Search
2025 Volume 5
Article Contents
ARTICLE   Open Access    

Unveiling the role of the TIR1/AFB gene family in Capsicum annuum: genome-wide identification, characterization, and transcriptomic analysis under Ralstonia solanacearum infection

  • # Authors contributed equally: Wenchao Du, Zhiyu Yu

More Information
  • TIR1/AFB proteins play key roles in plants responding to Solanum crop disease. In this study, five putative TIR1/AFB proteins were identified in pepper (Capsicum annuum). Chromosomal mapping revealed that CaTIR1/AFB genes are located on chromosomes DH02, DH03, DH04, and DH06. Phylogenetic tree analysis based on TIR1/AFB protein sequences from Arabidopsis, tomato, and eggplant revealed that the CaTIR1/AFB genes in pepper belong to three of the four identified subgroups. One collinear gene pair (CaTIR1C and CaTIR1B) was identified in pepper, suggesting the TIR1/AFB gene family in pepper experienced segmental duplications in evolution. Seventy-eight cis-elements related to the hormone, defense, drought, and light response were identified. RNA-sequencing data and expression profiling showed that TIR1/AFB genes were expressed differently under R. solanacearum treatment; downregulated expression of all CaTIR1/AFBs was observed in the resistant variety, while CaTIR1C was upregulated in the susceptible variety. Subcellular localization analysis showed that CaTIR1C is a nuclear-localized protein. Further virus-induced gene silencing (VIGS) assays revealed that knocking down CaTIR1C enhanced pepper resistance to R. solanacearum, indicating that CaTIR1C is a negative regulator in response to R. solanacearum. These findings contribute to the molecular breeding of bacterial wilt-resistant pepper and hold the potential to improve disease control efficiency by further exploring the function of CaTIR1C.
  • 加载中
  • Supplementary Table S1 Primers for qRT-PCR analysis.
    Supplementary Table S2 Primers for VIGS and subcellular location analysis.
    Supplementary Table S3 TIR1/AFB genes identified in Capsicum annuum.
    Supplementary Fig. S1 Transcriptomic expression analysis of the CaTIR1/AFB genes of resistant variety (R) and susceptible variety (S) under control (C) or R. Solanacearum treatment (T).
    Supplementary Fig. S2 Representative seedlings six days after 0.1 mL R. solanacearum injection (+R. solanacearum).
  • [1] Dezfulian MH, Jalili E, Roberto DKA, Moss BL, Khoo K, et al. 2016. Oligomerization of SCFTIR1 is essential for Aux/IAA degradation and auxin signaling in Arabidopsis. PLoS Genetics 12(9):e1006301 doi: 10.1371/journal.pgen.1006301

    CrossRef   Google Scholar

    [2] Takahashi K, Hayashi KI, Kinoshita T. 2012. Auxin activates the plasma membrane H+-ATPase by phosphorylation during hypocotyl elongation in Arabidopsis. Plant Physiology 159(2):632−41 doi: 10.1104/pp.112.196428

    CrossRef   Google Scholar

    [3] Jiang J, Zhu H, Li N, Batley J, Wang Y. 2022. The miR393-target module regulates plant development and responses to biotic and abiotic stresses. International Journal of Molecular Sciences 23(16):9477 doi: 10.3390/ijms23169477

    CrossRef   Google Scholar

    [4] Fendrych M, Akhmanova M, Merrin J, Glanc M, Hagihara S, et al. 2018. Rapid and reversible root growth inhibition by TIR1 auxin signalling. Nature Plants 4(7):453−59 doi: 10.1038/s41477-018-0190-1

    CrossRef   Google Scholar

    [5] Qi L, Kwiatkowski M, Chen H, Hoermayer L, Sinclair S, et al. 2022. Adenylate cyclase activity of TIR1/AFB auxin receptors in plants. Nature 611(7934):133−38 doi: 10.1038/s41586-022-05369-7

    CrossRef   Google Scholar

    [6] Du W, Lu Y, Li Q, Luo S, Shen S, et al. 2022. TIR1/AFB proteins: active players in abiotic and biotic stress signaling. Frontiers in Plant Science 13:1083409 doi: 10.3389/fpls.2022.1083409

    CrossRef   Google Scholar

    [7] Du W, Karamat U, Cao L, Li Y, Li H, et al. 2024. The TIR1/AFB family in Solanum Melongena: genome-wide identification and expression profiling under stresses and picloram treatment. Agronomy 14(7):1413 doi: 10.3390/agronomy14071413

    CrossRef   Google Scholar

    [8] Prigge MJ, Platre M, Kadakia N, Zhang Y, Greenham K, et al. 2020. Genetic analysis of the Arabidopsis TIR1/AFB auxin receptors reveals both overlapping and specialized functions. eLife 9:e54740 doi: 10.7554/elife.54740

    CrossRef   Google Scholar

    [9] Cui L, Zhang T, Li J, Lou Q, Chen J. 2014. Cloning and expression analysis of Cs-TIR1/AFB2: the fruit development-related genes of cucumber (Cucumis Sativus L.). Acta Physiologiae Plantarum 36(1):139−49 doi: 10.1007/s11738-013-1394-7

    CrossRef   Google Scholar

    [10] Ozga JA, Jayasinghege CPA, Kaur H, Gao LC, Nadeau CD, et al. 2022. Auxin receptors as integrators of developmental and hormonal signals during reproductive development in pea. Journal of Experimental Botany 73(12):4094−112 doi: 10.1093/jxb/erac152

    CrossRef   Google Scholar

    [11] Garrido-Vargas F, Godoy T, Tejos R, O'Brien JA. 2020. Overexpression of the auxin receptor AFB3 in Arabidopsis results in salt stress resistance and the modulation of NAC4 and SZF1. International Journal of Molecular Sciences 21(24):9528 doi: 10.3390/ijms21249528

    CrossRef   Google Scholar

    [12] Wojcik AM, Gaj MD. 2014. miR393 controls somatic embryogenesis in Arabidopsis through regulation of auxin signaling components (TIR1 and AFB2). Biotechnologia 95(1):123−24

    Google Scholar

    [13] Djami-Tchatchou AT, Harrison GA, Harper CP, Wang R, Prigge MJ, et al. 2020. Dual role of auxin in regulating plant defense and bacterial virulence gene expression during Pseudomonas Syringae PtoDC3000 pathogenesis. Molecular Plant-Microbe Interactions 33(8):1059−71 doi: 10.1094/MPMI-02-20-0047-R

    CrossRef   Google Scholar

    [14] Fousia S, Tsafouros A, Roussos PA, Tjamos SE. 2018. Increased resistance to Verticillium Dahliae in Arabidopsis plants defective in auxin signalling. Plant Pathology 67(8):1749−57 doi: 10.1111/ppa.12881

    CrossRef   Google Scholar

    [15] Su P, Zhao L, Li W, Zhao J, Yan J, et al. 2021. Integrated metabolo-transcriptomics and functional characterization reveals that the wheat auxin receptor TIR1 negatively regulates defense against Fusarium graminearum. Journal of Integrative Plant Biology 63(2):340−52 doi: 10.1111/jipb.12992

    CrossRef   Google Scholar

    [16] Kim T, Al Mijan M, Lee J, Yun J, Chung JH, et al. 2024. Essential oils for the treatment and management of nonalcoholic fatty liver disease (NAFLD). Natural Product Communications 19(4):1−7 doi: 10.1177/1934578x241250248

    CrossRef   Google Scholar

    [17] Sood T, Sood S, Sood VK, Badiyal A, Anuradha, et al. 2023. Assessment and validation of resistance to bacterial wilt (Ralstonia Solanacearum) through field and molecular studies in bell pepper. Journal of Plant Pathology 105(3):849−57 doi: 10.1007/s42161-023-01378-1

    CrossRef   Google Scholar

    [18] Shen L, Yang S, Yang F, Guan D, He S. 2020. CaCBL1 acts as a positive regulator in pepper response to Ralstonia Solanacearum. Molecular Plant-Microbe Interactions 33(7):945−57 doi: 10.1094/MPMI-08-19-0241-R

    CrossRef   Google Scholar

    [19] Shi L, Li X, Weng Y, Cai H, Liu K, et al. 2022. The CaPti1–CaERF3 module positively regulates resistance of Capsicum Annuum to bacterial wilt disease by coupling enhanced immunity and dehydration tolerance. The Plant Journal 111(1):250−68 doi: 10.1111/tpj.15790

    CrossRef   Google Scholar

    [20] Yang S, Cai W, Wu R, Huang Y, Lu Q, et al. 2023. Differential CaKAN3-CaHSF8 associations underlie distinct immune and heat responses under high temperature and high humidity conditions. Nature Communications 14(1):4477 doi: 10.1038/s41467-023-40251-8

    CrossRef   Google Scholar

    [21] Yang S, Shi Y, Zou L, Huang J, Shen L, et al. 2020. Pepper CaMLO6 negatively regulates Ralstonia Solanacearum resistance and positively regulates high temperature and high humidity responses. Plant and Cell Physiology 61(7):1223−38 doi: 10.1093/pcp/pcaa052

    CrossRef   Google Scholar

    [22] Yang S, Cai W, Shen L, Cao J, Liu C, et al. 2022. A CaCDPK29–CaWRKY27b module promotes CaWRKY40-mediated thermotolerance and immunity to Ralstonia solanacearum in pepper. New Phytologist 233(4):1843−63 doi: 10.1111/nph.17891

    CrossRef   Google Scholar

    [23] Zhang K, Wang X, Chen S, Liu Y, Zhang L, et al. 2025. The gap-free assembly of pepper genome reveals transposable-element-driven expansion and rapid evolution of pericentromeres. Plant Communications 6(2):101177 doi: 10.1016/j.xplc.2024.101177

    CrossRef   Google Scholar

    [24] Chen W, Wang X, Sun J, Wang X, Zhu Z, et al. 2024. Two telomere-to-telomere gapless genomes reveal insights into Capsicum evolution and capsaicinoid biosynthesis. Nature Communications 15(1):4295 doi: 10.1038/s41467-024-48643-0

    CrossRef   Google Scholar

    [25] Kwon JS, Nam JY, Yeom SI, Kang WH. 2021. Leaf-to-whole plant spread bioassay for pepper and Ralstonia solanacearum interaction determines inheritance of resistance to bacterial wilt for further breeding. International Journal of Molecular Sciences 22(5):2279 doi: 10.3390/ijms22052279

    CrossRef   Google Scholar

    [26] Mazumder R, Natale DA, Murthy S, Thiagarajan R, Wu CH. 2005. Computational identification of strain-, species- and genus-specific proteins. BMC Bioinformatics 6:279 doi: 10.1186/1471-2105-6-279

    CrossRef   Google Scholar

    [27] Waterhouse A, Bertoni M, Bienert S, Studer G, Tauriello G, et al. 2018. SWISS-MODEL: homology modelling of protein structures and complexes. Nucleic Acids Research 46(W1):W296−W303 doi: 10.1093/nar/gky427

    CrossRef   Google Scholar

    [28] Liu Y, Gao Y, Chen M, Jin Y, Qin Y, et al. 2023. GIFTdb: a useful gene database for plant fruit traits improving. The Plant Journal 116(4):1030−40 doi: 10.1111/tpj.16506

    CrossRef   Google Scholar

    [29] Kumar S, Stecher G, Li M, Knyaz C, Tamura K. 2018. MEGA X: molecular evolutionary genetics analysis across computing platforms. Molecular Biology and Evolution 35(6):1547−49 doi: 10.1093/molbev/msy096

    CrossRef   Google Scholar

    [30] Chen C, Chen H, Zhang Y, Thomas HR, Frank MH, et al. 2020. TBtools: an integrative toolkit developed for interactive analyses of big biological data. Molecular Plant 13(8):1194−202 doi: 10.1016/j.molp.2020.06.009

    CrossRef   Google Scholar

    [31] Wan H, Ni Z, Wang Y, Yu Y. 2025. The gma-miR164a/GmNAC115 module participates in the adaptation of soybean to drought and salt stress by influencing reactive oxygen species scavenging. Plant Physiology and Biochemistry 227:110191 doi: 10.1016/j.plaphy.2025.110191

    CrossRef   Google Scholar

    [32] Bailey TL, Boden M, Buske FA, Frith M, Grant CE, et al. 2009. MEME Suite: tools for motif discovery and searching. Nucleic Acids Research 37:W202−W208 doi: 10.1093/nar/gkp335

    CrossRef   Google Scholar

    [33] Lu Y, Irshad A, Rehman SU, Wang Y, Zhou B, et al. 2024. Connecting the dots between GmPERK-1 and enhanced grain weight in Glycine max. Agronomy 14(8):1679 doi: 10.3390/agronomy14081679

    CrossRef   Google Scholar

    [34] Wang X, Luo S, Li Q, Song L, Zhang W, et al. 2022. Delphinidins and naringenin chalcone underlying the fruit color changes during maturity stages in eggplant. Agronomy 12(5):1036 doi: 10.3390/agronomy12051036

    CrossRef   Google Scholar

    [35] Santangelo KS, Nuovo GJ, Bertone AL. 2012. In vivo reduction or blockade of interleukin-1β in primary osteoarthritis influences expression of mediators implicated in pathogenesis. Osteoarthritis and Cartilage 20(12):1610−18 doi: 10.1016/j.joca.2012.08.011

    CrossRef   Google Scholar

    [36] Choi HW, Hwang BK. 2015. Molecular and cellular control of cell death and defense signaling in pepper. Planta 241(1):1−27 doi: 10.1007/s00425-014-2171-6

    CrossRef   Google Scholar

    [37] Hong JK, Hwang IS, Hwang BK. 2017. Functional roles of the pepper leucine-rich repeat protein and its interactions with pathogenesis-related and hypersensitive-induced proteins in plant cell death and immunity. Planta 246(3):351−64 doi: 10.1007/s00425-017-2709-5

    CrossRef   Google Scholar

    [38] Singh AK, Ghosh D, Chakraborty S. 2022. Optimization of Tobacco Rattle Virus (TRV)-based virus-induced gene silencing (VIGS) in tomato. In Plant Gene Silencing, eds Mysore KS, Senthil-Kumar M. New York, NY: Humana. Volume 2408. pp. 133−45 doi: 10.1007/978-1-0716-1875-2_9
    [39] Ho F, Chen Y, Lin Y, Cheng C, Wang J. 2009. A tobacco rattle virus-induced gene silencing system for a soil-borne vascular pathogen Ralstonia solanacearum. Botanical Studies 50(4):413−24

    Google Scholar

    [40] Senthil-Kumar M, Mysore KS. 2011. Virus-induced gene silencing can persist for more than 2 years and also be transmitted to progeny seedlings in Nicotiana benthamiana and tomato. Plant Biotechnology Journal 9(7):797−806 doi: 10.1111/j.1467-7652.2011.00589.x

    CrossRef   Google Scholar

    [41] Guo F, Huang Y, Qi P, Lian G, Hu X, et al. 2021. Functional analysis of auxin receptor OsTIR1/OsAFB family members in rice grain yield, tillering, plant height, root system, germination, and auxinic herbicide resistance. New Phytologist 229(5):2676−92 doi: 10.1111/nph.17061

    CrossRef   Google Scholar

    [42] Fang YN, Yang XM, Jiang N, Wu XM, Guo WW. 2020. Genome-wide identification and expression profiles of phased siRNAs in a male-sterile somatic cybrid of pummelo (Citrus grandis). Tree Genetics & Genomes 16(3):46 doi: 10.1007/s11295-020-01437-z

    CrossRef   Google Scholar

    [43] Yang C, Deng W, Tang N, Wang X, Yan F, et al. 2013. Overexpression of ZmAFB2, the maize homologue of AFB2 gene, enhances salt tolerance in transgenic tobacco. Plant Cell, Tissue and Organ Culture 112(2):171−79 doi: 10.1007/s11240-012-0219-5

    CrossRef   Google Scholar

    [44] Cai Z, Zeng DE, Liao J, Cheng C, Ali Sahito Z, et al. 2019. Genome-wide analysis of auxin receptor family genes in Brassica juncea var. tumida. Genes 10(2):165 doi: 10.3390/genes10020165

    CrossRef   Google Scholar

    [45] Greenspan NS. 2011. Attributing functions to genes and gene products. Trends in Biochemical Sciences 36(6):293−97 doi: 10.1016/j.tibs.2010.12.005

    CrossRef   Google Scholar

    [46] Eitas TK, Dangl JL. 2010. NB-LRR proteins: pairs, pieces, perception, partners, and pathways. Current Opinion in Plant Biology 13(4):472−77 doi: 10.1016/j.pbi.2010.04.007

    CrossRef   Google Scholar

    [47] Elliott KT, Cuff LE, Neidle EL. 2013. Copy number change: evolving views on gene amplification. Future Microbiology 8(7):887−99 doi: 10.2217/fmb.13.53

    CrossRef   Google Scholar

    [48] Ronald PC. 1998. Resistance gene evolution. Current Opinion in Plant Biology 1(4):294−98 doi: 10.1016/1369-5266(88)80049-9

    CrossRef   Google Scholar

    [49] To CC, Vohradsky J. 2007. A parallel genetic algorithm for single class pattern classification and its application for gene expression profiling in Streptomyces coelicolor. BMC Genomics 8:49 doi: 10.1186/1471-2164-8-49

    CrossRef   Google Scholar

    [50] Wang S, Bai Y, Li P, Yang L, Wang X. 2019. Genome-wide identification and expression analysis of the dof (DNA binding with one finger) protein family in monocot and dicot species. Physiological and Molecular Plant Pathology 108:101431 doi: 10.1016/j.pmpp.2019.101431

    CrossRef   Google Scholar

    [51] Yamauchi Y, Matsuda A, Matsuura N, Mizutani M, Sugimoto Y. 2018. Transcriptome analysis of Arabidopsis Thaliana treated with green leaf volatiles: possible role of green leaf volatiles as self-made damage-associated molecular patterns. Journal of Pesticide Science 43(3−4):207−13 doi: 10.1584/jpestics.D18-020

    CrossRef   Google Scholar

    [52] Qiao Z, Li H, Wang X, Ji X, You C. 2023. Genome-wide identification of apple auxin receptor family genes and functional characterization of MdAFB1. Horticultural Plant Journal 9(4):645−58 doi: 10.1016/j.hpj.2023.02.001

    CrossRef   Google Scholar

    [53] Zhang L, Yu G, Xue H, Li M, Lozano-Durán R, et al. 2024. Ralstonia solanacearum alters root developmental programmes in auxin-dependent and -independent manners. Molecular Plant Pathology 25(12):e700743 doi: 10.1111/mpp.70043

    CrossRef   Google Scholar

    [54] French E, Kim BS, Rivera Zuluaga K, Iyer-Pascuzzi AS. 2018. Whole root transcriptomic analysis suggests a role for auxin pathways in resistance to Ralstonia solanacearum in tomato. Molecular Plant-Microbe Interactions 31(4):432−44 doi: 10.1094/MPMI-08-17-0209-R

    CrossRef   Google Scholar

    [55] Garcia AL, Lima WG, Souza EB, Michereff SJ, Mariano RLR. 2013. Characterization of Ralstonia solanacearum causing bacterial wilt in bell pepper in the State of Pernambuco, Brazil. Journal of Plant Pathology 95(2):237−45

    Google Scholar

  • Cite this article

    Du W, Yu Z, Miao X, Teng Z, Zhang C, et al. 2025. Unveiling the role of the TIR1/AFB gene family in Capsicum annuum: genome-wide identification, characterization, and transcriptomic analysis under Ralstonia solanacearum infection. Vegetable Research 5: e034 doi: 10.48130/vegres-0025-0028
    Du W, Yu Z, Miao X, Teng Z, Zhang C, et al. 2025. Unveiling the role of the TIR1/AFB gene family in Capsicum annuum: genome-wide identification, characterization, and transcriptomic analysis under Ralstonia solanacearum infection. Vegetable Research 5: e034 doi: 10.48130/vegres-0025-0028

Figures(8)  /  Tables(1)

Article Metrics

Article views(1013) PDF downloads(343)

ARTICLE   Open Access    

Unveiling the role of the TIR1/AFB gene family in Capsicum annuum: genome-wide identification, characterization, and transcriptomic analysis under Ralstonia solanacearum infection

Vegetable Research  5 Article number: e034  (2025)  |  Cite this article

Abstract: TIR1/AFB proteins play key roles in plants responding to Solanum crop disease. In this study, five putative TIR1/AFB proteins were identified in pepper (Capsicum annuum). Chromosomal mapping revealed that CaTIR1/AFB genes are located on chromosomes DH02, DH03, DH04, and DH06. Phylogenetic tree analysis based on TIR1/AFB protein sequences from Arabidopsis, tomato, and eggplant revealed that the CaTIR1/AFB genes in pepper belong to three of the four identified subgroups. One collinear gene pair (CaTIR1C and CaTIR1B) was identified in pepper, suggesting the TIR1/AFB gene family in pepper experienced segmental duplications in evolution. Seventy-eight cis-elements related to the hormone, defense, drought, and light response were identified. RNA-sequencing data and expression profiling showed that TIR1/AFB genes were expressed differently under R. solanacearum treatment; downregulated expression of all CaTIR1/AFBs was observed in the resistant variety, while CaTIR1C was upregulated in the susceptible variety. Subcellular localization analysis showed that CaTIR1C is a nuclear-localized protein. Further virus-induced gene silencing (VIGS) assays revealed that knocking down CaTIR1C enhanced pepper resistance to R. solanacearum, indicating that CaTIR1C is a negative regulator in response to R. solanacearum. These findings contribute to the molecular breeding of bacterial wilt-resistant pepper and hold the potential to improve disease control efficiency by further exploring the function of CaTIR1C.

    • The SCF (SKP1-CUL1-F-box protein) E3 ubiquitin-ligase complexes, in combination with F-box proteins TRANSPORT INHIBITOR RESPONSE 1/AUXIN-SIGNALING F-BOX (TIR1/AFB), play a critical role in the bio-process of auxin signal transduction in plants, involving many auxin-mediated responses through transcriptional regulation[1]. The substrate specificity of SCFTIR1/AFB is conferred by the interchangeable F-box protein subunits of the TIR1/AFB family, functioning as auxin coreceptors that interact with the auxin/indole-3-acetic acid (Aux/IAA) transcriptional regulators[2,3]. Auxin acts as a 'molecular glue', strengthening its interaction with auxin receptors TIR1/AFBs, degrading Aux/IAA, releasing auxin-responsive factors (ARFs), and activating the expression of genes involved in auxin-dependent processes[4,5].

      Characterized by the structures of F-Box domain and a transport inhibitor response 1 protein domain, TIR1/AFB proteins are identified widely in planta, while phylogenetic studies have found that the TIR1/AFB proteins are conserved across land plant lineages and can be categorized into four clades: TIR1/AFB1, AFB2/AFB3, AFB4/5, and AFB6[6]. Among these TIR1/AFB proteins, AFB6 was identified in Solanaceae species such as tomato (Solanum lycopersicum) and eggplant (S. melongena) but not in Arabidopsis[7,8], while AFB2/3 proteins have not yet been identified or characterized in Solanaceae species.

      The TIR1/AFB proteins mediate diverse responses to plant growth and abiotic stress, including embryogenesis, leaf development, fruit development, and salt stress[912]. In terms of biotic stress, TIR1/AFB mutants have also been reported with varied responses to disease infestation. The Arabidopsis quadruple mutant tir1 afb1 afb4 afb5 was observed with increased IAA levels along with higher susceptibility to Pseudomonas Syringae PtoDC3000[13]. Compared with Col-0, the afb1 and afb3 mutants of Arabidopsis showed partial resistance to Verticillium dahliae, with defense-related genes such as PR1 and PDF1.2 upregulated[14]. In addition, RNAi-mediated gene knockdown of TaTIR1 increased wheat resistance to Fusarium head blight[15].

      Pepper (Capsicum annuum) is one of the most economically important vegetable crops of the Solanaceae family, traditionally cultivated and consumed worldwide both as a vegetable and as a seasoning ingredient owing to its health benefits derived from abundant and diverse bioactive compounds[16]. Despite its economic value, the pepper industry is severely threatened by bacterial wilt caused by Ralstonia solanacearum[17]. Several genes, such as CaWRKY40, have been identified for their roles in pepper's response to R. solanacearum infestation[1822], but none of the CaTIR1/AFBs have been reported to be involved in response to R. solanacearum in pepper. Research in the pepper genome over the past decade has led to significant insights in this field, while the recent advances in the leaf-injection method for inoculating R. solanacearum and the leaf-to-whole-plant spread bioassay have enabled the accurate timing of inoculation, making it possible to investigate the in-time molecular responses in pepper during R. solanacearum infection[2325].

      In this study, a genome-wide analysis of TIR1/AFB genes in resistant and susceptible pepper varieties was performed to evaluate their expression under R. solanacearum infestation. The CaTIR1/AFBs were subsequently subjected to bioinformatic analysis, qRT-PCR experiment, transient expression in tobacco, and virus-induced gene silencing (VIGS) assays. All the results provided a theoretical basis for studying the characteristics and functions of TIR1/AFBs proteins in pepper when responding to bacterial wilt disease, and could support further research efforts utilizing these key auxin-response genes in pepper breeding.

    • Putative TIR1/AFB genes (PepperBase (bioinformaticslab.cn)) were retrieved from the pepper genome database[24] based on BLASTP analyses with the criteria of e-value < 0.01, while the previously identified six Arabidopsis TIR1/AFB genes (At1g12820, At3g26810, At3g26980, At4g03190, At4g24390, At5g49980), four tomato TIR1/AFB genes, and five eggplant TIR1/AFB were adopted as reference queries[7,8,26]. Six putative CaTIR1/AFB genes were found (Supplementary Table S1). Further screening was done based on the kinds of Protein Family Analysis and Modeling (PF, PF18791, transport inhibitor response 1 protein domain; PF18511, F-box) (http://pfam.xfam.org/) in these six members. Parameters including molecular weight, amino acid length, isoelectric point, and grand average of hydropathicity (GRAVY) were used for further characterization of CaTIR1/AFB proteins through the ExPASy-ProtParam tool[27]. In silico protein subcellular localization was predicted with WoLF PSORT (https://wolfpsort.hgc.jp/)[28]. The TBtools 'Show Genes on Chromosome' function was used to visualize chromosomal locations of all obtained TIR1/AFB genes.

    • Protein sequences of the CaTIR1/AFB genes in Arabidopsis, tomato, eggplant, and pepper were adopted to generate phylogenetic trees using the MEGA X (www.megasoftware.net) software[29]. Sequences were aligned using multiple sequence alignment and clustered by the neighbor-joining algorithm, with 1,000 bootstrap replicates.

    • The duplicated genes were retrieved and analyzed on the TBtools platform (https://github.com/CJ-Chen/TBtools), using BLAST, MCScanX, and Advanced Circos features. The synteny relationships within pepper were established by retrieving collinearity files with MCScanX, followed by generating dual synteny graphics through TBtools[30].

    • The cis-elements of CaTIR1/CaAFBs were identified from the sequences at 2 kb upstream of the start codon using the web tools in the pepper genome database (PepperBase; www.bioinformaticslab.cn/PepperBase), and analyzed using the PlantCARE web-based tool (http://bioinformatics.psb.ugent.be/webtools/plantcare/html/)[31], with the findings visualized by TBtools.

    • The identification of protein-conserved domains was performed on the NCBI conserved domain online server (www.ncbi.nlm.nih.gov), while their motifs were analyzed using the MEME Suite (meme-suite.org)[32]. The exon and intron structural prediction of the CaTIR1/AFB gene was conducted with the Gene Structure Display Server (GSDS) (http://gsds.cbi.pku.edu.cn/)[33], using the genomic sequences and CDS of five CaTIR1/AFB genes.

    • The inbred pepper lines 'HP-19' (resistant) and 'HP-21' (susceptible) were obtained from the State Key Laboratory for Vegetable Germplasm Enhancement and Utilization of Hebei Vegetable Germplasm Resource Centre, Baoding, China, and R. solanacearum strains were provided by the Fujian Academy of Agricultural Sciences, Fuzhou, China. Pepper seedlings were prepared and inoculated at the six-leaf stage in the plant-growing facilities at Baoding University. The seedlings were grown in an incubator (PGX-350L, Ningbo Saifu Experimental Instrument Co., Ltd, China) at 37 °C with a 16/8 h light/dark photoperiod under 80% humidity. The fourth leaves were inoculated through inocula injection with 0.05 mL R. solanacearum (OD600 = 0.3) suspension in 10 mm/L MgCl2 or the MgCl2 solution as a control, and sampled at 6 h after inoculation for transcriptomic analyses. Leaf samples were collected from at least three plants and instantly frozen in liquid nitrogen, then stored at −80 °C until RNA extraction. Three technical replicates were performed for each morphotype.

    • Three grams of each leaf sample were collected for RNA extraction using poly-T oligo-attached magnetic beads, and a NEBNext® Ultra™ RNA Library Prep Kit for Illumina® was used to construct libraries (NEB, Ipswich, MA, USA). Sequencing was performed on an Illumina HiSeq 4000 platform to obtain the 150 bp paired-end reads. Differentially expressed genes (DEGs) were identified from the differential expression analysis using the DESeq R package and the hierarchical cluster analysis. The significant DEGs were filtered based on false discovery rate (FDR) 0.01 and |log2 fold change| 1 in each pairwise comparison, and a heat map of expression profiles was constructed using TBtools[34]. Each qRT-PCR analysis included three biological and four technical replicates, respectively, while the 2−ΔΔCᴛ value was determined for the calculation of gene expression[35]. The gene-specific primers for qRT-PCR, subcellular location, and VIGS are listed in Supplementary Tables S1 and S2, annotated and designed based on the pepper genome database (PepperBase; bioinformaticslab.cn).

    • The 35S::CaTIR1C-GFP expression vector was constructed by cloning the full-length coding sequence of CaTIR1C into the pMDC43 plasmid, which was subsequently transferred into Agrobacterium tumefaciens strain GV3101 (Weidi, Shanghai, China) for transient transformation of young tobacco leaves. The subcellular localization of CaTIR1C-GFP was examined using a confocal laser scanning microscope (FV10i, Olympus; Leica Microsystems, Germany).

    • In previous studies on pepper resistance to diseases, virus-induced gene silencing (VIGS) showed promising results for the validation of gene functions[36,37]. In this experiment, pTRV1 and pTRV2 VIGS vectors were adopted for gene silencing[3840], with CaCHlH employed as the gene marker in VIGS assays. About 300 bp synthetic sequence designed by the VIGS tool (https://vigs.solgenomics.net/) corresponding to part of the CaCHlH and CaTIR1C genes was selected and cloned into the pTRV2 vector. The empty vector pTRV2 was used as the control. The constructed TRV-mediated VIGS were then applied to infect one-week-old pepper seedlings, followed by 0.1 mL R. solanacearum (OD600 = 0.3) injected into the fourth leaf in six-leaf stage pepper plants. The experiment included three independent replicates and was repeated twice.

    • Using the six Arabidopsis genes, four tomato genes, and five eggplant genes as reference queries[4], six putative TIR1/AFB genes were identified in pepper (Supplementary Table S3). Candidate genes containing at least one F-Box (PF18511) and one TIR1 protein domain (PF18791) were considered as 'true' TIR1/AFBs. Therefore, five genes were kept for further analysis. Predicted genes and related information were presented in Table 1.

      Table 1.  Predicted pepper TIR1/AFBs genes and related information.

      Gene ID Instability index Amino acids GRAVY score Predicted subcellular location Aliphatic index Annotation
      CaT2T02g02335 37.89 583 −0.058 Nucleus 91.97 CaAFB6
      CaT2T03g04037 53.39 581 −0.026 Nucleus 95.51 CaTIR1C
      CaT2T04g01000 42.25 646 0.051 Nucleus 91.02 CaAFB4/5
      CaT2T06g00038 51.41 580 0.015 Nucleus 96.50 CaTIR1B
      CaT2T06g03723 47.65 658 0.047 Nucleus 98.80 CaTIR1A

      The average length of these CaTIR1/AFB proteins was 609 amino acids (aa), ranging from 580 aa (CaT2T06g00038) to 658 aa (CaT2T06g03723). These putative TIR1/AFB genes are located on four chromosomes, while two of them were identified on DH06 (Fig. 1).

      Figure 1. 

      Chromosomal location of the five CaTIR1/AFB genes.

    • Phylogenetic studies based on the maximum likelihood algorithm were conducted to investigate the evolutionary relationship of CaTIR1/AFB genes from pepper. The phylogenetic tree was constructed using five TIR1/AFB genes from pepper, five from eggplant, five from Arabidopsis, and four from tomato, while these genes were assigned to four groups: TIR1/AFB1 (Group 1), AFB2/3 (Group 2), AFB4/5 (Group 3), and AFB6 (Group 4) (Fig. 2). The TIR1/AFB1 clade included three CaTIR1/AFB genes from pepper, while AFB4/5 and AFB6 each included one CaTIR1/AFB. Notably, the AFB6 subgroup was present in pepper, eggplant, and tomato but absent in Arabidopsis, suggesting that this subgroup evolved specifically within Solanaceae species. Furthermore, none of the genes from Solanaceae species were identified as closely related to the AFB2/3 subgroup, indicating a possible genetic loss of this subgroup in their genome.

      Figure 2. 

      Phylogenetic analyses of TIR1/AFB proteins identified in pepper, Arabidopsis, tomato, and eggplant. Protein sequences of the five TIR1/AFBs from pepper were used to construct a phylogenetic tree under the neighbor-joining algorism with 1,000 bootstrap replicates. The TIR1/AFBs were clustered into four groups (Groups I, II, III, and IV) with each highlighted in a different color.

    • The exon-intron structure analyses of the CaTIR1/AFB genes showed that, although these genes exhibited structural divergence, members within the same phylogenetically identified subgroup shared similar gene structures (Fig. 3a). CaTIR1B and CaTIR1C within the TIR1/AFB subgroup are notably similar, particularly in their CDSs. To further investigate functional characteristics, ten conserved motifs were identified in the CaTIR1/AFBs using the MEME algorithm. While all genes contain motifs of F-box, TIR1, and leucine-rich repeat (LRR), CaAFB6 uniquely featured a motif with an LRR_CC_2 domain (Fig. 3b). Additionally, untranslated regions (UTRs) were identified in CaTIR1C, CaAFB4/5, and CaAFB6, which could potentially differentiate their regulatory or functional roles.

      Figure 3. 

      Gene (left) structure and conserved domain (right) of CaTIR1/AFBs.

    • Tandem and segmental gene duplication in CaTIR1/AFBs was subject to BLASTP searches and analyses on TBtools. Out of five CaTIR1/AFB genes, one collinear gene pair, CaTIR1C and CaTIR1B, was identified, which is consistent with the result of the gene structure analysis (Fig. 4), indicating segmental duplication could be important in the expansion of the CaTIR1/AFB family genes.

      Figure 4. 

      Circos figure for the chromosome locations with the CaTIR1/AFBs segmental duplication links. Red lines indicate the syntenic cyclin gene pairs between the mentioned genes. Gray lines in the backdrop represent the syntenic cyclin gene pairs of other genes in pepper.

    • Sequences from the 2 kb upstream of the translation start sites of CaTIR1/AFBs were analyzed on the PlantCARE web tool, and the cis-elements in the promoters of these genes were identified and compared to explore the potential roles of CaTIR1/AFBs in regulatory mechanisms such as plant hormone signaling, defense, stress, and light responses. Six classes of cis-elements were identified, including responsive elements of light, drought, defense, auxin, abscisic acid (ABA), and gibberellin (GA). Light-responsive elements were identified in the upstream region of all five genes. However, auxin-responsive elements were found only in the upstream of CaAFB4/5 and CaTIR1A, while defense-responsive elements were identified only in the upstream of CaTIR1C and CaAFB6. Similarly, ABA-responsive and drought-responsive elements were only identified in the upstream of CaTIR1C and CaAFB6, respectively (Fig. 5).

      Figure 5. 

      Identified cis-elements in the promoters at 2 kb upstream of the CaTIR1/AFBs. Hormone, drought, light, and stress-responsive elements were identified and linked to cis-elements of the TIR1/AFBs genes.

    • Transcriptomic analyses showed that CaTIR1/AFB genes expressed differently in resistant and susceptible pepper varieties in response to R. solanacearum infection (Supplementary Fig. S1). All CaTIR1/AFB genes were downregulated in the resistant variety, whereas in the susceptible variety, expression levels varied among the genes (Supplementary Fig. S1; Fig. 6). Notably, CaTIR1C was significantly upregulated in the susceptible variety, suggesting that CaTIR1C may act as a positive regulator in response to R. solanacearum. These expression trends were further validated by quantitative real-time PCR (qRT-PCR), which showed general consistency with the RNA-seq data (Fig. 6). Significant expression difference of CaTIR1C between the R. solanacearum and control treatments in susceptible pepper was also confirmed (Fig. 6). The UBI quitin-like gene CaT2T06g03488 was used as the housekeeping gene for normalization.

      Figure 6. 

      Expression of CaTIR1/AFB genes under control or R. solanacearum treatment as measured by qRT-PCR (orange symbols) and RNA-seq (bars). For qRT-PCR, gene expression levels were normalized to those of CaUBI, and qRT-PCR expression levels in the resistant variety treated with control (R-C) were set to 1. Results were presented as the means ± SE of three biological replicates. * p < 0.05 (Student's t-test, qRT-PCR data).

    • Previous studies in Arabidopsis have shown that TIR1/AFB family members are localized to the nucleus[8]. To determine the subcellular location of CaTIR1C, the full-length coding region of CaTIR1C was cloned into pMDC43. The plasmids were transformed into Agrobacterium strain GV3101 (Weidi, Shanghai, China). The GV3101 with target plasmids was incubated with Kanamycin at 28 °C overnight in an infiltration buffer consisting of 10 mM MgCl2 (pH 5.7) and 150 mM acetosyringone. The 0.5 mL bacterial fluid (OD600 = 0.5) was injected into a tobacco leaf epidermal cell. After 48–72 h of incubation in darkness, the cells were observed using a confocal laser scanning microscope (Olympus FV10I; Leica Microsystems, Germany). The result revealed that the constructed 35S::CaTIR1C-GFP was selectively enriched in the nuclei of tobacco leaf epidermal cells (Fig. 7), suggesting that CaTIR1C may perform important nuclear functions related to its regulatory role.

      Figure 7. 

      Subcellular location and expression patterns of CaTIR1C in tobacco epidermal cells. The first column shows the fluorescence of the 35S::GFP, 35S::CaTIR1C-GFP proteins; the second column shows a bright-field image of the 35S::GFP, 35S::CaTIR1C-GFP proteins; the third column shows a bright-field image of the proteins; the fourth column represents an overlay of the fluorescent images. Bar = 20 μm.

    • The VIGS experiment showed that plants transfected with TRV2::CaTIR1C effectively silenced CaTIR1C, with significantly decreased CaTIR1C expression compared to those transfected with the TRV2 empty vector (TRV2::00). Seedlings transfected with TRV2::CaChlH turned yellow-white 25 d after injection (Fig. 8a, b). Six days after R. solanacearum inoculation, control plants exhibited the typical symptom of wilted leaves, whereas CaTIR1C-silenced plants showed resistance with reduced disease symptoms (Fig. 8b; Supplementary Fig. S2). These results showed that CaTIR1C plays a negative role in regulating R. solanacearum resistance.

      Figure 8. 

      Effects of CaTIR1C on pepper susceptibility to R. solanacearum. (a) Expression of CaTIR1C in leaves at 25 d after VIGS infection. Asterisks indicate statistical significance as determined by the two-tailed Student's t-test: **p < 0.01. (b) Young pepper plants at 6 d after the inoculation with 0.1 mL R. solanacearum injection (+ R. solanacearum). Bar = 0.5 cm.

    • With the development of whole genome sequencing, TIR1/AFB families have been identified in more plant species in recent years[6]. As auxin co-receptor proteins, TIR1/AFBs were reported to regulate many aspects of plant growth, development, and responses to salt stress[4143]. The present study conducted the first genome-wide analysis on TIR1/AFB family in pepper.

      In plant genomes, the number of TIR1/AFB genes varies. The Arabidopsis genome contains six TIR1/AFBs, divided into three subgroups: TIR1/AFB1 (TIR1, AFB1), AFB2/3 (AFB2, AFB3), and AFB4/5 (AFB4, AFB5)[8]. In Brassica juncea var. tumida, 18 TIR1/AFB genes have been identified and could be clustered into four subgroups, including TIR1/AFB1, AFB2/3, AFB4/5, and AFB6[44]. In Solanaceae crops, four TIR1/AFB genes have been reported in tomatoes, and five TIR1/AFB genes in eggplant. The present analysis identified five TIR1/AFB genes in the pepper genome (Table 1), which is similar to eggplant but one more than in tomato. These findings suggest that while Solanaceae species share a relatively conserved TIR1/AFB gene count, lineage-specific expansion and diversification.

      Extensive functional overlapping and specification have been reported among TIR1/AFB genes[6], with their varied functions due to structural divergence[45]. In this study, differences in gene structure within the CaTIR1/AFB gene family were observed. At least two untranslated regions (UTRs) were detected in CaTIR1C, CaAFB4/5, and CaAFB6, with CaAFB6 containing four UTRs. CaTIR1A contained four coding sequences (CDSs), compared with three CDSs in each of the other four members (Fig. 3), which indicates the unique function of CaTIR1A. While the five CaTIR1/AFB proteins shared similar motif compositions, which are characteristic of resistance proteins known to play crucial roles in plant innate immunity by mediating pathogen recognition and initiating defense responses[46]. A notable LRR_CC_2 domain was observed in CaAFB6, highlighting its potential involvement in specialized signaling functions.

      Interestingly, duplication links of the TIR1/AFB genes were found in eggplant (SmTIR1B and SmTIR1C) and pepper (CaTIR1B and CaTIR1C), which are not found in tomato (Fig. 2). This may be connected to the genes' segmental duplication[7]. Duplications of genes involved in pathogenicity or stress responses are widely recognized as a mechanism that enables plants to acquire novel functions or regulatory complexity, facilitating adaptation to diverse biotic challenges[47,48]. This observation suggests that CaTIR1B and CaTIR1C may be functionally related to plant resistance. Notably, differences were detected in the UTR region and defense-responsive cis-elements in the promoters of these two genes, which may lead to the expression differences.

      Gene expression patterns are usually associated with their functions[4951], while genome-wide analyses on these patterns and functions have been utilized to investigate transcript levels. For example, studies have examined the expression profiles of TIR1/AFB under salt stress in Brassica juncea var. tumida and Malus pumila, revealing condition-specific regulatory patterns[44,52]. In this study, expression profiling under the R. solanacearum inoculation treatment demonstrated three distinct expression clusters among the CaTIR1/AFB genes based on their expression patterns.

      Notably, CaTIR1B and CaTIR1C are similar in CDS and exon-intron structures; however, CaTIR1C was significantly upregulated in the susceptible pepper variety. This upregulation may be attributed to the presence of unique defense-responsive cis-elements in the promoter region of CaTIR1C. Virus-induced gene silencing (VIGS) of CaTIR1C revealed that its downregulation enhanced resistance to R. solanacearum. These findings suggest that CaTIR1C promotes susceptibility to R. solanacearum, potentially through modulating of auxin-related signaling pathways. Nevertheless, the exact molecular mechanisms by which CaTIR1C contributes to pathogen response require further investigation.

      Bacterial wilt poses a major threat to global crop production, particularly affecting Solanaceae crops, including pepper[17]. Research showed that the accumulation of auxin is promoted in Arabidopsis under bacterial wilt infection[53], and genome-wide transcriptomic changes of auxin pathways in the resistant tomato variety were also observed[54]. However, the relationship between bacterial wilt and auxin has not been clearly elucidated. The findings of this study suggest that functional diversification within the CaTIR1/AFB family contributes to variety-specific responses to R. solanacearum infection among resistant and susceptible pepper varieties. Given the physiological diversity of R. solanacearum[55], this gene-level specificity may confer a selective advantage, possibly explaining the conserved nature of CaTIR1/AFB genes within the pepper genome. This is the first study to identify roles for CaTIR1/AFB genes in pepper in the context of R. solanacearum infection. While conclusions are primarily drawn from bioinformatic and transcriptome analyses, further experimental validation is necessary to elucidate the precise roles of TIR1/AFB genes within the broader regulatory network of plant growth and immune responses.

      In conclusion, five TIR1/AFB genes were identified in the pepper genome. Bioinformatic and transcriptome analysis of gene structure and conserved motifs revealed both functional redundancy and diversification in CaTIR1/AFBs under R. solanacearum. Notably, CaAFB6 was identified in pepper, together with AFB6 in tomato and eggplant, suggesting lineage-specific expansion in Solanaceae. Among the identified genes, CaTIR1C uniquely harbored defense-responsive cis-elements and displayed distinct expression patterns between resistant and susceptible pepper varieties upon R. solanacearum infection. Further functional evidence from expression profiling, subcellular localization, and VIGS assays highlighted CaTIR1C as a key regulator in pepper's response to R. solanacearum. These findings provide valuable insights into the roles of TIR1/AFB genes in pepper, establishing a foundation for future functional studies with approaches such as gene overexpression, RNA interference, and genome editing.

      • This work was supported by the Doctoral Fund of Baoding University (Grant No. 2023B03), and the Found of Innovation Capability Enhancement Special Project of Baoding Science and Technology Bureau (Grant No. 2494N005).

      • The authors confirm contributions to the paper as follows: methodology: Liu P, Jia Z, Xia M, Hua J; formal analysis, investigation: Liu P, Jia Z, Karamat U, Zhang C, Teng Z, Zhang B; preparation, validation: Karamat U, Zhang C, Teng Z, Zhang B; writing, original draft: Du W, Yu Z, Miao X; manuscript editing: Liu P, Jia Z, Du W, Yu Z, Miao X. All authors reviewed the results and approved the final version of the manuscript.

      • The datasets generated during and/or analyzed during the current study are available from the corresponding author on reasonable request.

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

      • # Authors contributed equally: Wenchao Du, Zhiyu Yu

      • Copyright: © 2025 by the author(s). Published by Maximum Academic Press, Fayetteville, GA. This article is an open access article distributed under Creative Commons Attribution License (CC BY 4.0), visit https://creativecommons.org/licenses/by/4.0/.
    Figure (8)  Table (1) References (55)
  • About this article
    Cite this article
    Du W, Yu Z, Miao X, Teng Z, Zhang C, et al. 2025. Unveiling the role of the TIR1/AFB gene family in Capsicum annuum: genome-wide identification, characterization, and transcriptomic analysis under Ralstonia solanacearum infection. Vegetable Research 5: e034 doi: 10.48130/vegres-0025-0028
    Du W, Yu Z, Miao X, Teng Z, Zhang C, et al. 2025. Unveiling the role of the TIR1/AFB gene family in Capsicum annuum: genome-wide identification, characterization, and transcriptomic analysis under Ralstonia solanacearum infection. Vegetable Research 5: e034 doi: 10.48130/vegres-0025-0028

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return