eISSN: 2353-9461
ISSN: 0860-7796
BioTechnologia
Current issue Archive About the journal Editorial board Abstracting and indexing Subscription Contact Instructions for authors Publication charge Ethical standards and procedures
Editorial System
Submit your Manuscript
2/2022
vol. 103
 
Share:
Share:
RESEARCH PAPERS

Physicochemical properties and homology studies of the floral meristem identity gene LFY in nonflowering and flowering plants

Roshni Pulukkunadu Thekkeveedu
1
,
Smitha Hegde
2

1.
Department of Post Graduate Studies and Research in Biotechnology, St. Aloysius College, Mangalore, India
2.
Nitte University Centre for Science Education and Research, Mangalore, India
BioTechnologia vol. 103(2) ∙ pp. 113–129 ∙ 2022
Online publish date: 2022/06/29
Article file
Get citation
 
PlumX metrics:
 

Introduction

The homologous gene LFY (LEAFY) regulates cell proliferation and flower development in plants. It is widely distributed in algae, mosses, ferns, gymnosperms, and flowering plants (Villimová, 2012). LFY homologs reported in aquatic Charophyte green algae are closely related to those found in land plants. LFY encodes a plant-specific transcription factor that functions as an activator or a repressor, depending on the cofactor it interacts with (Siriwardana and Lamb, 2012). In Physcomitrella patens, LFY regulates cell division in gametophytes and sporophyte (Tanahashi et al., 2005), whereas LFY homologs in the fern Ceratopteris richardii function in shoot development (Plackett et al., 2018). LFY is a floral meristem identity gene that controls multiple aspects of inflorescence development in the flowering plant Arabidopsis thaliana (Weigel and Nilsson, 1995), and it is active during reproductive structure development in gymnosperms (Dornelas and Rodriguez, 2005; Moyroud et al., 2017). An increase in the expression of LFY results in early flowering, and a mutation in LFY causes a transition of flowers into leaves and shoots (Weigel et al., 1992). Charophytes (algae) (Domozych et al., 2016), P. patens (moss) (Cove et al., 2009), C. richardii (fern) (Hickok et al., 1995; Renzaglia and Warne, 1995), Picea abies (Spruce) (Nystedt et al., 2013), and Arabidopsis thaliana (flowering plant) (Ezhova, 1999) are model organisms for genetic, developmental, and evolutionary studies. In the present study, LFY homologs of few model plants were analyzed to determine molecular differences, i.e., their transition changes during evolution from simple to complex structures in plants. The diverging lineage of the LFY gene, which is modified and altered during the evolution of floral meristems, will help us to understand the origin of flower development or the lack of it in plants. The present study attempted to corroborate the molecular changes of LFY and the evolution of flowering in plants from Charophyte green algae to angiosperms.

Prediction of structure is imperative to study the biochemical and cellular functions of proteins. X-ray crystallography, NMR spectroscopy, and electron microscopy are the techniques currently used for protein structure prediction; however, these methods are time-consuming and require expensive wet lab tools (Venkatesan et al., 2013). Computational tools have been used for the past 30 years in protein structure prediction and continue to help researchers in experimental investigations on a large scale (Nagano, 1973; Gupta et al., 2014; Kc, 2017; Kuhlman and Bradley, 2019). Computational techniques have improved the success rate in protein prediction methods in the last decade. The prediction methods are categorized into comparative modeling (homology modeling) (Šali and Blundell, 1993; Lam et al., 2017), threading (Panchenko et al., 2000; Skolnick and Kihara, 2001; Xu et al., 2007), and ab initio modeling (free modeling) (Ortiz et al., 1998; Simons et al., 2001; Lee et al., 2017). The most accurate method for protein structure prediction is homology modeling, which is used to construct 3D models of unknown target sequences based on known structures (templates) with sequence similarity >30% (Cavasotto and Phatak, 2009) collected from databases by using software or web servers (Eswar et al., 2003; Pieper et al., 2006). The process of homology modeling involves template search of related structures for query sequence, multiple sequence alignment of targets and template structures, construction of a 3D model, and finally, evaluation of the model (Hasani and Barakat, 2017; Studer et al., 2019). The steps are repeated to obtain an optimum model (Martí-Renom et al., 2000). The 3D structures of proteins are more conserved than their amino acid sequences, and minor changes in sequences usually result in a slight variation in their 3D structure (Lesk and Chothia, 1986).

In the present study, the homologous sequences of the LFY gene were compared to analyze its phylogenesis and physicochemical properties of the gene product; moreover, protein structure prediction and construction of 3D models of protein and their evaluation were performed. The LFY homologous sequences from algae to flowering plant model systems were elaborated. A comparative analysis of the homologous genes was conducted virtually to study their structure, function, phylogeny, and proteins.

Materials and methods

Data mining from database and phylogeny construction

Complete and partial LEAFY protein sequences of the LFY gene of all available plant families (supplementary Table 1) were collected from GenBank NCBI (National Centre for Biotechnology Information) (http://www.ncbi.nlm.nih.gov) in FASTA format after clicking protein search, followed by building datasets of sequences using Notepad from the Protein database of related organisms in NCBI [LEAFY in plants – Protein – NCBI (nih.gov)]. A phylogenetic tree was constructed using MEGA 7 (Molecular Evolutionary Genetic Analysis version 7) software with the neighbor-joining method with bootstrapping values for 1000 replicates. The evolutionary distance was computed using the p-distance method. All the LEAFY protein sequences were copied onto MEGA 7 software and aligned using ClustalW. Complete and partial sequences of the LEAFY protein of charophyte green algae – Klebsormidium subtile and Coleochaete scutata (2 sequences), P. patens (2 sequences), Ceratopteris sp. (3 sequences), Picea sp. (2 sequences), and A. thaliana (3 sequences) were used (supplementary Table 1 and Table 2). The aligned sequences were then exported to conduct phylogenetic analysis. The phylogeny analysis was conducted using Neighbor Joining Tree method. It was further tested with the Bootstrap method. A total of 1000 bootstrap replications were used to compute and construct a phylogenetic tree.

Motif search

The Motif Finder server (https://www.genome.jp/tools/motif/) was used to analyze the family or the protein domain of the protein sequences. FASTA sequences of the protein were entered as a query sequence in the Motif Finder server. The results showed LEAFY protein sequences of all model organisms with conserved domains from Pfam databases. These sequences were confirmed in the CDD database (NCBI) webserver (https://www.ncbi.nlm.nih.gov/Structure/cdd/wrpsb.cgi) by using the protein accession number of the respective organisms.

Multiple sequence alignment

FASTA sequences of LEAFY proteins were aligned using the Clustal Omega Multiple Sequence Alignment Program [CLUSTAL O (1.2.4)] (https://www.ebi.ac.uk/Tools/msa/clustalo/) using complete and partial sequences of LEAFY proteins. FASTA sequences of LEAFY proteins were also used for screening sequence alignment of conserved, conservative mutated, semi-conservative mutated, and nonconservative mutated sequences among the species analyzed.

Physicochemical analysis

The amino acid sequences of LEAFY proteins were determined for analyzing physicochemical properties in the ProtParam tool – ExPASy (http://web.expasy.org/protparam) that computes various parameters such as molecular weight (MW), isoelectric point (pI), instability index (II), aliphatic index (AI), and grand average of hydropathicity (GRAVY). Each amino acid sequence was imported into the ProtParam tool ExPASy and computed for analyzing various parameters of LEAFY protein sequences.

Structure prediction and analysis

The secondary structure of the LEAFY protein was predicted using the PSI-blast-based secondary structure tool PREDiction (PSIPRED 4.0), and the 3D model was generated using PHYRE 2 (ProteinHomology/analogy RecognitionEngineV2.0) (http://www.sbg.bio.ic.ac.uk/~phyre2/html/page.cgi?id=index) program. Multiple sequences of each model plant were batch-processed while uploading onto the PHYRE2 server. The LEAFY protein sequence was submitted in FASTA format and processed for the following information: 1) summary and sequence analysis details, 2) secondary structure and disorder prediction, 3) domain analysis, and 4) detailed template alignment information with figures. Secondary structure was predicted as α-helix, β-strand, coils, and disorder in the structure analysis report.

Protein modeling and validation

Protein homology modeling for LEAFY proteins was performed with the SWISS-MODEL https://swissmodel.expasy.org/ web server. Protein modeling involved the following: 1) identification of template structure, 2) alignment of the target sequence and template structure, 3) model building, and 4) model quality evaluation.

The best models of LEAFY protein structures for all the plant models were evaluated for quality by using the structure validation tool MolProbity (http://molprobity.biochem.duke.edu/). The PDB code of the protein structure was further validated using their web server. The analysis revealed the distribution of residues in the torsion angles (φ and ψ) of Ramachandran plot with the φ and ψ values between +180 and –180 on the x-axis and y-axis, respectively. The percentage of residues in the favored, allowed region and outliers from the Ramachandran plot analysis were also detected.

Results and discussion

Evolution of the LFY gene in the plant family

Forty-one LEAFY protein sequences of different plant families were downloaded from GenBank (Table 1), and phylogenetic analysis was conducted. The phylogenetic tree (Fig. 1) illustrates the relationship among 41 amino acid sequences of LEAFY proteins analyzed in MEGA7 with 1000 bootstrap replicates using the neighborjoining method. The bootstrap percentage specifies the reliability of each node of the phylogenetic tree, and an estimate of < 70% reliability on tree topology is not considered to be acceptable (Hall, 2013). The p-distance method was used to compute evolutionary distances with the differences in the number of amino acids per site.

Table 1

List of LEAFY homolog of different plant families from GenBank NCBI

No.OrganismsProtein accession numberNumber of amino acid
1Ceratozamia mexicanaAIG12601.1375
2Lepidozamia peroffskyanaAIG12598.1380
3Dioon spinulosumAIG12603.1380
4Microcycas calocomaAIG12609.1376
5Encephalartos arenariusAIG12597.1380
6Bowenia spectabilisAIG12608.1375
7Macrozamia lucidaAIG12599.1377
8Stangeria eriopusAIG12610.1363
9Cycas revolutaAIG12606.1377
10Ginkgo bilobaADD64700.1402
11Picea abiesAAV49504.1386
12Picea sitchensisAKA55658.1380
13Angiopteris lygodiifoliaBAB93543.1344
14Sceptridium robustumBAB88864.1350
15Psilotum nudumBAB88863.1372
16Ceratopteris thalictroidesABF74516.1237
17Ceratopteris pteridoidesABF74512.1237
18Ceratopteris richardiiABF74513.1237
19Physcomitrella patensBAD91044.1349
20Physcomitrella patensBAD91043.1348
21Vanilla planifoliaAOA52645.1491
22Phalaenopsis hybrid cultivarACS94257.1437
23Tricyrtis formosanaBAN62610.1411
24Oryza sativa Japonica GroupAHX83809.1389
25Zea maysAAO43173.1393
26Allium cepaAFR67540.1372
27Allium cepaAVT42847.1370
28Amborella trichopodaAGV98899.1391
29Nymphaea odorataAAF77609.1387
30Chrysanthemum indicumARR73986.1412
31Litchi chinensisAGR45584.1388
32Mangifera indicaADX97320.1383
33Populus balsamiferaAEK06015.1377
34Magnolia virginianaACV88634.1389
35Annona squamosaAKV57239.1411
36Brassica rapaANJ12320.1417
37Arabidopsis thalianaAAM27932.1424
38Arabidopsis thalianaAAM27931.1424
39Arabidopsis thalianaAAM27941.1424
40Coleochaete scutataAHJ90705.1328
41Klebsormidium subtileAHJ90707.1495
Fig. 1

Phylogenetic tree constructed in MEGA 7 with 41 LEAFY protein sequences; bootstrap values are listed next to the branch

/f/fulltexts/BTA/47020/BTA-103-2-47020-g001_min.jpg

LEAFY protein sequences of Charophyte green algae were placed at the base of the phylogenetic tree and grouped into two distinct clusters. One cluster of LEAFY protein sequences was present in flowering plants starting from orchids to Arabidopsis, such as Vanilla planifolia, Phalaenopsis hybrid cultivar, Tricyrtis formosana, Oryza sativa Japonica Group, Zea mays, Allium cepa, Amborella trichopoda, Nymphaea odorata, Chrysanthemum indicum, Litchi chinensis, Mangifera indica, Populus balsamifera, Magnolia virginiana, Annona squamosa, Brassica rapa, and A. thaliana. The second cluster contained all nonflowering plants such as mosses, pteridophytes, gymnosperms, and cycads. It is reported that the plant-specific transcription factor LFY evolved in Streptophyte algae (Wilhelmsson et al., 2017). The LEAFY protein sequence of pteridophytes is closer to that of cycads and gymnosperms than to that of orchids, monocots, and other angiosperms, which indicates structural and functional similarity to LEAFY from gymnosperms. The tree shows an early alteration in the LFY homolog, which segregated and evolved into two clusters of flowering and nonflowering plants. It is also reported that ancient gene duplication and sub-functionalization processes influenced the evolution of the LEAFY gene (Gao et al., 2019).

Description of the candidate LEAFY protein

The LEAFY protein sequences from 2 species of Charophyte green algae (K. subtile – 495 aa and C. scutata – 328 aa), 2 moss species (P. patens – 349 aa, P. patens – 348 aa), 2 gymnosperm species (P. abies – 386 aa and P. sitchensis – 380 aa), 3 pteridophytes species (C. thalictroides – 237 aa, C. pteridoides – 237 aa, C. richardii – 237 aa), and 3 angiosperm species (A. thaliana – 424 aa) were collected from the NCBI protein database (Table 2) and analyzed.

Table 2

Description and list of LEAFY homolog of Charophyte green algae, Physcomitrella sp., Ceratopteris sp., Picea sp., and Arabidopsis sp. from GenBank NCBI

OrganismProtein accession numberProteinNumber of amino acid
Klebsormidium subtileAHJ90707.1KsLFY495
Coleochaete scutataAHJ90705.1CsLFY328
Physcomitrella patensBAD91044.1FLORICAULA/LEAFY homolog2349
Physcomitrella patensBAD91043.1FLORICAULA/LEAFY homolog2348
Ceratopteris thalictroidesABF74516.1LEAFY/FLORICAULA237
Ceratopteris richardiiABF74513.1LEAFY/FLORICAULA237
Ceratopteris pteridoidesABF74512.1LEAFY/FLORICAULA237
Picea abiesAAV49504.1PaLFY386
Picea sitchensisAKA55658.1LEAFY380
Arabidopsis thalianaAAM27931.1Leafy424
Arabidopsis thalianaAAM27932.1Leafy424
Arabidopsis thalianaAAM27941.1Leafy424

The LFY transcription factor gene evolved from algae (charophyte) (Sayou et al., 2014; Brunkard et al., 2015; Gao et al., 2019). LEAFY/LFY homologs in different model organisms, such as FLO/LFY genes (PpLFY1, PpLFY2), regulate the first zygotic cell division in P. patens (Tanahashi et al., 2005). LFY maintains apical stem cell activity in gametophyte and sporophyte during shoot development in C. richardii (Plackett et al., 2018). In gymnosperms, LFY and the paralog of LFY – NEEDLY (NLY) regulate male and female reproductive structures (Silva et al., 2016), and their expression levels were characterized in Picea sp. (Vázquez-Lobo et al., 2007). Moreover, LFY, the plant-specific transcription factors, are conserved as floral meristem identity genes, which control inflorescence and floral organ development in Arabidopsis (Wang et al., 2004).

Domain analysis

Motif search of protein sequences was conducted (Table 3), and it was found that the LEAFY protein sequences shared two domains, namely N-terminal Sterile Alpha Motif (SAM_ LFY) and C-terminal DNA binding domain (C_LFY_FLO) (GenomeNet bioinformatics tool). The result was confirmed using the CDD database (NCBI) web server (Fig. 2). The SAM domain mediates LFY oligomerization that helps to access low-affinity binding sites or closed chromatin regions (Sayou et al., 2016), and the biochemical properties of SAM domains are conserved throughout the evolution of all plant species. The crystal structure of the LFY- DNA binding domain resembles that of helix-turn-helix proteins and dimerizes on DNA, which triggers major developmental switches in plants (Hamès et al., 2008). The domains bind to short stretches of DNA called transcription factor binding sites (TFBS) that regulate gene expression. Domain analysis reported two domains of LEAFY proteins in all model organisms, except in fern species. In ferns, it was noted that only the C-terminal C_LFY_FLO, DNA binding domain is in partial confidence with other LEAFY protein sequences screened. The Pfam report in CDD revealed that these development LEAFY proteins are homologs of floricaula (FLO) and LEAFY (LFY), which function in floral meristem identity (Table 4). A mutation in these protein sequences affected the flower and leaf formation (Weigel et al., 1992; Hofer et al., 1997; Grandi et al., 2012; Monniaux et al., 2017).

Table 3

Domain and functional analysis of LFY/LEAFY homologous using motif find server

Protein IDPfamPosition (independent E-value)Description
AHJ90707.11.C_ LFY_FLO311–471 (7.7e-49)PF17538, DNA Binding Domain (C-terminal) Leafy/Floricaula
2.SAM_LFY162–238 (3.6e-19)PF01698, Floricaula /Leafy protein SAM domain
AHJ90705.11.C_ LFY_FLO189–287 (3.6e-46)PF17538, DNA Binding Domain (C-terminal) Leafy/Floricaula
2.SAM_LFY5–82 (3.5e-27)PF01698, Floricaula /Leafy protein SAM domain
BAD91044.11.C_ LFY_FLO183–347 (5.8e-86)PF17538, DNA Binding Domain (C – terminal) Leafy/Floricaula
2.SAM_ LFY38–115 (1e-27)PF01698, Floricaula/Leafy protein SAM domain
BAD91043.11.C_ LFY_FLO182–346 (6.9e-86)PF17538, DNA Binding Domain (C – terminal) Leafy/Floricaula
2.SAM_ LFY37–114 (2.4e-28)PF01698, Floricaula/Leafy protein SAM domain
ABF74516.11.C_ LFY_FLO119–237 (8.6e-68)PF17538, DNA Binding Domain (C – terminal) Leafy/Floricaula
ABF74512.11.C_ LFY_FLO119–237 (8.6e-68)PF17538, DNA Binding Domain (C – terminal) Leafy/Floricaula
ABF74513.11.C_ LFY_FLO119–237 (8.6e-68)PF17538, DNA Binding Domain (C – terminal) Leafy/Floricaula
AAV49504.11.C_ LFY_FLO218–383 (3.6e-95)PF17538, DNA Binding Domain (C – terminal) Leafy/Floricaula
2.SAM_ LFY35–112 (1.4e-37)PF01698, Floricaula/Leafy protein SAM domain
AKA55658.11.C_ LFY_FLO235–380 (4.8e-83)PF17538, DNA Binding Domain (C – terminal) Leafy/Floricaula
2.SAM_ LFY52–129 (1.4e-37)PF01698, Floricaula/Leafy protein SAM domain
AAM27932.11.C_ LFY_FLO229–393 (1.3e-107)PF17538, DNA Binding Domain (C – terminal) Leafy/Floricaula
2.SAM_ LFY45–123 (4.6e-43)PF01698, Floricaula/Leafy protein SAM domain
AAM27931.11.C_ LFY_FLO229–393 (1.3e-107)PF17538, DNA Binding Domain (C – terminal) Leafy/Floricaula
2.SAM_ LFY45–123 (4.6e-43)PF01698, Floricaula/Leafy protein SAM domain
AAM27941.11.C_ LFY_FLO229–393 (1.3e-107)PF17538, DNA Binding Domain (C – terminal) Leafy/Floricaula
2.SAM_ LFY45–123 (4.6e-43)PF01698, Floricaula/Leafy protein SAM domain
Table 4

Details of LEAFY protein sequences

OrganismProtein IDNumber of amino acidSequence
Klebsormidium subtileAHJ90707.1495 aacomplete sequence
Coleochaete scutataAHJ90705.1328 aacomplete sequence
Physcomitrella patensBAD91044.1349 aacomplete sequence
Physcomitrella patensBAD91043.1348 aacomplete sequence
Ceratopteris thalictroidesABF74516.1237 aapartial sequence
Ceratopteris richardiiABF74513.1237 aapartial sequence
Ceratopteris pteridoidesABF74512.1237 aapartial sequence
Picea abiesAAV49504.1386 aacomplete sequence
Picea sitchensisAKA55658.1380 aapartial sequence
Arabidopsis thalianaAAM27931.1424 aacomplete sequence
Arabidopsis thalianaAAM27932.1424 aacomplete sequence
Arabidopsis thalianaAAM27941.1424 aacomplete sequence
Fig. 2

Conserved domain analysis from Conserved Domain Database – NCBI of: A1 – Klebsormidium subtile, A2 – Coleochaete scutata, B1 – Physcomitrella patens, C1 – Ceratopteris sp., D1 – Picea abies, D2 – Picea sitchensis, E1 – Arabidopsis thaliana

/f/fulltexts/BTA/47020/BTA-103-2-47020-g002_min.jpg

Analysis of sequence similarity

Multiple sequence alignment methods align and compare DNA, RNA, or protein sequences for evolutionarily relatedness. The aligned sequences provide valuable information regarding the structural, functional, and evolutionary history, often leading to a common ancestor (Edgar and Batzoglou, 2006; Chatzou et al., 2016). LFY primary sequences were aligned using the Clustal Omega [CLUSTAL O (1.2.4)] program to find the conserved region(s). The sequences were aligned by inserting a gap or space into the sequence to extend to the same length after alignment (Wang and Jiang, 1994; Tran and Wallinga, 2017). Charophyte green algae shared 38–46% sequence similarity with Physcomitrella sp., 37–46% similarity with Ceratopteris sp., 33–41% similarity with Picea sp., and 32–38% similarity with Arabidopsis sp. Fifty conserved (similar amino acid sequences), 22 conservative mutated (mutation results in the replacement of amino acid with a similar biochemical property), and 9 semi-conservative mutated (mutation results in the replacement of amino acid with a similar shape but dissimilar biochemical property) amino acids were identified from the sequence alignment of LEAFY proteins. In Arabidopsis sp., in conservative mutated amino acid sequence, aspartic acid (D) was replaced with histidine (H) at position 296 and alanine (A) was replaced with serine (S) at position 343, while in semi-conservative mutated amino acid sequence, histidine was replaced with tyrosine (T) at position 262, which were different from the respective amino acid sequences in the other non-flowering plant species tested (Fig. 3).

Fig. 3

Multiple Sequence Alignment of FLO_LFY protein sequences of length 237, 348, 349 and 424 residues obtained using Clustal Omega; (*) – denotes conserved sequence which is highlighted in blue, (:) – denotes conservative mutation which is highlighted in yellow [variation in sequence marked in red], (.) – denotes semi-conservative mutation which is highlighted in green [variation in sequence marked in red] and ( ) – denotes non-conservative mutation

/f/fulltexts/BTA/47020/BTA-103-2-47020-g003_min.jpg

LFY orthologs are found in all land plants, and the LFY gene performs various functions in multiple species as it evolves after gene duplication events (Silva et al., 2016). LFY homologs are involved in regulating cell division, and expansion and arrangement in free-sporing land plants such as ferns or fern allies and bryophytes. They also regulate both floral identity and cell division in gymnosperms and angiosperms (Moyroud et al., 2010).

Physicochemical properties of LEAFY proteins

The physicochemical properties of proteins influence their affinity, interaction, and adaptability to a biological system (Panda and Chandra, 2012; Dhar et al., 2020). Physicochemical characterization of the amino acid sequence includes MW, pI, II, AI, and GRAVY (Kaur et al., 2020), which were estimated using Expasy’s ProtParam (Table 5). This comparative analysis helped us to identify the occurrence of diversity of LFY protein sequences across Charophyte green algae, Physcomitrella sp., Ceratopteris sp., Picea sp., and Arabidopsis sp. pI is the pH value at which there are no net charges on the protein, and the protein remains stable without migration in the electric field and remains firm and stable at this pI. pI is crucial in protein separation and characterization (Pergande and Cologna, 2017). The pI of the LEAFY protein was lower and acidic for Arabidopsis sp., Physcomitrella sp., and K. subtile, whereas it was higher and alkaline for Coleochaete scutata, Ceratopteris sp., and Picea sp. II reveals the stability of the protein in both in vivo and in vitro conditions. The II value above 40 indicates that the protein is unstable, while the value below 40 indicates that it is stable (Guruprasad et al., 1990; Gamage et al., 2019). Our findings reveal that the II value of LEAFY proteins ranged from 39.58 to 62.51, which indicates that the structure of proteins is unstable. The AI index is essential to determine the thermal stability potential of the amino acid sequence, and thermal stability increases with a higher AI value (Panda and Chandra, 2012). The AI of LEAFY proteins ranged from 62.57 to 74.87, which indicates that the protein is thermally stable at a wide range of temperature (20–45 °C) (Enany, 2014; Ikai, 1980). Similar to stability studies, it is essential to evaluate the hydrophobic and hydrophilic nature of proteins by using the GRAVY score. The negative GRAVY value indicates that the protein is hydrophilic, and a positive value indicates they are hydrophobic; the value usually ranges from –2 to +2 (Kyte and Doolittle, 1982; Kaur and Pati, 2018). The GRAVY score of the LEAFY proteins for all model organisms was negative, which indicated a high number of interactions with water.

Table 5

Physicochemical characterization of LEAFY proteins in the ProtParam tool

OrganismProtein IDNumber of amino acidMolecular weightIsoelectric point (pI)Instability index (II)Aliphatic index (AI)Grand average of hydropathicity (GRAVY)
Klebsormidium subtileAHJ90707.1495 aa55211.846.5162.5172.18–0.754
Coleochaete scutataAHJ90705.1328 aa37142.578.6747.3967.47–0.694
Physcomitrella patensBAD91044.1349 aa40090.816.4044.9074.87–0.702
Physcomitrella patensBAD91043.1348 aa40134.896.7847.6576.47–0.727
Ceratopteris thalictroidesABF74516.1237 aa27286.989.4440.3962.57–1.080
Ceratopteris richardiiABF74513.1237 aa27259.909.2740.3162.57–1.076
Ceratopteris pteridoidesABF74512.1237 aa27187.849.3639.5862.57–1.063
Picea abiesAAV49504.1386 aa44116.158.5548.1576.01–0.674
Picea sitchensisAKA55658.1380 aa43161.958.0545.8374.16–0.665
Arabidopsis thalianaAAM27931.1424 aa47168.866.4855.4766.08–0.685
Arabidopsis thalianaAAM27932.1424 aa47157.796.2255.5266.08–0.683
Arabidopsis thalianaAAM27941.1424 aa47099.756.3455.6866.08–0.676

Secondary structure prediction and analysis

Secondary structure formation is the initial step in protein folding to attain its functional shape (Pirovano and Heringa, 2010). The most accurate and reliable prediction of protein sequence structure is a challenging aspect of computational biology. The PHYRE2 program uses homology modeling techniques that help in structure prediction, function prediction, domain analysis, and mutation analysis (Kelley et al., 2015). As an essential step in the prediction of tertiary structures, PHYRE2 was first used for determining secondary structures such as an alpha helix, beta-strands, and irregular coil regions in the polypeptide chain of amino acids, which determine protein activity, interactions, and functions at the molecular level (Kelley et al., 2015). Homologous sequences for the LEAFY protein of each model organism (query sequence) were detected from multiple sequence alignment using PSI-Blast from the PHYRE2 server. The secondary structure prediction and disorder region prediction was made by Psi-Pred and Diso-Pred programs. In the predicted secondary structures of LFY proteins, it was found that the percentage of alpha-helix (α) structure ranged from 58 to 72% and that of beta-strands (β) ranged from 0 to 4% (Table 6). Proteins with alpha-helix (α) ≥ 40% and beta-strands (β) ≤ 5% were categorized as alpha protein class (Chou, 1995). Thus, these secondary structures belong to alpha protein classes. In alpha helices and beta strands, the potential to tolerate mutation differs significantly. Helices are more robust to mutation than strands or coils due to the noncovalent interactions of residues in the secondary structure units without a structural change (Abrusán and Marsh, 2016). The contact density among residues determines the acceptance of mutation without destabilizing the protein fold (England and Shakhnovich, 2003; Shakhnovich et al., 2005; Nemtseva et al., 2019). Thus, mutations result in lesser structural change.

Table 6

Secondary structure prediction of LEAFY proteins using the PHYRE 2 programme

Protein IDAlpha helix [%]Beta strand [%]Disordered [%]
AHJ90707.162235
AHJ90705.163143
BAD91044.166331
BAD91043.166432
ABF74516.172058
ABF74513.171150
ABF74512.172150
AAV49504.161241
AKA55658.160147
AAM27931.158244
AAM27932.158344
AAM27941.158344

The protein region without a secondary structure is a disordered region that affects the 3D structure of a protein. The disordered region binds with the partner molecule (nucleic acid, another protein, etc.) and thus exists as a structured protein (Dyson and Wright, 2005; Ishida and Kinoshita, 2008; Uversky, 2019). It often plays a functional role and is commonly involved in transcription, translation, and cell signaling (Van Der Lee et al., 2014; Hsu et al., 2020). Mutations in the disordered regions result in inappropriate protein folding (Uversky et al., 2005; Dyson, 2011); thus, the prediction of disordered regions is pivotal for the structure and function analysis of a protein sequence. The Diso-Pred server predicted the presence of 31% to 58% disordered regions in the tested LEAFY homologs, with the highest value found for C. thalictroides and the lowest value for P. patens, indicating that the LEAFY homolog is dynamic in the fern C. thalictroides and stable in the moss P. patens (Table 6).

Homology modeling

The determination of three-dimensional structure is essential as it provides insights into biochemical functions and protein interactions (Ittisoponpisan et al., 2019). The model was constructed by identifying sequence similarity (homologous sequence) with a target sequence and alignment with a suitable template from PDB (Fiser, 2010). The protein model was constructed using SWISS-MODEL based on the sequence and alignment with the most appropriate structural template for the LEAFY protein of model organisms from the PDB database, with GMQE and QMEAN Z-score values (Biasini et al., 2014). GMQE (Global Model Quality Estimation) estimates the model’s accuracy and is expressed as a number between 0 and 1. The QMEAN Z-score reports the reliability of the model quality estimation, and a QMEAN Z-score of around 0 indicates a good quality model (Benkert et al., 2011; Biasini et al., 2014; Waterhouse et al., 2018). Protein structure with a sequence homology of > 40% shares similarity with other protein structures, whereas sequence homology < 25% results in significant structural differences. Thus, a reliable protein structure cannot be predicted based on homology modeling when sequence homology is < 25% (Venclovas, 2011). Here, 32 template matches of the LEAFY protein (target sequence) for K. subtile and 43 template matches for Arabidopsis sp. were reported, in which the best and highest sequence similarity was reported for the template 2vy2.1A, which is the LEAFY protein structure of A. thaliana complex with DNA from Ag – I promoter (Hamès et al., 2008). Twenty-eight templates match of the LEAFY protein for C. scutata, 50 templates match for Physcomitrella sp., 8–9 templates match for Ceratopteris sp., 30 templates match for P. abies, and 50 templates match for P. sitchensis were reported, in which the best and highest sequence similarity was reported for the template 4bhk.1. A FLORICAULA/LEAFY HOMOLOG 1 codes for the transcription factor LEAFY in mosses, which interacts with DNA (Sayou et al., 2014) (Table 7). The 3D models for the LEAFY protein in each model organism were constructed based on the template, represented by rainbow color from N-terminal to C-terminal (Fig. 4).

Table 7

Template identification results for each LEAFY protein sequence in the SWISS-MODEL tool

Protein IDGMQEQMEANTemplateSequence identity [%]Description
AHJ90707.10.21–1.222vy2.1A41.07Protein Leafy
AHJ90705.10.24–1.354bhk.1.A54.48Floricaula/Leafy homolog1
BAD91044.10.26–0.634bhk.1.A97.62Floricaula/Leafy homolog1
BAD91043.10.26–0.544bhk.1.A100Floricaula/Leafy homolog1
ABF74516.10.30–0.244bhk.1.A80.17Floricaula/Leafy homolog1
ABF74513.10.29–0.284bhk.1.A80.17Floricaula/Leafy homolog1
ABF74512.10.30–0.244bhk.1.A80.17Floricaula/Leafy homolog1
AAV49504.10.26–0.424bhk.1.A79.17Floricaula/Leafy homolog1
AKA55658.10.26–0.424bhk.1.A79.59Floricaula/Leafy homolog1
AAM27931.10.24–0.282vy2.1.A100Protein Leafy
AAM27932.10.24–0.282vy2.1.A100Protein Leafy
AAM27941.10.24–0.282vy2.1.A100Protein Leafy
Fig. 4

Constructed 3D models of LEAFY protein: A1 – Klebsormidium subtile, A2 – Coleochaete scutata, B1–B2 – Physcomitrella patens, C1 – Ceratopteris thalictroides, C2 – Ceratopteris richardii, C3 – Ceratopteris pteridoides, D1 – Picea abies, D2 – Picea sitchensis, E1–E3 – Arabidopsis thaliana using SWISS-MODEL (represented in rainbow colour from N → C)

/f/fulltexts/BTA/47020/BTA-103-2-47020-g004_min.jpg

Structure evaluation

The structural validation of the predicted protein models is crucial as the predicted structures may contain substantial errors. Because the structure is related to function, the generated model should be error-free. The structure evaluation was conclusively performed by Ramachandran plot analysis (Carugo and Djinović-Carugo, 2013). The MolProbity tool was accessed with the PDB files of protein structures for Ramachandran analysis, which helps to determine the protein geometry (Chen et al., 2010). Ramachandran plot generates the graphical representation of the allowed and forbidden regions of torsion angles, phi (φ) and psi (ψ), by plotting phi (φ) on the x-axis and psi (ψ) on the y-axis. Torsion angles of amino acid residues in the protein structure form secondary structures corresponding to the allowed and disallowed regions (Saravanan and Selvaraj, 2017). The dark-colored region in the Ramachandran plot is considered as the most favorable, the light-colored region as favorable, and the white region is disallowed and regarded as forbidden in the four quadrants of the Ramachandran plot structure. The four-quadrant plot helps in analysis of possible combination of torsion angles of the proposed protein. An optimal quality structure contains all the combinations of torsion angles in the allowed region, whereas if all sets of torsion angles occupy a forbidden region, it reflects a poor-quality homology model, resulting in steric hinderance (Røgen, 2021). The conformation of phi-psi torsion angles of the predicted LEAFY protein structure was satisfactory, as > 96% of all residues were present in the allowed region (Table 8), indicating a good quality model. There were no outliers in the Ramachandran plot for all the plant species, except for C. scutata (Fig. 5); thus, it can be considered as a good quality model suitable for further application (Muhammed and Aki Yalcin, 2019).

Table 8

MolProbity results of 3D models of LEAFY protein generated after structure assessment

Protein IDRamachandran favoured [%]Ramachandran outliers [%]
AHJ90707.197.520.00
AHJ90705.196.750.81
BAD91044.198.080.00
BAD91043.198.080.00
ABF74516.198.180.00
ABF74513.198.180.00
ABF74512.198.180.00
AAV49504.198.720.00
AKA55658.197.790.00
AAM27931.198.140.00
AAM27932.198.140.00
AAM27941.198.140.00
Fig. 5

Ramachandran plot analysis of protein model of: A1 – Klebsormidium subtile, A2 – Coleochaete scutate, B1–B2 – Physcomitrella patens, C1 – Ceratopteris thalictroides, C2 – Ceratopteris richardii, C3 – Ceratopteris pteridoides, D1 – Picea abies, D2 – Picea sitchensis, E1–E3 – Arabidopsis thaliana

/f/fulltexts/BTA/47020/BTA-103-2-47020-g005_min.jpg

Conclusions

The present study revealed that LFY genes are conserved in Charophyte green algae, moss, fern, gymnosperms, and angiosperms. Domain analysis showed that the LEAFY proteins in all plant species shared two conserved domains, namely C_LFY_FLO and SAM_LFY. The physicochemical characterization reported that the LEAFY protein has an unstable structure, indicating its dynamic nature. The protein is thermally stable and hydrophilic in nature. In LEAFY protein sequences, most conserved, conservative mutated, and semi-conservative mutated sequences were predicted as helical structures. Beta strands were conserved in all plant species with only sequence differences in charophyte green algae, which is a unique variation in LFY evolution. The 3D models generated from the LEAFY protein sequences were of good quality and will help to corroborate structural and functional analysis. The results of phylogenetic analysis indicated a very early mutation that led to the formation of two distinct clusters, one leading to angiosperms and the other to gymnosperms. The LFY gene of the gymnosperms showed homology with that of mosses and pteridophytes as compared to that of orchids, monocots, and other flowering plants.

Acknowledgments

Our sincere thanks to Rev Fr Dr Praveen Martis SJ, Principal; Dr Shreelalitha Suvarna, HOD, Department of Post Graduate Studies and Research in Biotechnology; and Rev Fr Dr Leo D’Souza SJ, Director, Dr. Küppers Laboratory of Applied Biology, St Aloysius College (Autonomous) for providing the research facilities.

References

1 

Abrusán G., Marsh J.A. (2016) Alpha helices are more robust to mutations than beta-strands. PLoS Comput. Biol. 12(12): e1005242.

2 

Benkert P., Biasini M., Schwede T. (2011) Toward the estimation of the absolute quality of individual protein structure models. Bioinformatics 27(3): 343–350.

3 

Biasini M., Bienert S., Waterhouse A., Arnold K., Studer G., Schmidt T., Schwede T. (2014) SWISS-MODEL: modelling protein tertiary and quaternary structure using evolutionary information. Nucl. Acids Res. 42(W1): W252–W258.

4 

Brunkard J.O., Runkel A.M., Zambryski P.C. (2015) Comment on “A promiscuous intermediate underlies the evolution of LEAFY DNA binding specificity”. Science 347(6222): 621–621.

5 

Carugo O., Djinović-Carugo K. (2013) A proteomic Ramachandran plot (PRplot). Amino Acids 44(2): 781–790.

6 

Cavasotto C.N., Phatak S.S. (2009) Homology modelling in drug discovery: current trends and applications. Drug Discov. Today 14(13–14): 676–683.

7 

Chatzou M., Magis C., Chang J.M., Kemena C., Bussotti G., Erb I., Notredame C. (2016) Multiple sequence alignment modeling: methods and applications. Brief. Bioinformat. 17(6): 1009–1023.

8 

Chen V.B., Arendall W.B. 3rd, Headd J.J., Keedy D.A., Immormino R.M., Kapral G.J., Murray L.W., Richardson J.S., Richardson D.C. (2010) MolProbity: all-atom structure validation for macromolecular crystallography. Acta Crystallogr. D: Biol. Crystallogr. 66 (Pt 1): 12–21.

9 

Chou K.C. (1995) A novel approach to predicting structural protein classes in a (20–1) – D amino acid composition space. Proteins. 21(4): 319–344.

10 

Cove D.J., Perroud P.F., Charron A.J., McDaniel S.F., Khandelwal A., Quatrano R.S. (2009) The moss Physcomitrella patens: a novel model system for plant development and genomic studies. [in:] Emerging Model Organisms: a laboratory manual, vol. 1, NY: CSHL Press: 592.

11 

Dhar S., Sood V., Lohiya G., Devenderan H., Katti D.S. (2020) Role of physicochemical properties of protein in modulating the nanoparticle-bio interface. J. Biomed. Nanotechnol. 16(8): 1276–1295. http://doi.org/10.1166/jbn.2020.2958

12 

Domozych D., Popper Z.A., Sorensen I. (2016) Charophytes: evolutionary giants and emerging model organisms. Front. Plant Sci. 7: 1470.

13 

Dornelas M.C., Rodriguez A.P.M. (2005) A FLORICAULA/LEAFY gene homolog is preferentially expressed in developing female cones of the tropical pine Pinus caribaea var. caribaea. Genet. Mol. Biol. 28(2): 299–307.

14 

Dyson H.J. (2011) Expanding the proteome: disordered and alternatively folded proteins. Q. Rev. Biophys. 44(4): 467–518.

15 

Dyson H.J., Wright P.E. (2005) Intrinsically unstructured proteins and their functions. Nat. Rev. Mol. Cell Biol. 6(3): 197–208.

16 

Edgar R.C., Batzoglou S. (2006) Multiple sequence alignment. Curr. Opin. Struct. Biol. 16(3): 368–373.

17 

Enany S. (2014) Structural and functional analysis of hypothetical and conserved proteins of Clostridium tetani. J. Infect. Public Health 7(4): 296–307.

18 

England J.L., Shakhnovich E.I. (2003) Structural determinant of protein designability. Phys. Rev. Lett. 90(21): 218101.

19 

Eswar N., John B., Mirkovic N., Fiser A., Ilyin V.A., Pieper U., Stuart A.C., Marti-Renom M.A., Madhusudhan M.S., Yerkovich B., Sali A. (2003) Tools for comparative protein structure modeling and analysis. Nucl. Acids Res. 31(13): 3375–3380.

20 

Ezhova T.A. (1999) Arabidopsis thaliana (L.) Heynh. kak model'ny0 ob''ekt dlia izycheniiageneticheskogo kontrolia morfogeneza [Arabidopsis thaliana (L.) Heynh. as a model object for studying genetic control of morphogenesis]. Genetika 35(11): 1522–1537.

21 

Fiser A. (2010) Template-based protein structure modeling. [in:] Methods in molecular biology (methods and protocols). Totowa: Humana Press.

22 

Gamage D.G., Gunaratne A., Periyannan G.R., Russell T.G. (2019) Applicability of instability index for in vitro protein stability prediction. Protein Pept. Lett. 26(5): 339–347.

23 

Gao B., Chen M., Li X., Zhang J. (2019) Ancient duplications and grass-specific transposition influenced the evolution of LEAFY transcription factor genes. Commun. Biol. 2(1): 1–10.

24 

Grandi V., Gregis V., Kater M.M. (2012) Uncovering genetic and molecular interactions among floral meristem identity genes in Arabidopsis thaliana. Plant J. 69(5): 881–893.

25 

Gupta C.L., Akhtar S., Bajpai P. (2014) In silico protein modeling: possibilities and Gupta limitations. EXCLI J. 13: 513–515.

26 

Guruprasad K., Reddy B.B., Pandit M.W. (1990) Correlation between stability of a protein and its dipeptide composition: a novel approach for predicting in vivo stability of a protein from its primary sequence. Protein Eng. Des. Sel. 4(2): 155–161.

27 

Hall B.G. (2013) Building phylogenetic trees from molecular data with MEGA. Mol. Biol. Evol. 30(5): 1229–1235.

28 

Hamès C., Ptchelkine D., Grimm C., Thevenon E., Moyroud E., Gérard F., Martiel J.L., Benlloch R., Parcy F., Müller C.W. (2008) Structural basis for LEAFY floral switch function and similarity with helix turn helix proteins. EMBO J. 27(19): 2628–2637.

29 

Hasani H.J., Barakat K. (2017) Homology modeling: an overview of fundamentals and tools. Int. Rev. Model. Simul. 10(2): 1–14.

30 

Hickok L.G., Warne T.R., Fribourg R.S. (1995) The biology of the fern Ceratopteris and its use as a model system. Int. J. Plant Sci. 156(3): 332–345.

31 

Hofer J., Turner L., Hellens R., Ambrose M., Matthews P., Michael A., Ellis N. (1997) UNIFOLIATA regulates leaf and flower morphogenesis in pea. Curr. Biol. 7(8): 581–587.

32 

Hsu C.C., Buehler M.J., Tarakanova A. (2020) The orderdisorder continuum: linking predictions of protein structure and disorder through molecular simulation. Sci. Rep. 10(1): 1–14.

33 

Ikai A. (1980) Thermostability and aliphatic index of globular proteins. J. Biochem. 88(6): 1895–1898.

34 

Ishida T., Kinoshita K. (2008) Prediction of disordered regions in proteins based on the meta approach. Bioinformatics 24(11): 1344–1348.

35 

Ittisoponpisan S., Islam S.A., Khanna T., Alhuzimi E., David A., Sternberg M.J. (2019) Can predicted protein 3D structures provide reliable insights into whether missense variants are disease associated? J. Mol. Biol. 431(11): 2197–2212.

36 

Kaur A., Pati P.K., Pati A.M., Nagpal A.K. (2020) Physicochemical characterization and topological analysis of pathogenesis-related proteins from Arabidopsis thaliana and Oryza sativa using in-silico approaches. PLoS One 15(9): e0239836.

37 

Kaur G., Pati P.K. (2018) In silico physicochemical characterization and topology analysis of respiratory burst oxidase homolog (Rboh) proteins from Arabidopsis and rice. Bioinformation 14(3): 93.

38 

Kc D.B. (2017) Recent advances in sequence-based protein structure prediction. Brief Bioinform. 18(6): 1021–1032.

39 

Kelley L.A., Mezulis S., Yates C.M., Wass M.N., Sternberg M.J. (2015) The Phyre2 web portal for protein modeling, prediction and analysis. Nat. Protoc. 10(6): 845–858.

40 

Kuhlman B., Bradley P. (2019) Advances in protein structure prediction and design. Nat. Rev. Mol. Cell Biol. 20(11): 681–697.

41 

Kumar S., Stecher G., Tamura K. (2016) MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 33: 1870–1874.

42 

Kyte J., Doolittle R.F. (1982) A simple method for displaying the hydropathic character of a protein. J. Mol. Biol. 157(1): 105–132.

43 

Lam S.D., Das S., Sillitoe I., Orengo C. (2017) An overview of comparative modelling and resources dedicated to large-scale modelling of genome sequences. Acta Crystallogr. D: Struct. Biol. 73(8): 628–640.

44 

Lee J., Freddolino P.L., Zhang Y. (2017) Ab initio protein structure prediction. [in:] From protein structure to function with bioinformatics. Ed. Rigden D.J., Dordrecht: Springer: 3–35.

45 

Lesk A.M., Chothia C.H. (1986) The response of protein structures to amino-acid sequence changes. Phil. Trans. R. Soc. A. 317(1540): 345–356.

46 

Martí-Renom M.A., Stuart A.C., Fiser A., Sánchez R., Melo F., Šali A. (2000) Comparative protein structure modeling of genes and genomes. Annu. Rev. Biophys. Biomol. Struct. 29(1): 291–325.

47 

Monniaux M., McKim S.M., Cartolano M., Thévenon E., Parcy F., Tsiantis M., Hay A. (2017) Conservation vs divergence in LEAFY and APETALA1 functions between Arabidopsis thaliana and Cardamine hirsuta. New Phytol. 216(2): 549–561.

48 

Moyroud E., Kusters E., Monniaux M., Koes R., Parcy F. (2010) LEAFY blossoms. Trends Plant Sci. 15(6): 346–352.

49 

Moyroud E., Monniaux M., Thévenon E., Dumas R., Scutt C.P., Frohlich M.W., Parcy F. (2017) A link between LEAFY and B gene homologues in Welwitschia mirabilis sheds light on ancestral mechanisms prefiguring floral development. New Phytol. 216(2): 469-481.

50 

Muhammed M.T., Aki Yalcin E. (2019) Homology modeling in drug discovery: overview, current applications, and future perspectives. Chem. Biol. Drug Des. 93(1): 12–20.

51 

Nagano K. (1973) Logical analysis of the mechanism of protein folding. I. Predictions of helices, loops and beta-structures from primary structure. J. Mol. Biol. 75: 401–420.

52 

Nemtseva E.V., Gerasimova M.A., Melnik T.N., Melnik B.S. (2019) Experimental approach to study the effect of mutations on the protein folding pathway. PloS One 14(1): e0210361.

53 

Nystedt B., Street N.R., Wetterbom A., Zuccolo A., Lin Y.C., Scofield D.G., Vezzi F., Delhomme N., Giacomello S., Alexeyenko A., Vicedomini R. (2013) The Norway spruce genome sequence and conifer genome evolution. Nature 497(7451): 579–584.

54 

Ortiz A.R., Kolinski A., Skolnick J. (1998) Fold assembly of small proteins using Monte Carlo simulations driven by restraints derived from multiple sequence alignments. J. Mol. Biol. 277(2): 419–448.

55 

Panchenko A.R., Marchler-Bauer A., Bryant S.H. (2000) Combination of threading potentials and sequence profiles improves fold recognition. J. Mol. Biol. 296(5): 1319–1331.

56 

Panda S., Chandra G. (2012) Physicochemical characterization and functional analysis of some snake venom toxin proteins and related non-toxin proteins of other chordates. Bioinformation 8(18): 891.

57 

Pergande M.R., Cologna S.M. (2017) Isoelectric point separations of peptides and proteins. Proteomes. 5(1): 4.

58 

Pieper U., Eswar N., Davis F.P., Braberg H., Madhusudhan M.S., Rossi A., Marti-Renom M., Karchin R., Webb B.M., Eramian D. Shen M.Y., (2006) MODBASE: a database of annotated comparative protein structure models and associated resources. Nucl. Acids Res. 34(suppl_1): D291–D295.

59 

Pirovano W., Heringa J. (2010) Protein secondary structure prediction. [in:] Data mining techniques for the life sciences. Ed. Carugo O., Eisenhaber F., Humana Press: Springer Protocols: 327–348.

60 

Plackett A.R., Conway S.J., Hazelton K.D.H., Rabbinowitsch E.H., Langdale J.A., Di Stilio V.S. (2018) LEAFY maintains apical stem cell activity during shoot development in the fern Ceratopteris richardii. eLife 7: e39625.

61 

Renzaglia K.S., Warne T.R. (1995) Ceratopteris: an ideal model system for teaching plant biology. Int. J. Plant Sci. 156(3): 385–392.

62 

Røgen P. (2021) Quantifying steric hindrance and topological obstruction to protein structure superposition. Algorithms Mol. Biol. 16(1): 1–19.

63 

Šali A., Blundell T.L. (1993) Comparative protein modelling by satisfaction of spatial restraints. J. Mol. Biol. 234(3): 779–815.

64 

Saravanan K.M., Selvaraj S. (2017) Dihedral angle preferences of amino acid residues forming various non-local interactions in proteins. J. Biol. Phys. 43(2): 265–278.

65 

Sayou C., Monniaux M., Nanao M.H., Moyroud E., Brockington S.F., Thévenon E., Wong G.K.S. (2014) A promiscuous intermediate underlies the evolution of LEAFY DNA binding specificity. Science 343(6171): 645–648.

66 

Sayou C., Nanao M.H., Jamin M., Posé D., Thévenon E., Grégoire L., Schmid M. (2016) A SAM oligomerization domain shapes the genomic binding landscape of the LEAFY transcription factor. Nat. Commun. 7: 11222.

67 

Shakhnovich B.E., Deeds E., Delisi C., Shakhnovich E. (2005) Protein structure and evolutionary history determine sequence space topology. Genome Res. 15(3): 385–392.

68 

Silva C.S., Puranik S., Round A., Brennich M., Jourdain A., Parcy F., Zubieta C. (2016) Evolution of the plant reproduction master regulators LFY and the MADS transcription factors: the role of protein structure in the evolutionary development of the flower. Front. Plant Sci. 6: 1193.

69 

Simons K.T., Strauss C., Baker D. (2001) Prospects for ab initio protein structural genomics. J. Mol. Biol. 306(5): 1191–1199.

70 

Siriwardana N.S., Lamb R.S. (2012) The poetry of reproduction: the role of LEAFY in Arabidopsis thaliana flower formation. Int. J. Dev. Biol. 56(4): 207–221.

71 

Skolnick J., Kihara D. (2001) Defrosting the frozen approximation: PROSPECTOR — a new approach to threading. Proteins: Struct. Funct. Bioinf. 42(3): 319–331.

72 

Studer G., Tauriello G., Bienert S., Waterhouse A.M., Bertoni M., Bordoli L., Schwede T., Lepore R. (2019) Modeling of protein tertiary and quaternary structures based on evolutionary information. [in:] Computational methods in protein evolution. Ed. Sikosek T., Humana Press, New York: Springer Protocols: 301–316

73 

Tanahashi T., Sumikawa N., Kato M., Hasebe M. (2005) Diversification of gene function: homologs of the floral regulator FLO/LFY control the first zygotic cell division in the moss Physcomitrella patens. Development 132(7): 1727–1736.

74 

Tran Q., Wallinga M. (2017) A novel method for multiple sequence alignment using morphing techniques. J. Health Inform. Manag. 1(2): 2.

75 

Uversky V.N. (2019) Intrinsically disordered proteins and their “mysterious” (meta) physics. Front. Phys. 7: 10.

76 

Uversky V.N., Oldfield C.J., Dunker A.K. (2005) Showing your ID: intrinsic disorder as an ID for recognition, regulation and cell signalling. J. Mol. Recognit. 18(5): 343–384.

77 

Van Der Lee R., Buljan M., Lang B., Weatheritt R.J., Daughdrill G.W., Dunker A.K., Fuxreiter M., Gough J., Gsponer J., Jones D.T., et al. (2014) Classification of intrinsically disordered regions and proteins. Chem. Rev. 114(13): 6589–6631.

78 

Vázquez-Lobo A., Carlsbecker A., Vergara-Silva F., Alvarez-Buylla E.R., Piñero D., Engström P. (2007) Characterization of the expression patterns of LEAFY/FLORICAULA and NEEDLY orthologs in female and male cones of the conifer genera Picea, Podocarpus, and Taxus: implications for current evo-devo hypotheses for gymnosperms. Evol. Dev. 9(5): 446–459.

79 

Venclovas Č. (2011) Methods for sequence–structure alignment. [in:] Homology modeling. Ed. Orry A., Abagyan R. Humana Press: Springer Protocols: 55–82.

80 

Venkatesan A., Gopal J., Candavelou M., Gollapalli, S., Karthikeyan K. (2013) A computational approach for protein structure prediction. Healthc. Inform. Res. 19(2): 137–147.

81 

Villimová V. (2012) Evolution of gene network controlling plant reproductive development. Biosci. Master Rev. Univ. Lyon.

82 

Wang L.L., Liang H.M., Pang J.L., Zhu M.Y. (2004) [Regulation network and biological roles of LEAFY in Arabidopsis thaliana in floral development ] Yi Chuan; Hereditas 26(1): 137–142.

83 

Wang L., Jiang T. (1994) On the complexity of multiple sequence alignment. J. Comput. Biol. 1(4): 337–348.

84 

Waterhouse A., Bertoni M., Bienert S., Studer G., Tauriello G., Gumienny R., Heer F.T., de Beer T.A.P., Rempfer C., Bordoli L., Schwede T. (2018) SWISS-MODEL: homology modelling of protein structures and complexes. Nucl. Acids Res. 46(W1): W296–W303.

85 

Weigel D., Alvarez J., Smyth D.R., Yanofsky M.F., Meyerowitz E.M. (1992) LEAFY controls floral meristem identity in Arabidopsis. Cell. 69(5): 843–859.

86 

Weigel D., Nilsson O. (1995) A developmental switch sufficient for flower initiation in diverse plants. Nature 377(6549): 495–500.

87 

Wilhelmsson P.K., Mühlich C., Ullrich K.K., Rensing S.A. (2017) Comprehensive genome-wide classification reveals that many plant-specific transcription factors evolved in streptophyte algae. Genome Biol. Evol. 9(12): 3384–3397.

88 

Xu Y., Liu Z., Cai L., Xu D. (2007) Protein structure prediction by protein threading. [in:] Computational methods for protein structure prediction and modeling. Ed. Xu Y., Xu D., Liang J. New York: Springer: 1–42.

Copyright: © 2022 Institute of Bioorganic Chemistry, Polish Academy of Sciences This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivs (CC BY-NC-ND) (https://creativecommons.org/licenses/by-nc-nd/3.0/legalcode),.allowing third parties to download and share its works but not commercially purposes or to create derivative works.
 
Stosujemy się do standardu HONcode dla wiarygodnej informacji zdrowotnej This site complies with the HONcode standard for trustworthy health information: verify here