Next Article in Journal
The Oligomeric State of Vasorin in the Plasma Membrane Measured Non-Invasively by Quantitative Fluorescence Fluctuation Spectroscopy
Next Article in Special Issue
Physiological, Transcriptome, and Metabolome Analyses Reveal the Tolerance to Cu Toxicity in Red Macroalgae Gracilariopsis lemaneiformis
Previous Article in Journal
Biological Basis of Breast Cancer-Related Disparities in Precision Oncology Era
Previous Article in Special Issue
Overexpression of a Fragaria vesca NAM, ATAF, and CUC (NAC) Transcription Factor Gene (FvNAC29) Increases Salt and Cold Tolerance in Arabidopsis thaliana
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress of Small Plant Peptides on the Regulation of Plant Growth, Development, and Abiotic Stress

1
Shandong Provincial Key Laboratory of Plant Stress, College of Life Sciences, Shandong Normal University, Jinan 250014, China
2
Dongying Key Laboratory of Salt Tolerance Mechanism and Application of Halophytes, Dongying Institute, Shandong Normal University, No. 2 Kangyang Road, Dongying 257000, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(7), 4114; https://doi.org/10.3390/ijms25074114
Submission received: 1 March 2024 / Revised: 2 April 2024 / Accepted: 4 April 2024 / Published: 8 April 2024
(This article belongs to the Special Issue Advance in Plant Abiotic Stress)

Abstract

:
Small peptides in plants are typically characterized as being shorter than 120 amino acids, with their biologically active variants comprising fewer than 20 amino acids. These peptides are instrumental in regulating plant growth, development, and physiological processes, even at minimal concentrations. They play a critical role in long-distance signal transduction within plants and act as primary responders to a range of stress conditions, including salinity, alkalinity, drought, high temperatures, and cold. This review highlights the crucial roles of various small peptides in plant growth and development, plant resistance to abiotic stress, and their involvement in long-distance transport. Furthermore, it elaborates their roles in the regulation of plant hormone biosynthesis. Special emphasis is given to the functions and mechanisms of small peptides in plants responding to abiotic stress conditions, aiming to provide valuable insights for researchers working on the comprehensive study and practical application of small peptides.

1. Introduction

Based on their N-terminal signal sequences, small peptides are classified into secretory and non-secretory types. Secretory peptides include post-translationally modified peptides (PTMs) and cysteine-rich peptides (CRPs). The majority of them are secreted into the extracellular matrix through an endoplasmic-reticulum-dependent pathway, contributing to the regulation of extracellular growth and developmental activities [1]. Sometimes, they have an effect on the entire plant, functioning throughout the whole plant. Non-secretory peptides, characterized by their encoding through short open reading frames, differ significantly from secretory peptides. Non-secretory peptides work inside cells and often participate in cellular defense mechanisms. Since the identification of the first small plant peptide, systemin, a diverse array of peptide hormones has been discovered. These include plant sulfonamide (PSK) [2], C-terminally encoded peptides (CEP), CLAVATA3/Embryo Surrounding Region-related peptides (CLE), Inflorescence Deficient in Abscission (IDA), Rapid Alkalinization Factor (RALF), and Root Growth Factor (RGF), among others. These small peptides play a crucial role in the physiological activities of plants, encompassing cell communication [3], root development, microtubule organization, meristem maintenance, floral organ abscission, leaf senescence, nutrient assimilation, responses to diseases and pests, and fertilization. It is noteworthy that peptides also play an indispensable role in enabling plant responses to abiotic stressors such as salinity, drought, and extreme temperatures. This underscores the vital contribution of peptides to the adaptability and resilience of plants in diverse environmental conditions.
In the past, it was generally believed that small peptides were only involved in short-distance signal transduction in plants. However, recent research has shown that an increasing number of small peptides also play an important role in long-distance signal transduction [4]. Knowledge about the long-distance transport of peptides in response to abiotic stress, however, remains limited. The long-distance signal transduction of small peptides is accomplished through the peptide receptor system. Roots detect changes in the external environment, such as elevated salt levels or drought conditions, triggering the production of peptides. These peptides are then transported via the xylem to the stems, where they engage with specific receptors. Then, through a series of reactions, the plants’ tolerance to the environment is improved.
Despite notable advances in the study of small peptides, significant hurdles remain. Among these, a primary challenge is deciphering the signal transduction mechanisms between the upstream and downstream components of the signaling networks established by small peptides and their receptors. In this paper, we provide a comprehensive review of the diverse roles of small peptides in plant growth, development, and responses to abiotic stress. We delve into the involvement of small peptides in long-distance signaling pathways and examine their pivotal functions in the regulation of plant hormone biosynthesis.

2. Types and Sources of Small Peptides in Plants

The small peptides are commonly classified into three categories: cysteine-rich peptides (CRP), post-translationally modified peptides (PTM), and non-secretory small peptides [5,6,7,8,9] (Table 1). PTM and CRP, collectively referred to as secretory small peptides, share common characteristics, including N-terminal secretion signals and a central variable region. Peptides rich in cysteine and post-translationally modified peptides differ significantly in their origin, synthesis, and function. The formation of PTMs requires subsequent processing of proteins that have already been synthesized, while CRPs are formed directly after being encoded by genomic DNA and do not require post-translational modification. Mature PTM and CRP are produced through enzyme-mediated processing and modification of their precursor peptides [10].

2.1. Post-Translationally Modified Peptides

The process of post-translationally modified peptides involves tyrosine sulfation, proline hydroxylation, or arabinosylation, ultimately resulting in the formation of biologically active small peptides consisting of approximately 10 amino acid residues [29]. Tyrosine sulfation involves adding a sulfate group to the tyrosine residue, proline hydroxylation adds a hydroxyl group to the proline residue, and arabinosylation adds an arabinose molecule to the amino acid residue of a protein. Most of these small peptides are secreted into the extracellular space predominantly via passive diffusion [30]. Typical examples of post-translationally modified peptides include phytosulfokine (PSK), Tracheary Element Differentiation Inhibitory Factor (TDIF), CLAVATA3/EMBRYO Surrounding Region-Related (CLE), C-terminally Encoded Peptide (CEP), and Plant Peptide Containing Sulfated Tyrosine 1 (PSY) [23]. Since the discovery of the CLV3 gene, many other CLE genes have been successively identified. CLE peptides include A-type CLE peptides that promote root and shoot development, such as CLAVATA 3 (CLV3), which is involved in apical meristem differentiation, and B-type CLE peptides such as TDIF [13,14,31]. CLE peptides are mainly produced through post-translational arabinoxylan modification of precursor peptides. Their receptors are primarily leucine-rich repeat receptor-like kinases (LRR/RLKs), including CLAVATA1 (CLV1), CLAVATA2 (CLV2), CORYNE (CRN) [32], Receptor-Like Protein Kinase 2 (BAM) [33], HAESA-LIKE 1 (HSL1), and Somatic Embryogenesis Receptor-Like Kinase (SKM) [34]. The WUSCHEL (WUS) gene can induce the expression of CLV3, and CLV3 acts on CLV1 to inhibit the activity of WUS, establishing a negative feedback loop essential for the maintenance of shoot apical meristem (SAM) homeostasis [8,35,36]. In addition to interacting with the receptor CLV1, CLV3 also interacts with other receptors such as CLV2, CRN, receptor-like protein kinase 2 (RPK2) [37], and auxiliary receptors like somatic embryogenesis receptor kinase (SERK) [38]. PSK, another post-translationally modified peptide, becomes biologically active through tyrosine sulfation of its precursor [39]. Initially, PSK was isolated and identified from Asparagus officinalis L.

2.2. Cysteine-Rich Peptides

CRPs are distinguished by their multiple disulfide bonds and are named accordingly. The N-terminus of CRPs typically have a conserved signal peptide, while the C-terminus possesses an even number of cysteine residues, enabling the formation of disulfide bonds. Examples of CRPs include the RALF with four cysteine residues, late embryogenesis abundant-like protein 2-regulated (LURE), taximin (TAX) [24], lilly anther protein 2 (LAT52) [25], the epidermal patterning factor (EPF), and inflorescence deficient in abscission (IDA), among others. RALFs represent a category of cysteine-rich peptides derived from precursor peptides through cleavage by the S1P protein [22]. They are characterized by having four cysteine amino acid residues and are implicated in a variety of biological processes, including root development, cell wall remodeling, maintaining pollen tube integrity, facilitating fertilization, mediating immune responses, and modulating plant responses to environmental stress [40,41,42,43,44]. Furthermore, the discovery of new cysteine-rich peptides continues to expand our understanding of plant biology. For instance, a novel RALF-like peptide, EaF82, was recently identified and has been shown to reduce pollen and seed production [45].

2.3. The Relationship between Non-Secretory Small Peptides, Post-Translationally Modified Peptides, and Cysteine-Rich Peptides

Both PTMs and CRPs significantly contribute to the growth and developmental processes of plants. For instance, CLE40p activates CLV1 on the membrane, leading to the rapid phosphorylation of Cyclic Nucleotide-Gated Channel 9 and Cyclic Nucleotide-Gated Channel 6. This results in an elevated cytoplasmic Ca2+ concentration within the proximal meristematic tissue, thereby regulating the differentiation of the proximal root meristem [46]. Similarly, TAX, a membrane-localized peptide, plays a crucial role in regulating plant-specific metabolic pathways. When combined with methyl jasmonate (MeJA), TAX notably increases paclitaxel production and enhances the biosynthesis of alkaloids in tobacco roots [24].
Current research on non-secretory small peptides remains limited, leading to a relatively superficial comprehension of their interaction with secretory small peptides within signal transduction pathways. Nevertheless, recent discoveries indicate that in Arabidopsis, the non-secretory small peptide rotundifolia 4 (ROT4) enhances its interaction with brassinosteroid-signaling kinase 5 (BSK5) through S-acylation. This heightened interaction disrupts the binding of receptor kinase PEP receptor 1 (PEPR1) to BSK5, which is crucial for signaling of the secretory peptide plant elicitor peptide 1 (PEP1), thereby playing a significant role in regulating plant immunity [11]. This implies that secretory and non-secretory peptides do not act independently but interact with each other to collectively regulate plant growth and development processes. Non-secretory small peptides may serve as “switches”, modulating the activities of secretory small peptides and, thus, maintaining a dynamic equilibrium within the plant’s physiological processes. Such research sheds light on the possible regulatory roles of non-secretory small peptides in plant signal transduction and opens new avenues for understanding the complexity of plant physiological mechanisms.

3. The Role of Small Peptides in Plant Growth and Development

Small plant peptides play pivotal roles in regulating a diverse range of growth and developmental processes in plants. These roles encompass facilitating intercellular communication, promoting root development, influencing microtubule formation, maintaining meristem activity, orchestrating floral organ abscission, driving leaf senescence, optimizing nutrient uptake, and supporting fertilization, among others.

3.1. The Role of Small Peptides in Root Development

Small plant peptides are vital to multiple facets of plant root development, impacting virtually every phase of root formation. They play a significant role in nodulation, primary root growth, and the initiation of lateral roots. Furthermore, small plant peptides also influence the division, elongation, and differentiation of root apex cells [47,48].
Recent research has indicated that nodulation can promote plant flowering, thereby linking nodulation with various stages of plant growth and development [49]. Peptides such as CLE [50,51], CEP [16], and microRNA-encoded peptides (miPEPs) have been identified as key regulators of plant root nodulation [17]. It has been found that under conditions of ample soil nitrogen, CLE peptides produced in the roots transport over long distances to the stem, where they interact with the peptide receptor-like kinase HAR1, resulting in the suppression of miR2111 expression and, thus, controlling the number of root nodules in roots [52]. In contrast, under low-nitrogen conditions, CEP signaling peptides can trigger the production of miR2111, facilitating nodule formation even in the absence of root nodule bacteria [53]. These findings suggest a complex negative feedback loop between CLE and CEP peptides, which plays a crucial role in modulating root nodule formation. Nonetheless, the specific interactions between CLE and CEP peptides in the regulation of root nodulation warrant further investigation.
Small plant peptides fulfill diverse roles in regulating root growth, with specific peptides influencing root development in distinct ways. For instance, small plant peptides might contribute to root development and adaptation to flooding [54].
The regulatory role of CEP in root growth has been investigated in cucumber, Arabidopsis, Brassica rapa L., and other plants. For example, knocking out the CEP gene in Arabidopsis resulted in increased primary root length and number of lateral roots [55,56]. Similarly, peptides such as BrCEP3 and BrCEP26 have been demonstrated to significantly enhance primary root growth in moss [57].
In grapevines, the peptide vvi-miPEP171d has been identified as a regulator of adventitious root formation [58]. In rice, overexpression of OsPEP1 results in a shorter root phenotype, highlighting PEP1’s regulatory role in root development (Figure 1). The root meristem growth factor (RGF), a 13-amino acid sulfated peptide, contributes to the maintenance of root apical meristem activity [18]. However, the specific action site of RGF at the root apex and its regulation of PLETHORA proteins are yet to be clarified.
Serine Rich Endogenous Peptides (SCOOPs), encoded by the Brassicaceae-specific PROSCOOP family, are commonly believed to participate in defense signal transduction when plants are subjected to insect attacks or mechanical damage [26]. Recent research indicates that SCOOP12 regulates defense responses and root elongation in Arabidopsis. SCOOP12 reduces the size of the meristem and cell length in the root, suggesting a role in shortening the root under specific conditions. SCOOPs are induced when plants are invaded by a pathogen or microbe and act as a mediator of intermediate filament formation kinase 2 (MIK2), forming a co-receptor with BRI1-associated receptor kinase 1 (BAK1) to regulate downstream signaling pathways (Figure 1) [27].
Furthermore, small peptides like TDIF are key to lateral root development, binding to their receptor TDIF initiated (TOR) and activating BIN, which in turn promotes the phosphorylation of AUX/IAA regulatory genes to facilitate lateral root development (Figure 1). AtRALF34 is considered to be part of the gene regulatory network for lateral root initiation in Arabidopsis. In contrast to AtRALF34, its homologous gene CsRALF34 in cucumber does not participate in lateral root initiation but instead leads to G2/mitosis transition block [59,60,61].
In rice, the cleavage inducing factors OsCIF1 and OsCIF2 play a role in forming Casparian Strips in both endodermal and non-endodermal cell layers of the root, illustrating peptides’ regulatory capacity on root pericycle development as well [15].

3.2. The Role of Small Peptides in Leaf Senescence

In addition to their role in root development, peptides also influence leaf senescence. Studies have shown that the small peptide CLE14 is involved in regulating plant leaf senescence [62]. Treating detached Arabidopsis leaves with CLE14 peptide significantly reduces senescence. When plants undergo stress-induced leaf senescence or age-induced leaf senescence, CLE14 targets JUB1 to enhance the expression of genes involved in the removal of reactive oxygen species (ROS), thus mitigating leaf senescence (Figure 1). This study also indirectly indicates the role of CLE peptides in long-distance signal transduction from roots to leaves. Furthermore, the study found that overexpression and exogenous application of the peptide CLE42 can also delay leaf senescence. CLE42 counters the ethylene pathway by inhibiting ethylene biosynthesis, increasing levels of the ethylene-insensitive 3 (EIN3)-binding F-box protein (EBF) and reducing the activity of ethylene-related transcription factor EIN3 (Figure 1) [63]. However, the redundancy in CLE peptide functions, such as simultaneous expression of CLE1 and CLE4 in identical tissues, presents challenges in fully understanding their roles [64]. The involvement of CLE peptides in leaf processes underscores their potential as long-distance signaling molecules.

3.3. Other Functions of Small Peptides in Plant Growth and Development

In addition to the aforementioned functions, peptides significantly contribute to various aspects of plant physiology, including nutrient uptake, circadian rhythms, photomorphogenesis, resistance against pests and diseases, as well as the regulation of plant immunity [65,66,67,68]. For instance, synthetic CEP peptides have been shown to improve the uptake of nitrates, phosphates, and sulfates in plant roots [69]. The regulation of CLE peptide gene expression is influenced by the phosphorus and sulfur statuses in Arabidopsis. The antimicrobial peptide DmAMP1W from wheat curtails the proliferation of fungi, including fungi Rhizoctonia cerealis and Bipolaris sorokiniana, effectively combating wheat sharp eyespot and root rot diseases [70]. Additionally, a variety of small peptides, such as those released by papain-like cysteine proteases, Defensin-Derived Antifungal Peptides, and small cysteine-rich (SCR) secreted proteins like RsMf8HN, play an integral role in plants’ defense against fungal and bacterial pathogens [71,72,73]. The peptide VISP1, acting as a selective autophagy receptor, plays an essential role in plant immunity [74]. The external application of RALFs peptide holds promise for surmounting interspecific hybridization barriers within the Brassicaceae family [75]. Peptides are also pivotal in controlling flower organ abscission. By using CRISPR/Cas technology to simultaneously knock out the BnaIDA-A07 and BnaIDA-C06 genes in rapeseed, the shedding of floral organs is effectively inhibited [19].
In summary, peptides exhibit diverse functions across various plant parts, including roots, stems, leaves, flowers, and more.

4. The Role of Small Peptides in Plant under Abiotic Stress

The global warming of the climate and salinization pose significant constraints on the development of agricultural production. These conditions significantly constrain agricultural development and plant growth. Unlike animals, plants cannot freely move, rendering them vulnerable to various environmental stressors including temperature extremes, water scarcity, and saline–alkali conditions, which severely impact their growth and development. In coping with these challenges, small peptides, acting as signaling molecules, play a crucial role in plants’ response to abiotic stress [76].

4.1. The Role of Small Plant Peptides under Drought Stress

In recent years, the escalation of drought conditions globally has markedly affected plant growth, leading to substantial reductions in crop yields. The adverse effects of drought on plant life are comparable to those caused by pathogens. Extensive research has demonstrated the critical role of small peptides in plants’ responses to drought stress. Researchers have discovered a number of small peptides that assist plants in resisting drought stress. Specifically, the overexpression of OSS1Fa1 from rice in Arabidopsis substantially enhances the drought tolerance of Arabidopsis (Figure 2). Furthermore, S1Fa has been shown to bolster the tolerance of various other plants, including corn and Chinese cabbage, to a range of abiotic stresses. Nevertheless, a comprehensive understanding of the biochemical, molecular, and physiological functions of S1Fa remains elusive. Further investigation of its interacting proteins is essential to unravel its contribution to plant stress resistance [77]. In rice, a novel signaling peptide, OsDSSR1, has been discovered, triggering the expression of ROS scavenging-related genes, OsSodCc2 and OscAPX, under drought conditions (Figure 2). This peptide boosts the activity of superoxide dismutase (SOD) and ascorbate peroxidase (APX), facilitating ROS clearance and enhancing rice’s drought tolerance [78]. Additionally, OsRALF45/46 in rice interacts with the receptor-like kinase OsMRLK63, leading to improved drought resistance [79].
Research on CEP regulation of plant drought tolerance has been extensive, and studies in crops such as wheat and tomato have revealed the crucial role of CEP in enhancing plant drought tolerance [80]. Application of the TaCEP1D peptide on wheat leaves has been shown to improve drought resistance [81]. Some researchers have found that the receptors of CEP peptides are CEPR1 and CEPR; these two, along with RLK7, share high similarity in protein sequences and belong to the same LRR-RLK subfamily XI. Therefore, the CEP peptides involved in wheat drought and salt stress tolerance may also rely on RLK7, and further research is needed to elucidate the receptors and mechanisms of action.
CLE25, synthesized in the roots, targets membrane receptors BAM1 and BAM3 in the shoots. This interaction stimulates the expression of NCED3 and promotes ABA production, subsequently inducing stomatal closure in response to drought stress (Figure 2) [82,83]. This elucidates the capability of the CLE25 peptide to regulate plant growth and development through systemic signaling.
The Small Paraquat resistance protein (SPQ), LcSPQ, equipped with an N-terminal signal peptide, has been found to enhance plant drought resistance upon overexpression [84]; however, its specific receptors and underlying mechanisms require additional study.
Recent research has expanded the investigation of PSK (phytosulfokine) responses to stress from model organisms, such as Arabidopsis, to staple crops and even tall trees. After conducting a bioinformatics analysis of the rice PSK family, some researchers found that when plants are subjected to heat, salt, drought, and cold stress, the OsPSKR gene functions to enhance the plant’s resistance to various adverse environmental conditions [85]. The action of PSK under stress conditions is mediated by various subtilisin-like proteases of the phytaspase subtype, which cleave the PSK precursor, enabling its functionality under diverse stress conditions. In Arabidopsis, the subtilisin-like protease SBT3.8 cleaves the C-terminal end of the PSK precursor, enhancing drought resistance [2,39]. In tomato, phytaspase 2 processes the aspartic acid in the PSK precursor, facilitating flower and fruit abscission under drought stress [86,87]. PSK has also been shown to contribute to ROS homeostasis by modulating the activity of peroxidase in Cunninghamia lanceolata [83]. Current research on PSK precursor processing has predominantly focused on the C-terminal end, with further studies needed to explore the processing at the N-terminal end.

4.2. The Role of Small Peptides under Salt Stress in Plants

Globally, salinization affects over 10% of land, significantly hindering agricultural productivity. Salt stress triggers ion imbalance, osmotic shifts, oxidative stress, and secondary stresses in plants [88].
The peptide PIP3, activated by Pathogen-Associated Molecular Patterns (PAMPs), targets the receptor-like kinase 7 (RLK7), characterized by leucine-rich repeat motifs, under salt conditions (Figure 2). This activation of RLK7 enhances its activity, resulting in the amplification of salt stress signals by activating MPK3/6, maintaining cellular ion homeostasis, and thereby enhancing plant salt tolerance [20].
Additionally, AtPEP3, a hormone-like peptide produced by the AtPROPEP3 gene, augments plant salt tolerance through its interaction with the PEPR1 receptor (Figure 2). When plants experience salt stress, exacerbating cell damage leads to the release of AtPEP3, which in turn induces downstream signaling responses. However, the intricacies of AtPEP3’s signal transduction pathway in salt tolerance remain elusive. While pathogen-induced AtPEP3 acts on both PEPR1 and PEPR2 receptors, salt-induced AtPEP3 specifically targets PEPR1, hinting at potential salt-induced conformational changes in PEPR2 that require further investigation [89].
The RALF-FER pathway plays a pivotal role in plant stress adaptation (Figure 2) [90]. Under normal conditions, LRX3/4/5 proteins, which contribute to plant growth and cell wall integrity, bind RALF22/23 peptides. This interaction prevents RALF peptides from binding to FER proteins, thus inhibiting FER internalization (Figure 2). In contrast, under salt stress, LRX3/4/5 proteins perceive changes in the cell wall, leading to the dissociation of RALF peptides from LRX. This promotes the binding of RALF22/23 peptides to the FER protein, resulting in the loss of its salt tolerance capability and the manifestation of a salt-sensitive phenotype [91].
Research on CYSTM (Cys-rich Peptide of the Seed Tryptic Mixture) peptides, particularly regarding stress responses, is nascent, with a focus on bioinformatics analyses. CYSTM, identified as a novel cysteine-rich transmembrane module, plays a role in stress tolerance across eukaryotes [28]. CYSTM can modulate the plant’s tolerance to heat and ultraviolet stress. Additionally, CYSTM3 negatively regulates Arabidopsis’ salt stress tolerance [92]. Notably, CYSTM3 has been shown to diminish salt tolerance in Arabidopsis by impeding Na+ efflux and inhibiting ROS-scavenging enzymes, alongside downregulating salt-responsive gene expression (Figure 2) [93]. Given its mitochondrial localization, further research is required to elucidate CYSTM3’s impact on Ca2+ ion homeostasis and plant salt tolerance. The relationship between cysteine residues and transmembrane helices in CYSTM peptides also remains to be explored.

4.3. The Role of Small Plant Peptides under Temperature Stress

The fluctuation in temperature exerts a significant influence on plants, yet the role of small peptides in modulating temperature stress responses remains underexplored. It is established that low-temperature stress alters membrane composition, reactive oxygen species (ROS) levels, osmoregulation substances levels, and antioxidant capacities [94,95]. Beyond their known functions in growth and development, microprotein-encoded peptides (miPEPs) also contribute to abiotic stress resilience. Specifically, under cold conditions, plants produce miPEPs encoded by pri-miRNAs, such as vvi-miPEP172b and vvi-miPEP3635b (Figure 2). Studies have demonstrated that these peptides bolster plant cold resistance by modulating miRNA gene activity [96]. In contrast, during high-temperature exposure, the CLE45 peptide, which originates in the stigma and extends to the transmitting tract, suggests a potential function in heat stress adaptation [97].

4.4. The Role of Small Peptides under Other Stress Conditions and Biological Stress in Plants

Peptides play a pivotal role in augmenting plant tolerance to heavy metals. The long-distance signaling peptide IMA, which mediates root-to-shoot communication, is instrumental in this context. Specifically, IMA1 and IMA3 have been shown to enhance tolerance to cadmium (Cd) toxicity through the promotion of iron (Fe) accumulation (Figure 2) [98]. Furthermore, exposure to cadmium triggers an increase in the expression of BrCEP8 and BrCEP19, indicating that BrCEPs may facilitate cadmium absorption in Brassicaceae, thereby augmenting Brassicaceae’s tolerance to elevated Cd levels (Figure 2) [57]. Additionally, microRNA-encoded peptides (miPEPs) have been recognized for their regulatory role in plant tolerance to cadmium [99].
Additionally, under conditions of salt stress, PAMP-induced secreted peptides (PIP3) target receptor-like kinase 7 (RLK7), enhancing salt tolerance [20]. RLKs are pivotal in mediating plant stress resistance, with IDA peptides acting on receptors within the RLK family, hinting at IDA’s potential involvement in plant stress resilience. Recent research has unveiled that IDA family members, specifically IDL6 and IDL7, play a contrasting role by negatively affecting stress resistance in Arabidopsis through the acceleration of reactive oxygen species (ROS) production [100].
Small peptides play a regulatory role in modulating immune responses within plants [101]. Moreover, research has revealed that under phosphate-deficient conditions, PHR1 activates the RALF peptide, which in turn triggers the FERONIA receptor. This activation sequence suppresses pathogen-induced immunity, thereby facilitating the proliferation and development of root microbiota. Consequently, this mechanism contributes to mitigating the effects of phosphate deficiency in plants [102]. ROOT MERISTEM GROWTH FACTOR 7 is regulated by perception through the RGI4/RGI5-BAK1/SERK4 receptor complex in the immune response of Arabidopsis [103]. In addition, virus-induced small peptide 1 (VISP1) [104] and the small peptide StPIP1 in potato [105] also participate in the immune response of plants.
Under phosphate scarcity, Arabidopsis PHR1 directly binds to and activates the promoter of the Rapid Alkalinization Factor (RALF) gene. Subsequently, RALF inhibits the formation of the immune receptor complex triggered by pathogen-associated molecular patterns (PAMPs), via FERONIA. This PHR1-RALF-FERONIA pathway dampens immune responses, facilitating specific root microbiota colonization. This colonization aids in mitigating phosphate starvation by boosting the expression of phosphate starvation response (PSR) genes.

5. Long-Distance Signal Transduction Mediated by Small Peptides under Stress

Small peptides can function as both local signaling molecules in roots and long-distance signals, acting on receptors in various plant tissues [4]. The presence of peptides such as CLE, CEP, and CIF in tomato xylem sap highlights their role in root-to-shoot signal transmission [106]. Furthermore, recent discoveries of CEP and CLE peptides in the phloem suggest their involvement in shoot-to-root signaling [107]. The PSY family’s participation in root-to-shoot transport supports the potential for investigating sulfated peptides’ responses to drought stress [21]. During stress conditions, these peptides regulate the plant’s physiological state as long-distance signals. For instance, under drought, Arabidopsis CLE25, produced in the roots, travels to the leaves to activate the BAM1/3 receptor, triggering the synthesis of abscisic acid by upregulating NINE-CIS-EPOXYCAROTENOID DIOXYGENASE 3 (NCED3) [83]. Similarly, CEP peptides mediate nitrogen demand and support root nodule and lateral root formation. Nitrogen scarcity prompts CEP expression in roots, with the signal reaching the stem and activating LRR-RKs [3]. Beyond CLE, advancements in understanding long-distance signaling have been underscored by grafting experiments, which demonstrate the necessity of CEPR1 activity in both roots and shoots for modulating lateral root growth [108,109]. The comprehensive effects of CEP peptides on plant parts like leaves, shoots, and roots via long-distance transport are areas for ongoing research.

6. Plant Peptides Control the Biosynthesis of Hormones

The discovery of peptides involved in long-distance signal transduction has underscored their similar regulatory pathways to those of hormones. Hormones are known to facilitate plant growth and development through diverse interactions, highlighting the emerging interest in peptides’ regulatory roles within these hormone interactions among the scientific community.
The overexpression of RsCLE22a, for instance, elevates auxin levels in Raphanus sativus L., consequently inhibiting the growth of the primary root [71]. FER (FERONIA) is regulated by various hormones, such as ABA, ethylene, and jasmonic acid. For example, treating apple (Malus domestica) roots with ABA significantly induces the expression of MdFER [110,111]. Furthermore, overexpressing MdFER in apples can negatively regulate their sensitivity to ABA. Ethylene can also promote the transcription of FER in the roots. Due to the interaction between RALF and FER under salt stress, we speculate that RALF may also have intricate connections with plant hormones, akin to other plant peptides [112]. Recent findings indicate that ethylene downregulates the expression of CsRALF34 in cucumber [71], whereas in Arabidopsis, AtRALF34 is modulated by both auxin and ethylene [59]. Furthermore, the external application of SpPIP1 modulates jasmonic acid levels, thereby enhancing tomato resistance to stress [113]. GhRALF1 influences cotton’s diurnal rhythms and fiber development through modulating auxin signaling and proton pump activity [114].
Notably, auxin accumulation in the root system has been associated with increased wheat yields under drought conditions. Drought-responsive transcription factors regulate auxin signaling by increasing AUX/IAA expression, and CEP5 signaling has been shown to stabilize the AUX/IAA transcriptional repressor [115]. DEVIL-like (DVL) peptides play a pivotal role in sustaining root growth under abiotic stress conditions by affecting ABA-related gene expression [116]. The synergistic interaction between ABA-INDUCED EXPRESSION 1 (AIN1) and PIP2 significantly bolsters plant drought resistance, underscoring the critical role of ABA [117]. CEP peptides indirectly regulate plant responses to cytokinins by affecting components within the cytokinin signaling pathway [118]. Finally, the TDIF peptide targets auxin-related genes, altering internal auxin concentrations and, thus, orchestrating the growth of both primary and lateral roots [12].
In conclusion, peptides significantly affect plant growth and development by modulating plant hormone concentrations. Importantly, a growing body of research identifies peptides as key regulators in plants’ responses to abiotic stresses, alongside crucial hormones such as ABA, which are known to play a pivotal role in plant stress resistance. Thus, investigating the effects of peptides on plant hormones is of paramount importance for a deeper understanding of how plants respond to abiotic stress.

7. Discussion and Prospects

Research on the diverse families of plant peptides is still in its infancy, partly due to the challenges associated with isolating these molecules. However, advancements in peptideomics are increasingly enabling the identification of new peptides. Techniques such as Capillary Electrophoresis (CE) and Microchip Electrophoresis (MCE) have improved the efficiency of peptide separation, addressing the research difficulties posed by their small size. Recent technological breakthroughs, including nanoscale liquid chromatography-tandem mass spectrometry (nano-LC–MS/MS), alongside predictive tools like SignalP 6.0 and ExamPle, have facilitated the discovery and analysis of novel peptides [119,120,121]. To tackle functional redundancy, antagonistic peptide technology offers a promising solution [122], while CRISPR gene editing presents a way to obtain mutants for studying functional loss. Most small peptides undergo post-translational modifications, with a multitude of enzymes involved in their processing and modification. Further research is needed to understand how these enzymes influence the biological activity and signal transduction of peptides. Small peptides play crucial roles in various aspects of plant growth and development. Despite the critical signaling roles of peptides, many of their receptors remain unidentified, and the mechanisms of receptor perception and the role of auxiliary receptors in peptide signaling are still under-researched. The “one-step” nanoimprint method has emerged as a novel approach for receptor identification [120]. While exogenous peptide application is common in functional studies, differences from endogenous peptides due to processing and modifications are crucial considerations. Recent research utilizing isotope-labeled peptides has shed light on the effects of post-translational modifications, underscoring the complexity of peptide signaling pathways [123].
Through research, scientists have discovered that an increasing number of small peptides co-regulate plant growth and development with hormones. For example, there is a significant overlap in the signal transduction between CLE and hormones such as auxin, cytokinin, gibberellin, and abscisic acid [31]. Peptides are also known to collaborate with hormones in responding to environmental changes [97]. Although the hormonal aspects of plant stress responses are well-studied, the intricate regulatory network between peptides and plant hormones during such responses still demands further clarification. A notable gap in current research is a molecular-level understanding of how peptides respond to environmental stressors. For instance, the mechanism by which the rice OSS1Fa1 gene enhances drought tolerance in Arabidopsis remains poorly defined, indicating a general shortfall in exploring the proteins that interact with these genes. Despite these challenges, significant progress has been made in certain areas, such as the study of Arabidopsis CLE14, which has been shown to respond to drought and salt stress [62]. This research discovered that CLE14 activates JUB1, leading to increased expression of genes involved in ROS clearance and delayed leaf senescence. Such findings are pivotal, advancing the research on CLE peptides from a focus on development to include stress response and shedding light on the relationships between CLE genes and their downstream effects.
In recent years, the study of peptide-mediated long-distance signal transduction has emerged as a significant area of research. For instance, it has been discovered that overexpressing CLE3 in Arabidopsis roots leads to its transport to the shoots via long-distance signaling, where it influences WRKY33, thereby modulating the plant’s immune responses [124]. Additionally, the detection of the root-derived sugar peptide CLE-RS in the plant xylem points to its potential role in facilitating long-distance communication [125]. However, despite being acknowledged, the dynamics of peptide signaling under stress conditions remain largely unexplored. This research gap underscores the need for further investigation into how peptides contribute to plant adaptation and survival under adverse environmental conditions.
The primary objective of researching plant peptides is to benefit agriculture and foster agricultural advancement. Notably, the use of synthetically produced peptides externally applied to plants has been shown to regulate growth, development, and immune responses effectively. This indicates that peptides have considerable potential to serve as alternatives to traditional pesticides and chemical fertilizer. By harnessing the regulatory capabilities of peptides, we can develop safer, more sustainable agricultural practices that reduce reliance on chemical pesticides and enhance crop resilience.

Author Contributions

H.F. conceived the idea of the review, organized its plan and reviewed the study; G.R. and Y.Z. wrote major parts of the manuscript; G.R., Y.Z. and Z.C. performed literature search and prepared the tables; Z.C. and X.X. made the figures and prepared all references; H.F., G.R. and Y.Z. reviewed, edited and supervised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data will be made available on request.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could appear to influence the work reported in this paper.

References

  1. Djordjevic, M.A.; Mohd-Radzman, N.A.; Imin, N. Small-peptide signals that control root nodule number, development, and symbiosis. J. Exp. Bot. 2015, 17, 5171–5181. [Google Scholar] [CrossRef] [PubMed]
  2. Matsubayashi, Y.; Sakagami, Y. Phytosulfokine, sulfated peptides that induce the proliferation of single mesophyll cells of Asparagus officinalis L. Proc. Natl. Acad. Sci. USA 1996, 15, 7623–7627. [Google Scholar] [CrossRef] [PubMed]
  3. Tabata, R.; Sumida, K.; Yoshii, T.; Ohyama, K.; Shinohara, H.; Matsubayashi, Y. Perception of root-derived peptides by shoot LRR-RKs mediates systemic N-demand signaling. Science 2014, 6207, 343–346. [Google Scholar] [CrossRef] [PubMed]
  4. Oh, E.; Seo, P.J.; Kim, J. Signaling Peptides and Receptors Coordinating Plant Root Development. Trends Plant Sci. 2018, 4, 337–351. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, W.C.; Katinakis, P.; Hendriks, P.; Smolders, A.; de Vries, F.; Spee, J.; van Kammen, A.; Bisseling, T.; Franssen, H. Characterization of GmENOD40, a gene showing novel patterns of cell-specific expression during soybean nodule development. Plant J. 1993, 4, 573–585. [Google Scholar] [CrossRef] [PubMed]
  6. Frueauf, J.B.; Dolata, M.; Leykam, J.F.; Lloyd, E.A.; Gonzales, M.; VandenBosch, K.; Kieliszewski, M.J. Peptides isolated from cell walls of Medicago truncatula nodules and uninfected root. Phytochemistry 2000, 5, 429–438. [Google Scholar] [CrossRef] [PubMed]
  7. Ganguly, P.; Roy, D.; Das, T.; Kundu, A.; Cartieaux, F.; Ghosh, Z.; DasGupta, M. The Natural Antisense Transcript DONE40 Derived from the lncRNA ENOD40 Locus Interacts with SET Domain Protein ASHR3 During Inception of Symbiosis in Arachis hypogaea. Mol. Plant-Microbe Interact. 2021, 9, 1057–1070. [Google Scholar] [CrossRef] [PubMed]
  8. Crespi, M.D.; Jurkevitch, E.; Poiret, M.; d’Aubenton-Carafa, Y.; Petrovics, G.; Kondorosi, E.; Kondorosi, A. enod40, a gene expressed during nodule organogenesis, codes for a non-translatable RNA involved in plant growth. EMBO J. 1994, 21, 5099–5112. [Google Scholar] [CrossRef] [PubMed]
  9. Campalans, A.; Kondorosi, A.; Crespi, M. Enod40, a short open reading frame-containing mRNA, induces cytoplasmic localization of a nuclear RNA binding protein in Medicago truncatula. Plant Cell 2004, 4, 1047–1059. [Google Scholar] [CrossRef]
  10. Xie, H.P.; Zhao, W.; Li, W.L.; Zhang, Y.Z.; Hajny, J.; Han, H.B. Small signaling peptides mediate plant adaptions to abiotic environmental stress. Planta 2022, 4, 12. [Google Scholar] [CrossRef]
  11. Li, W.; Ye, T.; Ye, W.; Liang, J.; Wang, W.; Han, D.; Liu, X.; Huang, L.; Ouyang, Y.; Liao, J.; et al. S-acylation of a non-secreted peptide controls plant immunity via secreted-peptide signal activation. EMBO Rep. 2024, 25, 489–505. [Google Scholar] [CrossRef] [PubMed]
  12. Yang, S.H.; Bai, J.P.; Wang, J.H. TDIF peptides regulate root growth by affecting auxin homeostasis and PINs expression in Arabidopsis thaliana. Planta 2020, 6, 109. [Google Scholar] [CrossRef] [PubMed]
  13. Hsiao, Y.C.; Yamada, M. The Roles of Peptide Hormones and Their Receptors during Plant Root Development. Genes 2021, 1, 22. [Google Scholar] [CrossRef]
  14. Horak, H. As above, so below: CLE peptide signaling in shoot and root apical meristems. Plant Cell 2022, 4, 1159–1160. [Google Scholar] [CrossRef]
  15. Altmann, M.; Altmann, S.; Rodriguez, P.A.; Weller, B.; Vergara, L.E.; Palme, J.; la Rosa, N.M.D.; Sauer, M.; Wenig, M.; Villaécija-Aguilar, J.A.; et al. Extensive signal integration by the phytohormone protein network. Nature 2020, 583, 271–276. [Google Scholar] [CrossRef]
  16. Luo, Z.P.; Wang, J.; Li, F.Y.; Lu, Y.T.; Fang, Z.J.; Fu, M.D.; Mysore, K.S.; Wen, J.Q.; Gong, J.M.; Murray, J.D.; et al. The small peptide CEP1 and the NIN-like protein NLP1 regulate NRT2.1 to mediate root nodule formation across nitrate concentrations. Plant Cell 2023, 2, 776–794. [Google Scholar] [CrossRef] [PubMed]
  17. Couzigou, J.-M.; Andre, O.; Guillotin, B.; Alexandre, M.; Combier, J.-P. Use of microRNA-encoded peptide miPEP172c to stimulate nodulation in soybean. New Phytol. 2016, 2, 379–381. [Google Scholar] [CrossRef]
  18. Shinohara, H. Root meristem growth factor RGF, a sulfated peptide hormone in plants. Peptides 2021, 142, 170556. [Google Scholar] [CrossRef]
  19. Geng, R.; Shan, Y.; Li, L.; Shi, C.L.; Zhang, W.; Wang, J.; Sarwar, R.; Xue, Y.X.; Li, Y.L.; Zhu, K.M.; et al. CRISPR-mediated BnaIDA editing prevents silique shattering, floral organ abscission, and spreading of Sclerotinia sclerotiorum in Brassica napus. Plant Commun. 2022, 6, 4. [Google Scholar] [CrossRef]
  20. Zhou, H.P.; Xiao, F.; Zheng, Y.; Liu, G.Y.; Zhuang, Y.F.; Wang, Z.Y.; Zhang, Y.Y.; He, J.X.; Fu, C.X.; Lin, H.H. PAMP-INDUCED SECRETED PEPTIDE 3 modulates salt tolerance through RECEPTOR-LIKE KINASE 7 in plants. Plant Cell 2022, 2, 927–944. [Google Scholar] [CrossRef]
  21. Sin, W.C.; Lam, H.M.; Ngai, S.M. Identification of Diverse Stress-Responsive Xylem Sap Peptides in Soybean. Int. J. Mol. Sci. 2022, 15, 8641. [Google Scholar] [CrossRef]
  22. Stegmann, M.; Monaghan, J.; Smakowska-Luzan, E.; Rovenich, H.; Lehner, A.; Holton, N.; Belkhadir, Y.; Zipfel, C. The receptor kinase FER is a RALF-regulated scaffold controlling plant immune signaling. Science 2017, 6322, 287–289. [Google Scholar] [CrossRef]
  23. Matsubayashi, Y. Recent progress in research on small post-translationally modified peptide signals in plants. Genes. Cells 2012, 1, 1–10. [Google Scholar] [CrossRef]
  24. Onrubia, M.; Pollier, J.; Vanden Bossche, R.; Goethals, M.; Gevaert, K.; Moyano, E.; Vidal-Limon, H.; Cusido, R.M.; Palazon, J.; Goossens, A. Taximin, a conserved plant-specific peptide is involved in the modulation of plant-specialized metabolism. Plant Biotechnol. J. 2015, 12, 971–983. [Google Scholar] [CrossRef]
  25. Tang, W.H.; Ezcurra, I.; Muschietti, J.; McCormick, S. A cysteine-rich extracellular protein, LAT52, interacts with the extracellular domain of the pollen receptor kinase LePRK2. Plant Cell 2002, 9, 2277–2287. [Google Scholar] [CrossRef] [PubMed]
  26. Guillou, M.C.; Balliau, T.; Vergne, E.; Canut, H.; Chourre, J.; Herrera-Leon, C.; Ramos-Martin, F.; Ahmadi-Afzadi, M.; D’Amelio, N.; Ruelland, E.; et al. The PROSCOOP10 Gene Encodes Two Extracellular Hydroxylated Peptides and Impacts Flowering Time in Arabidopsis. Plants 2022, 24, 3554. [Google Scholar] [CrossRef] [PubMed]
  27. Guillou, M.C.; Vergne, E.; Aligon, S.; Pelletier, S.; Simonneau, F.; Rolland, A.; Chabout, S.; Mouille, G.; Gully, K.; Grappin, P.; et al. The peptide SCOOP12 acts on reactive oxygen species homeostasis to modulate cell division and elongation in Arabidopsis primary root. J. Exp. Bot. 2022, 18, 6115–6132. [Google Scholar] [CrossRef]
  28. Venancio, T.M.; Aravind, L. CYSTM, a novel cysteine-rich transmembrane module with a role in stress tolerance across eukaryotes. Bioinformatics 2010, 2, 149–152. [Google Scholar] [CrossRef]
  29. Matsubayashi, Y. Posttranslationally Modified Small-Peptide Signals in Plants. Annu. Rev. Plant Biol. 2014, 65, 385–413. [Google Scholar] [CrossRef] [PubMed]
  30. Stuhrwohldt, N.; Schaller, A. Regulation of plant peptide hormones and growth factors by post-translational modification. Plant Biol. 2019, 21, 49–63. [Google Scholar] [CrossRef]
  31. Whitewoods, C.D. Evolution of CLE peptide signalling. Semin. Cell Dev. Biol. 2021, 109, 12–19. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, G.D.; Fiers, M. CLE peptide signaling during plant development. Protoplasma 2010, 240, 33–43. [Google Scholar] [CrossRef] [PubMed]
  33. Willoughby, A.C.; Nimchuk, Z.L. WOX going on: CLE peptides in plant development. Curr. Opin. Plant Biol. 2021, 63, 102056. [Google Scholar] [CrossRef] [PubMed]
  34. Kucukoglu, M.; Nilsson, O. CLE peptide signaling in plants-the power of moving around. Physiol. Plant 2015, 1, 74–87. [Google Scholar] [CrossRef] [PubMed]
  35. Hirakawa, Y. CLAVATA3, a plant peptide controlling stem cell fate in the meristem. Peptides 2021, 142, 170579. [Google Scholar] [CrossRef] [PubMed]
  36. Wu, Q.Y.; Schmidt, W.; Aalen, R.B.; Xu, C.; Takahashi, F. Editorial: Peptide Signaling in Plants. Front. Plant Sci. 2022, 13, 3918. [Google Scholar] [CrossRef]
  37. Yamaguchi, Y.L.; Ishida, T.; Sawa, S. CLE peptides and their signaling pathways in plant development. J. Exp. Bot. 2016, 16, 4813–4826. [Google Scholar] [CrossRef]
  38. Fletcher, J.C. Recent Advances in Arabidopsis CLE Peptide Signaling. Trends Plant Sci. 2020, 10, 1005–1016. [Google Scholar] [CrossRef]
  39. Stuhrwohldt, N.; Buhler, E.; Sauter, M.; Schaller, A. Phytosulfokine (PSK) precursor processing by subtilase SBT3.8 and PSK signaling improve drought stress tolerance in Arabidopsis. J. Exp. Bot. 2021, 9, 3427–3440. [Google Scholar] [CrossRef]
  40. Frederick, R.O.; Haruta, M.; Tonelli, M.; Lee, W.; Cornilescu, G.; Cornilescu, C.C.; Sussman, M.R.; Markley, J.L. Function and solution structure of the Arabidopsis thaliana RALF8 peptide. Protein Sci. 2019, 6, 1115–1126. [Google Scholar] [CrossRef]
  41. Gao, Q.F.; Wang, C.; Xi, Y.S.; Shao, Q.L.; Hou, C.C.; Li, L.G.; Luan, S. RALF signaling pathway activates MLO calcium channels to maintain pollen tube integrity. Cell Res. 2023, 33, 71–79. [Google Scholar] [CrossRef] [PubMed]
  42. Zhu, S.R.; Pacheco, J.M.; Estevez, J.M.; Yu, F. Autocrine regulation of root hair size by the RALF-FERONIA-RSL4 signaling pathway. New Phytol. 2020, 1, 45–49. [Google Scholar] [CrossRef] [PubMed]
  43. Zhong, S.; Li, L.; Wang, Z.J.; Ge, Z.X.; Li, Q.Y.; Bleckmann, A.; Wang, J.Z.; Song, Z.H.; Shi, Y.H.; Liu, T.X.; et al. RALF peptide signaling controls the polytubey block in Arabidopsis. Science 2022, 6578, 290–296. [Google Scholar] [CrossRef] [PubMed]
  44. Mamaeva, A.; Lyapina, I.; Knyazev, A.; Golub, N.; Mollaev, T.; Chudinova, E.; Elansky, S.; Babenko, V.V.; Veselovsky, V.A.; Klimina, K.M.; et al. RALF peptides modulate immune response in the moss Physcomitrium patens. Front. Plant Sci. 2023, 14, 1077301. [Google Scholar] [CrossRef] [PubMed]
  45. Hung, C.Y.; Kittur, F.S.; Wharton, K.N.; Umstead, M.L.; Burwell, D.B.; Thomas, M.; Qi, Q.; Zhang, J.H.; Oldham, C.E.; Burkey, K.O.; et al. A Rapid Alkalinization Factor-like Peptide EaF82 Impairs Tapetum Degeneration during Pollen Development through Induced ATP Deficiency. Cells 2023, 11, 1542. [Google Scholar] [CrossRef] [PubMed]
  46. Breiden, M.; Olsson, V.; Blumke, P.; Schlegel, J.; Gustavo-Pinto, K.; Dietrich, P.; Butenko, M.A.; Simon, R. The Cell Fate Controlling CLE40 Peptide Requires CNGCs to Trigger Highly Localized Ca2+ Transients in Arabidopsis thaliana Root Meristems. Plant Cell Physiol. 2021, 8, 1290–1301. [Google Scholar] [CrossRef]
  47. Fu, B.L.; Xu, Z.P.; Lei, Y.T.; Dong, R.; Wang, Y.A.; Guo, X.L.; Zhu, H.; Cao, Y.R.; Yan, Z. A novel secreted protein, NISP1, is phosphorylated by soybean Nodulation Receptor Kinase to promote nodule symbiosis. J. Integr. Plant Biol. 2023, 5, 1297–1311. [Google Scholar] [CrossRef] [PubMed]
  48. Solís-Miranda, J.; Juárez-Verdayes, M.A.; Nava, N.; Rosas, P.; Leija-Salas, A.; Cárdenas, L.; Quinto, C. The Phaseolus vulgaris Receptor-Like Kinase PvFER1 and the Small Peptides PvRALF1 and PvRALF6 Regulate Nodule Number as a Function of Nitrate Availability. Int. J. Mol. Sci. 2023, 6, 5230. [Google Scholar] [CrossRef] [PubMed]
  49. Yun, J.X.; Wang, C.; Zhang, F.R.; Chen, L.; Sun, Z.X.; Cai, Y.P.; Luo, Y.Q.; Liao, J.W.; Wang, Y.L.; Cha, Y.Y.; et al. A nitrogen fixing symbiosis-specific pathway required for legume flowering. Sci. Adv. 2023, 2, eade1150. [Google Scholar] [CrossRef]
  50. Zhang, M.B.; Su, H.N.; Gresshoff, P.M.; Ferguson, B.J. Shoot-derived miR2111 controls legume root and nodule development. Plant Cell Environ. 2021, 5, 1627–1641. [Google Scholar] [CrossRef]
  51. Moreau, C.; Gautrat, P.; Frugier, F. Nitrate-induced CLE35 signaling peptides inhibit nodulation through the SUNN receptor and miR2111 repression. Plant Physiol. 2021, 3, 1216–1228. [Google Scholar] [CrossRef]
  52. Tsikou, D.; Yan, Z.; Holt, D.B.; Abel, N.B.; Reid, D.E.; Madsen, L.H.; Bhasin, H.; Sexauer, M.; Stougaard, J.; Markmann, K. Systemic control of legume susceptibility to rhizobial infection by a mobile microRNA. Science 2018, 6411, 233–235. [Google Scholar] [CrossRef]
  53. Gautrat, P.; Laffont, C.; Frugier, F. Compact Root Architecture 2 Promotes Root Competence for Nodulation through the miR2111 Systemic Effector. Curr. Biol. 2020, 7, 1339–1345. [Google Scholar] [CrossRef]
  54. Nanjo, Y.; Maruyama, K.; Yasue, H.; Yamaguchi-Shinozaki, K.; Shinozaki, K.; Komatsu, S. Transcriptional responses to flooding stress in roots including hypocotyl of soybean seedlings. Plant Mol. Biol. 2011, 77, 129–144. [Google Scholar] [CrossRef]
  55. Liu, Y.T.; Zuo, T.T.; Qiu, Z.W.; Zhuang, K.Q.; Hu, S.P.; Han, H.B. Genome-wide identification reveals the function of CEP peptide in cucumber root development. Plant Physiol. Biochem. 2021, 169, 119–126. [Google Scholar] [CrossRef]
  56. Huang, A.X.; Cui, T.T.; Zhang, Y.; Ren, X.F.; Wang, M.F.; Jia, L.Y.; Zhang, Y.H.; Wang, G.D. CRISPR/Cas9-Engineered Large Fragment Deletion Mutations in Arabidopsis CEP Peptide-Encoding Genes Reveal Their Role in Primary and Lateral Root Formation. Plant Cell Physiol. 2023, 1, 19–26. [Google Scholar] [CrossRef]
  57. Qiu, Z.W.; Zhuang, K.Q.; Liu, Y.T.; Ge, X.M.; Chen, C.; Hu, S.P.; Han, H.B. Functional characterization of C-TERMINALLY ENCODED PEPTIDE (CEP) family in Brassica rapa L. Plant Signal. Behav. 2022, 1, 10. [Google Scholar] [CrossRef]
  58. Xiang, D.; Meng, F.N.; Wang, A.D.; Wu, Y.R.; Wang, Z.Y.; Zheng, S.J.; Mao, C.A.Z. Root-secreted peptide OsPEP1 regulates primary root elongation in rice. Plant J. 2021, 2, 480–492. [Google Scholar] [CrossRef]
  59. De Rybel, B.; Vassileva, V.; Parizot, B.; Demeulenaere, M.; Grunewald, W.; Audenaert, D.; Van Campenhout, J.; Overvoorde, P.; Jansen, L.; Vanneste, S.; et al. A Novel Aux/IAA28 Signaling Cascade Activates GATA23-Dependent Specification of Lateral Root Founder Cell Identity. Curr. Biol. 2010, 19, 1697–1706. [Google Scholar] [CrossRef]
  60. Shumilina, J.; Kiryushkin, A.S.; Frolova, N.; Mashkina, V.; Ilina, E.L.; Puchkova, V.A.; Danko, K.; Silinskaya, S.; Serebryakov, E.B.; Soboleva, A.; et al. Integrative Proteomics and Metabolomics Analysis Reveals the Role of Small Signaling Peptide Rapid Alkalinization Factor 34 (RALF34) in Cucumber Roots. Int. J. Mol. Sci. 2023, 8, 7654. [Google Scholar] [CrossRef]
  61. Zhang, B.L.; Xin, B.N.; Sun, X.Q.; Chao, D.; Zheng, H.W.; Peng, L.Y.; Chen, X.X.; Zhang, L.; Yu, J.Y.; Ma, D.; et al. Small peptide signaling via OsCIF1/2 mediates Casparian strip formation at the root endodermal and nonendodermal cell layers in rice. Plant Cell 2023, 36, 383–403. [Google Scholar] [CrossRef]
  62. Zhang, Z.L.; Liu, C.; Li, K.; Li, X.X.; Xu, M.M.; Guo, Y.F. CLE14 functions as a “brake signal”to suppress age-dependent and stress-induced leaf senescence by promoting JUB1-mediated ROS scavenging in Arabidopsis. Mol. Plant 2022, 1, 179–188. [Google Scholar] [CrossRef]
  63. Zhang, Y.; Tan, S.Y.; Gao, Y.H.; Kan, C.C.; Wang, H.L.; Yang, Q.; Xia, X.L.; Ishida, T.; Sawa, S.; Guo, H.W.; et al. CLE42 delays leaf senescence by antagonizing ethylene pathway in Arabidopsis. New Phytol. 2022, 2, 550–562. [Google Scholar] [CrossRef]
  64. Czyzewicz, N.; Yue, K.; Beeckman, T.; De Smet, I. Message in a bottle: Small signalling peptide outputs during growth and development. J. Exp. Bot. 2013, 17, 5281–5296. [Google Scholar] [CrossRef]
  65. Djami-Tchatchou, A.T.; Tetorya, M.; Godwin, J.; Codjoe, J.M.; Li, H.; Shah, D.M. Small Cationic Cysteine-Rich Defensin-Derived Antifungal Peptide Controls White Mold in Soybean. J. Fungi 2023, 9, 873. [Google Scholar] [CrossRef]
  66. Wang, P.Y.; Jia, H.M.; Guo, T.; Zhang, Y.Y.; Wang, W.Q.; Nishimura, H.; Li, Z.G.; Kawano, Y. The secreted immune response peptide 1 functions as a phytocytokine in rice immunity. J. Exp. Bot. 2023, 3, 1059–1073. [Google Scholar] [CrossRef]
  67. Wang, J.; Wang, D.; Ji, X.B.; Wang, J.; Klosterman, S.J.; Dai, X.F.; Chen, J.Y.; Subbarao, K.V.; Hao, X.J.; Zhang, D.D. The Verticillium dahliae Small Cysteine-Rich Protein VdSCP23 Manipulates Host Immunity. Int. J. Mol. Sci. 2023, 11, 9403. [Google Scholar] [CrossRef]
  68. Wu, Q.Q.; Kuang, K.Y.; Lyu, M.H.; Zhao, Y.; Li, Y.; Li, J.; Pan, Y.; Shi, H.; Zhong, S.W. Allosteric deactivation of PIFs and EIN3 by microproteins in light control of plant development. Proc. Natl. Acad. Sci. USA 2020, 31, 18858–18868. [Google Scholar] [CrossRef]
  69. Roy, S.; Griffiths, M.; Torres-Jerez, I.; Sanchez, B.; Antonelli, E.; Jain, D.; Krom, N.; Zhang, S.L.; York, L.M.; Scheible, W.R.; et al. Application of Synthetic Peptide CEP1 Increases Nutrient Uptake Rates Along Plant Roots. Front. Plant Sci. 2022, 12, 793145. [Google Scholar] [CrossRef]
  70. Su, Q.; Wang, K.; Zhang, Z.Y. Ecotopic Expression of the Antimicrobial Peptide DmAMP1W Improves Resistance of Transgenic Wheat to Two Diseases: Sharp Eyespot and Common Root Rot. Int. J. Mol. Sci. 2020, 2, 647. [Google Scholar] [CrossRef]
  71. Dong, J.H.; Wang, Y.; Xu, L.; Li, B.S.; Wang, K.; Ying, J.L.; He, Q.; Liu, L.W. RsCLE22a regulates taproot growth through an auxin signaling-related pathway in radish (Raphanus sativus L.). J. Exp. Bot. 2023, 1, 233–250. [Google Scholar] [CrossRef]
  72. Liu, P.; Shi, C.; Liu, S.; Lei, J.; Lu, Q.; Hu, H.; Ren, Y.; Zhang, N.; Sun, C.; Chen, L.; et al. A papain-like cysteine protease-released small signal peptide confers wheat resistance to wheat yellow mosaic virus. Nat. Commun. 2023, 1, 7773. [Google Scholar] [CrossRef]
  73. Niu, X.Y.; Yamamoto, N.; Yang, G.J.; Lin, H.; Jiang, L.J.; Liu, Y.; Zheng, A.P. A small secreted protein, RsMf8HN, in Rhizoctonia solani triggers plant immune response, which interacts with rice OsHIPP28. Microbiol. Res. 2023, 266, 127219. [Google Scholar] [CrossRef]
  74. Tong, X.; Zhao, J.J.; Feng, Y.L.; Wang, X.B. The small peptide VISP1 acts as a selective autophagy receptor regulating plant-virus interactions. Autophagy 2023, 12, 3246–3247. [Google Scholar] [CrossRef]
  75. Lan, Z.J.; Song, Z.H.; Wang, Z.J.; Li, L.; Liu, Y.Q.; Zhi, S.H.; Wang, R.H.; Wang, J.Z.; Li, Q.Y.; Bleckmann, A.; et al. Antagonistic RALF peptides control an intergeneric hybridization barrier on Brassicaceae stigmas. Cell 2023, 22, 4773–4787. [Google Scholar] [CrossRef]
  76. Takahashi, F.; Hanada, K.; Kondo, T.; Shinozaki, K. Hormone-like peptides and small coding genes in plant stress signaling and development. Curr. Opin. Plant Biol. 2019, 51, 88–95. [Google Scholar] [CrossRef]
  77. Kim, S.I.; Lee, K.H.; Kwak, J.S.; Kwon, D.H.; Song, J.T.; Seo, H.S. Overexpression of Rice OsS1Fa1 Gene Confers Drought Tolerance in Arabidopsis. Plants 2021, 10, 2181. [Google Scholar] [CrossRef]
  78. Cui, Y.C.; Li, M.J.; Yin, X.M.; Song, S.F.; Xu, G.Y.; Wang, M.L.; Li, C.Y.; Peng, C.; Xia, X.J. OsDSSR1, a novel small peptide, enhances drought tolerance in transgenic rice. Plant Sci. 2018, 270, 85–96. [Google Scholar] [CrossRef]
  79. Jing, X.-Q.; Shi, P.-T.; Zhang, R.; Zhou, M.-R.; Shalmani, A.; Wang, G.-F.; Liu, W.-T.; Li, W.-Q.; Chen, K.-M. Rice kinase OsMRLK63 contributes to drought tolerance by regulating reactive oxygen species production. Plant Physiol. 2023, 194, 2679–2696. [Google Scholar] [CrossRef]
  80. Xu, K.X.; Tian, D.D.; Wang, T.J.; Zhang, A.J.; Elsadek, M.A.Y.; Liu, W.H.; Chen, L.P.; Guo, Y.F. Small secreted peptides (SSPs) in tomato and their potential roles in drought stress response. Mol. Hortic. 2023, 1, 17. [Google Scholar] [CrossRef]
  81. Tian, D.D.; Xie, Q.; Deng, Z.C.; Xue, J.; Li, W.; Zhang, Z.L.; Dai, Y.F.; Zheng, B.; Lu, T.G.; De Smet, I.; et al. Small secreted peptides encoded on the wheat (Triticum aestivum L.) genome and their potential roles in stress responses. Front. Plant Sci. 2022, 18, 1000297. [Google Scholar] [CrossRef]
  82. Christmann, A.; Grill, E. Peptide signal alerts plants to drought. Nature 2018, 7700, 178–179. [Google Scholar] [CrossRef]
  83. Takahashi, F.; Suzuki, T.; Osakabe, Y.; Betsuyaku, S.; Kondo, Y.; Dohmae, N.; Fukuda, H.; Yamaguchi-Shinozaki, K.; Shinozaki, K. A small peptide modulates stomatal control via abscisic acid in long-distance signalling. Nature 2018, 7700, 235–238. [Google Scholar] [CrossRef]
  84. Farago, D.; Zsigmond, L.; Benyo, D.; Alcazar, R.; Rigo, G.; Ayaydin, F.; Rabilu, S.A.; Hunyadi-Gulyas, E.; Szabados, L. Small paraquat resistance proteins modulate paraquat and ABA responses and confer drought tolerance to overexpressing Arabidopsis plants. Plant Cell Environ. 2022, 7, 1985–2003. [Google Scholar] [CrossRef]
  85. Nagar, P.; Kumar, A.; Jain, M.; Kumari, S.; Mustafiz, A. Genome-wide analysis and transcript profiling of PSKR gene family members in Oryza sativa. PLoS ONE 2020, 7, e0236349. [Google Scholar] [CrossRef]
  86. Reichardt, S.; Piepho, H.P.; Stintzi, A.; Schaller, A. Peptide signaling for drought-induced tomato flower drop. Science 2020, 6485, 1482–1485. [Google Scholar] [CrossRef]
  87. Hao, Z.D.; Wu, H.; Zheng, R.H.; Li, R.; Zhu, Z.L.; Chen, Y.; Lu, Y.; Cheng, T.L.; Shi, J.S.; Chen, J.H. The plant peptide hormone phytosulfokine promotes somatic embryogenesis by maintaining redox homeostasis in Cunninghamia lanceolata. Plant J. 2023, 113, 716–733. [Google Scholar] [CrossRef]
  88. Zhang, J.; Yuan, J.; Li, Z.; Fu, C.; Xu, M.; Yang, J.; Jiang, X.; Zhou, B.; Ye, X.; Xu, C. Exploring and exploiting plant cyclic peptides for drug discovery and development. Med. Res. Rev. 2021, 6, 3096–3117. [Google Scholar] [CrossRef]
  89. Nakaminami, K.; Okamoto, M.; Higuchi-Takeuchi, M.; Yoshizumi, T.; Yamaguchi, Y.; Fukao, Y.; Shimizu, M.; Ohashi, C.; Tanaka, M.; Matsui, M.; et al. AtPep3 is a hormone-like peptide that plays a role in the salinity stress tolerance of plants. Proc. Natl. Acad. Sci. USA 2018, 22, 5810–5815. [Google Scholar] [CrossRef]
  90. Zhang, X.; Yang, Z.H.; Wu, D.S.; Yu, F. RALF-FERONIA Signaling: Linking Plant Immune Response with Cell Growth. Plant Commun. 2020, 4, 100084. [Google Scholar] [CrossRef]
  91. Zhao, C.Z.; Zayed, O.; Yu, Z.P.; Jiang, W.; Zhu, P.P.; Hsu, C.C.; Zhang, L.R.; Tao, W.A.; Lozano-Duran, R.; Zhu, J.K. Leucine-rich repeat extensin proteins regulate plant salt tolerance in Arabidopsis. Proc. Natl. Acad. Sci. USA 2018, 51, 13123–13128. [Google Scholar] [CrossRef] [PubMed]
  92. Xu, Y.; Yu, Z.P.; Zhang, D.; Huang, J.G.; Wu, C.G.; Yang, G.D.; Yan, K.; Zhang, S.Z.; Zheng, C.C. CYSTM, a Novel Non-Secreted Cysteine-Rich Peptide Family, Involved in Environmental Stresses in Arabidopsis thaliana. Plant Cell Physiol. 2018, 2, 423–438. [Google Scholar] [CrossRef] [PubMed]
  93. Xu, Y.; Yu, Z.P.; Zhang, S.Z.; Wu, C.A.; Yang, G.D.; Yan, K.; Zheng, C.C.; Huang, J.G. CYSTM3 negatively regulates salt stress tolerance in Arabidopsis. Plant Mol. Biol. 2019, 99, 395–406. [Google Scholar] [CrossRef]
  94. Guo, Z.H.; Cai, L.J.; Liu, C.X.; Chen, Z.Q.; Guan, S.W.; Ma, W.D.; Pan, G.J. Low-temperature stress affects reactive oxygen species, osmotic adjustment substances, and antioxidants in rice (Oryza sativa L.) at the reproductive stage. Sci. Rep. 2022, 1, 13. [Google Scholar] [CrossRef] [PubMed]
  95. Khan, T.A.; Fariduddin, Q.; Yusuf, M. Low-temperature stress: Is. phytohormones application a remedy? Environ. Sci. Pollut. Res. 2017, 27, 21574–21590. [Google Scholar] [CrossRef] [PubMed]
  96. Chen, Q.J.; Zhang, L.P.; Song, S.R.; Wang, L.; Xu, W.P.; Zhang, C.X.; Wang, S.P.; Liu, H.F.; Ma, C. vvi-miPEP172b and vvi-miPEP3635b increase cold tolerance of grapevine by regulating the corresponding MIRNA genes. Plant Sci. 2022, 10, 111450. [Google Scholar] [CrossRef] [PubMed]
  97. Wang, G.; Zhang, G.; Wu, M. CLE Peptide Signaling and Crosstalk with Phytohormones and Environmental Stimuli. Front. Plant Sci. 2015, 1211, 167813. [Google Scholar] [CrossRef]
  98. Meng, X.X.; Li, W.F.; Shen, R.F.; Lan, P. Ectopic expression of IMA small peptide genes confers tolerance to cadmium stress in Arabidopsis through activating the iron deficiency response. J. Hazard. Mater. 2022, 14, 126913. [Google Scholar] [CrossRef]
  99. Lu, L.; Chen, X.; Chen, J.; Zhang, Z.; Zhang, Z.; Sun, Y.; Wang, Y.; Xie, S.; Ma, Y.; Song, Y.; et al. MicroRNA-encoded regulatory peptides modulate cadmium tolerance and accumulation in rice. Plant Cell Environ. 2024. [Google Scholar] [CrossRef]
  100. Vie, A.K.; Najafi, J.; Winge, P.; Cattan, E.; Wrzaczek, M.; Kangasjarvi, J.; Miller, G.; Brembu, T.; Bones, A.M. The IDA-LIKE peptides IDL6 and IDL7 are negative modulators of stress responses in Arabidopsis thaliana. J. Exp. Bot. 2017, 13, 3557–3571. [Google Scholar] [CrossRef]
  101. Segonzac, C.; Monaghan, J. Modulation of plant innate immune signaling by small peptides. Curr. Opin. Plant Biol. 2019, 51, 22–28. [Google Scholar] [CrossRef]
  102. Tang, J.; Wu, D.S.; Li, X.X.; Wang, L.F.; Xu, L.; Zhang, Y.; Xu, F.; Liu, H.B.; Xie, Q.J.; Dai, S.J.; et al. Plant immunity suppression via PHR1-RALF-FERONIA shapes the root microbiome to alleviate phosphate starvation. EMBO J. 2022, 6, e109102. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, X.Y.; Zhang, N.; Zhang, L.N.; He, Y.X.; Cai, C.; Zhou, J.G.; Li, J.; Meng, X.Z. Perception of the pathogen-induced peptide RGF7 by the receptor-like kinases RGI4 and RGI5 triggers innate immunity in Arabidopsis thaliana. New Phytol. 2021, 3, 1110–1125. [Google Scholar] [CrossRef] [PubMed]
  104. Tong, X.; Liu, S.Y.; Zou, J.Z.; Zhao, J.J.; Zhu, F.F.; Chai, L.X.; Wang, Y.; Han, C.G.; Wang, X.B. A small peptide inhibits siRNA amplification in plants by mediating autophagic degradation of SGS3/RDR6 bodies. EMBO J. 2021, 15, e108050. [Google Scholar] [CrossRef] [PubMed]
  105. Combest, M.M.; Moroz, N.; Tanaka, K.; Rogan, C.J.; Anderson, J.C.; Thura, L.; Rakotondrafara, A.M.; Goyer, A. StPIP1, a PAMP-induced peptide in potato, elicits plant defenses and is associated with disease symptom severity in a compatible interaction with Potato virus Y. J. Exp. Bot. 2021, 12, 4472–4488. [Google Scholar] [CrossRef] [PubMed]
  106. Okamoto, S.; Kawasaki, A.; Makino, Y. Characterization of Oligopeptides in Solanum lycopersicum Xylem Exudates. Life 2022, 4, 592. [Google Scholar] [CrossRef] [PubMed]
  107. Koenig, A.M.; Hoffmann-Benning, S. The interplay of phloem-mobile signals in plant development and stress response. Biosci. Rep. 2020, 20, BSR20193329. [Google Scholar] [CrossRef]
  108. Chapman, K.; Ivanovici, A.; Taleski, M.; Sturrock, C.J.; Ng, J.L.P.; Mohd-Radzman, N.A.; Frugier, F.; Bennett, M.J.; Mathesius, U.; Djordjevic, M.A. CEP receptor signalling controls root system architecture in Arabidopsis and Medicago. New Phytol. 2020, 6, 1809–1821. [Google Scholar] [CrossRef] [PubMed]
  109. Chapman, K.; Taleski, M.; Ogilvie, H.A.; Imin, N.; Djordjevic, M.A. CEP-CEPR1 signalling inhibits the sucrose-dependent enhancement of lateral root growth. J. Exp. Bot. 2019, 15, 3955–3967. [Google Scholar] [CrossRef]
  110. Xie, Y.H.; Zhang, F.J.; Sun, P.; Li, Z.Y.; Zheng, P.F.; Gu, K.D.; Hao, Y.J.; Zhang, Z.L.; You, C.X. Apple receptor-like kinase FERONIA regulates salt tolerance and ABA sensitivity in Malus domestica. J. Plant Physiol. 2022, 270, 1536161. [Google Scholar] [CrossRef]
  111. Zhao, C.Z.; Jiang, W.; Zayed, O.; Liu, X.; Tang, K.; Nie, W.F.; Li, Y.L.; Xie, S.J.; Li, Y.; Long, T.D.; et al. The LRXs-RALFs-FER module controls plant growth and salt stress responses by modulating multiple plant hormones. Natl. Sci. Rev. 2021, 1, nwaa149. [Google Scholar] [CrossRef] [PubMed]
  112. Lucena, C.; Waters, B.M.; Romera, F.J.; García, M.J.; Morales, M.; Alcántara, E.; Pérez-Vicente, R. Ethylene could influence ferric reductase, iron transporter, and H-ATPase gene expression by affecting FER or FER-like) gene activity. J. Exp. Bot. 2006, 15, 4145–4154. [Google Scholar] [CrossRef]
  113. Yang, R.R.; Wang, Z.C.; Zhao, L.; Liu, J.; Meng, J.; Luan, Y.S. Secreted Peptide SpPIP1 Modulates Disease Resistance and Salt Tolerance in Tomato. J. Agric. Food Chem. 2023, 32, 12264–12279. [Google Scholar] [CrossRef] [PubMed]
  114. Wang, D.H.; Hu, X.; Ye, H.Z.; Wang, Y.; Yang, Q.; Liang, X.D.; Wang, Z.L.; Zhou, Y.F.; Wen, M.M.; Yuan, X.Y.; et al. Cell-specific clock-controlled gene expression program regulates rhythmic fiber cell growth in cotton. Genome Biol. 2023, 1, 49. [Google Scholar] [CrossRef] [PubMed]
  115. Smith, S.; Zhu, S.S.; Joos, L.; Roberts, I.; Nikonorova, N.; Vu, L.D.; Stes, E.; Cho, H.; Larrieu, A.; Xuan, W.; et al. The CEP5 Peptide Promotes Abiotic Stress Tolerance, As Revealed by Quantitative Proteomics, and Attenuates the AUX/IAA Equilibrium in Arabidopsis. Mol. Cell. Proteom. 2020, 8, 1248–1262. [Google Scholar] [CrossRef] [PubMed]
  116. Yang, L.Y.; Hu, X.C.; Ren, M.F.; Ma, F.; Fu, J.; Cui, H.C. Stem-cell-expressed DEVIL-like small peptides maintain root growth under abiotic stress via abscisic acid signaling. Plant Physiol. 2023, 194, 2372–2386. [Google Scholar] [CrossRef] [PubMed]
  117. Liu, M.A.; Wang, C.Y.; Ji, Z.; Lu, J.Y.; Zhang, L.; Li, C.L.; Huang, J.G.; Yang, G.D.; Yan, K.; Zhang, S.Z.; et al. Regulation of drought tolerance in Arabidopsis involves the PLATZ4-mediated transcriptional repression of plasma membrane aquaporin PIP2;8. Plant J. 2023, 115, 434–451. [Google Scholar] [CrossRef] [PubMed]
  118. Chapman, K.; Taleski, M.; Frank, M.; Djordjevic, M.A. C-TERMINALLY ENCODED PEPTIDE (CEP) and cytokinin hormone signaling intersect to promote shallow lateral root angles. J. Exp. Bot. 2023, 75, 631–641. [Google Scholar] [CrossRef] [PubMed]
  119. Song, H.W.; Wang, X.F.; Hu, W.C.; Yang, X.H.; Diao, E.J.; Shen, T.; Qiang, Q. A cold-induced phytosulfokine peptide is related to the improvement of loquat fruit chilling tolerance. Food Chem. 2017, 232, 434–442. [Google Scholar] [CrossRef]
  120. Li, Z.S.; Jin, J.R.; Wang, Y.; Long, W.T.; Ding, Y.H.; Hu, H.Y.; Wei, L.Y. ExamPle: Explainable deep learning framework for the prediction of plant small secreted peptides. Bioinformatics 2023, 3, btad108. [Google Scholar] [CrossRef]
  121. Teufel, F.; Armenteros, J.J.A.; Johansen, A.R.; Gíslason, M.H.; Pihl, S.I.; Tsirigos, K.D.; Winther, O.; Brunak, S.; von Heijne, G.; Nielsen, H. SignalP 6.0 predicts all five types of signal peptides using protein language models. Nat. Biotechnol. 2022, 7, 1023–1025. [Google Scholar] [CrossRef] [PubMed]
  122. Song, X.F.; Guo, P.; Ren, S.C.; Xu, T.T.; Liu, C.M. Antagonistic Peptide Technology for Functional Dissection of CLV3/ESR Genes in Arabidopsis. Plant Physiol. 2013, 3, 1076–1085. [Google Scholar] [CrossRef] [PubMed]
  123. Bruck, S.; Pfannstiel, J.; Ingram, G.; Stintzi, A.; Schaller, A. Analysis of peptide hormone maturation and processing specificity using isotope-labeled peptides. In Methods in Molecular Biology; Springer: Clifton, NJ, USA, 2023; pp. 323–335. [Google Scholar]
  124. Ma, D.C.; Endo, S.; Betsuyaku, E.; Fujiwara, T.; Betsuyaku, S.; Fukuda, H. Root-specific CLE3 expression is required for WRKY33 activation in Arabidopsis shoots. Plant Mol. Biol. 2022, 3, 225–239. [Google Scholar] [CrossRef] [PubMed]
  125. Okamoto, S.; Shinohara, H.; Mori, T.; Matsubayashi, Y.; Kawaguchi, M. Root-derived CLE glycopeptides control nodulation by direct binding to HAR1 receptor kinase. Nat. Commun. 2013, 4, 2191. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The function of small plant peptide in regulating growth and development. Small plant peptides exert different effects in various parts of the plant, including roots, stems, leaves, and flowers. In root regulation, both CLE and CEP can be transported from the roots to the stem, acting on mir2111. CLE has an inhibitory effect on root nodules, while CEP promotes the formation of plant root nodules. TDIF positively regulates lateral root development by promoting the phosphorylation of ARF receptors. Additionally, the peptides OsPEP1 and RALF34 have promoting effects on lateral root development in rice and in Arabidopsis, respectively. SCOOP stimulates root growth by interacting with MIK2, leading to the formation of a co-receptor complex with BAK1. In leaf regulation, CLE14 delays leaf senescence by inhibiting the expression of genes related to reactive oxygen species (ROS) synthesis, while CLE42 slows down leaf senescence by closing the ethylene synthesis pathway. Regarding flower organ regulation, IDA inhibits the abscission of flower organs by acting on the receptor HAE/HSL2. The arrow and T-line without arrowheads respectively represent positive and negative regulation in the signaling pathway. Solid lines indicate direct regulation, while dashed lines indicate the presence of multiple steps or unknown components. Cross mark represents suppression.
Figure 1. The function of small plant peptide in regulating growth and development. Small plant peptides exert different effects in various parts of the plant, including roots, stems, leaves, and flowers. In root regulation, both CLE and CEP can be transported from the roots to the stem, acting on mir2111. CLE has an inhibitory effect on root nodules, while CEP promotes the formation of plant root nodules. TDIF positively regulates lateral root development by promoting the phosphorylation of ARF receptors. Additionally, the peptides OsPEP1 and RALF34 have promoting effects on lateral root development in rice and in Arabidopsis, respectively. SCOOP stimulates root growth by interacting with MIK2, leading to the formation of a co-receptor complex with BAK1. In leaf regulation, CLE14 delays leaf senescence by inhibiting the expression of genes related to reactive oxygen species (ROS) synthesis, while CLE42 slows down leaf senescence by closing the ethylene synthesis pathway. Regarding flower organ regulation, IDA inhibits the abscission of flower organs by acting on the receptor HAE/HSL2. The arrow and T-line without arrowheads respectively represent positive and negative regulation in the signaling pathway. Solid lines indicate direct regulation, while dashed lines indicate the presence of multiple steps or unknown components. Cross mark represents suppression.
Ijms 25 04114 g001
Figure 2. Related signal pathways of small peptides in drought, salt, temperature, and heavy metals stress. The arrow and T-line without arrowheads respectively represent positive and negative regulation in the signaling pathway. Solid lines indicate direct regulation, while dashed lines indicate the presence of multiple steps or unknown components.
Figure 2. Related signal pathways of small peptides in drought, salt, temperature, and heavy metals stress. The arrow and T-line without arrowheads respectively represent positive and negative regulation in the signaling pathway. Solid lines indicate direct regulation, while dashed lines indicate the presence of multiple steps or unknown components.
Ijms 25 04114 g002
Table 1. Different types of small peptides and their functions in plant.
Table 1. Different types of small peptides and their functions in plant.
TypesSmall PeptidesFunctionReferences
Non-secretory small peptidesearly nodulinroot nodule development, intracellular localization of proteins, etc.[5,6,7,8,9]
ROT4regulates plant immunity, modulates cell proliferation, etc.[11]
Post-translationally modified peptidesPSKabiotic stress, cell division, root development, etc.[2]
TDIFtracheary element differentiation inhibitory factor, root development, etc.[12]
CLEabiotic stress, root development, leaf senescence, etc.[13]
CLV3differentiation of the apical meristem, etc.[14]
CIFformation of the Casparian strip, etc.[15]
CEPabiotic stress, root development, nodulation, etc.[16]
miPEPabiotic stress, fruit ripening, flower development, photosynthesis, nodulation, etc.[17]
RGFmaintaining the activity of the root apical meristem, abiotic stress, etc.[18]
IDAfloral organ abscission, etc.[19]
PIPabiotic stress, etc.[20]
PSYabiotic stress, root development, etc.[21]
Cysteine-rich peptidesRALFsroot development, cell wall remodeling, immune response, abiotic stress, etc.[22]
LUREpromoting pollen tube germination, etc.[23]
TAXregulating plant-specific metabolism, etc.[24]
EPFregulating epidermal cell patterns, etc.[23]
LAT52pollen hydration process, pollen tube growth, etc.[25]
SCOOPsroot development, immune response, etc.[26,27]
CYSTMabiotic stress, etc.[28]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ren, G.; Zhang, Y.; Chen, Z.; Xue, X.; Fan, H. Research Progress of Small Plant Peptides on the Regulation of Plant Growth, Development, and Abiotic Stress. Int. J. Mol. Sci. 2024, 25, 4114. https://doi.org/10.3390/ijms25074114

AMA Style

Ren G, Zhang Y, Chen Z, Xue X, Fan H. Research Progress of Small Plant Peptides on the Regulation of Plant Growth, Development, and Abiotic Stress. International Journal of Molecular Sciences. 2024; 25(7):4114. https://doi.org/10.3390/ijms25074114

Chicago/Turabian Style

Ren, Guocheng, Yanling Zhang, Zengting Chen, Xin Xue, and Hai Fan. 2024. "Research Progress of Small Plant Peptides on the Regulation of Plant Growth, Development, and Abiotic Stress" International Journal of Molecular Sciences 25, no. 7: 4114. https://doi.org/10.3390/ijms25074114

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

Article Metrics

Back to TopTop