Next Article in Journal
Exploring Metabolomic Patterns in Type 2 Diabetes Mellitus and Response to Glucose-Lowering Medications—Review
Previous Article in Journal
Nutrient Metabolism Pathways Analysis and Key Candidate Genes Identification Corresponding to Cadmium Stress in Buckwheat through Multiomics Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring the Potential Role of Ribosomal Proteins to Enhance Potato Resilience in the Face of Changing Climatic Conditions

by
Eliana Valencia-Lozano
1,†,
Lisset Herrera-Isidrón
2,†,
Jorge Abraham Flores-López
2,
Osiel Salvador Recoder-Meléndez
2,
Braulio Uribe-López
2,
Aarón Barraza
3 and
José Luis Cabrera-Ponce
1,*
1
Departamento de Ingeniería Genética, Centro de Investigación y de Estudios Avanzados del IPN, Unidad Irapuato, Irapuato 36824, Guanajuato, Mexico
2
Unidad Profesional Interdisciplinaria de Ingeniería Campus Guanajuato (UPIIG), Instituto Politécnico Nacional, Av. Mineral de Valenciana 200, Puerto Interior, Silao de la Victoria 36275, Guanajuato, Mexico
3
CONACYT-Centro de Investigaciones Biológicas del Noreste, SC., Instituto Politécnico Nacional 195, Playa Palo de Santa Rita Sur, La Paz CP 23096, Baja California Sur, Mexico
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Genes 2023, 14(7), 1463; https://doi.org/10.3390/genes14071463
Submission received: 9 June 2023 / Revised: 5 July 2023 / Accepted: 14 July 2023 / Published: 18 July 2023
(This article belongs to the Section Plant Genetics and Genomics)

Abstract

:
Potatoes have emerged as a key non-grain crop for food security worldwide. However, the looming threat of climate change poses significant risks to this vital food source, particularly through the projected reduction in crop yields under warmer temperatures. To mitigate potential crises, the development of potato varieties through genome editing holds great promise. In this study, we performed a comprehensive transcriptomic analysis to investigate microtuber development and identified several differentially expressed genes, with a particular focus on ribosomal proteins—RPL11, RPL29, RPL40 and RPL17. Our results reveal, by protein–protein interaction (PPI) network analyses, performed with the highest confidence in the STRING database platform (v11.5), the critical involvement of these ribosomal proteins in microtuber development, and highlighted their interaction with PEBP family members as potential microtuber activators. The elucidation of the molecular biological mechanisms governing ribosomal proteins will help improve the resilience of potato crops in the face of today’s changing climatic conditions.

1. Introduction

Potato (Solanum tuberosum L.) is a staple food that produces tubers with high nutritional value and energy. It is the fourth most important crop and its yield will be affected by extreme drought, heat, and the relocation of potato diseases and pests. Food security depends on the development of strategies to overcome the effects of climate variability on the agriculture system [1]. Potatoes are propagated vegetatively from harvested tubers and certified seeds from suppliers. An understanding of the tuberization process is needed to boost resilience in the face of changing climatic conditions.
The molecular mechanisms involved in potato tuber development under greenhouse conditions have been widely studied for review; see [2]. An alternative technology is the in vitro induction of potato tubers called microtubers (MTs). MTs induction occurs when axillary buds are cultured in a medium containing high sucrose content, plant growth regulators, and different light quality/darkness. MTs production has advantages in terms of storage, transport, and cultivation due to their reduced size and weight, requiring less space than seedlings, with higher multiplication rates producing seed potatoes faster and cheaper than other methods [3,4,5].
In our previous work, we developed a protocol for MTs induction in potato S. tuberosum var α using stolons cultured in MS medium supplemented with 8% sucrose, gelrite 6.0 g/L, 2iP 10 mg/L under darkness conditions. The rationale for this protocol was that CK signaling interacts with homeobox transcription factors, RPs, cell cycle, carbon metabolism, auxin-responsive factors and stem cell genes in the microtuberization process [6].
In our following work, we reported a transcriptome analysis of the MTs revealing that 1715 up-regulated and 1624 down-regulated genes were involved in this biological process. The protein–protein interaction (PPI) network analyses, performed with the highest confidence in the STRING database platform (v11.5, www.string-db.org accessed on 15 September 2022), showed that 299 genes were tightly associated in 14 clusters including a main group of essential life proteins.
Ribosomal proteins (RPL11, RPL29, RPL40, RPL17) interact with several gene clusters, like the tuberigen, proteasome, immunophilins and oxidative stress. RPL11 interacts with thioredoxins, thylakoids, fatty acid biosynthesis, one-carbon metabolism, carbon metabolism, flowering (six PEBP family members), and the cell cycle [7].
Through a yeast two-hybrid screening approach, the interaction of PEBP-StSP6A (a positive regulator of tuber development) [8,9,10] with RPs and others involved in protein synthesis, RNA and DNA binding proteins, histones, initiation factors, signaling and carbon metabolism has been demonstrated [11,12].
Several transcriptomic analyses of potato tuberization have revealed the presence of RPs during the process [13,14,15,16,17]. From the RNA-seq database published by Sharma and Hannapel 2016 [13], we performed a PPI network analysis with the highest confidence (0.900), revealing that RPs interact with SP6A, RP40SA, RP60S, RPS8, RPS4A, RPL10, BEL5, and RPL14 [13].
In this manuscript, we will discuss the feasibility of activating RPs to induce MTs with several trait advantages such as size, number, stress tolerance, and protein content. We will also discuss the function and potential use of several genes interacting with RPs derived from the transcriptome analysis performed by Valencia-Lozano et al. in 2022 [7].

2. Materials and Methods

2.1. Potato MTs Induction

Potato MTs induction was induced by culture of stolon explants, about 3 cm in length, in flasks containing medium: MS medium, supplemented with 2iP 10 mg/L, 8% sucrose, 6 g/L gelrite, activated charcoal (0.3%), and pH 5.8 identified as MR8-G6-2iP. In the control medium, MS medium was supplemented with 1% sucrose, 3 g/L gelrite and 10 mg/L 2iP identified as MR1-G3-2iP. Flasks were sealed with plastic and incubated in the dark at 25/17 °C for 15 days [6].

2.2. RNA Isolation and qPCR Analysis

Total RNA derived from 15-day-old explants was isolated according to the methods of Valencia-Lozano et al. (2021) [7]. To validate the genome-wide analysis of MTs, oligonucleotides were designed (Table 1), and the expressions of EF1 and SEC3 were used as reference genes for calculating the relative amount of target gene expression using the 2−ΔΔCT method [18,19]. qPCR analysis was based on at least five biological replicates for each sample with three technical replicates.
The sequencing of cDNA derived from potato MTs and controls was undertaken in GENEWIZ, Plainfield, NJ, USA. To sequence, the Illumina HiSeq 4000 (Illumina, San Diego, CA, USA) was applied. The quality of sequence reads was assessed by the software package FastQC (http://www.bioinformatics.babraham.ac.uk/projects/fastqc/ accessed on 2 June 2023) and to remove sequence adapters and low-quality bases we used the software Trimmomatics [20].

2.3. Transcriptome and Interaction Analysis of Proteins Involved in Microuberization

RNA-seq reads aligned to the potato S. tuberosum reference genome were made according to the methods set out by Valencia-Lozano et al. (2022) [7]. The quantification and differential analysis of the transcripts was performed using the DESeq2 v1.12.4 program. Finally, an ontology analysis was performed using Blast2GO. The PPI analysis of microtuberization was performed with the STRING database v11.5 [21], based on S. tuberosum and the highest confidence (0.860).

3. Results

3.1. Transcriptome Analysis of MTs Induction

The transcriptome analysis of the MTs induction process was analyzed according to Valencia-Lozano et al. (2022) [7]. The study revealed 1699 up-regulated genes, and when analyzed in a PPI network with the highest confidence, 299 were tightly associated with two fundamental biological processes essential for life and highly conserved through organisms: RPs comprising 29 proteins and CC containing 117 proteins (Figure 1).
RPs interact with proteins that sense the environment: six PEBP family members, twenty-one involved in osmotic stress, nine related to oxidative stress, twenty three involved in CK response, six related to one-carbon metabolism, thirty eight associated with carbon metabolism, sixteen with TCA cycle, six with acyl carrier proteins, fourteen involved in fatty acid metabolism, thirteen in thylakoid, nine redoxins, eight involved in sulfur metabolism, five involved in disulfide isomerase activity and twelve related to immunophilins (Figure 1).

3.2. DEG in MTs Development Involved in Immunophilins, Redoxins, Oxidative Stress, Carbon Metabolism and One-Carbon Metabolism

In the PPI network, 11 genes were involved in immunophilins (Figure 2), 8 in redoxins (Figure 3), 8 in oxidative stress (Figure 4), 38 in carbon metabolism (Figure 5) and 6 in one-carbon metabolism (Figure 6). This cluster interacts with the cell cycle cluster through the interaction of the dihydrofolate reductase gene and RPs.

3.3. Validation of the Transcriptome-Wide Analysis

Validation in the transcriptomic-wide analysis was achieved by selecting the following genes: StSP6A (PGSC0003DMT400060057), RPL11 (PGSC0003DMT400031869), RPL29 (PGSC0003DMT400069470), RPL40 (PGSC0003DMT400047686) and RPL17 (PGSC0003DMT400060127). For DEG validation, EFa1 and SEC3 were used as endogen controls. The results of the validation indicate that the values are consistent with those obtained in the transcriptomic-wide analysis (Figure 7).

4. Discussion

4.1. What Is the Importance of Ribosome Biogenesis?

Ribosome biogenesis is tightly associated with plant growth, development, and reproduction. Genetic mutations related to ribosomal proteins (RPs) or ribosome biogenesis factors (RBFs) result in retarded growth, delayed flowering, and in more severe cases, they are lethal. In total, 19 ribosomal proteins resulted in loss of function and 26 ribosome biogenesis factors are seedling/embryo-lethal (Table 2).

Some RPs Are Lethal Mutants

The small-subunit RPs exhibit lethal effects; for example, some are seedling-lethal, such as RPS1 and RPS20 in rice [22,29], and RPS18A in tobacco [28], and some are embryo-lethal, such as RPS9 and RPS17 in maize [24,27], and RPS16 and RPS27 in Arabidopsis [26,31] (Table 2).
Large-subunits RPs, such as RPL13, RPL12, RPL13 and RPL21C, are seedling-/embryo-lethal in rice [36,37,77,78] (Table 2). In contrast, RPL5C, RPL9C,D, RPL10, RPL21C, and RPL28-1 are embryo-lethal in Arabidopsis [32,33,34,35,79] (Table 2). It has been shown that at least 29 ribosome biogenesis are lethal, including proteins involved in chloroplast development, ribosome biogenesis, nucleolar organization, and chlorophyll biosynthesis (Table 2).
In our transcriptome analysis, five RPLs (RPL1, 12, 13, 27 and 35), and four RPSs (RPS1, 9, 16 and 17) are embryo-lethal.

4.2. Overexpression of Ribosomal Proteins and Ribosome Biogenesis Factors

Horvath et al. (2006) [80] demonstrated the highest expression of EBP1 in the developing organs and its correlation with genes involved in ribosome biogenesis in potatoes. The EBP1 gene regulates the intermediate and late steps of rRNA processing. Silenced potato lines showed reduced size and tuber yield, and an abnormal morphology.
Transformed potato plants with the RP StoL13a from Solanum torvum, SW, a highly resistant plant to Verticillium dahliae infection, were more resistant to V. dahliae infection than the control plants. The transgenic plants showed lower levels of reactive oxygen species and attenuated oxidative damage. In addition, six defense and antioxidant enzyme genes were up-regulated in the StoL13a ectopic expression plants. These results suggest that StoL13a plays a role in plant defense against V. dahliae infection [81].
The overexpression of RPL6 in rice resulted in salt tolerance [82]. Transgenic rice plants overexpressing RPL23A showed resistance to water use efficiency, and suitable growth and yield parameters, compared to the negative control [83].
OLI2/NOP2A encodes a nucleolar methyltransferase required to mature the 25S ribosomal RNA of the 60S large ribosomal subunit. These seeds were lighter and heavier than wild-type seeds produced by oli2 mutant and OLI2 overexpressor plants respectively. The seeds from the oli2 mutant showed delayed germination, while OLI2 overexpressor lines germinated earlier than the wild type. The wild type had a greater number and longer length of lateral roots than the oli2 mutant. The lateral root development phenotype of the oli2 mutant resembles auxin-related mutants, but was not enhanced by exogenously supplied auxin. Furthermore, high concentrations of sugar induced hypersensitivity and less sensitivity in oli2 mutant and OLI2 overexpressor lines, respectively [84].
STCH4/REIL2 encodes a ribosomal biogenesis factor up-regulated during cold stress. When STCH4 is overexpressed, it confers chilling and freezing tolerance to Arabidopsis, although its mutation reduces CBF protein levels, resulting in the delayed induction of C-repeat-binding factor (CBF) regulon genes [85].

4.3. Are RPs Good Candidate Genes for Improving of Multiple Abiotic Stress Tolerance in Potato?

Drought stress affects potato plants at all stages of the crop’s growth, from seedling emergence to tuber initiation, and bulking, ultimately resulting in a reduction in tuber yield. Prolonged water scarcity induces several physiological disorders in potato tubers, such as tuber cracking, tuber malformation, hollow heart, vascular discoloration, and reduction in dry matter accumulation.
Kappachery et al. (2013) [86] identified potential drought tolerance genes in potato using a yeast functional screening method. A cDNA expression library was constructed from hyperosmotic stressed potato plants, and identified yeast transformants expressing different cDNAs for survival under hyperosmotic stress conditions.
Sixty-nine genes were identified to grow under drought, salt, and heat stress. Of these, eight were RPs (RPS7, RPL12, RPL10, RPL27, RPS11, RPL18a, RPL1 and RPL36) [86]. The potential of RPs to obtain potato MTs lines with desirable traits is very promising.

4.4. Interaction of RPs Cluster with Immunophilins

In the PPI network, 11 genes affected immunophilins. RPL29 (PGSC0003DMT400069847) interacts with peptidyl-prolyl cis-trans isomerase FKBP12. This group of proteins includes peptidyl-prolyl cis-trans-isomerases (PPIs) (immunophilins) and protein disulfide isomerases (Figure 8).
The cluster consists of four PPI genes interacting with SOS3 (PGSC0003DMT400023568), a calcium sensor calcineurin B essential for the transduction of the salt stress-induced Ca2+ signal and salt tolerance in Arabidopsis. A loss-of-function mutation that reduces the Ca2+ binding capacity of SOS3 (sos3-1) renders the mutant hypersensitive to salt [87].
FKBP12, overexpressed in Arabidopsis, has been reported to be directly involved in abiotic stress responses and in promoting growth under normal conditions [88].
The overexpression in potatoes of cyclophilin CYP21-4 involved in oxidative stress formed longer plants and heavier tubers, and when microtubers were induced, they yielded more in a shorter time [89].
Three genes with disulfide isomerase-like proteins are present in this cluster—STPDI1, 2 and 3. They interact with calnexin (PGSC0003DMT400036920). StPDI1, accumulated upon salt exposure, catalyzes the formation of disulfide bonds and confers tertiary and quaternary structures to the proteins. Thus, they may act as chaperones during salt stress, as previously described for HSPs [90].
The silencing of StPDI1 expression affects the tolerance of abiotic stress in transgenic potato plants. The amount of malate, succinate and 2-oxoglutarate, and the content of the reducing equivalent NADH, were decreased significantly in StPDI1-inhibited potato plants. In contrast, amino acids such as serine and threonine were upregulated compared to wild-type plants [91].

4.5. Redoxins

To minimize the adverse effects of reactive oxygen species (ROS), aerobic organisms have evolved defense systems, catalases (CAT), superoxide dismutase (SOD), ascorbate peroxidase (APX), glutathione peroxidase (GPX) and GST, as well as the capacity for the production of low-molecular weight antioxidants such as ascorbic acid (AA) and glutathione (GSH).
In the PPI network, 9 genes affected redoxins. RPs interact with the thioredoxin cluster through the interaction of the PRXQ peroxiredoxin (PGSC0003DMT400035271) with RPL11 (PGSC0003DMT400031869) and PRL18 (PGSC0003DMT400011931). PRXQ plays a role in protecting cells from oxidative stress by detoxifying peroxides. Also, this protein is involved in photosystem II’s protection against hydrogen peroxide.
Transgenic potato plants expressing 2-cysteine peroxiredoxin exhibit enhanced tolerance to environmental stresses, including MV-induced oxidative stress and high temperature, with plants under the control of the SWPA2 promoter exhibiting the best tolerance [92].

4.6. Response to Oxidative Stress

Within a cell, the superoxide dismutases (SODs) form the primary defense against ROS. In the PPI network, nine genes were involved in oxidative stress. Reactive O2 species (ROS) are produced in cells under both unstressed and stressed conditions.
The superoxidase dismutase 1 gene is essential to the potato’s response to low temperatures. The overexpression of StSOD1 increases low-temperature tolerance in potatoes, while interference expression decreases low-temperature tolerance in potatoes [93]. The overexpression of a cytosolic copper-zinc superoxide dismutase, from Potentilla atrosanguinea (PaSOD), in potato (S. tuberosum ssp. tuberosum L. cv. Kufri Sutlej) enhanced net photosynthetic rates (PN) and stomatal conductance (gs) compared to those in the wild-type (WT) plants under control (irrigated) as well as drought stress conditions [94]. Transgenic potato plants overexpressing SOD and ascorbate peroxidase (APX) promoted the enhancement of lignification and starch biosynthesis, and improved salt tolerance [95]. The overexpression in potato plants of SOD, APX and the bacterial choline oxidase (oda) led to enhanced protection against various abiotic stresses [96].

4.7. RPs Interacting with Carbon Metabolism

Carbon metabolism (CM) transforms carbon into energy at different amounts through glycolysis, gluconeogenesis, the pentose phosphate pathway, carbon fixation pathways, and the TCA pathway. In carbon metabolism, 38 genes were implicated in the PPI network. Through the interaction of the DHFR and RPs clusters, the CC cluster interacts with this cluster (Figure 9).
Watkinson et al. (2006) [97] assessed three accession genotypes of drought-stressed S. tuberosum ssp. andigena and made a transcriptomic analysis of genes associated with carbon metabolism, citrate cycle and oxidative stress. In agreement with their results, we found similar genes compared with “intermediate” genotypes, with 20 genes up-regulated under stress conditions, similar to our analysis. The term “intermediate” means plants that recovered their photosynthetic index after one cycle of stress and performed even better in the second, yielding similar results to the control plants unaffected by the stress. This may explain why molecular mechanisms under field conditions are very similar to those under in vitro conditions.
This includes PGK, phosphoglycerate kinase, cytosolic; TPI, triosephosphate isomerase, cytosolic; IAR4, pyruvate dehydrogenase e1 component subunit α-3, GAPC2, glyceraldehyde 3-phosphate dehydrogenase; LOS2, enolase and PKP3, pyruvate kinase. However, they showed that genes like Hexokinase-1 and Fructose-bisphosphate aldolase were down-regulated. In contrast, pyruvate dehydrogenase e1 component subunit β-3 did not show significant changes in expression, and these also did not occur when we induced stress. Regarding the TCA cycle, we found, according to Watkinson et al. (2006) [97], that the up-regulated genes are MMDH1 (malate dehydrogenase) and MDH (NAD-malate dehydrogenase). The overexpression of these genes plays a crucial role in different plant breeding programs. The phosphoglycerate kinase gene promotes biomass and yields in tobacco under salt stress conditions [98], while in rice, it improves thermotolerance [99]. Similarly, the overexpression of the pyruvate dehydrogenase gene is involved in grain size and weight in rice [100], and drought stress in barley [101] and rice [102]. Moreover, the overexpression of pyruvate kinase negatively affects root growth in maize [103], while when it is silenced in rice, this results in sucrose translocation defects, the inhibition of grain filling [104], and reduced grain starch content [105]. Additionally, the triosephosphate isomerase gene enhances photosynthesis under elevated CO2 levels in rice [99], makes pigeon peas more resilient to salt stress [106], and improves drought stress tolerance in both rice [107] and maize [108].
Salt stress tolerance is found to be influenced by enolase in Mesembryanthemum crystallinum L. Similarly, the salt stress response is tied to the fructose-bisphosphate aldose gene in Ulva compressa [109] and mango [110], as well as biomass accumulation in tobacco [111]. The overexpression of the acetyl-CoA carboxylase gene leads to an increased lipid content in microalgae, including Dunaliella sp. [112], Chlamydomonas reinhardtii [113], and Scenedesmus sp. [114], and increased seed yield in tobacco [115].
The overexpression of the sucrose phosphate synthase gene enhances growth, thermotolerance [116], sink strength in tomatoes [117], potato yield characteristics [118], and cold tolerance in chrysanthemum [119]. It also positively impacts biomass production in sugarcane [120] and in tomatoes under saturated light and CO2 conditions [121], and enhances foliar sucrose/starch levels in Arabidopsis [122]. Similarly, the ATP-citrate synthase gene plays a role in the salt stress response in Halogeton glomeratus [123] and sugar beet [124].
The overexpression of the glyceraldehyde 3-phosphate dehydrogenase gene enhances drought tolerance in potato [125] and salt tolerance in soybean [126], rice [127], and potato [128]. The 6-phosphogluconate dehydrogenase overexpression gene contributes to resistance against Nilaparvata lugens in rice [129], as well as salt tolerance in barley [130], and is also involved in starch accumulation in maize [131]. Additionally, malate dehydrogenase gene overexpression increases the production of organic acids and aluminum tolerance in alfalfa [132], as well as salt tolerance in rice [133], apple, and tomato, and also impacts cold tolerance [134,135]. However, the gene has been shown to be embryo-lethal in Arabidopsis [136].
Lastly, the overexpression of the phosphoenolpyruvate carboxylase gene increases photosynthetic efficiency in rice [137], fatty acid production in Nicotiana tabacum [138] and protein content in Vicia narbonensis [139], and affects dark and light respiration in potato [140].
In cerium stress-treated microalgae, Nannocloropsis oculata led to an increase in lipid content with the carbon metabolism and ribosome biogenesis genes prominently activated [141].

One-Carbon Metabolism

One-carbon metabolism is a critical metabolic process in which multiple enzymatic reactions provide methyl groups (one carbon) for nucleotide metabolism, purine and pyrimidine synthesis, and amino acid metabolism. These effects involve many cellular activities, such as cell growth, differentiation, and development. One-carbon metabolism transfers a carbon unit from serine or glycine to tetrahydrofolate (THF) to form methylene-THF for DNA synthesis. In the PPI network, seven genes were found to be involved in one-carbon metabolism.
The overexpression of the S-adenosylmethionine synthetase gene (SHM4 and SHM1) enhances cold and salt tolerance in tobacco [142,143], lipid production in Chlamydomonas [144], salt, H2O2 and drought tolerance in Arabidopsis [145,146], and alkali tolerance in tomato [147]. The overexpression of the adenosylhomocysteinase gene (HOG1) results in early flowering and reduced biomass in Arabidopsis [148], and it increases lycopene and reduces ripening time in tomatoes [149]. The serine hydroxymethyltransferase gene (MAT3), when overexpressed in rice, confers salt tolerance [150], increases root growth and sugar levels, and decreases H2O2 levels in Arabidopsis [151], while also limiting cold tolerance [152] and antioxidant ability in rice [153].
Lastly, the bifunctional dihydrofolate reductase-thymidylate synthase (DHFR-TS/THY-1) can regulate folate abundance in Arabidopsis [154], which is essential for the biosynthesis of nucleotide precursors of DNA [155] and also involved somatic embryogenesis in carrot [156].

5. Conclusions

The gene modulation of ribosomal proteins by genome editing can increase the genetic variability to enhance potato resistance to biotic and abiotic stress, and also increase nutritional value.
The genes involved in carbon metabolism—PGK, TPI, IAR4, GAPC2, LOS2, and PKP3—in both plant field experiments and under our conditions have the potential to improve biomass and yield under stress conditions.
The cluster of immunophilins and disulfide isomerase interacting with RPs will allow for the activation of alternative mechanisms of survival enhancement under adverse conditions.
The gene modulation of one-carbon metabolism pathways will favor survival in adverse environments.

Author Contributions

E.V.-L. and L.H.-I.: Conceptualization, methodology, validation, formal analysis, investigation, data curation, writing original draft preparation, review and editing, funding acquisition; J.A.F.-L., O.S.R.-M. and B.U.-L.: methodology, software, validation, investigation; A.B.: software, data curation, formal analysis, supervision, writing, review and editing; J.L.C.-P.: conceptualization, methodology, validation, formal analysis, investigation, data curation, writing original draft preparation, review and editing, funding acquisition, writing review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded, in part, by IPN (project SIP 20222061, proyectos de desarrollo tecnológico o inovación para alumnos del IPN # 120).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

BioProject: PRJNA898400 https://www.ncbi.nlm.nih.gov/bioproject/PRJNA898400 accessed on 2 June 2023.

Acknowledgments

We are very grateful to Diana Marcela Rivera Toro and Alex Ricardo Bermudez-Valle for the academic discussion during the writing of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. EIT Food. About EIT Food; EIT Food: Leuven, Belgium, 2019. [Google Scholar]
  2. Dutt, S.; Manjul, A.S.; Raigond, P.; Singh, B.; Siddappa, S.; Bhardwaj, V.; Kawar, P.G.; Patil, V.U.; Kardile, H.B. Key players associated with tuberization in potato: Potential candidates for genetic engineering. Crit. Rev. Biotechnol. 2017, 37, 942–957. [Google Scholar] [CrossRef] [PubMed]
  3. Donnelly, D.J.; Coleman, W.K.; Coleman, S.E. Potato microtuber production and performance: A review. Am. J. Potato Res. 2003, 80, 103–115. [Google Scholar] [CrossRef]
  4. Hannapel, D.J. Signalling the induction of tuber formation. In Potato Biology and Biotechnology; Elsevier Science BV: Amsterdam, The Netherlands, 2007; pp. 237–256. [Google Scholar]
  5. Vinterhalter, D.; Dragicevic, I.; Vinterhalter, B. Potato in vitro culture techniques and biotechnology. Fruit Veg. Cereal Sci. Biotech. 2008, 2, 16–45. [Google Scholar]
  6. Herrera-Isidron, L.; Valencia-Lozano, E.; Rosiles-Loeza, P.Y.; Robles-Hernández, M.G.; Napsuciale-Heredia, A.; Cabrera-Ponce, J. Gene expression analysis of microtubers of potato Solanum tuberosum L. induced in cytokinin containing medium and osmotic stress. Plants 2021, 10, 876. [Google Scholar] [CrossRef] [PubMed]
  7. Valencia-Lozano, E.; Herrera-Isidrón, L.; Flores-López, J.A.; Recoder-Meléndez, O.S.; Barraza, A.; Cabrera-Ponce, J.L. Solanum tuberosum Microtuber Development under Darkness Unveiled through RNAseq Transcriptomic Analysis. Int. J. Mol. Sci. 2022, 23, 13835. [Google Scholar] [CrossRef]
  8. Abelenda, J.A.; Navarro, C.; Prat, S. From the model to the crop: Genes controlling tuber formation in potato. Curr. Opin. Biotechnol. 2011, 22, 287–292. [Google Scholar] [CrossRef]
  9. Abelenda, J.A.; Bergonzi, S.; Oortwijn, M.; Sonnewald, S.; Du, M.; Visser, R.G.; Sonnewald, U.; Bachem, C.W. Source-sink regulation is mediated by interaction of an FT homolog with a SWEET protein in potato. Curr. Biol. 2019, 29, 1178–1186. [Google Scholar] [CrossRef] [Green Version]
  10. Teo, C.J.; Takahashi, K.; Shimizu, K.; Shimamoto, K.; Taoka, K.I. Potato tuber induction is regulated by interactions between components of a tuberigen complex. Plant Cell Physiol. 2017, 58, 365–374. [Google Scholar] [CrossRef]
  11. Purwestri, Y.A.; Susanto, F.A.; Tsuji, H. Hd3a florigen recruits different proteins to reveal its function in plant growth and development. In Plant Engineering; IntechOpen: London, UK, 2017. [Google Scholar]
  12. Wang, E.; Liu, T.; Sun, X.; Jing, S.; Zhou, T.; Liu, T.; Song, B. Profiling of the Candidate Interacting Proteins of SELF-PRUNING 6A (SP6A) in Solanum tuberosum. Int. J. Mol. Sci. 2022, 23, 9126. [Google Scholar] [CrossRef]
  13. Sharma, P.; Lin, T.; Hannapel, D.J. Targets of the StBEL5 transcription factor include the FT ortholog StSP6A. Plant Physiol. 2016, 170, 310–324. [Google Scholar] [CrossRef] [Green Version]
  14. Salvato, F.; Havelund, J.F.; Chen, M.; Rao, R.S.P.; Rogowska-Wrzesinska, A.; Jensen, O.N.; Gang, D.R.; Thelen, J.J.; Møller, I.M. The potato tuber mitochondrial proteome. Plant Physiol. 2014, 164, 637–653. [Google Scholar] [CrossRef] [Green Version]
  15. van Dijk, J.P.; Cankar, K.; Scheffer, S.J.; Beenen, H.G.; Shepherd, L.V.; Stewart, D.; Davies, H.V.; Wilkockson, S.J.; Leifert, C.; Gruden, K.; et al. Transcriptome Analysis of Potato Tubers Effects of Different Agricultural Practices. J. Agric. Food Chem. 2009, 57, 1612–1623. [Google Scholar] [CrossRef]
  16. Shan, J.; Song, W.; Zhou, J.; Wang, X.; Xie, C.; Gao, X.; Xie, T.; Liu, J. Transcriptome analysis reveals novel genes potentially involved in photoperiodic tuberization in potato. Genomics 2013, 102, 388–396. [Google Scholar] [CrossRef]
  17. Vulavala, V.K.; Fogelman, E.; Faigenboim, A.; Shoseyov, O.; Ginzberg, I. The transcriptome of potato tuber phellogen reveals cellular functions of cork cambium and genes involved in periderm formation and maturation. Sci. Rep. 2019, 9, 10216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Livak, K.J.; Schmittgen, T.D. Analysis of relative gene expression usingreal-time quantitative PCR and the 2−ΔΔCT method. Methods 2001, 25, 402–408. [Google Scholar] [CrossRef] [PubMed]
  19. Bustin, S.A.; Beaulieu, J.F.; Huggett, J.; Jaggi, R.; Kibenge, F.S.; Olsvik, P.A.; Penning, L.C.; Toegel, S. MIQE precis: Practical implementation of minimum standard guidelines for fluorescence-based quantitative real-time PCR exp. BMC Mol. Biol. 2010, 11, 74–80. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Szklarczyk, D.; Franceschini, A.; Wyder, S.; Forslund, K.; Heller, D.; Huerta-Cepas, J.; Von Mering, C. STRING v10: Protein–protein interaction networks, integrated over the tree of life. Nucleic Acids Res. 2015, 43, D447–D452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Zhou, K.; Zhang, C.; Xia, J.; Yun, P.; Wang, Y.; Ma, T.; Li, Z. Albino seedling lethality 4; Chloroplast 30S ribosomal protein S1 is required for chloroplast ribosome biogenesis and early chloroplast development in rice. Rice 2021, 14, 47. [Google Scholar] [CrossRef]
  23. Weijers, D.; Franke-van Dijk, M.; Vencken, R.J.; Quint, A.; Hooykaas, P.; Offringa, R. An Arabidopsis Minute-like phenotype caused by a semi-dominant mutation in a RIBOSOMAL PROTEIN S5 gene. Development 2001, 128, 4289–4299. [Google Scholar] [CrossRef]
  24. Ma, Z.; Dooner, H.K. A mutation in the nuclear-encoded plastid ribosomal protein S9 leads to early embryo lethality in maize. Plant J. 2004, 37, 92–103. [Google Scholar] [CrossRef] [PubMed]
  25. Ito, T.; Kim, G.T.; Shinozaki, K. Disruption of an Arabidopsis cytoplasmic ribosomal protein S13-homologous gene by transposon-mediated mutagenesis causes aberrant growth and development. Plant J. 2000, 22, 257–264. [Google Scholar] [CrossRef] [PubMed]
  26. Tsugeki, R.; Kochieva, E.Z.; Fedoroff, N.V. A transposon insertion in the Arabidopsis SSR16 gene causes an embryo-defective lethal mutation. Plant J. 1996, 10, 479–489. [Google Scholar] [CrossRef] [PubMed]
  27. Schultes, N.P.; Sawers, R.J.; Brutnell, T.P.; Krueger, R.W. Maize high chlorophyll fluorescent 60 mutation is caused by an Ac disruption of the gene encoding the chloroplast ribosomal small subunit protein 17. Plant J. 2000, 21, 317–327. [Google Scholar] [CrossRef] [PubMed]
  28. Rogalski, M.; Ruf, S.; Bock, R. Tobacco plastid ribosomal protein S18 is essential for cell survival. Nucleic Acids Res. 2006, 34, 4537–4545. [Google Scholar] [CrossRef] [Green Version]
  29. Gong, X.; Jiang, Q.; Xu, J.; Zhang, J.; Teng, S.; Lin, D.; Dong, Y. Disruption of the rice plastid ribosomal protein S20 leads to chloroplast developmental defects and seedling lethality. G3 Genes Genom. Genet. 2013, 3, 1769–1777. [Google Scholar] [CrossRef] [Green Version]
  30. Dong, X.; Duan, S.; Wang, H.B.; Jin, H.L. Plastid ribosomal protein LPE2 is involved in photosynthesis and the response to C/N balance in Arabidopsis thaliana. J. Integr. Plant Biol. 2020, 62, 1418–1432. [Google Scholar] [CrossRef] [Green Version]
  31. Revenkova, E.; Masson, J.; Koncz, C.; Afsar, K.; Jakovleva, L.; Paszkowski, J. Involvement of Arabidopsis thaliana ribosomal protein S27 in mRNA degradation triggered by genotoxic stress. EMBO J. 1999, 18, 490–499. [Google Scholar] [CrossRef] [Green Version]
  32. Dupouy, G.; McDermott, E.; Cashell, R.; Scian, A.; McHale, M.; Ryder, P.; Groot, J.; Lucca, N.; Brychkova, G.; McKeown, P.C.; et al. Plastid ribosome protein L5 is essential for post-globular embryo development in Arabidopsis thaliana. Plant Reprod. 2022, 35, 189–204. [Google Scholar] [CrossRef]
  33. Devis, D.; Firth, S.M.; Liang, Z.; Byrne, M.E. Dosage sensitivity of RPL9 and concerted evolution of ribosomal protein genes in plants. Front. Plant Sci. 2015, 6, 1102. [Google Scholar] [CrossRef] [Green Version]
  34. Falcone Ferreyra, M.L.; Biarc, J.; Burlingame, A.L.; Casati, P. Arabidopsis L10 ribosomal proteins in UV-B responses. Plant Signal. Behav. 2010, 5, 1222–1225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Pesaresi, P.; Varotto, C.; Meurer, J.; Jahns, P.; Salamini, F.; Leister, D. Knock-out of the plastid ribosomal protein L11 in Arabidopsis: Effects on mRNA translation and photosynthesis. Plant J. 2001, 27, 179–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Zhao, D.S.; Zhang, C.Q.; Li, Q.F.; Yang, Q.Q.; Gu, M.H.; Liu, Q.Q. A residue substitution in the plastid ribosomal protein L12/AL1 produces defective plastid ribosome and causes early seedling lethality in rice. Plant Mol. Biol. 2016, 91, 161–177. [Google Scholar] [CrossRef]
  37. Lee, J.; Jang, S.; Ryu, S.; Lee, S.; Park, J.; Lee, S.; An, G.; Park, S.K. Mutation of plastid ribosomal protein L13 results in an albino seedling-lethal phenotype in rice. Plant Breed. Biotechnol. 2019, 7, 395–404. [Google Scholar] [CrossRef]
  38. Bobik, K.; Fernandez, J.C.; Hardin, S.R.; Ernest, B.; Ganusova, E.E.; Staton, M.E.; Burch-Smith, T.M. The essential chloroplast ribosomal protein uL 15c interacts with the chloroplast RNA helicase ISE 2 and affects intercellular trafficking through plasmodesmata. New Phytol. 2019, 221, 850–865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Yin, T.; Liu, W.; Zhao, Y.; Wang, X.; Wang, K.; Shen, Y.; Ding, Y.; Tang, S. Effects of high temperature on rice grain development and quality formation based on proteomics comparative analysis under field warming. Front. Plant Sci. 2021, 12, 746180. [Google Scholar]
  40. Degenhardt, R.F.; Bonham-Smith, P.C. Arabidopsis ribosomal proteins RPL23aA and RPL23aB are differentially targeted to the nucleolus and are disparately required for normal development. Plant Physiol. 2008, 147, 128–142. [Google Scholar] [CrossRef] [Green Version]
  41. Zhou, F.; Roy, B.; Von Arnim, A.G. Translation reinitiation and development are compromised in similar ways by mutations in translation initiation factor eIF3h and the ribosomal protein RPL24. BMC Plant Biol. 2010, 10, 193. [Google Scholar] [CrossRef] [Green Version]
  42. Romani, I.; Tadini, L.; Rossi, F.; Masiero, S.; Pribil, M.; Jahns, P.; Kater, M.; Leister, D.; Pesaresi, P. Versatile roles of Arabidopsis plastid ribosomal proteins in plant growth and development. Plant J. 2012, 72, 922–934. [Google Scholar] [CrossRef]
  43. Magnard, J.L.; Heckel, T.; Massonneau, A.; Wisniewski, J.P.; Cordelier, S.; Lassagne, H.; Perez, P.; Dumas, C.; Rogowsky, P.M. Morphogenesis of maize embryos requires ZmPRPL35-1 encoding a plastid ribosomal protein. Plant Physiol. 2004, 134, 649–663. [Google Scholar] [CrossRef] [Green Version]
  44. Uwer, U.; Willmitzer, L.; Altmann, T. Inactivation of a glycyl-tRNA synthetase leads to an arrest in plant embryo development. Plant Cell 1998, 10, 1277–1294. [Google Scholar] [CrossRef] [Green Version]
  45. Zhu, R.M.; Chai, S.; Zhang, Z.Z.; Ma, C.L.; Zhang, Y.; Li, S. Arabidopsis Chloroplast protein for Growth and Fertility1 (CGF1) and CGF2 are essential for chloroplast development and female gametogenesis. BMC Plant Biol. 2020, 20, 172. [Google Scholar] [CrossRef] [Green Version]
  46. Bellaoui, M.; Keddie, J.S.; Gruissem, W. DCL is a plant-specific protein required for plastid ribosomal RNA processing and embryo development. Plant Mol. Biol. 2003, 53, 531–543. [Google Scholar] [CrossRef]
  47. Hanania, U.; Velcheva, M.; Or, E.; Flaishman, M.; Sahar, N.; Perl, A. Silencing of chaperonin 21; that was differentially expressed in inflorescence of seedless and seeded grapes; promoted seed abortion in tobacco and tomato fruits. Transgenic Res. 2007, 16, 515–525. [Google Scholar] [CrossRef]
  48. Weis, B.L.; Kovacevic, J.; Missbach, S.; Schleiff, E. Plant-specific features of ribosome biogenesis. Trends Plant Sci. 2015, 20, 729–740. [Google Scholar] [CrossRef] [PubMed]
  49. Durut, N.; Sáez-Vásquez, J. Nucleolin: Dual roles in rDNA chromatin transcription. Gene 2015, 556, 7–12. [Google Scholar] [CrossRef]
  50. Chen, S.; Blank, M.F.; Iyer, A.; Huang, B.; Wang, L.; Grummt, I.; Voit, R. SIRT7-dependent deacetylation of the U3-55k protein controls pre-rRNA processing. Nat. Commun. 2016, 7, 10734. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Chen, L.; Zhang, B.B.; Cheung, P.C. Comparative proteomic analysis of mushroom cell wall proteins among the different developmental stages of Pleurotus tuber-regium. J. Agric. Food Chem. 2012, 60, 6173–6182. [Google Scholar] [CrossRef] [PubMed]
  52. Ohtani, M.; Demura, T.; Sugiyama, M. Arabidopsis root initiation defective1; a DEAH-box RNA helicase involved in pre-mRNA splicing; is essential for plant development. Plant Cell 2013, 25, 2056–2069. [Google Scholar] [CrossRef] [Green Version]
  53. Strittmatter, P.; Soll, J.; Bölter, B. The chloroplast protein import machinery: A review. Methods Mol. Biol. 2010, 619, 307–321. [Google Scholar]
  54. Yu, B.; Wakao, S.; Fan, J.; Benning, C. Loss of plastidic lysophosphatidic acid acyltransferase causes embryo-lethality in Arabidopsis. Plant Cell Physiol. 2004, 45, 503–510. [Google Scholar] [CrossRef] [PubMed]
  55. Kovacheva, S.; Bédard, J.; Patel, R.; Dudley, P.; Twell, D.; Ríos, G.; Koncz, C.; Jarvis, P. In vivo studies on the roles of Tic110; Tic40 and Hsp93 during chloroplast protein import. Plant J. 2005, 41, 412–428. [Google Scholar] [CrossRef] [Green Version]
  56. Bang, W.Y.; Jeong, I.S.; Kim, D.W.; Im, C.H.; Ji, C.; Hwang, S.M.; Kim, S.W.; Son, Y.S.; Jeong, J.; Shiina, T.; et al. Role of Arabidopsis CHL27 protein for photosynthesis; chloroplast development and gene expression profiling. Plant Cell Physiol. 2008, 49, 1350–1363. [Google Scholar] [CrossRef]
  57. Chi, W.; Ma, J.; Zhang, D.; Guo, J.; Chen, F.; Lu, C.; Zhang, L. The pentratricopeptide repeat protein DELAYED GREENING1 is involved in the regulation of early chloroplast development and chloroplast gene expression in Arabidopsis. Plant Physiol. 2008, 147, 573–584. [Google Scholar] [CrossRef] [Green Version]
  58. Patel, R.; Hsu, S.C.; Bédard, J.; Inoue, K.; Jarvis, P. The Omp85-related chloroplast outer envelope protein OEP80 is essential for viability in Arabidopsis. Plant Physiol. 2008, 148, 235–245. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Garcion, C.; Guilleminot, J.; Kroj, T.; Parcy, F.; Giraudat, J.; Devic, M. AKRP and EMB506 are two ankyrin repeat proteins essential for plastid differentiation and plant development in Arabidopsis. Plant J. 2006, 48, 895–906. [Google Scholar] [CrossRef]
  60. Dong, H.; Deng, Y.; Mu, J.; Lu, Q.; Wang, Y.; Xu, Y.; Chu, C.; Chong, K.; Lu, C.; Zuo, J. The Arabidopsis Spontaneous Cell Death1 gene; encoding a ζ-carotene desaturase essential for carotenoid biosynthesis; is involved in chloroplast development; photoprotection and retrograde signalling. Cell Res. 2007, 17, 458–470. [Google Scholar] [CrossRef] [Green Version]
  61. Qian, W.; Yu, C.; Qin, H.; Liu, X.; Zhang, A.; Johansen, I.E.; Wang, D. Molecular and functional analysis of phosphomannomutase (PMM) from higher plants and genetic evidence for the involvement of PMM in ascorbic acid biosynthesis in Arabidopsis and Nicotiana benthamiana. Plant J. 2007, 49, 399–413. [Google Scholar] [CrossRef]
  62. Huang, X.; Zhang, X.; Yang, S. A novel chloroplast-localized protein EMB1303 is required for chloroplast development in Arabidopsis. Cell Res. 2009, 19, 1205–1216. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Liang, Q.; Lu, X.; Jiang, L.; Wang, C.; Fan, Y.; Zhang, C. EMB1211 is required for normal embryo development and influences chloroplast biogenesis in Arabidopsis. Physiol. Plant. 2010, 140, 380–394. [Google Scholar] [CrossRef] [PubMed]
  64. Komatsu, T.; Kawaide, H.; Saito, C.; Yamagami, A.; Shimada, S.; Nakazawa, M.; Matsui, M.; Nakano, A.; Tsujimoto, M.; Natsume, M.; et al. The chloroplast protein BPG2 functions in brassinosteroid-mediated post-transcriptional accumulation of chloroplast rRNA. Plant J. 2010, 61, 409–422. [Google Scholar] [CrossRef]
  65. Bryant, N.; Lloyd, J.; Sweeney, C.; Myouga, F.; Meinke, D. Identification of nuclear genes encoding chloroplast-localized proteins required for embryo development in Arabidopsis. Plant Physiol. 2011, 155, 1678–1689. [Google Scholar] [CrossRef] [Green Version]
  66. Palm, D.; Streit, D.; Shanmugam, T.; Weis, B.L.; Ruprecht, M.; Simm, S.; Schleiff, E. Plant-specific ribosome biogenesis factors in Arabidopsis thaliana with essential function in rRNA processing. Nucleic Acids Res. 2019, 47, 1880–1895. [Google Scholar] [CrossRef] [Green Version]
  67. Asakura, Y.; Galarneau, E.; Watkins, K.P.; Barkan, A.; van Wijk, K.J. Chloroplast RH3 DEAD box RNA helicases in maize and Arabidopsis function in splicing of specific group II introns and affect chloroplast ribosome biogenesis. Plant Physiol. 2012, 159, 961–974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Inoue, H.; Li, M.; Schnell, D.J. An essential role for chloroplast heat shock protein 90 (Hsp90C) in protein import into chloroplasts. Proc. Natl. Acad. Sci. USA 2013, 110, 3173–3178. [Google Scholar] [CrossRef]
  69. Lu, X.; Zhang, D.; Li, S.; Su, Y.; Liang, Q.; Meng, H.; Shen, S.; Fan, Y.; Liu, C.; Zhang, C. FtsHi4 is essential for embryogenesis due to its influence on chloroplast development in Arabidopsis. PLoS ONE 2014, 9, e99741. [Google Scholar] [CrossRef] [Green Version]
  70. Chen, H.; Zou, W.; Zhao, J. Ribonuclease J is required for chloroplast and embryo development in Arabidopsis. J. Exp. Bot. 2015, 66, 2079–2091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Jeon, Y.; Ahn, C.S.; Jung, H.J.; Kang, H.; Park, G.T.; Choi, Y.; Hwang, J.; Pai, H.S. DER containing two consecutive GTP-binding domains plays an essential role in chloroplast ribosomal RNA processing and ribosome biogenesis in higher plants. J. Exp. Bot. 2014, 65, 117–130. [Google Scholar] [CrossRef] [Green Version]
  72. Missbach, S.; Weis, B.L.; Martin, R.; Simm, S.; Bohnsack, M.T.; Schleiff, E. 40S ribosome biogenesis co-factors are essential for gametophyte and embryo development. PLoS ONE 2013, 8, e54084. [Google Scholar] [CrossRef] [Green Version]
  73. Liu, H.; Xiu, Z.; Yang, H.; Ma, Z.; Yang, D.; Wang, H.; Tan, B.C. Maize Shrek1 encodes a WD40 protein that regulates pre-rRNA processing in ribosome biogenesis. Plant Cell 2022, 34, 4028–4044. [Google Scholar] [CrossRef] [PubMed]
  74. Burgess, A.L.; David, R.; Searle, I.R. Conservation of tRNA and rRNA 5-methylcytosine in the kingdom Plantae. BMC Plant Biol. 2015, 15, 199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Chi, W.; He, B.; Mao, J.; Li, Q.; Ma, J.; Ji, D.; Zou, M.; Zhang, L. The function of RH22; a DEAD RNA helicase; in the biogenesis of the 50S ribosomal subunits of Arabidopsis chloroplasts. Plant Physiol. 2012, 158, 693–707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Li, P.C.; Li, K.; Wang, J.; Zhao, C.Z.; Zhao, S.Z.; Hou, L.; Xia, H.; Ma, C.L.; Wang, X.J. The AAA-ATPase MIDASIN 1 functions in ribosome biogenesis and is essential for embryo and root development. Plant Physiol. 2019, 180, 289–304. [Google Scholar] [CrossRef]
  77. Lee, J.; Jang, S.; Ryu, S.; Lee, S.; Park, J.; Lee, S.; An, G.; Park, S.K. Impaired plastid ribosomal protein L3 causes albino seedling lethal phenotype in rice. J. Plant Biol. 2019, 62, 419–428. [Google Scholar] [CrossRef]
  78. Lin, D.; Jiang, Q.; Zheng, K.; Chen, S.; Zhou, H.; Gong, X.; Xu, J.; Teng, S.; Dong, Y. Mutation of the rice ASL 2 gene encoding plastid ribosomal protein L21 causes chloroplast developmental defects and seedling death. Plant Biol. 2015, 17, 599–607. [Google Scholar] [CrossRef]
  79. Yin, J.; Ibrahim, S.; Petersen, F.; Yu, X. Autoimmunomic signatures of aging and age-related neurodegenerative diseases are associated with brain function and ribosomal proteins. Front. Aging Neurosci. 2021, 13, 679688. [Google Scholar] [CrossRef]
  80. Horvath, B.M.; Magyar, Z.; Zhang, Y.; Hamburger, A.W.; Bako, L.; Visser, R.G.; Bachem, C.W.; Bögre, L. EBP1 regulates organ size through cell growth and proliferation in plants. EMBO J. 2006, 25, 4909–4920. [Google Scholar] [CrossRef]
  81. Yang, L.; Xie, C.; Li, W.; Zhang, R.; Jue, D.; Yang, Q. Expression of a wild eggplant ribosomal protein L13a in potato enhances resistance to Verticillium dahliae. Plant Cell Tissue Organ Cult. 2013, 115, 329–340. [Google Scholar] [CrossRef]
  82. Moin, M.; Saha, A.; Bakshi, A.; Madhav, M.S.; Kirti, P.B. Constitutive expression of ribosomal protein L6 modulates salt tolerance in rice transgenic plants. Gene 2021, 789, 145670. [Google Scholar] [CrossRef] [PubMed]
  83. Moin, M.; Bakshi, A.; Madhav, M.S.; Kirti, P.B. Expression profiling of ribosomal protein gene family in dehydration stress responses and characterization of transgenic rice plants overexpressing RPL23A for water-use efficiency and tolerance to drought and salt stresses. Front. Chem. 2017, 5, 97. [Google Scholar] [CrossRef] [Green Version]
  84. Maekawa, S.; Yanagisawa, S. Ribosome biogenesis factor OLI2 and its interactor BRX1-2 are associated with morphogenesis and lifespan extension in Arabidopsis thaliana. Plant Biotechnol. 2021, 38, 117–125. [Google Scholar] [CrossRef]
  85. Yu, H.; Kong, X.; Huang, H.; Wu, W.; Park, J.; Yun, D.J.; Lee, B.H.; Shi, H.; Zhu, J.K. STCH4/REIL2 confers cold stress tolerance in Arabidopsis by promoting rRNA processing and CBF protein translation. Cell Rep. 2020, 30, 229–242. [Google Scholar] [CrossRef] [Green Version]
  86. Kappachery, S.; Yu, J.W.; Baniekal-Hiremath, G.; Park, S.W. Rapid identification of potential drought tolerance genes from Solanum tuberosum by using a yeast functional screening method. Comptes Rendus Biol. 2013, 336, 530–545. [Google Scholar] [CrossRef]
  87. Mahajan, S.; Pandey, G.K.; Tuteja, N. Calcium-and salt-stress signaling in plants: Shedding light on SOS pathway. Arch. Biochem. Biophys. 2008, 471, 146–158. [Google Scholar] [CrossRef] [PubMed]
  88. Alavilli, H.; Lee, H.; Park, M.; Yun, D.J.; Lee, B.H. Enhanced multiple stress tolerance in Arabidopsis by overexpression of the polar moss peptidyl prolyl isomerase FKBP12 gene. Plant Cell Rep. 2018, 37, 453–465. [Google Scholar] [CrossRef] [PubMed]
  89. Park, H.J.; Lee, A.; Lee, S.S.; An, D.J.; Moon, K.B.; Ahn, J.C.; Kim, H.S.; Cho, H.S. Overexpression of golgi protein CYP21-4s improves crop productivity in potato and rice by increasing the abundance of mannosidic glycoproteins. Front. Plant Sci. 2017, 8, 1250. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Andème Ondzighi, C.; Christopher, D.A.; Cho, E.J.; Chang, S.C.; Staehelin, L.A. Arabidopsis protein disulfide isomerase-5 inhibits cysteine proteases during trafficking to vacuoles before programmed cell death of the endothelium in developing seeds. Plant Cell 2008, 20, 2205–2220. [Google Scholar] [CrossRef] [Green Version]
  91. Eggert, E.; Obata, T.; Gerstenberger, A.; Gier, K.; Brandt, T.; Fernie, A.R.; Walttraud, S.; Kühn, C. A sucrose transporter-interacting protein disulphide isomerase affects redox homeostasis and links sucrose partitioning with abiotic stress tolerance. Plant Cell Environ. 2016, 39, 1366–1380. [Google Scholar] [CrossRef] [Green Version]
  92. Kim, M.D.; Kim, Y.H.; Kwon, S.Y.; Jang, B.Y.; Lee, S.Y.; Yun, D.J.; Cho, J.H.; Kwak, S.S.; Lee, H.S. Overexpression of 2-cysteine peroxiredoxin enhances tolerance to methyl viologen-mediated oxidative stress and high temperature in potato plants. Plant Physiol. Biochem. 2011, 49, 891–897. [Google Scholar] [CrossRef]
  93. Che, Y.; Zhang, N.; Zhu, X.; Li, S.; Wang, S.; Si, H. Enhanced tolerance of the transgenic potato plants overexpressing Cu/Zn superoxide dismutase to low temperature. Sci. Hortic. 2020, 261, 108949. [Google Scholar] [CrossRef]
  94. Pal, A.K.; Acharya, K.; Vats, S.K.; Kumar, S.; Ahuja, P.S. Over-expression of PaSOD in transgenic potato enhances photosynthetic performance under drought. Biol. Plant. 2013, 57, 359–364. [Google Scholar] [CrossRef]
  95. Shafi, A.; Pal, A.K.; Sharma, V.; Kalia, S.; Kumar, S.; Ahuja, P.S.; Singh, A.K. Transgenic potato plants overexpressing SOD and APX exhibit enhanced lignification and starch biosynthesis with improved salt stress tolerance. Plant Mol. Biol. Rep. 2017, 35, 504–518. [Google Scholar] [CrossRef]
  96. Ahmad, R.; Kim, Y.H.; Kim, M.D.; Kwon, S.Y.; Cho, K.; Lee, H.S.; Kwak, S.S. Simultaneous expression of choline oxidase; superoxide dismutase and ascorbate peroxidase in potato plant chloroplasts provides synergistically enhanced protection against various abiotic stresses. Physiol. Plant. 2010, 138, 520–533. [Google Scholar] [CrossRef] [PubMed]
  97. Watkinson, J.I.; Hendricks, L.; Sioson, A.A.; Vasquez-Robinet, C.; Stromberg, V.; Heath, L.S.; Schuler, M.; Bohnert, H.J.; Bonierbale, M.; Grene, R. Accessions of Solanum tuberosum ssp. andigena show differences in photosynthetic recovery after drought stress as reflected in gene expression profiles. Plant Sci. 2006, 171, 745–758. [Google Scholar]
  98. Joshi, R.; Karan, R.; Singla-Pareek, S.L.; Pareek, A. Ectopic expression of Pokkali phosphoglycerate kinase-2 (OsPGK2-P) improves yield in tobacco plants under salinity stress. Plant Cell Rep. 2016, 35, 27–41. [Google Scholar] [CrossRef] [PubMed]
  99. Suzuki, Y.; Konno, Y.; Takegahara-Tamakawa, Y.; Miyake, C.; Makino, A. Effects of suppression of chloroplast phosphoglycerate kinase on photosynthesis in rice. Photosynth. Res. 2022, 153, 83–91. [Google Scholar] [CrossRef]
  100. Lei, J.; Teng, X.; Wang, Y.; Jiang, X.; Zhao, H.; Zheng, X.; Ren, Y.; Dong, H.; Wang, Y.; Duan, E.; et al. Plastidic pyruvate dehydrogenase complex E1 component subunit Alpha1 is involved in galactolipid biosynthesis required for amyloplast development in rice. Plant Biotechnol. J. 2022, 20, 437–453. [Google Scholar] [CrossRef]
  101. Guo, P.; Baum, M.; Grando, S.; Ceccarelli, S.; Bai, G.; Li, R.; Korff, M.V.; Varshney, R.K.; Graner, A.; Valkoun, J. Differentially expressed genes between drought-tolerant and drought-sensitive barley genotypes in response to drought stress during the reproductive stage. J. Exp. Bot. 2009, 60, 3531–3544. [Google Scholar] [CrossRef]
  102. Huang, S.; Xin, S.; Xie, G.; Han, J.; Liu, Z.; Wang, B.; Zhang, S.; Wu, Q.; Cheng, X. Mutagenesis reveals that the rice OsMPT3 gene is an important osmotic regulatory factor. Crop. J. 2020, 8, 465–479. [Google Scholar] [CrossRef]
  103. He, K.; Zhao, Z.; Ren, W.; Chen, Z.; Chen, L.; Chen, F.; Mi, G.; Pan, Q.; Yuan, L. Mining genes regulating root system architecture in maize based on data integration analysis. Theor. Appl. Genet. 2023, 136, 127. [Google Scholar] [CrossRef]
  104. Hu, L.; Tu, B.; Yang, W.; Yuan, H.; Li, J.; Guo, L.; Zheng, L.; Chen, W.; Zhu, X.; Wang, Y.; et al. Mitochondria-associated pyruvate kinase complexes regulate grain filling in rice. Plant Physiol. 2020, 183, 1073–1087. [Google Scholar] [CrossRef]
  105. Cai, Y.; Li, S.; Jiao, G.; Sheng, Z.; Wu, Y.; Shao, G.; Xie, L.; Peng, C.; Xu, J.; Tang, S.; et al. Os PK 2 encodes a plastidic pyruvate kinase involved in rice endosperm starch synthesis; compound granule formation and grain filling. Plant Biotechnol. J. 2018, 16, 1878–1891. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Awana, M.; Jain, N.; Samota, M.K.; Rani, K.; Kumar, A.; Ray, M.; Gaikwad, K.; Praveen, S.; Singh, N.K.; Singh, A. Protein and gene integration analysis through proteome and transcriptome brings new insight into salt stress tolerance in pigeonpea (Cajanus cajan L.). Int. J. Biol. Macromol. 2020, 164, 3589–3602. [Google Scholar] [CrossRef]
  107. Salekdeh, G.H.; Siopongco, J.; Wade, L.J.; Ghareyazie, B.; Bennett, J. Proteomic analysis of rice leaves during drought stress and recovery. Proteomics 2002, 2, 1131–1145. [Google Scholar] [CrossRef]
  108. Riccardi, F.; Gazeau, P.; de Vienne, D.; Zivy, M. Protein changes in response to progressive water deficit in maize: Quantitative variation and polypeptide identification. Plant Physiol. 1998, 117, 1253–1263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Xing, Q.; Bi, G.; Cao, M.; Belcour, A.; Aite, M.; Mo, Z.; Mao, Y. Comparative transcriptome analysis provides insights into response of Ulva compressa to fluctuating salinity conditions. J. Phycol. 2021, 57, 1295–1308. [Google Scholar] [CrossRef]
  110. Perveen, N.; Dinesh, M.R.; Sankaran, M.; Ravishankar, K.V.; Krishnajee, H.G.; Hanur, V.S.; Alamri, S.; Kesawat, M.S.; Irfan, M. Comparative transcriptome analysis provides novel insights into molecular response of salt-tolerant and sensitive polyembryonic mango genotypes to salinity stress at seedling stage. Front. Plant Sci. 2023, 14, 1152485. [Google Scholar] [CrossRef]
  111. Uematsu, K.; Suzuki, N.; Iwamae, T.; Inui, M.; Yukawa, H. Increased fructose 1; 6-bisphosphate aldolase in plastids enhances growth and photosynthesis of tobacco plants. J. Exp. Bot. 2012, 63, 3001–3009. [Google Scholar] [CrossRef]
  112. Talebi, A.F.; Tohidfar, M.; Bagheri, A.; Lyon, S.R.; Salehi-Ashtiani, K.; Tabatabaei, M. Manipulation of carbon flux into fatty acid biosynthesis pathway in Dunaliella salina using AccD and ME genes to enhance lipid content and to improve produced biodiesel quality. Biofuel Res. J. 2014, 1, 91–97. [Google Scholar] [CrossRef]
  113. Chen, D.; Yuan, X.; Liang, L.; Liu, K.; Ye, H.; Liu, Z.; Liu, Y.; Huang, L.; He, W.; Chen, Y.; et al. Overexpression of acetyl-CoA carboxylase increases fatty acid production in the green alga Chlamydomonas reinhardtii. Biotechnol. Lett. 2019, 41, 1133–1145. [Google Scholar] [CrossRef] [PubMed]
  114. Ma, C.; Ren, H.; Xing, D.; Xie, G.; Ren, N.; Liu, B. Mechanistic understanding towards the effective lipid production of a microalgal mutant strain Scenedesmus sp. Z-4 by the whole genome bioinformation. J. Hazard. Mat. 2019, 375, 115–120. [Google Scholar] [CrossRef] [PubMed]
  115. Madoka, Y.; Tomizawa, K.I.; Mizoi, J.; Nishida, I.; Nagano, Y.; Sasaki, Y. Chloroplast transformation with modified accD operon increases acetyl-CoA carboxylase and causes extension of leaf longevity and increase in seed yield in tobacco. Plant Cell Physiol. 2002, 43, 1518–1525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Zhang, Y.; Zeng, D.; Liu, Y.; Zhu, W. SlSPS; a sucrose phosphate synthase gene; mediates plant growth and thermotolerance in tomato. Horticulturae 2022, 8, 491. [Google Scholar] [CrossRef]
  117. Nguyen-Quoc, B.; N’Tchobo, H.; Foyer, C.H.; Yelle, S. Overexpression of sucrose phosphate synthase increases sucrose unloading in transformed tomato fruit. J. Exp. Bot. 1999, 50, 785–791. [Google Scholar] [CrossRef] [Green Version]
  118. Ishimaru, K.; Hirotsu, N.; Kashiwagi, T.; Madoka, Y.; Nagasuga, K.; Ono, K.; Ohsugi, R. Overexpression of a maize SPS gene improves yield characters of potato under field conditions. Plant Prod. Sci. 2008, 11, 104–107. [Google Scholar] [CrossRef]
  119. Wang, K.; Bai, Z.Y.; Liang, Q.Y.; Liu, Q.L.; Zhang, L.; Pan, Y.Z.; Liu, G.L.; Jiang, B.B.; Zhang, F.; Jia, Y. Transcriptome analysis of chrysanthemum (Dendranthema grandiflorum) in response to low temperature stress. BMC Genom. 2018, 19, 319. [Google Scholar] [CrossRef] [Green Version]
  120. Anur, R.M.; Mufithah, N.; Sawitri, W.D.; Sakakibara, H.; Sugiharto, B. Overexpression of sucrose phosphate synthase enhanced sucrose content and biomass production in transgenic sugarcane. Plants 2020, 9, 200. [Google Scholar] [CrossRef] [Green Version]
  121. Galtier, N.; Foyer, C.H.; Murchie, E.; Aired, R.; Quick, P.; Voelker, T.A.; Thepenier, C.; Lasceve, G.; Betsche, T. Effects of light and atmospheric carbon dioxide enrichment on photosynthesis and carbon partitioning in the leaves of tomato (Lycopersicon esculentum L.) plants over-expressing sucrose phosphate synthase. J. Exp. Bot. 1995, 46, 1335–1344. [Google Scholar] [CrossRef]
  122. Signora, L.; Galtier, N.; Skøt, L.; Lucas, H.; Foyer, C.H. Over-expression of sucrose phosphate synthase in Arabidopsis thaliana results in increased foliar sucrose/starch ratios and favours decreased foliar carbohydrate accumulation in plants after prolonged growth with CO2 enrichment. J. Exp. Bot. 1998, 49, 669–680. [Google Scholar] [CrossRef] [Green Version]
  123. Wang, J.; Li, B.; Meng, Y.; Ma, X.; Lai, Y.; Si, E.; Yang, K.; Ren, P.; Shang, X.; Wang, H. Transcriptomic profiling of the salt-stress response in the halophyte Halogeton glomeratus. BMC Genom. 2015, 16, 169. [Google Scholar] [CrossRef] [Green Version]
  124. Wang, Y.; Stevanato, P.; Lv, C.; Li, R.; Geng, G. Comparative physiological and proteomic analysis of two sugar beet genotypes with contrasting salt tolerance. J. Agric. Food Chem. 2019, 67, 6056–6073. [Google Scholar] [CrossRef] [PubMed]
  125. Kappachery, S.; Baniekal-Hiremath, G.; Yu, J.W.; Park, S.W. Effect of over-and under-expression of glyceraldehyde 3-phosphate dehydrogenase on tolerance of plants to water-deficit stress. Plant Cell Tissue Organ Cult. 2015, 121, 97–107. [Google Scholar] [CrossRef]
  126. Zhao, X.; Hong, H.; Wang, J.; Zhan, Y.; Teng, W.; Li, H.; Li, W.; Li, Y.; Zhao, X.; Han, Y. Genome-wide identification and analysis of glyceraldehyde-3-phosphate dehydrogenase family reveals the role of GmGAPDH14 to improve salt tolerance in soybean (Glycine max L.). Front. Plant Sci. 2023, 14, 1193044. [Google Scholar] [CrossRef] [PubMed]
  127. Lim, H.; Hwang, H.; Kim, T.; Kim, S.; Chung, H.; Lee, D.; Kim, S.; Park, S.; Cho, W.; Ji, H.; et al. Transcriptomic analysis of rice plants overexpressing PsGAPDH in response to salinity stress. Genes 2021, 12, 641. [Google Scholar] [CrossRef]
  128. Jeong, M.J.; Park, S.C.; Byun, M.O. Improvement of salt tolerance in transgenic potato plants by glyceraldehyde-3 phosphate dehydrogenase gene transfer. Mol. Cells 2001, 12, 185–189. [Google Scholar] [CrossRef]
  129. Chen, L.; Kuai, P.; Ye, M.; Zhou, S.; Lu, J.; Lou, Y. Overexpression of a cytosolic 6-phosphogluconate dehydrogenase gene enhances the resistance of rice to Nilaparvata lugens. Plants 2020, 9, 1529. [Google Scholar] [CrossRef]
  130. Witzel, K.; Weidner, A.; Surabhi, G.K.; Varshney, R.K.; Kunze, G.; Buck-Sorlin, G.H.; Börner, A.; Mock, H.P. Comparative analysis of the grain proteome fraction in barley genotypes with contrasting salinity tolerance during germination. Plant Cell Environ. 2010, 33, 211–222. [Google Scholar] [CrossRef] [Green Version]
  131. Spielbauer, G.; Li, L.; Römisch-Margl, L.; Do, P.T.; Fouquet, R.; Fernie, A.R.; Eisenreich, W.; Gierl, A.; Settles, A.M. Chloroplast-localized 6-phosphogluconate dehydrogenase is critical for maize endosperm starch accumulation. J. Exp. Bot. 2013, 64, 2231–2242. [Google Scholar] [CrossRef] [Green Version]
  132. Tesfaye, M.; Temple, S.J.; Allan, D.L.; Vance, C.P.; Samac, D.A. Overexpression of malate dehydrogenase in transgenic alfalfa enhances organic acid synthesis and confers tolerance to aluminum. Plant Physiol. 2001, 127, 1836–1844. [Google Scholar] [CrossRef]
  133. Zhang, Y.; Wang, Y.; Sun, X.; Yuan, J.; Zhao, Z.; Gao, J.; Wen, X.; Tang, F.; Kang, M.; Abliz, B.; et al. Genome-Wide Identification of MDH Family Genes and Their Association with Salt Tolerance in Rice. Plants 2022, 11, 1498. [Google Scholar] [CrossRef]
  134. Yao, Y.X.; Dong, Q.L.; Zhai, H.; You, C.X.; Hao, Y.J. The functions of an apple cytosolic malate dehydrogenase gene in growth and tolerance to cold and salt stresses. Plant Physiol. Biochem. 2011, 49, 257–264. [Google Scholar] [CrossRef] [PubMed]
  135. Wang, Q.J.; Sun, H.; Dong, Q.L.; Sun, T.Y.; Jin, Z.X.; Hao, Y.J.; Yao, Y.X. The enhancement of tolerance to salt and cold stresses by modifying the redox state and salicylic acid content via the cytosolic malate dehydrogenase gene in transgenic apple plants. Plant Biotechnol. 2016, 14, 1986–1997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Beeler, S.; Liu, H.C.; Stadler, M.; Schreier, T.; Eicke, S.; Lue, W.L.; Truernit, E.; Zeeman, S.C.; Chen, J.; Kötting, O. Plastidial NAD-dependent malate dehydrogenase is critical for embryo development and heterotrophic metabolism in Arabidopsis. Plant Physiol. 2014, 164, 1175–1190. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Behera, D.; Swain, A.; Karmakar, S.; Dash, M.; Swain, P.; Baig, M.J.; Molla, K.A. Overexpression of Setaria italica phosphoenolpyruvate carboxylase gene in rice positively impacts photosynthesis and agronomic traits. Plant Physiol. Biochem. 2023, 194, 169–181. [Google Scholar] [CrossRef]
  138. Fan, Z.; Li, J.; Lu, M.; Li, X.; Yin, H. Overexpression of phosphoenolpyruvate carboxylase from Jatropha curcas increases fatty acid accumulation in Nicotiana tabacum. Acta Physiol. Plant. 2013, 35, 2269–2279. [Google Scholar] [CrossRef]
  139. Rolletschek, H.; Borisjuk, L.; Radchuk, R.; Miranda, M.; Heim, U.; Wobus, U.; Weber, H. Seed-specific expression of a bacterial phosphoenolpyruvate carboxylase in Vicia narbonensis increases protein content and improves carbon economy. Plant Biotechnol. J. 2004, 2, 211–219. [Google Scholar] [CrossRef]
  140. Häusler, R.E.; Kleines, M.; Uhrig, H.; Hirsch, H.J.; Smets, H. Overexpression of phosphoenol pyruvate carboxylase from Corynebacterium glutamicum lowers the CO2 compensation point (Γ*) and enhances dark and light respiration in transgenic potato. J. Exp. Bot. 1999, 50, 1231–1242. [Google Scholar] [CrossRef] [Green Version]
  141. Wu, D.; Hou, Y.; Cheng, J.; Han, T.; Hao, N.; Zhang, B.; Fan, X.; Ji, X.; Chen, F.; Gong, D.; et al. Transcriptome analysis of lipid metabolism in response to cerium stress in the oleaginous microalga Nannochloropsis oculata. Sci. Total Environ. 2022, 838, 156420. [Google Scholar] [CrossRef]
  142. Seong, E.S.; Jeon, M.R.; Choi, J.H.; Yoo, J.H.; Lee, J.G.; Na, J.K.; Kim, N.Y.; Yu, C.Y. Overexpression of S-adenosylmethionine synthetase enhances tolerance to cold stress in tobacco. Russ. J. Plant Physiol. 2020, 67, 242–249. [Google Scholar] [CrossRef]
  143. Zhu, H.; He, M.; Jahan, M.S.; Wu, J.; Gu, Q.; Shu, S.; Sun, J.; Guo, S. CsCDPK6; a CsSAMS1-interacting protein; affects polyamine/ethylene biosynthesis in Cucumber and enhances salt tolerance by overexpression in tobacco. Int. J. Mol. Sci. 2021, 22, 11133. [Google Scholar] [CrossRef]
  144. Kim, J.H.; Ahn, J.W.; Park, E.J.; Choi, J.I. Overexpression of S-Adenosylmethionine Synthetase in Recombinant Chlamydomonas for Enhanced Lipid Production. J. Microbiol. Biotechnol. 2023, 33, 310. [Google Scholar] [CrossRef] [PubMed]
  145. Ma, C.; Wang, Y.; Gu, D.; Nan, J.; Chen, S.; Li, H. Overexpression of S-adenosyl-L-methionine synthetase 2 from sugar beet M14 increased Arabidopsis tolerance to salt and oxidative stress. Int. J. Mol. Sci. 2017, 18, 847. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Kim, E.Y.; Park, K.Y.; Seo, Y.S.; Kim, W.T. Arabidopsis small rubber particle protein homolog SRPs play dual roles as positive factors for tissue growth and development and in drought stress responses. Plant Physiol. 2016, 170, 2494–2510. [Google Scholar] [CrossRef] [Green Version]
  147. Gong, B.; Li, X.; VandenLangenberg, K.M.; Wen, D.; Sun, S.; Wei, M.; Li, Y.; Yang, F.; Shi, Q.; Wang, X. Overexpression of S-adenosyl-l-methionine synthetase increased tomato tolerance to alkali stress through polyamine metabolism. Plant Biotechnol. J. 2014, 12, 694–708. [Google Scholar] [CrossRef] [PubMed]
  148. Godge, M.R.; Kumar, D.; Kumar, P.P. Arabidopsis HOG1 gene and its petunia homolog PETCBP act as key regulators of yield parameters. Plant Cell Rep. 2008, 27, 1497–1507. [Google Scholar] [CrossRef]
  149. Yang, Y.; Zhu, G.; Li, R.; Yan, S.; Fu, D.; Zhu, B.; Tian, H.; Luo, Y.; Zhu, H. The RNA editing factor SlORRM4 is required for normal fruit ripening in tomato. Plant Physiol. 2017, 175, 1690–1702. [Google Scholar] [CrossRef] [Green Version]
  150. Mishra, P.; Bhoomika, K.; Dubey, R.S. Differential responses of antioxidative defense system to prolonged salinity stress in salt-tolerant and salt-sensitive Indica rice (Oryza sativa L.) seedlings. Protoplasma 2013, 250, 3–19. [Google Scholar] [CrossRef]
  151. Yuan, Y.; Xu, D.; Xiang, D.; Jiang, L.; Hu, H. Serine Hydroxymethyltransferase 1 Is Essential for Primary-Root Growth at Low-Sucrose Conditions. Int. J. Mol. Sci. 2022, 23, 4540. [Google Scholar] [CrossRef]
  152. Fang, C.; Zhang, P.; Li, L.; Yang, L.; Mu, D.; Yan, X.; Li, Z.; Lin, W. Serine hydroxymethyltransferase localised in the endoplasmic reticulum plays a role in scavenging H 2 O 2 to enhance rice chilling tolerance. BMC Plant Biol. 2020, 20, 1–13. [Google Scholar] [CrossRef]
  153. Wang, S.; Li, X.; Zhu, J.; Liu, H.; Liu, T.; Yu, G.; Shao, M. Covalent interaction between high hydrostatic pressure-pretreated rice bran protein hydrolysates and ferulic acid: Focus on antioxidant activities and emulsifying properties. J. Agric. Food Chem. 2021, 69, 7777–7785. [Google Scholar] [CrossRef]
  154. Gorelova, V.; De Lepeleire, J.; Van Daele, J.; Pluim, D.; Meï, C.; Cuypers, A.; Leroux, O.; Rebeille, F.; Schellens, J.H.M.; Blancquaert, D.; et al. Dihydrofolate reductase/thymidylate synthase fine-tunes the folate status and controls redox homeostasis in plants. Plant Cell 2017, 29, 2831–2853. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Maniga, A. Studies on the meristematic and E2F-dependent gene expression in Arabidopsis thaliana plants. Cell Biol. 2017, 17, 5077–5086. [Google Scholar]
  156. Albani, D.; Giorgetti, L.; Pitto, L.; Luo, M.; Cantoni, R.M. Proliferation-dependent pattern of expression of a dihydrofolate reductase-thymidylate synthase gene from Daucus carota. Eur. J. Histochem. 2005, 49, 107–116. [Google Scholar] [PubMed]
Figure 1. PPI network of MTs development in MR8-G6-2iP medium under darkness. In the network there is a tight interaction of the RPs cluster with the tuberigen StSP6A and PEBP family members, thioredoxins, immunophilins, oxidative stress, one-carbon and carbon metabolism. The PPI network has the highest confidence (0.900).
Figure 1. PPI network of MTs development in MR8-G6-2iP medium under darkness. In the network there is a tight interaction of the RPs cluster with the tuberigen StSP6A and PEBP family members, thioredoxins, immunophilins, oxidative stress, one-carbon and carbon metabolism. The PPI network has the highest confidence (0.900).
Genes 14 01463 g001
Figure 2. Hierarchical clustering analyses (HCA) and heat map of up-regulated genes involved in immunophilins during MTs development under dark conditions; levels of up-regulation are presented in Log2.
Figure 2. Hierarchical clustering analyses (HCA) and heat map of up-regulated genes involved in immunophilins during MTs development under dark conditions; levels of up-regulation are presented in Log2.
Genes 14 01463 g002
Figure 3. Hierarchical clustering analyses (HCA) and heat map of up-regulated genes involved in redoxins during MTs development under dark conditions; levels of up-regulation are presented in Log2.
Figure 3. Hierarchical clustering analyses (HCA) and heat map of up-regulated genes involved in redoxins during MTs development under dark conditions; levels of up-regulation are presented in Log2.
Genes 14 01463 g003
Figure 4. Hierarchical clustering analyses (HCA) and heat map of up-regulated genes involved in oxidative stress during MTs development under dark conditions; levels of up-regulation are presented in Log2.
Figure 4. Hierarchical clustering analyses (HCA) and heat map of up-regulated genes involved in oxidative stress during MTs development under dark conditions; levels of up-regulation are presented in Log2.
Genes 14 01463 g004
Figure 5. Hierarchical clustering analyses (HCA) and heat map of up-regulated genes involved in carbon metabolism during MTs development under dark conditions; levels of up-regulation are presented in Log2.
Figure 5. Hierarchical clustering analyses (HCA) and heat map of up-regulated genes involved in carbon metabolism during MTs development under dark conditions; levels of up-regulation are presented in Log2.
Genes 14 01463 g005
Figure 6. Hierarchical clustering analyses (HCA) and heat map of up-regulated genes involved in one-carbon metabolism during MTs development under dark conditions; levels of up-regulation are presented in Log2.
Figure 6. Hierarchical clustering analyses (HCA) and heat map of up-regulated genes involved in one-carbon metabolism during MTs development under dark conditions; levels of up-regulation are presented in Log2.
Genes 14 01463 g006
Figure 7. Validation of the transcriptomic-wide analysis by quantitative reverse transcription PCR (qRT-PCR) of 5 DEG up-regulated genes involved in RPs and PEBP of potato S. tuberosum. Blue columns correspond to absolute gene expression derived from the genome-wide analysis. The green bars represent the expression of the transcriptome expressing the number of transcripts per million.
Figure 7. Validation of the transcriptomic-wide analysis by quantitative reverse transcription PCR (qRT-PCR) of 5 DEG up-regulated genes involved in RPs and PEBP of potato S. tuberosum. Blue columns correspond to absolute gene expression derived from the genome-wide analysis. The green bars represent the expression of the transcriptome expressing the number of transcripts per million.
Genes 14 01463 g007
Figure 8. PPI network of RPs interacting with immunophilins, disulfide isomerase, thioredoxins, and oxidative stress proteins during MTs development in potato. RPL29 (PGSC0003DMT400069847) interacts with peptidyl-prolyl cis-trans isomerase FKBP12. The PPI network has the highest confidence (0.900).
Figure 8. PPI network of RPs interacting with immunophilins, disulfide isomerase, thioredoxins, and oxidative stress proteins during MTs development in potato. RPL29 (PGSC0003DMT400069847) interacts with peptidyl-prolyl cis-trans isomerase FKBP12. The PPI network has the highest confidence (0.900).
Genes 14 01463 g008
Figure 9. Cluster of genes involved in carbon metabolism, one-carbon metabolism, TCA cycle interacting with the CC cluster and RPs.
Figure 9. Cluster of genes involved in carbon metabolism, one-carbon metabolism, TCA cycle interacting with the CC cluster and RPs.
Genes 14 01463 g009
Table 1. Primer design of DEG used to validate the genome-wide analysis of MTs development of potato S. tuberosum cv. α Oligonucleotides were designed for qPCR (2−ΔΔCT method analysis) gene expression or transcriptional analysis.
Table 1. Primer design of DEG used to validate the genome-wide analysis of MTs development of potato S. tuberosum cv. α Oligonucleotides were designed for qPCR (2−ΔΔCT method analysis) gene expression or transcriptional analysis.
Gen IDSequencesID NCBI
RPL29F: AATCAGTCGTACAAGGCTCACXM_015312492.1
R: GCATACCTCTGGTTCCTCAAG
RPL11F: GAGAAGCTGCCGGATCTTAATXM_006343698.2
R: TCGGAGGGTCAATGTCAATTC
RPL40F: AGCCAAGATTCAGGACAAGG XM_006350430.1
R: TCAGGACAAGATGCAGAGTTG
RPL17F: GTTTCCAATTACCTTGCCGAAG XM_006345811.2
R: ACCATTGTTCATCCCGTCTTC
SP6AF: TGCAACCTAGGGCTCATATTG NM_001287968.1
R: GCCAATGTAGATACTCCCTCAAG
EFα1F: TTTGGCCCTACTGGTTTGACNM_001288491.1
R: GCACTGGAGCATATCCGTTT
SEC3F: GCTTGCACACGCCATATCAATXM_006342542.2
R: TGGATTTTACCACCTTCCGCA
Table 2. Ribosomal proteins with loss of function and chloroplast/ribosome biogenesis factors in plants.
Table 2. Ribosomal proteins with loss of function and chloroplast/ribosome biogenesis factors in plants.
ProteinFunction/Process/OrganismLoss-of-Function PhenotypeReferences
RPsRibosomal Proteins
RPS1Ribosomal protein. RiceSeedling lethalityZhou et al., 2021 [22]
RPS5ARibosomal protein. ArabidopsisEmbryo lethalityWeijers et al., 2001 [23]
RPS9Ribosomal protein. MaizeEmbryo lethalMa and Dooner, 2004 [24]
RPS13ARibosomal proteinRoot growth retarded and late floweringIto et al., 2000 [25]
RPS16Ribosomal protein. ArabidopsisEmbryo lethalityTsugeki et al., 1996 [26]
RPS17Ribosomal protein. MaizeSeedling lethalitySchultes et al., 2000 [27]
RPS18ARibosomal protein. TobaccoSeedling lethalityRogalski et al., 2006 [28]
RPS20Ribosomal protein. RiceSeedling lethalityGong et al., 2013 [29]
RPS21Ribosomal protein. ArabidopsisReduced photosynthetic activityDong et al., 2020 [30]
RPS27Ribosomal protein. ArabidopsisEmbryo lethalityRevenkova et al., 1999 [31]
RPL5CRibosomal. ArabidopsisEmbryo lethalityDupouy et al., 2022 [32]
RPL9C, RPL9DRibosomal protein. ArabidopsisEmbryo lethalityDevis et al., 2015 [33]
RPL10Ribosomal protein. Arabidopsis, MaizeEmbryo lethalityFalcone et al., 2010 [34]
RPL11Ribsomal protein. ArabidopsisDecreased leaf pigmentation, plant growth and photosyntesisPesaresi et al., 2001 [35]
RPL12Ribosomal protein. RiceSeedling lethalityZhao et al., 2016 [36]
RPL13Ribsomal protein. RiceEmbryo lethalityLee et al., 2019 [37]
RPL15CRibosomal protein. ArabidopsisEmbryo lethalityBobik et al., 2019 [38]
RPL21CRibosomal protein. Arabidopsis and RiceEmbryo lethalityYin et al., 2021, Lin et al., 2015 [39]
RPL23aRibosomal protein, ArabidopsisAbnormal root and leaves, delayed transition to reproductive growth and reduced seed productionDegenhardt and Bonham-Smith, 2008 [40]
RPL24BRibosomal protein. ArabidopsisDefects in Auxin response related to ARF3 and ARF5Zhou et al., 2010 [41]
RPL28-1Ribosomal protein. ArabidopsisEmbryo lethalityRomani et al., 2012 [42]
RPL35-1Ribosomal protein. MaizeEmbryo lethalityMagnard et al., 2004 [43]
RPS20, RPL1, RPL4, RPL27 and RPL35Ribosomal proteins. ArabidopsisEmbryo lethalityRomani et al., 2012 [42]
Chloroplast/Ribosome biogenesis factors
EDD1 (GlyRS9)Glycyl tRNA synthetase. ArabidopsisEmbryo lethalityUwer et al., 1998 [44]
CFG1, CFG2Chloroplast development. ArabidopsisSeedling lethalityZhu et al., 2020 [45]
DCL-MDefective chloroplast and leaf-
mutable. Tomato
Embryo lethalityBellaoui et al., 2003 [46]
CPN21Chaperonin: Tomato, TobaccoSeed abortionHanania et al., 2006 [47]
AtBRX-1-1, AtBRX-1-2Maturation of the large pre-60S ribosomal subunitPointed leaves, delayed growthWeis et al., 2015 [48]
AtNuc-L1-AtNuc-L2Ribosome biogenesis. ArabidopsisSeedling lethalityDurut et al., 2014 [49]
AtTHALNucleolar organizationEmbryo lethalityChen et al., 2016 [50]
AtNMD3Nuclear export adaptor of 60S pre-ribosome export and maturationLethalChen et al., 2012 [51]
RID1DEAH-box RNA helicase, Pre-mRNA splicingAbnormal shoot and root apical meristem maintenance, leaf and root morphogenesisOhtani et al., 2013 [52]
TIC32Translocon of the inner envelope of chloroplastsEmbryo lethalityHörmann et al., 2004 [53]
ATS2Phosphatidic acid as intermediate for chloroplast membrane lipid biosynthesisEmbryo lethalityYu et al., 2004 [54]
TIC110Translocon of the inner envelope of chloroplastsEmbryo lethalityKovacheva et al., 2005 [55]
CHL27Chlorophyl biosynthesisRetarded growth and chloroplast developmental defectsBang et al., 2008 [56]
DG1Early chloroplast developmentDelayed greening phenotypeChi et al., 2008 [57]
OEP80Chloroplast outer envelope proteinEmbryo lethalityPatel et al., 2008 [58]
EMB5067/AKRPEmbryo development chloroplast proteinEmbryo lethalityGarcion et al., 2006 [59]
SPC1Carotenoid biosynthesisEmbryo lethalityDong et al., 2007 [60]
PDS3phytoene desaturase gene,Embryo lethalityQin et al., 2007 [61]
EMB1303-1Chloroplast biogenesisEmbryo lethalityHuang et al., 2009 [62]
EMB1211Chloroplast biogenesisSeedling lethalityLiang et al., 2010 [63]
BPG2Chloroplast protein accumulation induced by BrassinazoleDecreased number of stacked grana thylakoidsKomatsu et al., 2010 [64]
119 Nuclear genes-assoc. w/chloroplastEmbryo defective mutants/associated to chloroplastEmbryo lethalityBryant et al., 2011 [65]
IRMInvolved in RNA processingEmbryo lethalityPalm et al., 2019 [66]
ZMRH3The RH3 DEAD Box HelicaseEmbryo lethalityAsakura et al., 2012 [67]
HSP90CChloroplast biogenesisEmbryo lethalityInoue et al., 2013 [68]
FTSHI4Thylakoid membrane-associated proteinEmbryo lethalityLu et al., 2014 [69]
RNAJRibonuclease J (RNase J) required for chloroplast and embryo developmentEmbryo lethalityChen et al., 2015 [70]
DERChloroplast ribosomal RNA processingEmbryo lethalityJeon et al., 2014 [71]
Rrp5, Pwp2, Nob1, Enp1 and Noc4Ribosome biogenesis factorsEmbryo lethalityMissbach et al., 2013 [72]
SHREK1Ribosome biogenesis factorEmbryo lethalityLiu et al., 2022 [73]
NOP2A, NOP2BtRNA and rRNA methylation profilesEmbryo lethalityBurgess et al., 2015 [74]
RH22RNA helicase22Embryo lethalityChi et al., 2012 [75]
MDN1The AAA-ATPase MIDASIN 1 functions in ribosome biogenesisEmbryo lethalityLi et al., 2019 [76]
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

Valencia-Lozano, E.; Herrera-Isidrón, L.; Flores-López, J.A.; Recoder-Meléndez, O.S.; Uribe-López, B.; Barraza, A.; Cabrera-Ponce, J.L. Exploring the Potential Role of Ribosomal Proteins to Enhance Potato Resilience in the Face of Changing Climatic Conditions. Genes 2023, 14, 1463. https://doi.org/10.3390/genes14071463

AMA Style

Valencia-Lozano E, Herrera-Isidrón L, Flores-López JA, Recoder-Meléndez OS, Uribe-López B, Barraza A, Cabrera-Ponce JL. Exploring the Potential Role of Ribosomal Proteins to Enhance Potato Resilience in the Face of Changing Climatic Conditions. Genes. 2023; 14(7):1463. https://doi.org/10.3390/genes14071463

Chicago/Turabian Style

Valencia-Lozano, Eliana, Lisset Herrera-Isidrón, Jorge Abraham Flores-López, Osiel Salvador Recoder-Meléndez, Braulio Uribe-López, Aarón Barraza, and José Luis Cabrera-Ponce. 2023. "Exploring the Potential Role of Ribosomal Proteins to Enhance Potato Resilience in the Face of Changing Climatic Conditions" Genes 14, no. 7: 1463. https://doi.org/10.3390/genes14071463

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