3 relative to C4-like in older leaves depending on the growing conditions.

Regulation of photosynthesis and hormonal signaling in Sesuvium

Plants have the ability to choose different photosynthetic pathways, which is controlled by TFs. Six TF families i.e., C2C2, HOMEOBOX, NAC, WRKY, bHLH, and bZIP were found up-regulated in leaves and cotyledons of S. sesuvioides when compared leaves to cotyledons and also when compared to adult leaves of S. sesuvioides and to adult leaves of S. portulacastrum. These families have been hypothesized to be involved in the regulation of C4 and CAM photosynthesis in Chenopodiaceae, Aizoaceae, and Asteraceae [51,64]. However, TFs from the C2C2, and bHLH families were the most expressed in leaves, and cotyledons during day and night. Likewise, these TFs were predominant in the C3 and C4-like species. This may indicate the significant weight of the C2C2 and bHLH TF families in the regulation of cotyledons and leaf development in S. sesuvioides and S. portulacastrum. The binding site of TF C4ZINC FINGER-TYPE (DOF3.6) from the DOF/C2H2 was most enriched in C4 and CAM genes in all comparisons. In addition, at least three copies of DOFs were among the top five. Several copies of DOF proteins (DOF1 and DOF2) were found involved in the regulation of the light-dependent C4 gene PEPC in maize with antagonist effects. While DOF1 activates C4 genes, DOF2 can activate or repress them [65]. These results indicate that different copies of DOF genes are likely involved in the regulation of C4-like and CAM genes in S. sesuvioides and S. portulacastrum, as well. It is worth mentioning that transcription factors (TFs) from the MYB, NUCLEAR FACTOR Y (NF-Y) and NAC families have been suggested to play a role in regulating CAM in facultative CAM species such as Mesembryanthemum crystallinum [66] and T. triangulare [67]. While the isoforms of these genes may differ, these TF families were found to be up-regulated when comparing CAM-induced (CD) and non-induced (CN) conditions, as well as when comparing day-time leaves and cotyledons.

Transcription factors are regulated by phytohormones (signaling molecules) under environmental stresses [68]. Our data showed that phytohormones were clustered with C4 and CAM genes, as well as TFs which indicates a regulatory network involving TFs and phytohormones. Indeed, Ferrari et al. [51] found that ABA and CK-related genes regulate TFs connected to CAM and C4 photosynthesis in Portulaca oleracea. It has been suggested that ABA plays a role in responding to drought stress in facultative CAM species such as M. crystallinum [69] and T. triangulare [46,67]. In S. sesuvioides cotyledons, transcripts of genes that encode for ABA were found to be more enriched in CD as compared to CN. As the temperature at night was cooler than during the day, this could indicate that ABA might also play a role in CAM induction. While ABA and CKs have been studied intensely in CAM and C4 photosynthesis, several other phytohormones that regulate photosynthesis (reviewed by [50]) have received little attention. Here, we found that transcripts encoding signaling peptides were the most abundant plant hormones during the day and at night in cotyledons as opposed to leaves. Similarly, diurnal and nocturnal comparison expression in cotyledons showed that signaling peptides were predominantly accumulated. Transcripts of genes encoding for the signaling peptides Gibberellic acid (GA)-STIMULATED ARABIDOPSIS/GA-STIMULATED TRANSCRIPT (GAST) and the RAPID ALKALINIZATION FACTOR (RAFL) were predominantly expressed in cotyledons as compared to leaves. These plant hormones play important roles in plant growth, development, and stress responses ( [70, 71]), and may control cotyledon growth and response to environmental conditions in S. sesuvioides. Conversely, in leaves as opposed to CD, the most dominant hormone was jasmonic acid followed by auxin and cytokinin CKs. Jasmonic acid, auxin, and cytokinin are classical phytohormones that regulate various aspects of plant growth and abiotic and biotic stress responses. While jasmonic acid can regulate stomatal closure and opening under drought stress in Arabidopsis [72], auxin coordinates cell division, expansion, and differentiation [73], and CKs are implicated in cell cycle progression [74]. Several genes encoding for these phytohormones were found to be significantly accumulated (Additional file 3) and should be used as candidate genes involved in the regulation of CCMs in these species.

Conclusions

This study of gene expression profiles of S. sesuvioides provides evidence of extraordinary photosynthetic plasticity under adverse conditions with induced CAM in cotyledons and an integration of CAM and C4-like photosynthesis in adult leaves. However, the modularity of the co-occurrence of the two CCMs needs to be explored in future studies. We assume that further detection of co-occurring CCMs is just a matter of more experimental studies that explicitly look for this and we believe that it is more common in succulent C4 lineages than currently known. Our findings suggest a complex regulatory network involving TFs and phytohormones and underpin the regulation of CCMs and adaptation of Sesuvium species which grow in disturbed and highly dynamic environments.

Availability of data and materials

The datasets used and/or analysed during the current study are available in the NCBI Sequence Read Archive (SRA) under the BioProject ID: PRJNA1067387, http://www.ncbi.nlm.nih.gov/bioproject/1067387

References

  1. Berasategui J, Žerdoner Čalasan A, Zizka A, Kadereit G. Global distribution and climatic preferences of C4 eudicots and how they differ from those of C4 grasses. Ecol Evol. 2023;13 (11):e10720. https://doi.org/10.22541/au.168301289.96241129/v1.

  2. Kadereit G, Ackerly D, Pirie MD. A broader model for C4 photosynthesis evolution in plants inferred from the goosefoot family (chenopodiaceae s.s.). Proc R Soc B Biol Sci. 2012;279:3304–11.

    Article  Google Scholar 

  3. Bohley K, Joos O, Hartmann H, Sage R, Liede-Schumann S, Kadereit G. Phylogeny of sesuvioideae (Aizoaceae) - biogeography, leaf anatomy and the evolution of C4 photosynthesis. Perspect Plant Ecol Evol Syst. 2015;17:116–30.

    Article  Google Scholar 

  4. Ocampo G, Columbus JT. Molecular phylogenetics, historical biogeography, and chromosome number evolution of Portulaca (Portulacaceae). Mol Phylogenet Evol. 2012;63:97–112.

    Article  PubMed  Google Scholar 

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

    Article  Google Scholar 

  6. Koch K, Kennedy RA. Characteristics of crassulacean acid metabolism in the succulent C4 dicot. Portulaca oleracea L Plant Physiol. 1980;65:193–7.

    Article  CAS  PubMed  Google Scholar 

  7. Ho CL, Chiang JM, Lin TC, Martin CE. First report of C4/CAM-cycling photosynthetic pathway in a succulent grass, Spinifex littoreus (Brum. f.) Merr., in coastal regions of Taiwan. Flora. 2019;254:194–202.

    Article  Google Scholar 

  8. Han S, Xing Z, Li W, Huang W. Response of anatomy and CO2-concentrating mechanisms to variable CO2 in linear juvenile leaves of heterophyllous ottelia alismoides: comparisons with other leaf types. Environ Exp Bot. 2020;179 March:104194.

    Article  Google Scholar 

  9. Winter K, Garcia M, Virgo A, Ceballos J, Holtum JAM. Does the C4 plant Trianthema portulacastrum (Aizoaceae) exhibit weakly expressed crassulacean acid metabolism (CAM)? Funct Plant Biol. 2021;48:655–65.

    Article  CAS  PubMed  Google Scholar 

  10. Moreno-Villena JJ, Zhou H, Gilman IS, Tausta SL, Cheung CM, Edwards EJ. Spatial resolution of an integrated C4+CAM photosynthetic metabolism 1. Sci Adv. 2021;8:14.

    Google Scholar 

  11. Sage RF. The evolution of C4 photosynthesis. New Phytol. 2004;161:341–70.

    Article  CAS  PubMed  Google Scholar 

  12. Muhaidat R, Sage RF, Dengler NG. Diversity of Kranz anatomy and biochemistry in C4 eudicots. Am J Bot. 2007;94:362–81.

    Article  CAS  PubMed  Google Scholar 

  13. Schüssler C, Freitag H, Koteyeva N, Schmidt D, Edwards G, Voznesenskaya E, et al. Molecular phylogeny and forms of photosynthesis in tribe Salsoleae (Chenopodiaceae). J Exp Bot. 2017;68:207–23.

    Article  PubMed  Google Scholar 

  14. Gilman IS, Smith JAC, Holtum JAM, Sage RF, Silvera K, Winter K, Edwards EJ. The CAM lineages of planet Earth. Ann Bot. 2023;132(4):627–54. https://doi.org/10.1093/aob/mcad135.

    Article  CAS  PubMed  Google Scholar 

  15. Winter K, Garcia M, Virgo A, Holtum JAM. Operating at the very low end of the crassulacean acid metabolism spectrum: Sesuvium portulacastrum (Aizoaceae). J Exp Bot. 2019;70:6561–70.

    Article  CAS  PubMed  Google Scholar 

  16. Hartmann HE, Gerbaulet M. Aizoaceae. Berlin: Springer; 2017.

  17. Bohley K, Schröder T, Kesselmeier J, Ludwig M, Kadereit G. C4-like photosynthesis and the effects of leaf senescence on C 4-like physiology in Sesuvium sesuvioides (Aizoaceae). J Exp Bot. 2019;70:1567–80.

    Article  Google Scholar 

  18. Sage RF, Zhu XG. Exploiting the engine of C4 photosynthesis. J Exp Bot. 2011;62:2989–3000.

    Article  CAS  PubMed  Google Scholar 

  19. Cui H. Challenges and approaches to crop improvement through C3-to-C4 engineering. Front Plant Sci. 2021;12:715391.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Christin PA, Arakaki M, Osborne CP, Bräutigam A, Sage RF, Hibberd JM, et al. Shared origins of a key enzyme during the evolution of C4 and CAM metabolism. J Exp Bot. 2014;65:3609–21.

    Article  PubMed  PubMed Central  Google Scholar 

  21. Silvera K, Santiago LS, Winter K. Distribution of crassulacean acid metabolism in orchids of Panama: evidence of selection for weak and strong modes. Funct Plant Biol. 2005;32:397–407.

    Article  CAS  PubMed  Google Scholar 

  22. Siadjeu C, Mayland-quellhorst E, Laubinger S, Albach DC. Transcriptome sequence reveals candidate genes involving in the post-harvest hardening of trifoliate yam Dioscorea dumetorum. Plants. 2021;1–21.

  23. Ewels P, Magnusson M, Lundin S, Käller M. MultiQC: Summarize analysis results for multiple tools and samples in a single report. Bioinformatics. 2016;32:3047–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Song L, Florea L. Rcorrector: efficient and accurate error correction for Illumina RNA-seq reads. Gigascience. 2015;4:1–8.

    Article  Google Scholar 

  25. Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nat Methods. 2012;9:357–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, Adiconis X, Fan L, Raychowdhury R, Zeng Q, Chen Z, Mauceli E, Hacohen N, Gnirke A, Rhind N, Palma F di, Birren BW, Nusbaum C, Lindblad-Toh K, Friedman N, Regev A. Trinity: reconstructing a full-length transcriptome without a genome from RNA-Seq data. Nat Biotechnol. 2013;29:644–52.

    Article  Google Scholar 

  27. Manni M, Berkeley MR, Seppey M, Simão FA, Zdobnov EM. BUSCO update: novel and streamlined workflows along with broader and deeper phylogenetic coverage for scoring of eukaryotic, prokaryotic, and viral genomes. Mol Biol Evol. 2021;38:4647–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Fu L, Niu B, Zhu Z, Wu S, Li W. CD-HIT: accelerated for clustering the next-generation sequencing data. Bioinformatics. 2012;28:3150–2.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Soneson C, Love MI, Robinson MD. Differential analyses for RNA-seq: transcript-level estimates improve gene-level inferences. F1000Research. 2016;4:1–19.

    Article  Google Scholar 

  30. Emms DM, Kelly S. OrthoFinder: phylogenetic orthology inference for comparative genomics. Genome Biol. 2019;20:1–14.

    Article  Google Scholar 

  31. Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15:1–21.

    Article  Google Scholar 

  32. Schwacke R, Ponce-Soto GY, Krause K, Bolger AM, Arsova B, Hallab A, et al. MapMan4: a refined protein classification and annotation framework applicable to multi-omics data analysis. Mol Plant. 2019;12:879–92.

    Article  CAS  PubMed  Google Scholar 

  33. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol Biol Evol. 2013;30:772–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Rousseeuw PJ. Silhouettes: a graphical aid to the interpretation and validation of cluster analysis. J Comput Appl Math. 1987;20 C:53–65.

    Article  Google Scholar 

  36. Robert T, Walther Guenther HT. Estimating the number of clusters in a data set via the gap statistic. J R Stat Soc Ser B (Statistical Methodol. 2001;63:411–23.

    Article  Google Scholar 

  37. Bailey TL, Johnson J, Grant CE, Noble WS. The MEME Suite. Nucleic Acids Res. 2015;43:W39–49.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Gupta S, Stamatoyannopoulos JA, Bailey TL, Noble WS. Quantifying similarity between motifs. Genome Biol. 2007;8:R24.

    Article  PubMed  PubMed Central  Google Scholar 

  39. McLeay RC, Bailey TL. Motif Enrichment analysis: a unified framework and an evaluation on ChIP data. BMC Bioinformatics. 2010;11:165.

    Article  PubMed  PubMed Central  Google Scholar 

  40. Winter K. Ecophysiology of constitutive and facultative CAM photosynthesis. J Exp Bot. 2019;70:6495–508.

    Article  CAS  PubMed  Google Scholar 

  41. Gilman IS, Moreno-Villena JJ, Lewis ZR, Goolsby EW, Edwards EJ. Gene co-expression reveals the modularity and integration of C4 and CAM in Portulaca. Plant Physiol. 2022;189:735–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Borland AM, Griffiths H, Hartwell J, Smith JAC. Exploiting the potential of plants with crassulacean acid metabolism for bioenergy production on marginal lands. J Exp Bot. 2009;60:2879–96.

    Article  CAS  PubMed  Google Scholar 

  43. Smith JAC, Ingram J, Tsiantis MS, Barkla BJ, Bartholomew DM, Bettey M, et al. Transport across the vacuolar membrane in CAM plants. Crassulacean Acid Metab. 1996;114:53–71.

    Article  CAS  Google Scholar 

  44. Tronconi MA, Fahnenstich H, Gerrard Weehler MC, Andreo CS, Flügge UI, Drincovich MF, et al. Arabidopsis NAD-malic enzyme functions as a homodimer and heterodimer and has a major impact on nocturnal metabolism. Plant Physiol. 2008;146:1540–52.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Winter K, Smith JAC. CAM photosynthesis: the acid test. New Phytol. 2022;233:599–609.

    Article  CAS  PubMed  Google Scholar 

  46. Brilhaus D, Bräutigam A, Mettler-Altmann T, Winter K, Weber APM. Reversible burst of transcriptional changes during induction of crassulacean acid metabolism in talinum triangulare1[OPEN]. Plant Physiol. 2016;170:102–22.

    Article  CAS  PubMed  Google Scholar 

  47. Weise SE, Van Wijk KJ, Sharkey TD. The role of transitory starch in C 3, CAM, and C 4 metabolism and opportunities for engineering leaf starch accumulation. J Exp Bot. 2011;62:3109–18.

    Article  CAS  PubMed  Google Scholar 

  48. Heyduk K, Ray JN, Ayyampalayam S, Moledina N, Borland A, Harding SA, et al. Shared expression of crassulacean acid metabolism (CAM) genes pre-dates the origin of CAM in the genus Yucca. J Exp Bot. 2019;70:6597–609.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Chastain CJ, Failing CJ, Manandhar L, Zimmerman MA, Lakner MM, Nguyen THT. Functional evolution of C4 pyruvate, orthophosphate dikinase. J Exp Bot. 2011;62:3083–91.

    Article  CAS  PubMed  Google Scholar 

  50. Müller M, Munné-Bosch S. Hormonal impact on photosynthesis and photoprotection in plants. Plant Physiol. 2021;185:1500–22.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Ferrari RC, Kawabata AB, Ferreira SS, Hartwell J, Freschi L. A matter of time: regulatory events behind the synchronization of C4 and crassulacean acid metabolism in Portulaca oleracea. J Exp Bot. 2022;73:4867–85.

    Article  CAS  PubMed  Google Scholar 

  52. O’Leary B, Park J, Plaxton WC. The remarkable diversity of plant PEPC (phosphoenolpyruvate carboxylase): recent insights into the physiological functions and post-translational controls of non-photosynthetic PEPCs. Biochem J. 2011;436:15–34.

    Article  PubMed  Google Scholar 

  53. Xia H, Xu T, Zhang J, Shen K, Li Z, Liu J. Drought-induced responses of nitrogen metabolism in ipomoea batatas. Plants. 2020;9:1–16.

    Article  Google Scholar 

  54. Shi J, Yi K, Liu Y, Xie L, Zhou Z, Chen Y, et al. Phosphoenolpyruvate carboxylase in arabidopsis leaves plays a crucial role in carbon and nitrogen metabolism. Plant Physiol. 2015;167:671–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Ting IP. Photosynthesis of Arid and Subtropical Succulent Plants. Aliso: A Journal of Systematic and Floristic Botany. 1989;12.

  56. Hafke JB, Hafke Y, Smith JAC, Lüttge U, Thiel G. Vacuolar malate uptake is mediated by an anion-selective inward rectifier. Plant J. 2003;35:116–28.

    Article  CAS  PubMed  Google Scholar 

  57. Meyer S, Scholz-Starke J, De Angeli A, Kovermann P, Burla B, Gambale F, et al. Malate transport by the vacuolar AtALMT6 channel in guard cells is subject to multiple regulation. Plant J. 2011;67:247–57.

    Article  CAS  PubMed  Google Scholar 

  58. Sultan SE. Phenotypic plasticity for plant development, function and life history. Trends Plant Sci. 2000;5:537–42.

    Article  CAS  PubMed  Google Scholar 

  59. Lauterbach M, Billakurthi K, Kadereit G, Ludwig M, Westhoff P, Gowik U. C3 cotyledons are followed by C4 leaves: Intra-individual transcriptome analysis of Salsola soda (Chenopodiaceae). J Exp Bot. 2017;68:161–76.

    Article  CAS  PubMed  Google Scholar 

  60. Pyankov VI, Black CC, Artyusheva EG, Voznesenskaya EV, Ku MSB, Edwards GE. Features of photosynthesis in haloxylon species of chenopodiaceae that are dominant plants in central Asian deserts. Plant Cell Physiol. 1999;40:125–34.

    Article  CAS  Google Scholar 

  61. Pyankov VI, Voznesenskaya EV, Kuz’Min AN, Ku MSB, Ganko E, Franceschi VR, et al. Occurrence of C3 and C4 photosynthesis in cotyledons and leaves of Salsola species (Chenopodiaceae). Photosynth Res. 2000;63:69–84.

    Article  CAS  PubMed  Google Scholar 

  62. Ferrari RC, Bittencourt PP, Rodrigues MA, Moreno-Villena JJ, Alves FRR, Gastaldi VD, et al. C4 and crassulacean acid metabolism within a single leaf: deciphering key components behind a rare photosynthetic adaptation. New Phytol. 2020;225:1699–714.

    Article  CAS  PubMed  Google Scholar 

  63. Kovermann P, Meyer S, Hörtensteiner S, Picco C, Scholz-Starke J, Ravera S, et al. The arabidopsis vacuolar malate channel is a member of the ALMT family. Plant J. 2007;52:1169–80.

    Article  CAS  PubMed  Google Scholar 

  64. Siadjeu C, Lauterbach M, Kadereit G. Insights into regulation of C2 and C4 photosynthesis in amaranthaceae / chenopodiaceae using RNA-Seq. Int J Mol Sci. 2021;22:1–18.

    Article  Google Scholar 

  65. Yanagisawa S, Sheen J. Involvement of maize Dof zinc finger proteins in tissue-specific and light-regulated gene expression. Plant Cell. 1998;10:75–89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Schaeffer HJ, Forsthoefel NR, Cushman JC. Identification of enhancer and silencer regions involved in salt-responsive expression of Crassulacean acid metabolism (CAM) genes in the facultative halophyte Mesembryanthemum crystallinum. Plant Mol Biol. 1995;28:205–18.

    Article  CAS  PubMed  Google Scholar 

  67. Maleckova E, Brilhaus D, Wrobel TJ, Weber APM. Transcript and metabolite changes during the early phase of abscisic acid-mediated induction of crassulacean acid metabolism in Talinum triangulare. J Exp Bot. 2019;70:6581–96.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Saibo NJM, Lourenço T, Oliveira MM. Transcription factors and regulation of photosynthetic and related metabolism under environmental stresses. Ann Bot. 2009;103:609–23.

    Article  CAS  PubMed  Google Scholar 

  69. Taybi T, Cushman JC. Abscisic acid signaling and protein synthesis requirements for phosphoenolpyruvate carboxylase transcript induction in the common ice plant. J Plant Physiol. 2002;159:1235–43.

    Article  Google Scholar 

  70. Wu VW, Thieme N, Huberman LB, Dietschmann A, Kowbel DJ, Lee J, et al. The regulatory and transcriptional landscape associated with carbon utilization in a filamentous fungus. Proc Natl Acad Sci U S A. 2020;117:6003–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Lee SH, Yoon JS, Jung WJ, Kim DY, Seo YW. Genome-wide identification and characterization of the lettuce GASA family in response to abiotic stresses. BMC Plant Biol. 2023;23:1–14.

    Article  Google Scholar 

  72. Savchenko T, Kolla VA, Wang CQ, Nasafi Z, Hicks DR, Phadungchob B, et al. Functional convergence of oxylipin and abscisic acid pathways controls stomatal closure in response to drought. Plant Physiol. 2014;164:1151–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Perrot-Rechenmann C. Cellular responses to auxin: division versus expansion. Cold Spring Harb Perspect Biol. 2010;2:1–15.

    Article  Google Scholar 

  74. Park J, Lee S, Park G, Cho H, Choi D, Umeda M, et al. Cytokinin-responsive growth regulator regulates cell expansion and cytokinin-mediated cell cycle progression. Plant Physiol. 2021;186:1734–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We would like to thank Sebastian Walter for germinating the seeds of S. sesuvioides and growing the plant material. Additionally, we extend our thanks to Thibaud Messerschmid for reviewing and providing feedback on the initial manuscript. We thank the reviewers for their very insightful comments.

Funding

Open Access funding enabled and organized by Projekt DEAL. This work was supported by the Deutsche Forschungsgemeinschaft (DFG) with grants to GK (KA1816/7–3).

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: G.K. Data curation: C.S. Formal Analysis: C.S. Funding acquisition: G.K. Investigation: C.S. Methodology: C.S. Project administration: G.K. Resources: G.K. Software: C.S. Supervision: G.K. Validation: C.S., G.K. Visualization: C.S. Writing-original draft: C.S. Writing – review & editing: G.K., C.S.

Corresponding author

Correspondence to Christian Siadjeu.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

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

Supplementary Information

12864_2024_10553_MOESM1_ESM.pdf

Additional file 1: Figure S1. Images of potted S. sesuvioides plants growing in the climate chamber and in the uncontrolled greenhouse environment.

Additional file 2: Dataset S1. List of transcripts/genes belonging to each functional group in all comparisons.

Additional file 3: Dataset S2. DE transcripts in all comparisons.

Additional file 4: Figure S2. Gene ontology enrichment.

12864_2024_10553_MOESM5_ESM.pdf

Additional file 5: Fig. S3. Unrooted maximum likelihood tree of genes (PPC-1E1, PPC-1E2 and PPC-2) from 35 different species with 11 transcripts of genes encoding PEPC. The IDs of PPC-1E1, PPC-1E2, and PPC-2 in the tree are Uniprot IDs, these IDs can be used to retrieve the corresponding protein sequences.

12864_2024_10553_MOESM6_ESM.pdf

Additional file 6: Fig. S4. The best K values were determined using the Elbow and Silhouette methods, as well as the Gap Statistic, in all comparisons.

12864_2024_10553_MOESM7_ESM.pdf

Additional file 7: Fig. S5. K-means clustering of DE transcripts between leaves and cotyledons of S. seuvioides and between C3 and C4 species.

Additional file 8: Dataset S3. List of DE transcripts within clusters in all comparisons.

12864_2024_10553_MOESM9_ESM.xlsx

Additional file 9: Dataset S4. Motif enrichments of DE genes related to CCM (Carbon Concentrating Mechanism) and starch metabolism in Sesuvium species.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Siadjeu, C., Kadereit, G. C4-like Sesuvium sesuvioides (Aizoaceae) exhibits CAM in cotyledons and putative C4-like + CAM metabolism in adult leaves as revealed by transcriptome analysis. BMC Genomics 25, 688 (2024). https://doi.org/10.1186/s12864-024-10553-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12864-024-10553-2

Keywords