Next Article in Journal
Impact of Somatic DNA Repair Mutations on the Clinical Outcomes of Bone Metastases from Castration-Resistant Prostate Cancer
Next Article in Special Issue
Auxin Involvement in Ceratopteris Gametophyte Meristem Regeneration
Previous Article in Journal
Cancer-Associated Fibroblasts Influence Survival in Pleural Mesothelioma: Digital Gene Expression Analysis and Supervised Machine Learning Model
Previous Article in Special Issue
Comparative Transcriptomic Analysis of Genes in the 20-Hydroxyecdysone Biosynthesis in the Fern Microsorum scolopendria towards Challenges with Foliar Application of Chitosan
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Proteome and Interactome Linked to Metabolism, Genetic Information Processing, and Abiotic Stress in Gametophytes of Two Woodferns

by
Sara Ojosnegros
1,
José Manuel Alvarez
1,
Jonas Grossmann
2,3,
Valeria Gagliardini
4,
Luis G. Quintanilla
5,
Ueli Grossniklaus
4 and
Helena Fernández
1,*
1
Area of Plant Physiology, Department of Organisms and Systems Biology, University of Oviedo, 33071 Oviedo, Spain
2
Functional Genomic Center Zurich, University and ETH Zurich, 8092 Zurich, Switzerland
3
Swiss Institute of Bioinformatics, 1015 Lausanne, Switzerland
4
Department of Plant and Microbial Biology & Zurich-Basel Plant Science Center, University of Zurich, 8008 Zurich, Switzerland
5
Department of Biology and Geology, Physics and Inorganic Chemistry, University Rey Juan Carlos, 28933 Móstoles, Spain
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(15), 12429; https://doi.org/10.3390/ijms241512429
Submission received: 23 June 2023 / Revised: 27 July 2023 / Accepted: 31 July 2023 / Published: 4 August 2023
(This article belongs to the Special Issue Molecular Approach to Fern Development)

Abstract

:
Ferns and lycophytes have received scant molecular attention in comparison to angiosperms. The advent of high-throughput technologies allowed an advance towards a greater knowledge of their elusive genomes. In this work, proteomic analyses of heart-shaped gametophytes of two ferns were performed: the apomictic Dryopteris affinis ssp. affinis and its sexual relative Dryopteris oreades. In total, a set of 218 proteins shared by these two gametophytes were analyzed using the STRING database, and their proteome associated with metabolism, genetic information processing, and responses to abiotic stress is discussed. Specifically, we report proteins involved in the metabolism of carbohydrates, lipids, and nucleotides, the biosynthesis of amino acids and secondary compounds, energy, oxide-reduction, transcription, translation, protein folding, sorting and degradation, and responses to abiotic stresses. The interactome of this set of proteins represents a total network composed of 218 nodes and 1792 interactions, obtained mostly from databases and text mining. The interactions among the identified proteins of the ferns D. affinis and D. oreades, together with the description of their biological functions, might contribute to a better understanding of the function and development of ferns as well as fill knowledge gaps in plant evolution.

Graphical Abstract

1. Introduction

Ferns and lycophytes represent a genetic legacy of great value, being descendants of the first plants that evolved vascular tissues about 470 million years ago. They are distributed throughout the world and play an important role in ecosystem functioning. Compared to angiosperms, they have received scant attention, relegating them to the background after a splendid past. The aesthetic appeal of their leaves and their use to alleviate ailments in traditional medicine is all that these plant groups have traditionally inspired. Specifically, fern gametophytes are ideal organisms for research on plant growth and reproduction, which is facilitated by a simple in vitro culture system and their small size of a few millimeters. In relation to climate change and other environmental events, ferns can also provide insights into adaptation, for example, they survived periods of high CO2 levels. Only a handful of species have been used to delve into basic developmental processes, such as photomorphogenesis [1], spore germination [2,3,4], cell polarity [5], cell wall composition [6], or reproduction. These studies focused on the gametophyte generation, constituted by an autonomously growing organism, which is well-suited for in vitro culture and sample collection [7,8]. Although fern gametophytes possess a very simple structure consisting mostly of a one-cell-thick layer, they display some degree of complexity: apical-basal polarity, dorsoventral asymmetry, rhizoids, meristems in the apical or lateral parts, reproductive organs (male antheridia and female archegonia), and trichomes distributed over the entire surface.
From a metabolic point of view, ferns and lycophytes contain many secondary metabolites, such as flavonoids, alkaloids, phenols, steroids, etc., and exhibit various bioactivities, including antibacterial, antidiabetic, anticancer, antioxidant, etc. [9]. The therapeutic use of both plant groups is changing from its use in the traditional medicine of different peoples to current applications, in which these plants are used to generate nanoparticles [10]. Finally, the use of ferns and lycophytes was recently advocated to address problems caused by biotic and abiotic stresses. Drought is one of the most severe abiotic stresses affecting plant growth and productivity, and ferns and lycophytes could contribute to better understanding and managing it [11]. Other important adaptations of ferns to extreme environments, such as salinity, heavy metals, epiphytes, or a low invasion of their habitats, were summarized by Rathinasabapathi [12]. Likewise, Dhir [13] highlights the high efficiency of many species of aquatic and terrestrial ferns in extracting various organic and inorganic pollutants from the environment. Recently, researchers have become more interested in these plants, which has been made possible by the advent of high-throughput technologies, such as transcriptomics, proteomics, and metabolomics. In fact, performing molecular analyses in ferns has been elusive, as they exhibit higher chromosome numbers and larger genomes than mosses and seed plants [14], which made it difficult to obtain genomic data. Gene expression, induced by either environmental or developmental conditions, can now be examined in non-model organisms because the required techniques have become more affordable as automation and efficiency have reduced costs.
Some transcriptomic and proteomic datasets have been published for ferns during the last decade. First, physiological and proteome analyses were reported with regard to the mechanism of drought tolerance in the resurrection lycophyte Selaginella tamariscina [11]. In 2011, the transcriptome of Pteridium aquilinum gametophytes was characterized de novo by pyrosequencing, representing the first complete analysis of the transcriptome of a fern [15]. In the same year, a proteomic analysis of the roots of Pteris vittata under arsenic stress was carried out [16]. In Blechnum spicant, proteomic profiles of male and female gametophytes were reported, revealing an increase in the amount of defense and stress proteins and a decrease in protein synthesis and photosynthesis when inducing male gametophyte development with antheridiogen pheromones [17]. In 2014, a de novo transcriptome assembly of Lygodium japonicum was carried out, and with this information a public database was created: Ljtrans DB [18]. Likewise, proteomic and cytological studies associated with germination and growth of rhizoid tips in the fern Osmunda cinnamomea were conducted [4]. In the tree fern Cyathea delgadii, a proteomic analysis on stipe explants revealed differentially expressed proteins associated with direct somatic embryogenesis [19]. Overlapping patterns of gene expression in gametophytes and sporophytes of the species Polypodium amorphum were analyzed by Sigel and colleagues [20], and in the Himalayan fern Diplazium maximum, proteomic analysis revealed multiple adaptive response mechanisms to cope with abiotic stresses [21]. Concerning reproduction, sexual versus apomictic expression profiles were analyzed in Ceratopteris thalictroides [22], as well as in the species Adiantum reniforme var. sinense [23]. In addition, some molecular studies were performed on the fern Ceratopteris richardii [5,24]; in one of them, sex determination was found to be accompanied by changes in the transcriptome driving epigenetic reprogramming of the young gametophyte [25]; in another, the WUSCHEL-related homeobox (WOX) gene, which promotes cell division in gametophytes and organ development in sporophytes, was characterized for the first time in ferns [26]; and finally, transcriptional analysis of the young sporophyte showed the conservation of stem cell factors in the root apical meristem [27]. More recently, other studies on ferns were carried out. In 2022, the first complete transcriptome of multiple organs of the fern Alsophila spinulosa was obtained, highlighting genes resistant to light stress [28]; its genome was also assembled and characterized; stem anatomy and lignin biosynthesis were investigated [29]; and transcriptomes of several fern species were compared in order to study the adaptive evolution of leaf and root morphology [30]. With the ferns Dryopteris affinis ssp. affinis [31,32,33] and Dryopteris oreades [33,34], both transcriptomic and proteomic analyses were performed by RNA-sequencing and shotgun proteomics using tandem mass spectrometry.
The current work expands our knowledge of proteomic data in ferns, which are far less explored than in seed plants. We present a continuation of previous work [21] on D. affinis spp. affinis (referred to as D. affinis hereafter) and its relative D. oreades. We have chosen these species as study models because they represent the two types of reproduction present in fern gametophytes: sexual in D. oreades and apomictic in D. affinis. Proteins of heart-shaped gametophytes were extracted and identified using a species-specific transcriptome database established in a previous project [31,32,33]. The functional annotation was inferred by blasting identified full-length protein sequences. We report the categorization of proteins that are shared by both sexual and apomictic gametophytes. Specifically, our analysis reveals new proteomic information involved in the metabolism of carbohydrates and lipids, the biosynthesis of amino acids, the metabolism of nucleotides and energy, as well as of secondary compounds, such as flavonoids, terpenoids, lignans, etc., which are important in the plant’s defense against stress. In addition, proteins related to transcription, translation, as well as protein folding, sorting, transport, and degradation are reported.

2. Results

A set of 218 proteins shared by the gametophytes of the apomictic fern D. affinis (DA) and its sexual relative D. oreades (DO) were analyzed using the software programs STRING version 11.5 and CYTOSCAPE version 3.9.1. Proteomic data are available online: https://www.frontiersin.org/articles/10.3389/fpls.2021.718932/full#supplementary-material accessed on 23 June 2023). Therefore, the present work completes previous studies in which 206 and 166 proteins, upregulated in the apomictic and sexual gametophytes, respectively, were analyzed by means of bioinformatic tools, as well as 145 proteins of the remaining 417 that were present in equal amounts in both apomictic and sexual gametophytes. In the present work, the biological functions of the list of 218 proteins were analyzed by using informatic support such as Gene Ontology (GO) and the Kyoto Encyclopedia of Genes and Genomes (KEGG) classifications provided by the STRING platform (Figure 1 and Figure 2). Based on GO classification, most of the identified proteins are involved in known biological functions. In summary, the percentage of proteins associated with the following biological functions are as follows: metabolism of carbohydrates (11.5%), lipids (1.8%), nucleotides (2.2%), biosynthesis of amino acids (6.8%), energy (15.1%), genetic information processing (26.6%), protein folding, sorting and degradation (26.6%), nitrogen and sulfur metabolism (1.8%), and secondary metabolism, including proteins coping with abiotic stress (15.1%).
Some relevant proteins found that will be discussed below are: GLYCINE-RICH RNA-BINDING PROTEIN 3 (RBG3), classified in the group transcription and translation, which functions in RNA processing during stress; FERREDOXIN-NITRITE REDUCTASE 1 (NIR1), grouped in sulfur and nitrogen metabolism, the main activities of which are the reduction of nitrite to ammonium and the improvement of plant assimilation of NO2; the protein ENOYL-[ACYL-CARRIER-PROTEIN] REDUCTASE (MOD1), classified in the group metabolism of lipids, involved in fatty acid synthesis and plant morphology; CHALCONE-FLAVANONE ISOMERASE 1 (CHI1), related to the group metabolism of secondary compounds and involved in the flavonoid synthesis; and PROTEIN TRANSLOCASE SUBUNIT SECA1 (SECA1), classified in the group transport, active in protein transfer across thylakoid membrane and plant acclimation. In addition, we found the following proteins in the gametophytes: THYLAKOID LUMENAL 16.5 kDa, grouped in the metabolism of energy and involved in photosynthesis; PHOTOSYNTHETIC NDH SUBUNIT OF LUMENAL LOCATION 5 IC (PNSL5), a protein also classified in the group metabolism of energy that binds to the promoter of the gene FLOWERING LOCUS D and represses its expression; ELONGATION FACTOR 2 (LOS1), related to the group transcription and translation and involved in the response to cold; and the protein LEUCINE AMINOPEPTIDASE 1 (LAP1), a chaperone protecting proteins from heat-induced damage.
KEGG classification also revealed that common proteins are mostly associated with the biosynthesis of secondary metabolites, the ribosome, the biosynthesis of amino acids, and protein degradation. These processes include the building of cellular organelles, such as ribosomes or proteasomes (Table 1). Related to ribosomes, there were several protein classes, such as nucleic acid-binding proteins, ribosomal proteins, translation elongation factors, etc. On the other hand, proteasomes mediate the degradation of proteins, and we found proteins of the 20S particle, the proteolytic core, but also regulatory factors.
Protein domains that are abundant in the gametophytes of both ferns were the pyruvate dehydrogenase E1 component and the histidine and lysine active sites of the phosphoenolpyruvate carboxylase, which is involved in carbohydrate metabolism. Regarding the biosynthesis of amino acids, the most abundant domains were aspartate aminotransferase and pyridoxal phosphate-dependent transferase. Among proteins involved in the metabolism of energy, the HAS barrel domain and the F1 complex of the alpha and beta subunits of ATP synthase were abundant. Related to the metabolism of secondary compounds, aromatic amino acid lyase, phenylalanine ammonia-lyase, and the N-terminus of histidase were enriched. Finally, the frequently found domains in proteins involved in transcription and translation were the GTP-binding domain and domain 2 of elongation factor Tu, as well as conserved sites of the ribosomal proteins S10 and S4.
The interactome of the proteins common to D. affinis and D. oreades represented a network composed of 218 nodes and 1792 interactions (p-value < 0.0001). The proteins with the highest number of interactions among the identified proteins are shown in Table 2: LARGE RIBOSOMAL SUBUNIT PROTEIN UL4Z (RPL4A), SMALL RIBOSOMAL SUBUNIT PROTEIN US11X (RPS14C), and SMALL RIBOSOMAL SUBUNIT PROTEIN US17Y (RPS11B) each have 44 interactions.
The strength of the interactions can be weak or strong (Table S1), using a scale from 0 to 1 where a weak interaction will have a score close to 0 and a strong one a score close to 1. Taking only interactions with a score equal to or greater than 0.99 for each group of proteins studied into account, proteins involved in transcription and translation are those with the highest number of interactions (554), followed by proteins involved in energy (29), carbohydrate metabolism (16), biosynthesis of amino acids (5), and transport (3). According to the STRING software v.11, the evidence of interactions between proteins can be of various types: (a) Experiments: these refer to proteins that have been shown to have chemical, physical, or genetic interaction in laboratory experiments. (b) Databases: this describes interactions of proteins found in the same databases. (c) Text mining: the proteins are mentioned in the same PubMed abstract or the same article of an internal selection of the STRING software v.11. (d) Co-expression: this indicates that the expression patterns of the two proteins are similar. (e) Neighborhood: the genes encoding the proteins are close to each other in the genome. (f) Gene fusion: this indicates that at least in one organism the orthologous genes encoding the two proteins are fused into a single gene. (g) Co-occurrence: this refers to proteins that have a similar phylogenetic distribution. The interactome presented here uses the species Arabidopsis thaliana because all the proteins discussed later were found to be homologs of proteins in this species. Therefore, the protein–protein interactions are those expected to be found in A. thaliana. On the other hand, STRING collects information from several sources and proposes protein–protein interactions according to the deposited data. We chose to explore all the types of protein–protein interactions published in the STRING platform for our selected proteins.
Next, we consider some of these types of interaction between proteins (Figure 3: text mining, experiments, co-expression, and databases). Specifically, we focused on the group metabolism of carbohydrates (Figure 3a), metabolism of energy (Figure 3b), ribogenesis (Figure 3c), and protein degradation (Figure 3d). Paying attention only to the two main types of evidence for each of these groups, their relationships were analyzed (Figure 4). Evidence from databases and text mining were the most relevant for the metabolism of carbohydrates (Figure 4a), biosynthesis of amino acids (Figure 4b), the metabolism of secondary compounds, and transport, as well as text mining and co-expression data for the metabolism of energy (Figure 4c), and experiments and co-expression data for transcription and translation (Figure 4d). The significance of associations between variables were as follows: in the metabolism of carbohydrates and in the transcription and translation, highly significant in both (p-value < 0.001); in the biosynthesis of amino acids, not significant (p-value > 0.05); and in the metabolism of energy, marginally significant (p-value slightly greater than 0.05).
Alternatively, when comparing the same type of evidence among the different groups of proteins, we observed that the neighborhood interaction was the most important for the biosynthesis of amino acids and transcription and translation; gene fusion for the metabolism of carbohydrates and biosynthesis of amino acids; co-occurrence for the biosynthesis of amino acids and metabolism of secondary compounds; co-expression for the metabolism of energy and transcription and translation; experiments for transcription and translation and transport; evidence from databases for the metabolism of carbohydrates and transport; and, finally, text mining for the metabolism of secondary compounds and transport.

3. Discussion

Molecular research conducted in non-model species, such as ferns and lycophytes, is still scarce. These plant groups are the last major lineage of land plants without a reference genome. Their higher chromosome numbers and larger genomes compared to those of mosses and seed plants have contributed to the paucity of reports dealing with genomic or proteomic analyses. Much more effort is needed to broaden the number of species analyzed to complete our knowledge of plant development. Thus, the current work provides novel information on the proteome shared by gametophytes of the apomictic fern D. affinis and its sexual relative D. oreades, and extends and completes previous studies in these species [31,32,33,34].
Next, we discuss the biological functions and interactions between the set of proteins obtained, which are grouped into three major categories: metabolism, genetic information processing, and response to abiotic stress (Table 3).

3.1. Metabolism

Primary and secondary metabolism encompasses a great number of enzymes connecting all the chemical pathways associated with them. Therefore, proteomic analyses usually yield a lot of proteins linked to the biosynthesis or degradation of carbohydrates, lipids, proteins, and nucleotides, as well as others that, albeit being called secondary, are not less important. Additionally, proteins linked to the metabolism of energy mediated by processes such as photosynthesis or photorespiration are commonly reported as well.
  • Carbohydrates
Within this metabolic group, the process of glycolysis converts glucose into pyruvate, and in the gametophytes under study, we found enzymes, such as ATP-DEPENDENT 6-PHOSPHOFRUCTOKINASE 3 (PFK3), involved in the first reaction, two enzymes participating in glycolysis and gluconeogenesis, FRUCTOSE-BISPHOSPHATE ALDOLASE 3 (FBA3), and others catalyzing the decarboxylation of pyruvate to acetyl-CoA, such as PYRUVATE DEHYDROGENASE E1 COMPONENT SUBUNIT BETA-1 (PDH2). The protein FBA3 reported here and also the protein FBA8 were identified in studies of the fern C. delgadii [19]. Linked to pyruvate metabolism, we identified two phosphoenolpyruvate carboxylases (PPC2 and PPC3), which supply oxaloacetate for the tricarboxylic acid cycle, and the protein NAD-DEPENDENT MALIC ENZYME 1 (NAD-ME1), which is involved in regulating the metabolism of sugars and amino acids during the night [35]. Worth mentioning is also 2,3-BIPHOSPHOGLYCERATE-INDEPENDENT PHOSPHOGLYCERATE MUTASE 1 (PGM1), which is important for the functioning of stomatal guard cells and fertility in A. thaliana [36]. Gametophytes of D. affinis and D. oreades produce proteins involved in starch synthesis, including 1,4-ALPHA-GLUCAN-BRANCHING ENZYME 2-2 (SBE2.2) and GRANULE-BOUND STARCH SYNTHASE 1 (GBSS1).
  • Tricarboxylic acid cycle and pentose phosphate pathway
Likewise, we identified some proteins associated with the citrate/tricarboxylic acid cycle, SUCCINATE-CoA LIGASE [ADP-FORMING] SUBUNIT BETA (AT2G20420) and SUCCINATE-CoA LIGASE [ADP-FORMING] SUBUNIT ALPHA-1 (AT5G08300), which are involved in the only phosphorylation step at the substrate level of this cycle. Another protein is MALATE DEHYDROGENASE 1 (MDH1), which catalyzes a reversible NAD-dependent dehydrogenase reaction involved in central metabolism and redox homeostasis between organelle compartments [37]. This protein was also found in three ferns: P. vittata, when studying the response to arsenic stress in the roots with or without arbuscular mycorrhizal symbiosis [16]; in germinating spores of O. cinnamomea [4]; and in sexual and apomictic gametophytes of C. thalictroides [22]. In parallel to glycolysis, the pentose phosphate pathway generates NADPH and pentoses. This metabolic pathway is represented in our dataset by the proteins 6-PHOSPHOGLUCONATE DEHYDROGENASE, 1 DECARBOXYLATING 1 (PGD1), GLUCOSE-6-PHOSPHATE 1-DEHYDROGENASE 6 (G6PD6), and PHOSPHOGLYCERATE KINASE 1 (PGK1). Specifically, a mutation in the gene of the first protein may decrease cellulose synthesis, thus altering the structure and composition of the primary cell wall [38]. The enzyme G6PD6 is important for the synthesis of fatty acids and nucleic acids involved in membrane synthesis and cell division [39]. G6PD6 was also reported in the vegetative tissues of the lycophyte S. tamariscina, which was involved in the response to drought [11], and the protein PGD1 was identified by analyzing stipe explants of the fern C. delgadii, and found to be associated with direct somatic embryogenesis [19].
  • Metabolism of lipids
Regarding the metabolism of lipids, three proteins were identified in this study. The first protein is ENOYL-[ACYL-CARRIER-PROTEIN] REDUCTASE (MOD1), which catalyzes the last reduction step of the de novo fatty acid synthesis cycle and the fatty acid elongation cycle. A mutation causing a decreased activity of this protein reduces the number of fatty acids, which triggers mosaic premature cell death and changes in the plant’s morphology, such as chlorotic and curly leaves, distorted siliques, and dwarfism [40]. The second protein is ATP-CITRATE SYNTHASE ALPHA CHAIN PROTEIN 1 (ACLA-1), which is necessary for the normal growth and development of plants because it synthesizes acetyl-CoA, a key compound for many metabolic pathways (fatty acids and glucosinolates in chloroplasts; flavonoids, sterols, and phospholipids in the cytoplasm; and ATP and amino acid carbon skeletons in mitochondria). Moreover, it is the substrate for histone acetylation in the nucleus, affecting chromosome structure and regulating transcription [41,42]. The third protein is CITRATE SYNTHASE 2 (CSY2), which synthesizes citrate in peroxisomes for the respiration of fatty acids in seedlings and is required for seed germination [43].
  • Biosynthesis of amino acids and nucleotides
Involved in the biosynthesis of amino acids, we found the proteins ASPARTATE AMINOTRANSFERASE (ASP1); 3-ISOPROPYLMALATE DEHYDRATASE LARGE SUBUNIT (IIL1), which acts in glucosinolate biosynthesis involved in the defense against insects [44]; FERREDOXIN-DEPENDENT GLUTAMATE SYNTHASE 1 (GLU1), which is required for the re-assimilation of ammonium ions generated during photorespiration [45]; HISTIDINOL DEHYDROGENASE (HISN8); and S-ADENOSYLMETHIONINE SYNTHASE 4 (METK4) [46,47,48]. We also identified proteins associated with the metabolism of nucleotides, specifically with AMP syntheses, such as ADENOSINE KINASE 1 (ADK1) and ADENYLOSUCCINATE SYNTHETASE (PURA). Of note is the protein PROBABLE RIBOSE-5-PHOSPHATE ISOMERASE 3 (RPI3), which is essential for the synthesis of numerous compounds such as purines, pyrimidines, aromatic amino acids, NAD, and NADP [38]. Apart from the proteins mentioned above, we found some that are associated with the biosynthesis of nucleotide sugars, such as the two pyrophosphorylases GLUCOSE-1-PHOSPHATE ADENYLYLTRANSFERASE SMALL SUBUNIT (ADG1) and GLUCOSE-1-PHOSPHATE ADENYLYLTRANSFERASE LARGE SUBUNIT 1 (ADG2).
  • Metabolism of energy
The protein OXYGEN-EVOLVING ENHANCER PROTEIN 1-2 (PSBO2), which regulates the replacement of the protein D1 impaired by light [49], and the protein THYLAKOID LUMENAL 16.5 kDa were reported. The latter protein is necessary to carry out photosynthesis correctly and efficiently under two conditions: controlled photoinhibitory light and fluctuating light. In nature, plants experience rapid and extreme changes in sunlight, requiring rapid adaptation [50]. Involved in photosynthesis, we found the protein PHOTOSYNTHETIC NDH SUBUNIT OF LUMENAL LOCATION 5 (PNSL5), which modulates the conformation of the protein BRASSINAZOLE-RESISTANT 1 (BZR1) [51]. This protein binds to the promoter of the gene FLOWERING LOCUS D (FLD) and represses its expression, eventually leading to the expression of the FLOWERING LOCUS C (FLC) gene, which encodes a repressor of flowering [51]. Finally, CBBY-LIKE PROTEIN (CBBY) degrades xylulose-1,5-bisphosphate, a potent inhibitor of the protein RUBISCO [52]. On the other hand, photorespiration represents a waste of the energy produced by photosynthesis. The enzyme D-GLYCERATE 3-KINASE (GLYK) catalyzes the final reaction of photorespiration [53]. Another important protein for photorespiration is SERINE-GLYOXYLATE AMINOTRANSFERASE (AGT1), which also participates in primary and lateral root development [54]. The latter protein was also reported in the fern C. delgadii [19].
  • Sulfur and nitrogen metabolism
Proteins involved in sulfur metabolism are represented by UDP-SULFOQUINOVOSE SYNTHASE (SQD1), which converts UDP-glucose and sulfite to the precursor of the main group of sulfolipids, UDP-sulfoquinovose, thus preventing sulfite from accumulating as it is toxic to the cell [55], and SUFE-LIKE PROTEIN 1 (SUFE1), a sulfur acceptor that activates cysteine desulfurases in plastids and mitochondria, which is essential for embryogenesis [56]. Regarding nitrogen metabolism, there are the proteins NITROGEN REGULATORY PROTEIN P-II HOMOLOG (GLB1), which is a nitrogen regulatory protein and intervenes in glycosaminoglycan degradation [57], and FERREDOXIN-NITRITE REDUCTASE (NIR1), which catalyzes the reduction of nitrite to ammonium [58]. As the amount of this protein in the cell increases, the tolerance and assimilation of nitrogen dioxide by the plant improves. As nitrogen dioxide is an air pollutant produced largely by motorized vehicles, plants could act as a sink for this substance, i.e., this protein could be used in biotechnological applications for bioremediation [58].
  • Metabolism of secondary compounds
Several proteins related to flavonoid biosynthesis are represented in this work, such as CHALCONE-FLAVANONE ISOMERASE 1 (CHI1), which is responsible for the isomerization of chalcones into naringenin [59]. We also found enzymes involved in the biosynthesis of terpenoids, such as HETERODIMERIC GERANYLGERANYL PYROPHOSPHATE SYNTHASE LARGE SUBUNIT 1 (GGPPS1); the biosynthesis of lignans, PHENYLCOUMARAN BENZYLIC ETHER REDUCTASE 1 (PCBER1); and the biosynthesis of phenylpropanoids. We also found the protein 4-COUMARATE-COA LIGASE 3 (4CL3), which produces CoA-thioesters of hydroxy- and methoxy-substituted cinnamic acids, used to synthesize anthocyanins, flavonoids, isoflavonoids, coumarins, lignin, suberin, and phenols [60], and 3-PHOSPHOSHIKIMATE 1-CARBOXYVINYLTRANSFERASE (AT2G45300), involved in the synthesis of chorismate, which is the precursor of the amino acids phenylalanine, tryptophan, and tyrosine [61]. The proteins 4CL3 and the transferases GLUTATHIONE S-TRANSFERASE L2 (GSTL2) and GLUTATHIONE S-TRANSFERASE L3 (GSTL3) found in our species, was also studied in the fern A. spinulosa [29]. The last ones catalyze the glutathione-dependent reduction of S-glutathionyl quercetin to quercetin [62]. Besides, GSTL2 and GSTL3 proteins were also reported in a lycophyte, S. tamariscina, where they are required in the response to desiccation [11].

3.2. Genetic Information Processing

  • Transcription and translation
In the gametophytes of D. affinis and D. oreades, we identified two proteins involved in transcription, specifically the 14-3-3-like proteins 14-3-3-LIKE PROTEIN GF14 NU (GRF7) and 14-3-3-LIKE PROTEIN GF14 IOTA (GRF12), which are associated with a DNA-binding complex that binds to the G-box, a cis-regulatory DNA element [63]. These 14-3-3-like proteins were also studied in two species: the lycophyte S. tamariscina [11] and the fern C. delgadii [19]. Related to translation, we found GLYCINE-RICH RNA-BINDING PROTEIN 3 (RBG3), which has a role in RNA processing during stress, specifically in editing cytosine to uracil in mitochondrial RNA, thereby controlling 6% of all mitochondrial editing sites [64]. This protein was also identified in S. tamariscina when studying the response to drought [11], as well as other proteins, such as PUTATIVE PENTATRICOPEPTIDE REPEAT-CONTAINING PROTEIN AT1G03510 (PCMP-E3), POLYADENYLATE-BINDING PROTEIN RBP47B (RBP47B), and UBP1-ASSOCIATED PROTEIN 2A (UBA2A), which regulates mRNAs and stabilizes RNAs in the nucleus [65]. Apart from several ribosomal subunits, there are others linked to translation elongation, like the protein ELONGATION FACTOR 2 (LOS1), which is also involved in the response to cold [66].
  • Protein folding and sorting
Once the proteins have been formed, there is a quality check to ensure that they have been synthesized completely and folded correctly. Among the proteins playing a major role in the acceleration of folding or the degradation of misfolded proteins are CHAPERONIN 1 (CPN10-1) and PEPTIDYL-PROLYL CIS-TRANS ISOMERASE FKBP16-4 (FKBP16-4). The gametophyte of the ferns under study harbor proteins linked to the sorting or transport of molecules within the cell and between the inside and outside of cells. In line with this, we found PROTEIN TRANSLOCASE SUBUNIT SECA1 (SECA1), which has a role in coupling ATP hydrolysis to protein transfer across the thylakoid membrane, thus participating in photosynthetic acclimation and chloroplast formation [67]; IMPORTIN SUBUNIT ALPHA-2 (IMPA2), which acts in nuclear import [68]; the proteins ADP and ATP CARRIER (AAC2), which mediates the import of ADP into the mitochondrial matrix [69], and TIC62 (TIC62), which is involved in the import of nuclear-encoded proteins into chloroplasts [70]. The IMPA-2 protein was also identified when studying germinating spores of the fern O. cinnamomea [4]. In addition, we found proteins associated with the transport of water and small hydrophilic molecules through the cell membrane: PROBABLE AQUAPORIN PIP1-4 (PIP1.4) [71]. COATOMER SUBUNIT ALPHA-1 (AT1G62020) and COATOMER SUBUNIT GAMMA (AT4G34450) are associated with clathrin-uncoated vesicles that are transported from the endoplasmic reticulum to the Golgi apparatus and vice versa. In contrast, the proteins CLATHRIN INTERACTOR EPSIN 2 (EPSIN2) and DYNAMIN-2B (DRP2B) are related to clathrin-coated vesicles, with the latter participating in planar polarity formation to correctly positioning the root hairs [72].
  • Protein degradation
Among the proteases, we found ATP-DEPENDENT CLP PROTEASE PROTEOLYTIC SUBUNIT-RELATED PROTEIN 3 (CLPR3) and ATP-DEPENDENT CLP PROTEASE ATP-BINDING SUBUNIT CLPT2 (CLPR2). Plants need to cope with heat stress, and for this, the gametophytes studied here rely on the aminopeptidases LEUCINE AMINOPEPTIDASE 1 and LEUCINE AMINOPEPTIDASE 3 (LAP1 and LAP3), which are probably involved in the processing and turnover of intracellular proteins and function as molecular chaperones protecting proteins from heat-induced damage [73].

3.3. Protein–Protein Interactions

Using the STRING platform, we thoroughly analyzed—one by one—the interactions of the groups of proteins studied. We observed that for the metabolism of carbohydrates, evidence from co-expression, text mining, and experiments were stronger between the mitochondrial proteins SUCCINATE-CoA LIGASE [ADP-FORMING] SUBUNIT BETA and SUBUNIT ALPHA-1 than the others in this group. Both proteins are involved in the tricarboxylic acid cycle [74].
Among the proteins for biosynthesis of amino acids, evidence from co-expression was stronger between the proteins ASPARTATE-SEMIALDEHYDE DEHYDROGENASE and DIHYDROXY-ACID DEHYDRATASE (DHAD), while evidence from databases was stronger between DIHYDROXY-ACID DEHYDRATASE and 2-ISOPROPYLMALATE SYNTHASE 2 (IPMS2), 3-ISOPROPYLMALATE DEHYDRATASE LARGE SUBUNIT (IIL1) and 2-ISOPROPYLMALATE SYNTHASE 2, and 3-ISOPROPYLMALATE DEHYDRATASE LARGE SUBUNIT and 3-ISOPROPYLMALATE DEHYDROGENASE 2 (IMD2). In fact, these proteins are involved in the synthesis of numerous compounds necessary for plant growth and development: ASPARTATE-SEMIALDEHYDE DEHYDROGENASE for the biosynthesis of lysine, threonine, and methionine [75]; DIHYDROXY-ACID DEHYDRATASE for isoleucine and valine [76]; 2-ISOPROPYLMALATE SYNTHASE 2 and 3-ISOPROPYLMALATE DEHYDROGENASE 2 for leucine [77,78]; and 3-ISOPROPYLMALATE DEHYDRATASE LARGE SUBUNIT for glucosinolates [44].
For the metabolism of energy, evidence from co-expression was stronger between the proteins ATP SYNTHASE GAMMA CHAIN 1 (ATPC1) and GLYCERALDEHYDE-3-PHOSPHATE DEHYDROGENASE (GAPA-2), while experimental data provided strong evidence for the interaction between PHOTOSYSTEM I P700 CHLOROPHYLL A APOPROTEIN A1 (PSAA) and PHOTOSYSTEM I IRON-SULFUR CENTER (PSAC). In photosynthesis, the C-terminus of PSAC interacts with PSAA and other proteins, such as PHOTOSYSTEM I P700 CHLOROPHYLL A APOPROTEIN A2 (PSAB) and PHOTOSYSTEM I REACTION CENTER SUBUNIT II-1 (PSAD1), for its assembly into the photosystem I [79]. Evidence from databases indicated an interaction between SERINE-GLYOXYLATE AMINOTRANSFERASE (AGT1) and GLYCOLATE OXIDASE 2 (GLO2), both proteins being involved in photorespiration [54], while evidence from text mining suggested an interaction between OXYGEN-EVOLVING ENHANCER PROTEIN 1-2 (PSBO2) and OXYGEN-EVOLVING ENHANCER PROTEIN 2-1 (PSBP1), both being chloroplastic oxygen-evolving enhancer proteins that form part of photosystem II [49].
For the metabolism of secondary compounds, evidence from text mining was the strongest, indicating an interaction between PHENYLALANINE AMMONIA-LYASE 1 (PAL1) and PHENYLALANINE AMMONIA-LYASE 4 (PAL4). Both proteins participate in the synthesis from phenylalanine of numerous compounds based on the phenylpropane skeleton, which is fundamental to plant metabolism [80].
With respect to transcription and translation, co-expression evidence was stronger between ribosomal proteins.
Finally, for protein transport, co-expression data provided the strongest evidence for interactions between the proteins COATOMER SUBUNIT ALPHA-1 and COATOMER SUBUNIT GAMMA; experimental data for the interaction between COATOMER SUBUNIT ALPHA-1 and COATOMER SUBUNIT DELTA; and text mining data for the interactions between PROTEIN TRANSLOCASE SUBUNIT SECA1 and ATPase GET3B.
As indicated in the results, in the group related to the metabolism of carbohydrates, the protein with the most interactions was PHOSPHOGLYCERATE KINASE 1, which is involved in glycolysis [81]. For the biosynthesis of amino acids, ASPARTATE-SEMIALDEHYDE DEHYDROGENASE, DIHYDROXY-ACID DEHYDRATASE, and 3-ISOPROPYLMALATE DEHYDROGENASE 2 were involved in several biosynthetic pathways: lysine, leucine, valine, isoleucine, methionine, and threonine [77]. In the group metabolism of energy, ATP SYNTHASE GAMMA CHAIN 1 had the highest number of interactions, likely because it is part of a chloroplastic ATP synthase [82]. The protein 4-COUMARATE-COA LIGASE 3 had the most interactions in the group metabolism of secondary compounds. It plays a key role in the synthesis of numerous secondary metabolites, such as anthocyanins, flavonoids, isoflavonoids, coumarins, lignin, suberin, and phenols [60]. In transcription and translation, the ribosomal proteins LARGE RIBOSOMAL SUBUNIT PROTEIN UL4Z, SMALL RIBOSOMAL SUBUNIT PROTEIN US11X, and SMALL RIBOSOMAL SUBUNIT PROTEIN US17Y, which are necessary for the formation of ribosomes, had the highest number of interactions [83]. Finally, in transport, the proteins PROTEIN TRANSLOCASE SUBUNIT SECA1 and COATOMER SUBUNIT GAMMA participate in coupling ATP hydrolysis to protein transfer across the thylakoid membrane and in the transport of clathrin-uncoated vesicles from the endoplasmic reticulum to the Golgi apparatus and vice versa, respectively [67].
Regarding the statistical analysis of the two highest scoring types of interactions in the studied groups of metabolism of carbohydrates (database and text mining) and transcription and translation (experiments and co-expression), Pearson’s correlation coefficients, which measure the tendency of two vectors to increase or decrease together, were significant. One of the most popular types of data in databases is text, and the process of synthesizing information is known as text mining. In the case of proteins linked to carbohydrate metabolism, there seems to be a lot of information in databases about it, and therefore text mining could be enriched as well. Regarding transcription and translation, we speculate that most of the experiments on molecular biology cope with these processes, reporting more genes involved on them.

4. Materials and Methods

4.1. Plant Material and Growth Conditions

Spores of D. affinis were obtained from sporophytes growing in Turón valley (Asturias, Spain), 477 m a.s.l., 43°12′10″ N−5°43′43″ W. For D. oreades, spores were collected from sporophytes growing in Neila lagoons (Burgos, Spain), 1.920 m a.s.l., 42°02′48″ N−3°03′44″ W. Spores were released from sporangia, soaked in water for 2 h, and then washed for 10 min with a solution of NaClO (0.5%) and Tween 20 (0.1%). Then, they were rinsed three times with sterile, distilled water. Spores were centrifuged at 1300× g for 3 min between rinses and then cultured in 500 mL Erlenmeyer flasks containing 100 mL of liquid Murashige and Skoog (MS) medium [84]. Unless otherwise noted, media were supplemented with 2% sucrose (w/v), and the pH was adjusted to 5.7 with 1 or 0.1 N NaOH. The cultures were kept on an orbital shaker (75 rpm) at 25 °C under cool, white fluorescent light (70 µmol m−2s−1) with a 16 h photoperiod.
Following spore germination, filamentous gametophytes were subcultured into 200 mL flasks containing 25 mL of MS medium supplemented with 2% sucrose (w/v) and 0.7% agar. The gametophytes of D. affinis became two-dimensional, arriving at the spatulate and heart stage after 20 or 30 additional days, respectively. Gametophytes of D. oreades grew slower and needed around six months to become cordate and reach sexual maturity (Figure 5). Apomictic and sexual gametophytes were collected, and images were taken under a light microscope (Nikon Eclipse E600, Tokyo, Japan) using microphotographic equipment (DS Camera Control, Nikon, Tokyo, Japan). Gametophytes of D. oreades had only female reproductive organs (i.e., archegonia), while cordate, apomictic gametophytes of D. affinis had visible developing apogamic centers composed of smaller and darker isodiametric cells. Samples of apomictic and sexual cordate gametophytes were weighed before and after lyophilization for 48 h (Telstar-Cryodos, Terrassa, Spain) and stored in Eppendorf tubes in a freezer at −20 °C until use.

4.2. Protein Extraction, Separation, and In-Gel Digestion

The present work expands previous bioinformatic analyses, and, specifically, it is carried out with a set of 218 proteins, which had been extracted and annotated as follows.
The protocol used for protein extraction, separation, and in-gel digestion was reported earlier [31]. In brief, samples were solubilized with 800 μL of buffer A (0.5 M Tris-HCl (pH 8.0), 5 mM EDTA, 0.1 MHEPES-KOH, 4 mM DTT, 15 mM EGTA, 1 mM PMSF, 0.5% (w/v) PVP, and 1x protease inhibitor cocktail (Roche, Rotkreuz, Switzerland)) and homogenized, and proteins were extracted in two steps: first, the homogenate was subjected to centrifugation at 16,200× g for 10 min at 4 °C on a tabletop centrifuge, and, second, the supernatant was subjected to ultracentrifugation at 117–124 kPa (100,000× g) for 45 min at 4 °C in an Airfuge (Beckman Coulter, Pasadena, CA, USA), yielding the soluble protein fraction in the supernatant. In parallel, the pellet obtained from the first ultracentrifugation was re-dissolved in 200 μL of buffer B (40 mM Tris-base, 40 mM DTT, 4% (w/v) SDS, and 1× protease inhibitor cocktail (Roche, Rotkreuz, Switzerland)) to extract membrane proteins using the ultracentrifuge, as described above, in the supernatant. Protein concentrations were determined using a Qubit Fluorometer (Invitrogen, Carlsbad, CA, USA), and 1D gel electrophoresis was performed as follows: 1 mg protein was treated with sample loading buffer and 2 M DTT, heated at 99 °C for 5 min, followed by a short cooling period on ice, and then loaded separately onto a 0.75 mm thick, 12% SDS-PAGE mini-gel. Electrophoresis conditions were 150 V and 250 mA for 1 h in 1× running buffer.

4.3. Protein Separation and In-Gel Digestion

Each gel lane was cut into six 0.4 cm wide sections resulting in 48 slices, and then fragmented into smaller pieces and subjected to 10 mM DTT (in 25 mM AmBic, pH8) for 45 min at 56 °C and 50 mM iodoacetamide for 1 h at room temperature in the dark prior to trypsin digestion at 37 °C overnight. Subsequently, gel pieces were washed twice with 100 μL of 100 mM NH4HCO3/50% acetonitrile and washed once with 50 μL acetonitrile. At this point, the supernatants were discarded. Peptides were digested with 20 μL trypsin (5 ng/L in 10 mM Tris/2 mM CaCl2, pH 8.2) and 50 μL buffer (10 mM Tris/2 mM CaCl2, pH 8.2). After microwave-heating for 30 min at 60 °C, the supernatant was removed, and gel pieces were extracted once with 150 μL 0.1% TFA/50% acetonitrile. All supernatants were put together, then dried and dissolved in 15 μL 0.1% formic acid/3% acetonitrile, and, finally, transferred to auto-sampler vials for liquid chromatography (LC)-tandem mass spectrometry (MS/MS) for which 5 μL was injected.

4.4. Protein Identification, Verification, and Bioinformatic Downstream Analyses

MS/MS and peptide identification (Orbitrap XL) were performed according to [31]. Scaffold software (version Scaffold 4.2.1, Proteome Software Inc., Portland, OR, USA) was used to validate MS/MS-based peptide and protein identifications. Mascot results were analyzed together using the MudPIT option. Peptide identifications were accepted if they scored better than 95.0% probability as specified by the Peptide Prophet algorithm with delta mass correction, and protein identifications were accepted if the Protein Prophet probability was above 95%. Indeed, the unique peptide ≥ 2′ was considered. Proteins that contained the same peptides and could not be differentiated based on MS/MS alone were grouped to satisfy the principles of parsimony using the scaffolds cluster analysis option. Only proteins that met the above criteria were considered as positively identified for further analysis. The number of random matches was evaluated by performing the Mascot searches against a database containing decoy entries and checking how many decoy entries (proteins or peptides) passed the applied quality filters. The peptide FDR and protein FDR were estimated at 2% and 1%, respectively, indicating the stringency of the analyses. A semi-quantitative spectrum counting analysis was conducted. The “total spectrum count” for each protein and each sample was reported, and these spectrum counts were averaged for each species, D. affinis and D. oreades. Then, one “1” was added to each average in order to prevent division by zero, and a log2 ratio of the averaged spectral counts from D. affinis versus D. oreades was calculated. Proteins were considered as differentially expressed if this log2 ratio was above 0.99. This refers to at least twice as many peptide spectrum match (PSM) assignments in one group compared to the other. Also, to provide some functional understanding of the identified proteins, we blasted the whole protein sequences of all identified proteins against Sellginella moellendorfii and A. thaliana Uniprot sequences and retrieved the best matching identifier from each of them along with the corresponding e-value, accepting blast-hits with values below 1 × 10−7. These better-described ortholog identifiers were then used in further downstream analysis.

4.5. Protein Analysis Using the STRING Platform

The identifiers of the genes from apomictic and sexual gametophyte samples were used as input for STRING platform version 11.5 analysis, and a high threshold (0.700) was selected for a positive interaction between a pair of proteins.

4.6. Statistical Analyses

Regarding the two major protein–protein interactions highlighted for carbohydrate metabolism, amino acid biosynthesis, energy metabolism, and transcription and translation, a Pearson’s correlation test was performed using R software 4.2.0, and p-values lower than 0.05 were considered significant.

5. Conclusions

The analysis of a set of 218 proteins shared by the gametophytes of the apomictic fern D. affinis and its sexual relative D. oreades revealed the presence of proteins mostly involved in biological functions associated with metabolism, the processing of genetic information, and abiotic stresses. Some smaller protein groups were studied in detail: metabolism of carbohydrates; biosynthesis of amino acids; metabolism of energy; metabolism of secondary compounds; transcription, translation, and transport; and abiotic stress. Possible interactions between these proteins were identified, with the most common source of evidence for interactions stemming from databases and information from text mining. The proteins involved in transcription and translation exhibit the strongest interactions. The description of possible biological functions and the possible protein–protein interactions among the identified proteins expands our current knowledge about ferns and plants in general.

Supplementary Materials

The supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ijms241512429/s1.

Author Contributions

Conceptualization, H.F. and U.G.; methodology, H.F., J.G., V.G., J.M.A. and S.O.; formal analysis, J.G. and H.F.; writing—original draft preparation, H.F. and S.O., with help from U.G.; writing—review and editing, U.G., J.G., V.G., L.G.Q. and J.M.A.; funding acquisition, H.F. and U.G.; resources, L.G.Q. and U.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by core funding of the University of Zurich to U.G., the University of Oviedo through the Grant CESSTT1819 for the International Mobility of Research Staff to H.F., and the European Union’s 7th Framework Program: PRIME-XS-000252 to H.F. and U.G.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The concatenated dDB is available online at http://fgcz-r-021.uzh.ch/fasta/p1222_combo_NGS_n_Viridi_20160205.fasta (accessed on 9 November 2022).

Acknowledgments

We thank the University of Oviedo for a grant from the International Mobility of Research Staff, according to the collaboration agreement CESSTT1819, and the Functional Genomics Center Zurich for access to its excellent infrastructure. We also thank Hanspeter Schöb for his logistic support during H.F.’s visits to the Grossniklaus laboratory.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wada, M. The fern as a model system to study photomorphogenesis. J. Plant Res. 2007, 120, 3–16. [Google Scholar] [CrossRef]
  2. Salmi, M.L.; Bushart, T.; Stout, S.; Roux, S. Profile and analysis of gene expression changes during early development in germinating spores of Ceratopteris richardii. Plant Physiol. 2005, 138, 1734–1745. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Salmi, M.L.; Morris, K.E.; Roux, S.J.; Porterfield, D.M. Nitric oxide and CGMP signaling in calcium-dependent development of cell polarity in Ceratopteris richardii. Plant Physiol. 2007, 144, 94–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Suo, J.; Zhao, Q.; Zhang, Z.; Chen, S.; Cao, J.; Liu, G.; Wei, X.; Wang, T.; Yang, C.; Dai, S. Cytological and proteomic analyses of Osmunda cinnamomea germinating spores reveal characteristics of fern spore germination and rhizoid tip growth. Mol. Cell. Proteom. 2015, 14, 2510–2534. [Google Scholar] [CrossRef] [Green Version]
  5. Salmi, M.L.; Bushart, T.J. Cellular, molecular, and genetic changes during the development of Ceratopteris richardii gametophytes. In Working with Ferns: Issues and Applications; Fernández, H., Kumar, A., Revilla, M.A., Eds.; Springer International Publishing: New York, NY, USA, 2010; pp. 11–24. [Google Scholar]
  6. Eeckhout, S.; Leroux, O.; Willats, W.G.; Popper, Z.A.; Viane, R.L. Comparative glycan profiling of Ceratopteris richardii ‘C-fern’ gametophytes and sporophytes links cell-wall composition to functional specialization. Ann. Bot. 2014, 114, 1295–1307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Fernández, H.; Revilla, M.A. In vitro culture of ornamental ferns. Plant Cell. Tissue Organ Cult. 2003, 73, 1–13. [Google Scholar] [CrossRef]
  8. Rivera, A.; Cañal, M.J.; Grossniklaus, U.; Fernández, H. The gametophyte of fern: Born to reproduce. In Current Advances in Fern Research; Fernández, H., Ed.; Springer International Publishing: New York, NY, USA, 2018; pp. 3–19. [Google Scholar]
  9. Chen, C.-Y.; Chiu, F.-Y.; Lin, Y.; Huang, W.-J.; Hsieh, P.-S.; Hsu, F.-L. Chemical constituents analysis and antidiabetic activity validation of four fern species from Taiwan. Int. J. Mol. Sci 2015, 16, 2497–2516. [Google Scholar] [CrossRef] [Green Version]
  10. Femi-Adepoju, A.G.; Dada, A.O.; Otun, K.O.; Adepoju, A.O.; Fatoba, O.P. Green synthesis of silver nanoparticles using terrestrial fern (Gleichenia pectinata (Willd.) C. Presl.): Characterization and antimicrobial studies. Heliyon 2019, 5, e01543. [Google Scholar] [CrossRef] [Green Version]
  11. Wang, X.; Chen, S.; Zhang, H.; Shi, L.; Cao, F.; Guo, L.; Xie, Y.; Wang, T.; Yan, X.; Dai, S. Desiccation tolerance mechanism in resurrection fern-ally Selaginella tamariscina revealed by physiological and proteomic analysis. J. Proteome Res. 2010, 9, 6561–6577. [Google Scholar] [CrossRef]
  12. Rathinasabapathi, B. Ferns represent an untapped biodiversity for improving crops for environmental stress tolerance. New Phytol. 2006, 172, 385–390. [Google Scholar] [CrossRef]
  13. Dhir, B. Role of ferns in environmental cleanup. In Current Advances in Fern Research; Fernández, H., Ed.; Springer International Publishing: Cham, Switzerland, 2018; pp. 517–531. [Google Scholar]
  14. Barker, M.S.; Wolf, P.G. Unfurling fern biology in the genomics age. Bioscience 2010, 60, 177–185. [Google Scholar] [CrossRef] [Green Version]
  15. Der, J.P.; Barker, M.S.; Wickett, N.J.; de Pamphilis, C.W.; Wolf, P.G. De novo characterization of the gametophyte transcriptome in bracken fern, Pteridium aquilinum. BMC Genom. 2011, 12, 99. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Bona, E.; Marsano, F.; Massa, N.; Cattaneo, C. Proteomic analysis as a tool for investigating arsenic stress in Pteris vittata roots colonized or not by arbuscular mycorrhizal symbiosis. J. Proteom. 2011, 74, 1338–1350. [Google Scholar] [CrossRef]
  17. Valledor, L.; Menéndez, V.; Canal, M.J.; Revilla, A.; Fernández, H. Proteomic approaches to sexual development mediated by antheridiogen in the fern Blechnum spicant L. Proteomics 2014, 14, 2061–2071. [Google Scholar] [CrossRef] [PubMed]
  18. Aya, K.; Kobayashi, M.; Tanaka, J.; Ohyanagi, H.; Suzuki, T.; Yano, K.; Takano, T.; Yano, K.; Matsuoka, M. De novo transcriptome assembly of a fern, Lygodium japonicum, and a web resource database Ljtrans DB. Plant Cell Physiol. 2015, 56, e5. [Google Scholar] [CrossRef] [Green Version]
  19. Domżalska, L.; Kędracka-Krok, S.; Jankowska, U.; Grzyb, M.; Sobczak, M.; Rybczyński, J.J.; Mikuła, A. Proteomic analysis of stipe explants reveals differentially expressed proteins involved in early direct somatic embryogenesis of the tree fern Cyathea delgadii Sternb. Plant Sci. 2017, 258, 61–76. [Google Scholar] [CrossRef] [PubMed]
  20. Sigel, E.M.; Schuettpelz, E.; Pryer, K.M.; Der, J.P. Overlapping patterns of gene expression between gametophyte and sporophyte phases in the fern Polypodium amorphum (Polypodiales). Front. Plant Sci. 2018, 9, 1450. [Google Scholar] [CrossRef]
  21. Sareen, B.; Thapa, P.; Joshi, R.; Bhattacharya, A. Proteome analysis of the gametophytes of a western Himalayan fern Diplazium maximum reveals their adaptive responses to changes in their micro-environment. Front. Plant Sci. 2019, 10, 1623. [Google Scholar] [CrossRef]
  22. Chen, X.; Chen, Z.; Huang, W.; Fu, H.; Wang, Q.; Wang, Y.; Cao, J. Proteomic analysis of gametophytic sex expression in the fern Ceratopteris thalictroides. PLoS ONE 2019, 14, e0221470. [Google Scholar] [CrossRef] [Green Version]
  23. Fu, Q.; Chen, L. Comparative transcriptome analysis of two reproductive modes in Adiantum reniforme var. sinense targeted to explore possible mechanism of apogamy. BMC Genet. 2019, 20, 1–14. [Google Scholar] [CrossRef]
  24. Cordle, A.; Irish, E.; Cheng, C.L. Gene expression associated with apogamy commitment in Ceratopteris richardii. Sex. Plant Reprod. 2012, 25, 293–304. [Google Scholar] [CrossRef] [PubMed]
  25. Atallah, N.M.; Vitek, O.; Gaiti, F.; Tanurdzic, M.; Banks, J.A. Sex determination in Ceratopteris richardii is accompanied by transcriptome changes that drive epigenetic reprogramming of the young gametophyte. G3 2018, 8, 2205–2214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Youngstrom, C.E.; Geadelmann, L.F.; Irish, E.E.; Cheng, C.-L. A fern WUSCHEL-RELATED HOMEOBOX gene functions in both gametophyte and sporophyte generations. BMC Plant Biol. 2019, 19, 1–13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Aragón-Raygoza, A.; Herrera-Estrella, L.; Cruz-Ramirez, A. Transcriptional analysis of Ceratopteris richardii young sporophyte reveals conservation of stem cell factors in the root apical meristem. Front. Plant Sci. 2022, 13, 924660. [Google Scholar] [CrossRef]
  28. Hong, Y.; Wang, Z.; Li, M.; Su, Y.; Wang, T. First multi-organ full-length transcriptome of tree fern Alsophila spinulosa highlights the stress-resistant and light-adapted genes. Front. Plant Sci. 2022, 12, 784546. [Google Scholar] [CrossRef]
  29. Huang, X.; Wang, W.; Gong, T.; Wickell, D.; Kuo, L.-Y.; Zhang, X.; Wen, J.; Kim, H.; Lu, F.; Zhao, H.; et al. The flying spider-monkey tree fern genome provides insights into fern evolution and arborescence. Nat. Plants 2022, 8, 500–512. [Google Scholar] [CrossRef]
  30. Xia, Z.; Liu, L.; Wei, Z.; Wang, F.; Shen, H.; Yan, Y. Analysis of comparative transcriptome and positively selected genes reveal adaptive evolution in leaf-less and root-less whisk ferns. Plants 2022, 11, 1198. [Google Scholar] [CrossRef]
  31. Grossmann, J.; Fernández, H.; Chaubey, P.M.; Valdés, A.E.; Gagliardini, V.; Cañal, M.J.; Russo, G.; Grossniklaus, U. Proteogenomic analysis greatly expands the identification of proteins related to reproduction in the apogamous fern Dryopteris affinis ssp. affinis. Front. Plant Sci. 2017, 8, 336. [Google Scholar] [CrossRef] [Green Version]
  32. Wyder, S.; Rivera, A.; Valdés, A.E.; Cañal, M.J.; Gagliardini, V.; Fernández, H.; Grossniklaus, U. Differential gene expression profiling of one- and two-dimensional apogamous gametophytes of the fern Dryopteris affinis ssp. affinis. Plant Physiol. Biochem. 2020, 148, 302–311. [Google Scholar] [CrossRef]
  33. Fernández, H.; Grossmann, J.; Gagliardini, V.; Feito, I.; Rivera, A.; Rodríguez, L.; Quintanilla, L.G.; Quesada, V.; Cañal, M.J.; Grossniklaus, U. Sexual and apogamous species of woodferns show different protein and phytohormone profiles. Front. Plant Sci. 2021, 12, 718932. [Google Scholar] [CrossRef]
  34. Ojosnegros, S.; Alvarez, J.M.; Grossmann, J.; Gagliardini, V.; Quintanilla, L.G.; Grossniklaus, U.; Fernández, H. The shared proteome of the apomictic fern Dryopteris affinis ssp. affinis and its sexual relative Dryopteris oreades. Int. J. Mol. Sci. 2022, 23, 14027. [Google Scholar] [CrossRef]
  35. Tronconi, M.A.; Fahnenstich, H.; Gerrard Weehler, M.C.; Andreo, C.S.; Flügge, U.I.; Drincovich, M.F.; Maurino, V.G. Arabidopsis NAD-malic enzyme functions as a homodimer and heterodimer and has a major impact on nocturnal metabolism. Plant Physiol. 2008, 146, 1540. [Google Scholar] [CrossRef] [Green Version]
  36. Zhao, Z.; Assmann, S.M. The glycolytic enzyme, phosphoglycerate mutase, has critical roles in stomatal movement, vegetative growth, and pollen production in Arabidopsis thaliana. J. Exp. Bot. 2011, 62, 5179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Tomaz, T.; Bagard, M.; Pracharoenwattana, I.; Lindén, P.; Lee, C.P.; Carroll, A.J.; Ströher, E.; Smith, S.M.; Gardeström, P.; Millar, A.H. Mitochondrial malate dehydrogenase lowers leaf respiration and alters photorespiration and plant growth in Arabidopsis. Plant Physiol. 2010, 154, 1143–1157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Howles, P.A.; Birch, R.J.; Collings, D.A.; Gebbie, L.K.; Hurley, U.A.; Hocart, C.H.; Arioli, T.; Williamson, R.E. A mutation in an Arabidopsis ribose 5-phosphate isomerase reduces cellulose synthesis and is rescued by exogenous uridine. Plant J. 2006, 48, 606–618. [Google Scholar] [CrossRef]
  39. Wakao, S.; Benning, C. Genome-wide analysis of glucose-6-phosphate dehydrogenases in Arabidopsis. Plant J. 2005, 41, 243–256. [Google Scholar] [CrossRef] [PubMed]
  40. Mou, Z.; He, Y.; Dai, Y.; Liu, X.; Li, J. Deficiency in fatty acid synthase leads to premature cell death and dramatic alterations in plant morphology. Plant Cell 2000, 12, 405–417. [Google Scholar] [CrossRef] [Green Version]
  41. Fatland, B.L.; Ke, J.; Anderson, M.D.; Mentzen, W.I.; Wei Cui, L.; Christy Allred, C.; Johnston, J.L.; Nikolau, B.J.; Syrkin Wurtele, E.; Biology LWC, M. Molecular characterization of a heteromeric ATP-citrate lyase that generates cytosolic acetyl-coenzyme A in Arabidopsis. Plant Physiol. 2002, 130, 740–756. [Google Scholar] [CrossRef] [Green Version]
  42. Fatland, B.L.; Nikolau, B.J.; Wurtele, E.S. Reverse genetic characterization of cytosolic acetyl-CoA generation by ATP-citrate lyase in Arabidopsis. Plant Cell 2005, 17, 182–203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Pracharoenwattana, I.; Cornah, J.E.; Smith, S.M. Arabidopsis peroxisomal citrate synthase is required for fatty acid respiration and seed germination. Plant Cell 2005, 17, 2037–2048. [Google Scholar] [CrossRef] [Green Version]
  44. Knill, T.; Reichelt, M.; Paetz, C.; Gershenzon, J.; Binder, S. Arabidopsis thaliana encodes a bacterial-type heterodimeric isopropylmalate isomerase involved in both Leu biosynthesis and the Met chain elongation pathway of glucosinolate formation. Plant Mol. Biol. 2009, 71, 227–239. [Google Scholar] [CrossRef] [Green Version]
  45. Ishizaki, T.; Ohsumi, C.; Totsuka, K.; Igarashi, D. Analysis of glutamate homeostasis by overexpression of Fd-GOGAT gene in Arabidopsis thaliana. Amin. Acids 2009, 38, 943–950. [Google Scholar] [CrossRef] [PubMed]
  46. Zhang, Y.; Sun, K.; Sandoval, F.J.; Santiago, K.; Roje, S. One-carbon metabolism in plants: Characterization of a plastid serine hydroxymethyltransferase. Biochem. J. 2010, 430, 97–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Petersen, L.N.; Marineo, S.; Mandalà, S.; Davids, F.; Sewell, B.T.; Ingle, R.A. The missing link in plant histidine biosynthesis: Arabidopsis MYOINOSITOL MONOPHOSPHATASE-LIKE2 encodes a functional histidinol-phosphate phosphatase. Plant Physiol. 2010, 152, 1186–1196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Shen, B.; Li, C.; Tarczynski, M.C. High free-methionine and decreased lignin content result from a mutation in the Arabidopsis S-ADENOSYL-L-METHIONINE SYNTHETASE 3 gene. Plant J. 2002, 29, 371–380. [Google Scholar] [CrossRef] [PubMed]
  49. Lundin, B.; Hansson, M.; Schoefs, B.; Vener, A.V.; Spetea, C. The Arabidopsis PsbO2 protein regulates dephosphorylation and turnover of the photosystem II reaction centre D1 protein. Plant J. 2007, 49, 528–539. [Google Scholar] [CrossRef] [PubMed]
  50. Liu, J.; Last, R.L. A chloroplast thylakoid lumen protein is required for proper photosynthetic acclimation of plants under fluctuating light environments. Proc. Natl. Acad. Sci. USA 2017, 114, E8110–E8117. [Google Scholar] [CrossRef]
  51. Zhang, Y.; Li, B.; Xu, Y.; Li, H.; Li, S.; Zhang, D.; Mao, Z.; Guo, S.; Yang, C.; Weng, Y.; et al. The cyclophilin CYP20-2 modulates the conformation of BRASSINAZOLE-RESISTANT1, which binds the promoter of FLOWERING LOCUS D to regulate flowering in Arabidopsis. Plant Cell 2013, 25, 2504–2521. [Google Scholar] [CrossRef] [Green Version]
  52. Bracher, A.; Sharma, A.; Starling-Windhof, A.; Hartl, F.U.; Hayer-Hartl, M. Degradation of potent RUBISCO inhibitor by selective sugar phosphatase. Nat. Plants 2014, 112015, 1–7. [Google Scholar] [CrossRef]
  53. Boldt, R.; Edner, C.; Kolukisaoglu, Ü.; Hagemann, M.; Weckwerth, W.; Wienkoop, S.; Morgenthal, K.; Bauwe, H. D-glycerate 3-kinase, the last unknown enzyme in the photorespiratory cycle in Arabidopsis, belongs to a novel kinase family. Plant Cell 2005, 17, 2413–2420. [Google Scholar] [CrossRef] [Green Version]
  54. Wang, R.; Yang, L.; Han, X.; Zhao, Y.; Zhao, L.; Xiang, B.; Zhu, Y.; Bai, Y.; Wang, Y. Overexpression of AtAGT1 promoted root growth and development during seedling establishment. Plant Cell Rep. 2019, 38, 1165–1180. [Google Scholar] [CrossRef] [PubMed]
  55. Sanda, S.; Leustek, T.; Theisen, M.J.; Garavito, R.M.; Benning, C. Recombinant Arabidopsis SQD1 converts UDP-glucose and sulfite to the sulfolipid head group precursor UDP-sulfoquinovose in vitro. J. Biol. Chem. 2001, 276, 3941–3946. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Ye, H.; Abdel-Ghany, S.E.; Anderson, T.D.; Pilon-Smits, E.A.; Pilon, M. CpSufE activates the cysteine desulfurase CpNifS for chloroplastic Fe-S cluster formation. J. Biol. Chem. 2006, 281, 8958–8969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Ferrario-Méry, S.; Meyer, C.; Hodges, M. Chloroplast nitrite uptake is enhanced in Arabidopsis PII mutants. FEBS Lett. 2008, 582, 1061–1066. [Google Scholar] [CrossRef] [Green Version]
  58. Takahashi, M.; Sasaki, Y.; Ida, S.; Morikawa, H. Nitrite reductase gene enrichment improves assimilation of NO2 in Arabidopsis. Plant Physiol. 2001, 126, 731–741. [Google Scholar] [CrossRef] [Green Version]
  59. Shirley, B.W.; Hanley, S.; Goodman, H.M. Effects of ionizing radiation on a plant genome: Analysis of two Arabidopsis transparent testa mutations. Plant Cell 1992, 4, 333–347. [Google Scholar] [CrossRef]
  60. Ehlting, J.; Büttner, D.; Wang, Q.; Douglas, C.J.; Somssich, I.E.; Kombrink, E. Three 4-coumarate: Coenzyme A ligases in Arabidopsis thaliana represent two evolutionarily divergent classes in angiosperms. Plant J. 1999, 19, 9–20. [Google Scholar] [CrossRef]
  61. Klee, H.J.; Muskopf, Y.M.; Gasser, C.S. Cloning of an Arabidopsis thaliana gene encoding 5-enolpyruvyl shikimate-3-phosphate synthase: Sequence analysis and manipulation to obtain glyphosate-tolerant plants. Mol. Gen. Genet. 1987, 210, 437–442. [Google Scholar] [CrossRef]
  62. Dixon, D.P.; Edwards, R. Roles for stress-inducible lambda glutathione transferases in flavonoid metabolism in plants as identified by ligand fishing. J. Biol. Chem. 2010, 285, 36322–36329. [Google Scholar] [CrossRef] [Green Version]
  63. Rosenquist, M.; Alsterfjord, M.; Larsson, C.; Sommarin, M. Data mining the Arabidopsis genome reveals fifteen 14-3-3 genes: Expression is demonstrated for two out of five novel genes. Plant Physiol. 2001, 127, 142–149. [Google Scholar] [CrossRef] [Green Version]
  64. Shi, X.; Hanson, M.R.; Bentolila, S. Two RNA recognition motif-containing proteins are plant mitochondrial editing factors. Nucleic Acids Res. 2015, 43, 3814–3825. [Google Scholar] [CrossRef]
  65. Lambermon, M.H.L.; Fu, Y.; Kirk, D.A.W.; Dupasquier, M.; Filipowicz, W.; Lorković, Z.J. UBA1 and UBA2, two proteins that interact with UBP1, a multifunctional effector of pre-mRNA maturation in plants. Mol. Cell. Biol. 2002, 22, 4346–4357. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Guo, Y.; Xiong, L.; Ishitani, M.; Zhu, J.K. An Arabidopsis mutation in TRANSLATION ELONGATION FACTOR 2 causes superinduction of CBF/DREB1 transcription factor genes but blocks the induction of their downstream targets under low temperatures. Proc. Natl. Acad. Sci. USA 2002, 99, 7786–7791. [Google Scholar] [CrossRef]
  67. Skalitzky, C.A.; Martin, J.R.; Harwood, J.H.; Beirne, J.J.; Adamczyk, B.J.; Heck, G.R.; Cline, K.; Fernandez, D.E. Plastids contain a second Sec translocase system with essential functions. Plant Physiol. 2011, 155, 354–369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Bhattacharjee, S.; Lee, L.Y.; Oltmanns, H.; Cao, H.; Veena; Cuperus, J.; Gelvin, S.B. IMPa-4, an Arabidopsis importin α isoform, is preferentially involved in Agrobacterium-mediated plant transformation. Plant Cell 2008, 20, 2661–2680. [Google Scholar] [CrossRef] [Green Version]
  69. Haferkamp, I.; Hackstein, J.H.P.; Voncken, F.G.J.; Schmit, G.; Tjaden, J. Functional integration of mitochondrial and hydrogenosomal ADP/ATP carriers in the Escherichia coli membrane reveals different biochemical characteristics for plants, mammals and anaerobic chytrids. Eur. J. Biochem. 2002, 269, 3172–3181. [Google Scholar] [CrossRef]
  70. Küchler, M.; Decker, S.; Hörmann, F.; Soll, J.; Heins, L. Protein import into chloroplasts involves redox-regulated proteins. EMBO J. 2002, 21, 6136–6145. [Google Scholar] [CrossRef] [Green Version]
  71. Lee, S.H.; Chung, G.C.; Jang, J.Y.; Ahn, S.J.; Zwiazek, J.J. Overexpression of PIP2;5 aquaporin alleviates effects of low root temperature on cell hydraulic conductivity and growth in Arabidopsis. Plant Physiol. 2012, 159, 479–488. [Google Scholar] [CrossRef] [Green Version]
  72. Stanislas, T.; Hüser, A.; Barbosa, I.C.R.; Kiefer, C.S.; Brackmann, K.; Pietra, S.; Gustavsson, A.; Zourelidou, M.; Schwechheimer, C.; Grebe, M. Arabidopsis D6PK is a lipid domain-dependent mediator of root epidermal planar polarity. Nat. Plants 2015, 1, 15162. [Google Scholar] [CrossRef] [PubMed]
  73. Scranton, M.A.; Yee, A.; Park, S.Y.; Walling, L.L. Plant leucine aminopeptidases moonlight as molecular chaperones to alleviate stress-induced damage. J. Biol. Chem. 2012, 287, 18408–18417. [Google Scholar] [CrossRef] [Green Version]
  74. Tan, Y.-F.; O’Toole, N.; Taylor, N.L.; Millar, A.H. Divalent metal ions in plant mitochondria and their role in interactions with proteins and oxidative stress-induced damage to respiratory function. Plant Physiol. 2010, 152, 747–761. [Google Scholar] [CrossRef] [Green Version]
  75. Zhang, Y.; Swart, C.; Alseekh, S.; Scossa, F.; Jiang, L.; Obata, T.; Graf, A.; Fernie, A.R. The extra-pathway interactome of the TCA cycle: Expected and unexpected metabolic interactions. Plant Physiol. 2018, 177, 966–979. [Google Scholar] [CrossRef] [Green Version]
  76. Yan, Y.; Liu, Q.; Zang, X.; Yuan, S.; Bat-Erdene, U.; Nguyen, C.; Gan, J.; Zhou, J.; Jacobsen, S.E.; Tang, Y. Resistance-gene-directed discovery of a natural-product herbicide with a new mode of action. Nature 2018, 559, 415–418. [Google Scholar] [CrossRef]
  77. de Kraker, J.W.; Luck, K.; Textor, S.; Tokuhisa, J.G.; Gershenzon, J. Two Arabidopsis genes (IPMS1 and IPMS2) encode isopropylmalate synthase, the branchpoint step in the biosynthesis of leucine. Plant Physiol. 2007, 143, 970–986. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. He, Y.; Chen, L.; Zhou, Y.; Mawhinney, T.P.; Chen, B.; Kang, B.H.; Hauser, B.A.; Chen, S. Functional characterization of Arabidopsis thaliana isopropylmalate dehydrogenases reveals their important roles in gametophyte development. New Phytol. 2011, 189, 160–175. [Google Scholar] [CrossRef]
  79. Varotto, C.; Pesaresi, P.; Meurer, J.; Oelmüller, R.; Steiner-Lange, S.; Salamini, F.; Leister, D. Disruption of the Arabidopsis photosystem I gene psaE1 affects photosynthesis and impairs growth. Plant J. 2000, 22, 115–124. [Google Scholar] [CrossRef]
  80. Cochrane, F.C.; Davin, L.B.; Lewis, N.G. The Arabidopsis phenylalanine ammonia lyase gene family: Kinetic characterization of the four PAL isoforms. Phytochemistry 2004, 65, 1557–1564. [Google Scholar] [CrossRef]
  81. Gargano, D.; Maple-Groedem, J.; Moeller, S.G. In vivo phosphorylation of FtsZ2 in Arabidopsis thaliana. Biochem. J. 2012, 446, 517–521. [Google Scholar] [CrossRef]
  82. Takagi, D.; Amako, K.; Hashiguchi, M.; Fukaki, H.; Ishizaki, K.; Goh, T.; Fukao, Y.; Sano, R.; Kurata, T.; Demura, T.; et al. Chloroplastic ATP synthase builds up a proton motive force preventing production of reactive oxygen species in photosystem I. Plant J. 2017, 91, 306–324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Carroll, A.J.; Heazlewood, J.L.; Ito, J.; Millar, A.H. Analysis of the Arabidopsis cytosolic ribosome proteome provides detailed insights into its components and their post-translational modification. Mol. Cell Proteom. 2008, 7, 347–369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Murashige, T.; Skoog, F. A revised medium for rapid growth and bioassays with tobacco tissue cultures. Plant Physiol. 1962, 15, 473–497. [Google Scholar] [CrossRef]
Figure 1. GO enrichment terms of the proteomes shared by gametophytes of D. affinis and D. oreades according to the category biological function; analyzed by STRING and CYTOSCAPE. Turquoise indicates metabolism of carbohydrates, yellow metabolism of energy, pink transcription and translation, and green protein degradation.
Figure 1. GO enrichment terms of the proteomes shared by gametophytes of D. affinis and D. oreades according to the category biological function; analyzed by STRING and CYTOSCAPE. Turquoise indicates metabolism of carbohydrates, yellow metabolism of energy, pink transcription and translation, and green protein degradation.
Ijms 24 12429 g001
Figure 2. KEGG enrichment terms of the proteomes shared by gametophytes of D. affinis and D. oreades; analyzed using the STRING platform.
Figure 2. KEGG enrichment terms of the proteomes shared by gametophytes of D. affinis and D. oreades; analyzed using the STRING platform.
Ijms 24 12429 g002
Figure 3. Circular representations obtained by STRING and CYTOSCAPE for proteins detected in the gametophytes of both D. affinis and D. oreades: (a) metabolism of carbohydrates, (b) metabolism of energy, (c) ribogenesis, and (d) protein degradation. Pink lines refer to evidence from experiments, green lines from text mining, black lines from co-expression, and blue lines from databases.
Figure 3. Circular representations obtained by STRING and CYTOSCAPE for proteins detected in the gametophytes of both D. affinis and D. oreades: (a) metabolism of carbohydrates, (b) metabolism of energy, (c) ribogenesis, and (d) protein degradation. Pink lines refer to evidence from experiments, green lines from text mining, black lines from co-expression, and blue lines from databases.
Ijms 24 12429 g003
Figure 4. Plots of the two main types of evidence for interactions in the groups of proteins shared by the gametophytes of D. affinis and D. oreades: (a) metabolism of carbohydrates, (b) biosynthesis of amino acids, (c) metabolism of energy, and (d) transcription and translation. Each spot represents the intersection of the type of evidence for interactions between two proteins. The linear regression and the coefficient of correlation are provided for each pair of evidence for the interaction.
Figure 4. Plots of the two main types of evidence for interactions in the groups of proteins shared by the gametophytes of D. affinis and D. oreades: (a) metabolism of carbohydrates, (b) biosynthesis of amino acids, (c) metabolism of energy, and (d) transcription and translation. Each spot represents the intersection of the type of evidence for interactions between two proteins. The linear regression and the coefficient of correlation are provided for each pair of evidence for the interaction.
Ijms 24 12429 g004
Figure 5. Morphological traits in the apomictic fern D. affinis and its sexual relative D. oreades: (a) confocal image of spores of D. affinis; (b) gametophytes of D. affinis growing in a Petri dish; (c) image taken under a light microscope of a chordate gametophyte of D. affinis showing an apomictic center in the middle; and (d) archegonia in the gametophyte of D. oreades.
Figure 5. Morphological traits in the apomictic fern D. affinis and its sexual relative D. oreades: (a) confocal image of spores of D. affinis; (b) gametophytes of D. affinis growing in a Petri dish; (c) image taken under a light microscope of a chordate gametophyte of D. affinis showing an apomictic center in the middle; and (d) archegonia in the gametophyte of D. oreades.
Ijms 24 12429 g005
Table 1. Proteins involved in ribogenesis and protein degradation found in the gametophyte of the ferns D. affinis and D. oreades.
Table 1. Proteins involved in ribogenesis and protein degradation found in the gametophyte of the ferns D. affinis and D. oreades.
CategoryAccession NumberUniProtKB/
Swiss-Prot
Gene NameProtein NameMW (kDa)%
Coverage
Exclusive Unique
Peptides
Total SpectrumE-Value
Ribogenesis1-12177_3_ORF5P56799RPS4SMALL RIBOSOMAL SUBUNIT PROTEIN US4C25.60037.63 × 10−70
Ribogenesis16646-605_4_ORF2Q9SX68RPL18LARGE RIBOSOMAL SUBUNIT PROTEIN UL18C20.97137.35 × 10−57
Ribogenesis24557-506_3_ORF2Q9SKX4RPL3ALARGE RIBOSOMAL SUBUNIT PROTEIN UL3C29.314295.37 × 10−121
Ribogenesis139931-207_5_ORF1O04603RPL5LARGE RIBOSOMAL SUBUNIT PROTEIN UL5C32.482142.41 × 10−92
Ribogenesis26987-486_4_ORF2Q9M3C3RPL23ABLARGE RIBOSOMAL SUBUNIT PROTEIN UL23Y20.2311393.98 × 10−60
Ribogenesis6809-878_4_ORF2P49227RPL5BLARGE RIBOSOMAL SUBUNIT PROTEIN UL18Y34.7134121.09 × 10−53
Ribogenesis87113-269_4_ORF1Q9M9W1RPL22BLARGE RIBOSOMAL SUBUNIT PROTEIN EL22Z16.114131.35 × 10−52
Ribogenesis21899-533_1_ORF2P51419RPL27CLARGE RIBOSOMAL SUBUNIT PROTEIN EL27X18.2193227.19 × 10−65
Ribogenesis24282-509_2_ORF1P50883RPL12ALARGE RIBOSOMAL SUBUNIT PROTEIN UL11Z19.113061.24 × 10−104
Ribogenesis36225-421_1_ORF2Q9SIM4RPL14ALARGE RIBOSOMAL SUBUNIT PROTEIN EL14Z17.9192112.62 × 10−68
Ribogenesis10983-723_3_ORF2P51418RPL18ABLARGE RIBOSOMAL SUBUNIT PROTEIN EL20Y22.910191.99 × 10−119
Ribogenesis34727-430_6_ORF2P49637RPL27ACLARGE RIBOSOMAL SUBUNIT PROTEIN UL15X16.37148.73 × 10−78
Ribogenesis73912-293_6_ORF2Q42064RPL8CLARGE RIBOSOMAL SUBUNIT PROTEIN UL2X28.7254422.24 × 10−167
Ribogenesis37807-413_3_ORF1Q93VI3RPL17ALARGE RIBOSOMAL SUBUNIT PROTEIN UL22Z23.54175.22 × 10−107
Ribogenesis176389-160_3_ORF2Q42351RPL34ALARGE RIBOSOMAL SUBUNIT PROTEIN EL34Z18.2121151.23 × 10−65
Ribogenesis8788-797_5_ORF2Q9FZH0RPL35ABLARGE RIBOSOMAL SUBUNIT PROTEIN EL33Z13.6180176.54 × 10−60
Ribogenesis27503-481_2_ORF2O80929RPL36ALARGE RIBOSOMAL SUBUNIT PROTEIN EL36Z13161288.82 × 10−52
Ribogenesis75664-290_4_ORF2Q9SF40RPL4ALARGE RIBOSOMAL SUBUNIT PROTEIN UL4Z46.3123200
Ribogenesis45247-378_5_ORF2Q9SZX9RPL9DLARGE RIBOSOMAL SUBUNIT PROTEIN UL6X25.171111.63 × 10−105
Ribogenesis86488-270_3_ORF2Q9LZH9RPL7ABLARGE RIBOSOMAL SUBUNIT PROTEIN EL8Y32.5268337.63 × 10−154
Ribogenesis163051-176_2_ORF2Q8VZ19RPL30BLARGE RIBOSOMAL SUBUNIT PROTEIN EL30Y16.713291.91 × 10−61
Ribogenesis69050-304_1_ORF1P49200RPS20ASMALL RIBOSOMAL SUBUNIT PROTEIN US10Z21.1122143.32 × 10−72
Ribogenesis5816-941_6_ORF2P42036RPS14CSMALL RIBOSOMAL SUBUNIT PROTEIN US11X18.722372.52 × 10−87
Ribogenesis39126-407_1_ORF2Q8LC83RPS24BSMALL RIBOSOMAL SUBUNIT PROTEIN ES24Y20.77128.38 × 10−74
Ribogenesis108940-238_6_ORF2Q42262RPS3ABSMALL RIBOSOMAL SUBUNIT PROTEIN ES1Y32.7225218.87 × 10−155
Ribogenesis85164-272_6_ORF2Q8VYK6RPS4DSMALL RIBOSOMAL SUBUNIT PROTEIN ES4X3072189.28 × 10−165
Ribogenesis45770-376_1_ORF2Q9LXG1RPS9BSMALL RIBOSOMAL SUBUNIT PROTEIN US4Z24.74091.32 × 10−121
Ribogenesis140134-206_2_ORF2F4JB06MGH6.2RIBOSOMAL PROTEIN S5/ELONGATION FACTOR G/III/V FAMILY PROTEIN16.97012 × 10−50
Ribogenesis1627-1498_1_ORF1P61841RPS7-ASMALL RIBOSOMAL SUBUNIT PROTEIN US7CZ18.8133113.24 × 10−77
Ribogenesis31704-450_4_ORF1Q9FIF3ES8YRIBOSOMAL PROTEIN ES8Y16.8281191.93 × 10−73
Ribogenesis11320-714_2_ORF2Q9XJ27RPS9SMALL RIBOSOMAL SUBUNIT PROTEIN US9C25.93131.03 × 10−76
Ribogenesis21394-539_3_ORF2P42791RPL18BLARGE RIBOSOMAL SUBUNIT PROTEIN EL18Y23.4253211.03 × 10−105
Ribogenesis40578-399_1_ORF2P56798RPS3SMALL RIBOSOMAL SUBUNIT PROTEIN US3C25.36166.73 × 10−82
Ribogenesis10791-728_3_ORF2O65569RPS11BSMALL RIBOSOMAL SUBUNIT PROTEIN US17Y20.5122365.17 × 10−84
Ribogenesis66444-310_2_ORF2Q1PEP5NUCL2NUCLEOLIN 261.5126116.83 × 10−49
Ribogenesis260531-93_3_ORF2O04658NOP5-1PROBABLE NUCLEOLAR PROTEIN 5-162.5115100
Proteasome93086-259_4_ORF1Q9LT08RPN1126S PROTEASOME NON-ATPASE REGULATORY SUBUNIT 14 HOMOLOG34.93150
Proteasome78122-285_4_ORF2Q93Y35RPN726S PROTEASOME NON-ATPASE REGULATORY SUBUNIT 6 HOMOLOG44.84190
Proteasome353924-42_4_ORF1O23712PAF2PROTEASOME SUBUNIT ALPHA TYPE-1-B13,98122.8 × 10−136
Proteasome7073-864_3_ORF1O23708PAB1PROTEASOME SUBUNIT ALPHA TYPE-2-A26.492104.87 × 10−158
Proteasome326729-54_1_ORF2O81148PAC1PROTEASOME SUBUNIT ALPHA TYPE-4-A30.613393.69 × 10−154
Proteasome68626-305_4_ORF2Q42134PAE2PROTEASOME SUBUNIT ALPHA TYPE-5-B28.7283134.68 × 10−158
Proteasome23928-513_3_ORF1O81147PAA2PROTEASOME SUBUNIT ALPHA TYPE-6-B30.873129.67 × 10−153
Proteasome340935-47_4_ORF1O24616PAD2PROTEASOME SUBUNIT ALPHA TYPE-7-B27.4184113.86 × 10−148
Proteasome14462-637_1_ORF2O23714PBD1PROTEASOME SUBUNIT BETA TYPE-2-A24.34171.43 × 10−103
Table 2. Biological functions and number of interactions exhibited by proteins common to apomictic D. affinis and sexual D. oreades.
Table 2. Biological functions and number of interactions exhibited by proteins common to apomictic D. affinis and sexual D. oreades.
Biological FunctionProteinNº of Interactions
Metabolism of carbohydratesPHOSPHOGLYCERATE KINASE 111
Biosynthesis of amino acids3-ISOPROPYLMALATE DEHYDROGENASE5
DIHYDROXY-ACID DEHYDRATASE5
ASPARTATE-SEMIALDEHYDE DEHYDROGENASE5
Metabolism of energyATP SYNTHASE GAMMA CHAIN 118
Secondary metabolism4-COUMARATE-COA LIGASE 33
Transcription & TranslationLARGE RIBOSOMAL SUBUNIT PROTEIN UL4Z44
SMALL RIBOSOMAL SUBUNIT PROTEIN US11X44
SMALL RIBOSOMAL SUBUNIT PROTEIN US17Y44
TransportPROTEIN TRANSLOCASE SUBUNIT SECA12
COATOMER SUBUNIT GAMMA2
Table 3. Selected proteins found in gametophytes of both apomictic D. affinis and sexual D. oreades.
Table 3. Selected proteins found in gametophytes of both apomictic D. affinis and sexual D. oreades.
CategoryAccession NumberUniProtKB/
Swiss-Prot
Gene NameProtein NameMW (kDa)%
Coverage
Exclusive Unique PeptidesTotal SpectraE-Value
Carbohydrates58787-330_2_ORF2Q94AA4PFK3ATP-DEPENDENT 6-PHOSPHOFRUCTOKINASE 364.52110
Carbohydrates135690-210_1_ORF2Q9ZU52FBA3FRUCTOSE-BISPHOSPHATE ALDOLASE 342.7327380
Carbohydrates38153-411_5_ORF2Q38799PDH2PYRUVATE DEHYDROGENASE E1 COMPONENT SUBUNIT BETA-140.314460
Carbohydrates83096-276_3_ORF2Q5GM68PPC2PHOSPHOENOLPYRUVATE CARBOXYLASE 2112.86550
Carbohydrates54280-344_1_ORF1Q84VW9PPC3PHOSPHOENOLPYRUVATE CARBOXYLASE 3111.81010
Carbohydrates113756-233_2_ORF1Q9SIU0NAD-ME1NAD-DEPENDENT MALIC ENZYME 171.43230
Carbohydrates102811-246_6_ORF2O04499PGM12,3-BIPHOSPHOGLYCERATE-INDEPENDENT PHOSPHOGLYCERATE MUTASE 163.24120
Carbohydrates64100-316_1_ORF1O82662AT2G20420SUCCINATE-COA LIGASE [ADP-FORMING] SUBUNIT BETA509040
Carbohydrates8279-816_3_ORF2P68209AT5G08300SUCCINATE-COA LIGASE [ADP-FORMING] SUBUNIT ALPHA-134.69230
Carbohydrates222487-119_2_ORF2P93819MDH1MALATE DEHYDROGENASE 138.4202160
Carbohydrates156827-185_4_ORF1Q9SH69PGD16-PHOSPHOGLUCONATE DEHYDROGENASE, DECARBOXYLATING 158.816291.37 × 10−112
Carbohydrates12493-682_6_ORF2Q9FJI5G6PD6GLUCOSE-6-PHOSPHATE 1-DEHYDROGENASE 665.16330
Carbohydrates20760-547_4_ORF1Q9LD57PGK1PHOSPHOGLYCERATE KINASE 119.514035.63 × 10−43
Carbohydrates69882-302_6_ORF2Q9LZS3SBE2.21,4-ALPHA-GLUCAN-BRANCHING ENZYME 2-298.65460
Carbohydrates96049-255_6_ORF1Q9MAQ0GBSS1GRANULE BOUND STARCH SYNTHASE 170.161110
Lipids20213-554_2_ORF1Q9SLA8MOD1ENOYL-[ACYL-CARRIER-PROTEIN] REDUCTASE [NADH]41.87164.48 × 10−180
Lipids387953-27_4_ORF1Q9SGY2ACLA-1ATP-CITRATE SYNTHASE ALPHA CHAIN PROTEIN 146.810240
Lipids211149-128_1_ORF1Q9LXS6CSY2CITRATE SYNTHASE 257.92010
Amino acids47558-369_4_ORF2P46643ASP1ASPARTATE AMINOTRANSFERASE51.55050
Amino acids72506-296_4_ORF1Q94AR8IIL13-ISOPROPYLMALATE DEHYDRATASE LARGE SUBUNIT57.66340
Amino acids125905-219_3_ORF2Q9ZNZ7GLU1FERREDOXIN-DEPENDENT GLUTAMATE SYNTHASE 1181.365130
Amino acids14065-645_3_ORF2Q9C5U8HISN8HISTIDINOL DEHYDROGENASE55.21210
Amino acids294436-71_4_ORF2Q9LUT2METK4S-ADENOSYLMETHIONINE SYNTHASE 442.75020
Nucleotides2121-1366_3_ORF2Q9SF85ADK1ADENOSINE KINASE 139.214340
Nucleotides59309-329_5_ORF1Q96529PURAADENYLOSUCCINATE SYNTHETASE57.92110
Nucleotides152024-193_3_ORF2Q9S726RPI3PROBABLE RIBOSE-5-PHOSPHATE ISOMERASE 336.12025.33 × 10−120
Nucleotides27769-479_3_ORF1P55228ADG1GLUCOSE-1-PHOSPHATE ADENYLYLTRANSFERASE SMALL SUBUNIT12.57014.43 × 10−63
Nucleotides181563-155_3_ORF2P55229ADG2GLUCOSE-1-PHOSPHATE ADENYLYLTRANSFERASE LARGE SUBUNIT 157.46240
Energy154679-189_1_ORF2Q9S841PSBO2OXYGEN-EVOLVING ENHANCER PROTEIN 1-235.3357246.62 × 10−141
Energy218625-122_1_ORF2O22773AT4G02530THYLAKOID LUMENAL 16.5 kDa PROTEIN24.75118.55 × 10−47
CategoryAccession NumberUniProtKB/
Swiss-Prot
Gene NameProtein NameMW (kDa)%
Coverage
Exclusive Unique PeptidesTotal SpectraE-Value
Energy6036-926_2_ORF1Q9ASS6PNSL5PHOTOSYNTHETIC NDH SUBUNIT OF LUMENAL LOCATION 532.2154102.45 × 10−93
Energy250817-99_2_ORF2Q94K71CBBYCBBY-LIKE PROTEIN34.97236.24 × 10−131
Energy235330-110_2_ORF1Q944I4GLYKD-GLYCERATE 3-KINASE43.93012.82 × 10−153
Energy297118-70_2_ORF2Q56YA5AGT1SERINE-GLYOXYLATE AMINOTRANSFERASE47.82020
S&N metabolism33137-439_6_ORF2O48917SQD1UDP-SULFOQUINOVOSE SYNTHASE54.910340
S&N metabolism227095-115_1_ORF2Q84W65SUFE1SUFE-LIKE PROTEIN 140.72113.22 × 10−106
S&N metabolim311596-62_2_ORF2Q9ZST4GLB1NITROGEN REGULATORY PROTEIN P-II HOMOLOG23.415113.28 × 10−62
S&N metabolim318906-58_1_ORF1Q39161NIR1FERREDOXIN-NITRITE REDUCTASE69.67480
Secondary compounds156331-186_3_ORF2P41088CHI1CHALCONE-FLAVANONE ISOMERASE 126.26026.89 × 10−56
Secondary compounds230420-113_2_ORF2P34802GGPPS1HETERODIMERIC GERANYLGERANYL PYROPHOSPHATE SYNTHASE LARGE SUBUNIT 1415111.48 × 10−147
Secondary compounds85783-271_1_ORF2Q9T030PCBER1PHENYLCOUMARAN BENZYLIC ETHER REDUCTASE 134.9339234.03 × 10−115
Secondary compounds156554-185_2_ORF1Q9S7774CL34-COUMARATE-COA LIGASE 351.92120
Secondary compounds223603-118_1_ORF1P05466AT2G453003-PHOSPHOSHIKIMATE 1-CARBOXYVINYLTRANSFERASE44.32110
Oxido
-reduction
133847-212_2_ORF2Q9SID3AT2G31350HYDROXYACYLGLUTATHIONE HYDROLASE 233.16116.51 × 10−131
Oxido
-reduction
34437-432_2_ORF1Q9M2W2GSTL2GLUTATHIONE S-TRANSFERASE L216.115253.24 × 10−29
Oxido
-reduction
115571-230_4_ORF1Q9LZ06GSTL3GLUTATHIONE S-TRANSFERASE L335.31216.13 × 10−75
Transcription181200-155_2_ORF2Q96300GRF714-3-3-LIKE PROTEIN GF14 NU32.812191.11 × 10−161
Transcription287872-75_1_ORF1Q9C5W6GRF1214-3-3-LIKE PROTEIN GF14 IOTA32.812193.83 × 10−151
Translation209284-130_2_ORF2Q9FNR1RBG3GLYCINE-RICH RNA-BINDING PROTEIN 319.815241.45 × 10−31
Translation293356-72_1_ORF1Q9LR72PCMP-E3PUTATIVE PENTATRICOPEPTIDE REPEAT-CONTAINING PROTEIN AT1G03510 (POLIPASA)26.15110.4
Translation26795-487_6_ORF2Q0WW84RBP47BPOLYADENYLATE-BINDING PROTEIN RBP47B45.92013.27 × 10−136
Translation174433-162_1_ORF1Q9ASR1LOS1ELONGATION FACTOR 273.891120
Folding26640-489_1_ORF2Q9M1C2CPN10-110 kDa CHAPERONIN 119.415269.48 × 10−39
Folding189606-147_1_ORF2Q9SR70FKBP16-4PEPTIDYL-PROLYL CIS-TRANS ISOMERASE FKBP16-426.413361.35 × 10−89
Folding2524-1285_6_ORF2Q9SKQ0CYP19-2PEPTIDYL-PROLYL CIS-TRANS ISOMERASE CYP19-221.6275242.24 × 10−90
Sorting19573-562_5_ORF2Q9SYI0SECA1PROTEIN TRANSLOCASE SUBUNIT SECA1115.72120
Sorting146969-201_2_ORF1F4JL11IMPA2IMPORTIN SUBUNIT ALPHA-259.15026 × 10−61
Sorting151836-193_1_ORF2P40941AAC2ADP, ATP CARRIER PROTEIN 242.37150
Sorting161087-178_2_ORF2Q8H0U5TIC62PROTEIN TIC 6273.46465.77 × 10−115
Sorting82340-277_1_ORF2Q39196PIP1.4PROBABLE AQUAPORIN PIP1-4336233.75 × 10−170
Sorting272341-85_2_ORF2Q94A40AT1G62020COATOMER SUBUNIT ALPHA-11371010
Sorting29489-466_3_ORF1Q0WW26AT4G34450COATOMER SUBUNIT GAMMA1032110
Sorting43675-385_1_ORF2Q67YI9EPSIN2CLATHRIN INTERACTOR EPSIN 285.21111.51 × 10−117
Sorting68824-304_5_ORF2Q9LQ55DRP2BDYNAMIN-2B105.31110
Degradation141778-205_4_ORF2Q8L770CLPR3ATP-DEPENDENT CLP PROTEASE PROTEOLYTIC SUBUNIT-RELATED PROTEIN 338.92111.73 × 10−136
Degradation172993-163_5_ORF1Q9XJ36CLPR2ATP-DEPENDENT CLP PROTEASE PROTEOLYTIC SUBUNIT-RELATED PROTEIN 232.74114.49 × 10−126
Degradation170504-166_2_ORF2P30184LAP1LEUCINE AMINOPEPTIDASE 162.54140
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ojosnegros, S.; Alvarez, J.M.; Grossmann, J.; Gagliardini, V.; Quintanilla, L.G.; Grossniklaus, U.; Fernández, H. Proteome and Interactome Linked to Metabolism, Genetic Information Processing, and Abiotic Stress in Gametophytes of Two Woodferns. Int. J. Mol. Sci. 2023, 24, 12429. https://doi.org/10.3390/ijms241512429

AMA Style

Ojosnegros S, Alvarez JM, Grossmann J, Gagliardini V, Quintanilla LG, Grossniklaus U, Fernández H. Proteome and Interactome Linked to Metabolism, Genetic Information Processing, and Abiotic Stress in Gametophytes of Two Woodferns. International Journal of Molecular Sciences. 2023; 24(15):12429. https://doi.org/10.3390/ijms241512429

Chicago/Turabian Style

Ojosnegros, Sara, José Manuel Alvarez, Jonas Grossmann, Valeria Gagliardini, Luis G. Quintanilla, Ueli Grossniklaus, and Helena Fernández. 2023. "Proteome and Interactome Linked to Metabolism, Genetic Information Processing, and Abiotic Stress in Gametophytes of Two Woodferns" International Journal of Molecular Sciences 24, no. 15: 12429. https://doi.org/10.3390/ijms241512429

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop