Next Article in Journal
Sitagliptin Mitigates Diabetic Cardiomyopathy Through Oxidative Stress Reduction and Suppression of VEGF and FLT-1 Expression in Rats
Previous Article in Journal
Oral Pharmacokinetic Evaluation of a Microemulsion-Based Delivery System for Novel A190 Prodrugs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Involvement of Pathogenesis-Related Proteins and Their Roles in Abiotic Stress Responses in Plants

College of Life and Environmental Sciences, Minzu University of China, Beijing 100081, China
*
Author to whom correspondence should be addressed.
Biomolecules 2025, 15(8), 1103; https://doi.org/10.3390/biom15081103
Submission received: 26 June 2025 / Revised: 25 July 2025 / Accepted: 28 July 2025 / Published: 30 July 2025

Abstract

Plant pathogenesis-related (PR) proteins are a large and diverse family of proteins with antimicrobial activity, often induced by pathogen attack. Traditionally, PR proteins were thought to mainly participate in plant defense mechanisms against biotic stress. However, in recent years, increasing evidence has shown that these proteins also play important roles in the response to abiotic stress in plants. In the present review, we provide a summary of the latest findings on PR proteins and focus on their response to various abiotic stresses, the mechanism by which PR proteins are activated by external and internal signals, and their biological functions in plant responses to abiotic stresses. In addition, the existing challenges and future applications are also summarized, aiming to provide a reference for further research on PR proteins in the context of plant physiology.

1. Introduction

It is imperative to enhance the resilience of crops to abiotic stresses to meet the escalating global demand for food. Abiotic stresses, including drought, high salinity, heat, low temperature, heavy metals, ultraviolet light, and waterlogging, are prevalent unfavorable conditions for crop cultivation. Crops frequently exhibit diminished yield and compromised product quality under abiotic stresses. It is evident that climate change, characterized by rising global temperatures, will lead to an increased frequency of extreme weather events, such as heat waves and droughts, across the globe. Despite the considerable efforts of scientists to engineer abiotic stress tolerance in economic plants, there remains a paucity of robust crop varieties that demonstrate substantial resistance to abiotic stresses. In recent years, the field of plant research has seen a surge of interest in the study of the roles of plant pathogenesis-related (PR) proteins in plant responses to abiotic stresses. These proteins were originally recognized to be induced by a variety of abiotic stresses. Moreover, transgenic plants overexpressing PR genes exhibit significant abiotic stress resistance. To enhance the comprehension of the biological functions of PR proteins in relation to abiotic stress tolerance in plants, a comprehensive review of the existing literature was conducted. This review outlines the classification of PR proteins and reported abiotic stressors inducing PR genes and proteins, summarizes the impact of changes in PR genes expression on plant abiotic stress tolerance and the possible regulatory mechanisms underlying the response of PR proteins to abiotic stress, and points out the existing challenges and future applications, with the aim to provide a reference for further research on PR proteins in the context of plant physiology.

2. Overview of PR Proteins

PR proteins refer to a broad category of proteins induced by various biotic or abiotic stresses in plants [1]. They were initially detected in tobacco leaves infected with tobacco mosaic virus (TMV) [2]. The accumulation of PR proteins in both infected and non-infected parts of plants has been shown to induce a hypersensitive reaction (HR), cause local cell death, and result in the development of systemic acquired resistance (SAR) in the plant, thereby limiting the spread of pathogens [3]. The molecular weights of PRs range from 6 kDa to 43 kDa, and these proteins are characterized by their resistance to heat and proteases, as well as their solubility at low pH (<3) [4]. Currently, PR proteins are classified into 17 distinct families (PR-1 to PR-17) based on their plant origin, serological properties, and amino acid sequence homology [1]. These families includes one PR-1 family with unknown biochemical properties, one β-1,3-glucanase family (PR-2), four chitinase families (PR-3, PR-4, PR-8, and PR-11), one protease inhibitor family (PR-6), and one specific peroxidase family (PR-9), as well as the thaumatin-like protein (PR-5) family and ribonuclease-like protein (PR-10) families. The subsequent introduction will provide a brief overview of the 17 PR families, accompanied by a summary of their structural characteristics and primary biological functions.

2.1. PR-1 Proteins

The first member of the PR-1 family was the earliest identified PR protein [2], and its presence was subsequently detected in monocotyledonous and dicotyledonous plants. PR-1 proteins possess molecular weights ranging from 14 to 16 kDa [5], and they are classified into two groups: acidic and basic. Acidic PR-1 proteins are secreted into the extracellular space, while basic PR-1 proteins are predominantly localized within the vacuole [6]. All PR-1 proteins contain a CAP (cysteine-rich secretory protein, antigen 5, pathogenesis-related 1) domain. This CAP domain is folded into a distinctive α-β-α sandwich structure, which generally consists of four α-helices and a four-stranded β-sheet, stabilized by conserved disulfide bonds. The structure of the CAP domain determines the function of PR-1 proteins in plants, thereby conferring them antifungal properties [7]. It has been demonstrated that PR-1 proteins exert their antiviral activity by impeding the binding of viral capsid proteins to plant receptor molecules. This inhibition occurs through a multifaceted mechanism that involves the attack on pathogens, the degradation of cell wall macromolecules, and the prevention of viral capsid protein binding to plant receptor molecules [8]. Beyond their role in biotic stress responses, PR-1 proteins demonstrate responsiveness to various abiotic stresses, including drought, elevated salt levels, low temperatures, high temperatures, and heavy metals [9,10,11]. Furthermore, PR-1 proteins have been identified as playing a pivotal role in a variety of metabolic pathways, including the MAPK signaling pathway and phytohormone signaling [12].

2.2. PR-2 Proteins

PR-2 proteins are members of the glycoside hydrolase superfamily, specifically the β-1,3-glucanase subfamily. PR-2 functions as a catalyst, facilitating the intramolecular cleavage of the 1,3-β-D-glucosidic bond, thereby generating single β-1,3-glucan units. The PR-2 proteins are characterized by the presence of a glycoside hydrolase domain and a secreted signal peptide (NTS). PR-2 level is induced by pathogen infection, plant hormones such as salicylic acid (SA) and abscisic acid (ABA) [13], as well as by low temperatures. β-1,3-glucan is an important component of fungal cell walls. PR-2 proteins directly hydrolyze β-1,3-glucan in fungal cell walls, causing hyphal deformities and lysis, and exhibit antifungal activity both in vitro and in vivo [14]. Moreover, PR-2 degrades the cell walls of pathogens to produce various types of oligosaccharides, which can promote the release of defense elicitors from plants and activate plant immunity [15]. Many pieces of evidence indicate that PR-2 proteins have been implicated in the metabolism of plant callus, the regulation of plant development, and the intercellular transport of hormones and RNA [16,17]. Other studies have shown that PR-2 proteins participate in plant response to various abiotic stresses including drought [18], low temperature [19], osmotic stress [20], salt, and heavy metal stresses [21].

2.3. PR-3, PR-4, PR-8, and PR-11 Proteins

The PR-3, PR-4, PR-8, and PR-11 families are all distinct classes of chitinases. The molecular weights of these proteins range from 25 to 36 kDa, with the majority falling within the glycoside hydrolase GH18 and GH19 families. Chitinases were classified into seven distinct classes based on structural characteristics, substrate specificity, catalytic mechanisms, and sensitivity to inhibitors [22]. Class I chitinases are basic proteins that are secreted to vesicles and plasmodesmata. They have a C-terminal propeptide that helps to target vacuoles, and N-terminal chitin-binding domains (CBDs) that are proline- and glycine-rich. Class II chitinases are acidic proteins that are secreted to the plasmodesmata and they lack a CBD domain. Class III chitinases are acidic proteins that are secreted into vesicles and apoplast, and these proteins contain two N-terminal LysM domains, which contribute to antifungal activity and lack a CBD domain [23,24]. Class IV chitinases are also acidic proteins that are secreted into the apoplast, and their N-terminus contains Hevein domains that can bind to chitin [25]. Class V chitinases are secreted into the apoplast and have a C-terminal extension and two CBD domains [26]. Class VI chitinases are acidic proteins secreted into vesicles and the apoplast. Class VII chitinases do not have any CBD [23,27]. The PR-3 family comprises the class I/II/IV/VI/VII chitinases, which are characterized by the presence of a common CBD domain and a helical structure with the catalytic domain. The PR-4 family and the PR-3 family, despite their classification as class I and II chitinases, exhibit significant structural divergences. PR-4 family proteins contain three intramolecular disulfide bonds and a conserved C-terminal structural domain, BARWIN, which is implicated in antifungal defense [28]. The PR-8 family is classified as class III chitinases, with acidic and basic isoforms. The PR-8 family exhibits both lysozyme activity and the potential for antibacterial activity. The PR-11 family, classified as class V chitinase, lacks the typical CBD domain [24]. In general, PR-3, PR-4, PR-8, and PR-11 proteins have been shown to inhibit fungal spore germination, combat diseases by degrading β-1,4-chitin chains in fungal cell walls and exoskeletons of insects [29,30,31], and exhibit antiviral activity [24,32]. Beyond their antifungal and antiviral functions, chitinases have also been implicated in plant responses to abiotic stresses such as drought [33], low temperatures [34], heavy metals [35], and high temperatures [36].

2.4. PR-5 Proteins

PR-5 proteins are thaumatin-like proteins (TLPs) [37], initially identified in tobacco leaves infected with TMV [38]. Several members have been reported to possess endo-β-1,3-glucanase activity and α-amylase inhibitory activity [39]. The PR-5 proteins are divided into two classes: one class (large TLPs) with a molecular weight of 21–26 kDa and with 16 conserved cysteine residues. The molecular weight of the other group (small TLPs) has been determined to be between 15 and 18 kDa, and small TLPs contain 10 cysteine residues [40]. The core conserved domain of PR-5 proteins is the thaumatin fold, a conserved domain that consists of two antiparallel β-sheets arranged in a “β-sandwich” conformation. This structure typically encompasses approximately ten β-strands interconnected by α-helices and loop regions [41]. PR-5 proteins are involved in various physiological processes, including plant development, secondary wall synthesis [42], and developmental regulation and ripening of fruits [43]. Furthermore, PR-5 proteins are involved in auxin signaling pathways [44]. PR-5 proteins possess antimicrobial enzyme activity [45] and are also regulated by stress factors such as high salinity [46] and viral or bacterial infections [47].

2.5. PR-6 Proteins

PR-6 is a class of protease inhibitors (PIs) with a molecular weight ranging from 8 to 22 kDa. These proteins contain a conserved cysteine residue in the N-terminal, which is believed to form a disulfide bond with another cysteine at the C-terminus, thereby maintaining protein structural stability [48]. Based on the specificity for pathogen-derived proteases, PR-6 proteins are classified into three major categories, including serine protease inhibitors, cysteine protease inhibitors, and aspartate proteases or metal carboxypeptidase inhibitors [49]. PR-6 belongs to the group of highly stable, naturally occurring plant defense proteins that show potent antimicrobial activity against a wide range of bacterial and fungal pathogens [50]. PR-6 binds to the target pathogen proteases to form stable complexes [51,52] that protect plant proteases from destruction. In addition to their roles in biotic stress responses, PR-6 genes are also induced by abiotic stresses, such as low temperature [53], and by external application of phytohormones, such as jasmonic acid (JA) [54].

2.6. PR-7 Proteins

PR-7 proteins have a molecular weight of approximately 69 kDa and were first discovered in tomato. PR-7 has endonuclease activity and exists in a monomeric form, with its activity dependent on Ca2+ activation [55]. The PR-7 proteins are secreted and enriched in the extracellular space of leaf and stem cells in citrus exocortis viroid-infected tomato plants [56]. Structurally, PR-7 contains the catalytic triad, which is homologous to the Bacillus subtilis protease family. The catalytic triad consists of three conserved amino acid residues: Asp, His, and Ser, and it is responsible for hydrolyzing peptide bonds. PR-7 is a central β-barrel structure composed of 7–8 β-sheets, surrounded by multiple α-helices to form a stable three-dimensional conformation. The catalytic triad is located in the cleft between the β-barrel and the α-helices [57]. Functionally, PR-7 proteins have been shown to enhance the antifungal activity of plants by solubilizing the cell wall of pathogens, thereby increasing disease resistance. The induction of PR-7 proteins has been observed in response to various biotic stresses, including fungi [58] and bacteria [59], as well as abiotic stresses such as calcium deficiency [60]. Hormones like ethylene, SA, and JA [61,62] have also been implicated in the regulation of PR-7 expression.

2.7. PR-9 Proteins

PR-9 proteins are a class of peroxidases with a molecular weight ranging from 32–45 kDa, consisting of 10–12 conserved α-helices, 2 β-sheets, and 4 pairs of disulfide bonds [63]. PR-9 binds to the plant cell wall in two forms: an ion-bound form or a covalently bound form. PR-9 functions as a catalyst for the oxidation of substrates such as phenols and their peroxide derivatives. PR-9 plays a variety of roles in the defense of the host plant against necrotrophic or biotrophic pathogens by reinforcing cell walls and producing reactive oxygen species (ROS). In addition, PR-9 proteins are involved in seed germination, cell growth, and cell wall relaxation. Beyond their defensive and developmental functions, PR-9 proteins have demonstrated antiviral activity [64], trap heavy metals [65], and are also highly induced by a variety of environmental stresses, such as high salt and leaf tissue damage [66].

2.8. PR-10 Proteins

PR-10 proteins are ribonuclease-like proteins with molecular weights ranging from 16 to 19 kDa. PR-10 proteins have a similar three-dimensional structure [67] consisting of a 25-amino-acid C-terminal α-helix, a seven-stranded antiparallel β-sheets, and two N-terminal short α-helices, with a short-looped connecting sequence. The main feature of PR-10 proteins is the presence of a large Y-shaped hydrophobic lumen, a structure necessary for its transportation into the plant cell wall [68]. Most PR-10 proteins are comprised of two domains, the Bet_v_1 and P-loop domains. The Bet_v_1 domain participates in defense against pathogen infections. The P-loop domain contains a phosphate-binding loop motif that functions as a nucleotide binding site [69]. PR-10 proteins possess both antimicrobial and antiviral activities [70,71], and have been shown to play a regulatory role in plant growth, as well as in the modulation of endogenous auxin levels [72]. PR-10 proteins have been found to bind to a variety of physiological ligands, including auxins, and are implicated in plant defense responses and developmental regulation [67].

2.9. PR-12 Proteins

PR-12 proteins, also known as defensins, are a class of small-molecule antimicrobial peptides involved in plant defense. PR-12 proteins have at least eight highly conserved cysteine residues, which form at least four intramolecular disulfide bonds. The three-dimensional structure of PR-12 proteins includes an α-helix and an antiparallel triple-stranded β-sheet, and the main structural feature is the cysteine-stabilized α-helix β-fold motif (CSαβ) [73]. Functionally, PR-12 has been shown to possess both antifungal and antiviral properties [74]. PR-12 proteins inhibit cell–virus fusion and target viral envelopes, ultimately resulting in the lysis of viral pathogens [64,75]. Beyond their role in biotic stress responses, PR-12 has been observed to be induced by a variety of abiotic stresses, including cold [76], drought [77], and heavy metals [78].

2.10. PR-13 Proteins

PR-13 proteins, also known as thionins, are small cysteine-rich proteins with a molecular weight of 5 kDa. According to the three-dimensional structure, thionins can also be divided into two categories, namely α/β-thionins and γ-thionins. α/β-Thionins are composed of two α-helices and a double-stranded β-sheet. γ-Thionins are composed of one α-helix and three antiparallel β-sheets, which collectively form a two-layered α/β sandwich structure [79]. PR-13 has been shown to possess antiviral and antifungal activity. Tomato and potato plants overexpressing PR-13 showed enhanced resistance to fungal pathogens [80]. In addition to their role in biotic stress responses, PR-13 is induced by abiotic stresses such as drought [81] and heavy metals [82].

2.11. PR-14 Proteins

PR-14, a class of polypeptides with a length of 90–95 amino acids, possesses the characteristics of nonspecific lipid transfer proteins (LTPs) and is thus classified as a category of LTPs. PR-14 proteins are characterized by their involvement in the process of intercellular lipid shuttle transport. The center of the three-dimensional structure of PR-14 is a bundle-like structure composed of four α- helices and a hydrophobic cavity that can accommodate multiple fluids, which facilitates the transport and loading of lipid molecules [83]. PR-14 proteins exhibit antimicrobial activity [84] and are located in the extracellular space. In addition to their defensive functions, PR-14 has been demonstrated to play a role in various plant developmental processes, including seed germination, and responses to heat and drought [85]. Furthermore, a variety of signaling molecules, including phytohormones such as ABA and SA, have been identified as regulators of PR-14 expression.

2.12. PR-15 and PR-16 Proteins

PR-15 and PR-16 are classified as germins (oxalate oxidase) and germin-like proteins (oxalate oxidase like), respectively, and they are members of the Calycin superfamily of proteins. PR-16 proteins contain a conserved β-barrel domain, and metal ions like Mn2+ and Ca2+ are necessary for their activity. Some PR-16 members possess superoxide dismutase (SOD) or oxalate oxidase (OXO) activity, thereby playing a role in reactive oxygen species (ROS) metabolism [86]. PR-16 proteins are resistant to proteases, distributed in the extracellular space, and generally exist in the form of heteropentamers [87]. Both PR-15 and PR-16 are glycoproteins in nature, and are produced immediately after pathogen infection [88]. PR-15 proteins play a pivotal role in defense responses against abiotic and biotic stresses by catalyzing the aerobic oxidation of oxalic acid and oxygen to produce CO2 and H2O2 [89]. Functionally, PR-15 and PR-16 play a part in responses to biotic stresses, such as fungi, viruses, pests, and others, as well as being involved in seed germination and floral induction [90]. PR-15 and PR-16 are responsive to abiotic stresses such as drought, low temperature, and high temperature [91,92], and the PR-16 family also plays a key role in the response of plants to saline and oxidative stresses.

2.13. PR-17 Proteins

PR-17 proteins are a novel class of PR proteins first identified in barley in 2002 and they play important roles in plant defense against a variety of pathogens. PR-17 proteins contain active sites and peptide-binding grooves that are highly similar to those of aminopeptidase N and Thermolysin [93]. In the amino acid sequences of PR-17 family members, there are five highly conserved domains in the central and C-terminal parts, namely Box A to Box E. The PR-17 proteins exhibit zinc affinity and structural similarity to zinc metalloproteinases [93]. The PR-17 genes were found to be increased following treatments involving tobacco mosaic virus infection, mechanical damage, and drought conditions [94]. The overexpression of the PR-17 gene resulted in increased tolerance to molds in wheat [95]. Finally, Table 1 summarizes the properties, molecular functions, and responses to biotic stress and abiotic stress of different PR proteins.

3. Response of Plant PR Proteins to Abiotic Stress

Abiotic stress refers to adverse environmental physical or chemical factors that negatively impact plant growth, development, and survival [176]. Abiotic stress triggers significant alterations in gene expression and protein abundance within plants. Currently reported abiotic stresses that modulate the expression of plant PR genes and proteins include drought, high salt, low/high temperature, heavy metal, UV radiation, and waterlogging (Figure 1). The response of plant PR proteins to abiotic stress primarily occurs at the transcriptional level, the protein abundance level, the protein phosphorylation level, and the enzymatic activity level. Nevertheless, while transcriptional- and abundance-level changes are well-documented, there are comparatively fewer reports detailing the regulation of PR proteins at the phosphorylation and enzymatic activity levels.

3.1. Drought Stress or Osmotic Stress

Under drought or osmotic stress conditions, the induction of PR families, including PR-1, PR-2, and PR-3, has been widely reported. In wheat, osmotic stress induced the expression of TaPR-1 [177]. Similarly, PR-2 expression increased in sesame under drought stress [18]. Proteomic analysis has revealed that, in wheat, sunflower, and grapevine, PR-2, PR-5, and PR-10 proteins were significantly upregulated in response to drought stress [178]. Drought induced the expression of members from multiple PR families, including PR-3, PR-11, PR-12, PR-15, and PR-16 in citrus roots [59]. Drought significantly induced the expression of PR-5 in tobacco [132]. Compared with growth under normal conditions, the transcript abundance of PR-8 increased during drought stress in Lotus spp. [33]. Both the transcriptional level and protein level of PR-10 genes were upregulated in rice under drought stress [179], accompanied by significant changes in phosphorylation status [180]. In addition, drought induced PR-12 gene expression in pepper [77]. PR-13 expression increased in barley plants under drought conditions [81]. PR-15 proteins levels increased in wheat flag leaves under drought stress [169]. PR-16 transcript levels increased in soybean under drought stress [92]. The tobacco NtPRp27, belonging to the PR-17 family, exhibited elevated transcript levels after drought treatment [175]. Among various PR families responsive to drought stress, the PR-1 family is the most extensively studied across different plant species.

3.2. High Salinity Stress

Under conditions of high salinity stress, the induction of PR families including PR-1, PR-2, and PR-3 has been widely reported. For instance, PR-1 gene expression increased in rice under high salinity stress [9], while its protein level decreased in wheat after salt stress [178]. Similarly, PR-2 expression increased in sugarcane under high salinity [21]. PR-3 was significantly induced by high salt stress in rice [120]. PR-5 gene expression was upregulated in Arabidopsis thaliana under salt stress [181], and its protein abundance increased in soybean and wheat under salt stress [178]. PR-9 expression was induced by salt stress in sweet potato [66]. In rice, PR-10 expression was upregulated under salt-stress conditions [179]. PR-10 and PR-14 were accumulated in barley roots under salt stress [178]. PR-12 and PR-13 proteins levels increased in mustard seed coats subjected to salt stress treatment [162]. PR-16 transcript levels also rose in soybean leaves after high-salinity stress [92]. Among the various PR families responsive to salt stress, the PR-10 family is the most extensively studied in plants.

3.3. Low Temperature Stress

Under cold stress or cold acclimation conditions, the induction of PR families, including PR-1, PR-2, and PR-3, has been reported. PR-1 gene expression increased in tomato under low-temperature stress [102]. Similarly, PR-2 gene expression levels increased in grapevine subjected to low-temperature treatment [113]. The chitinase, PR-2, and PR-5 protein levels rose in overwintering grasses after low-temperature treatment [113]. The abundance of several PR proteins, including PR-3 and PR-5, increased in Ammopiptanthus mongolicus under low-temperature treatment [182]. PR-4 transcript levels increased in ginseng in response to low-temperature treatment [125]. PR-10, PR-12, and PR-13 expression levels increased in wheat under low-temperature treatment [165,183]. Proteomic analysis revealed that both PR-5 and PR-10 are accumulated in winter wheat and barley after low-temperature treatment [178]. PR-12 is similarly induced in winter wheat under low-temperature treatment [76]. PR-15 gene expression increases in barley subjected to low-temperature stress [91]. Among the various PR families responsive to low-temperature stress, the PR-1 family is the most extensively studied in plants, followed by the PR-10 family.

3.4. High-Temperature Stress

Under high-temperature stress, the induction of several PR families, such as PR-1 and chitinases, has been reported. PR-1 gene expression levels significantly increased in wheat plants after high-temperature treatment [103]. Similarly, chitinase expression increased in wheat under high-temperature conditions [36]. PR-5 expression also increased in wheat [184] and grapevine [185] in response to heat stress. In addition, PR-12 was induced in A. thaliana under high-temperature treatment [186].

3.5. Other Abiotic Stresses

3.5.1. Heavy Metal Stress

Nine PR families, including PR-1, PR-2, and PR-4, have been reported to be induced by heavy metal stress. Proteomic analysis revealed that PR-1, PR-2, chitinase, PR-5, and PR-17 levels in barley leaf increased in response to Cd exposure [187]. Copper stress induced the expression of PR-1 and PR-2 genes [121,188]. Similarly, PR-4 expression increased in maize after AgNO3 treatment [24]. In durum wheat, PR-5 was strongly induced by several heavy metals, such as Cd, Cu, and Zn [136]. Four defensin genes were induced by Zn exposure, but not by Cd exposure in A. thaliana [78]. The expression of PR-9 increased under heavy metal stress in tomato [189], and its enzymatic activity increased in water lily under similar conditions [65]. PR-10 proteins were accumulated in yellow lupin root tips after lead ion treatment, and PR-10 gene transcriptional activity increased under Pb, Cd, and As treatments [153]. Similarly, PR-10 proteins were upregulated in Cd-treated flax cells [178]. Copper stress induced increased PR-10 proteins expression in both the roots and leaves of copper/zinc-tolerant birch [190]. PR-13 gene expression increased in rice under Cd exposure [82]. In addition, Cd, Co, and Cu significantly induced PR-15 content in wheat plants [191]. Among the various PR families responsive to heavy metal stress, the PR-10 family is the most extensively studied in plants.

3.5.2. UV Radiation Stress

Under UV-B stress, the induction of families such as PR-1, PR-2, and PR-3 has been observed, involving responses at the gene transcription and protein accumulation levels. In tobacco, UV-B induced PR-1 with specific accumulation in irradiated leaf areas [105]. UV-B increased the transcript levels of PR-1, PR-2, and PR-5 in A. thaliana [115]. Similarly, PR-2 was significantly induced in grapevine in response to UV radiation [113]. The transcript levels of PR-3 and PR-5 in UV-treated strawberry leaves were significantly higher than those in non-irradiated controls [192]. PR-10 proteins increased in the leaves of white lupin after UV treatment [154]. Among the various PR families responsive to heavy metal stress, the PR-10 family is the most frequently analyzed in plants.

3.5.3. Waterlogging Stress

Under waterlogging stress, the induction of several families such as PR-1, PR-2, and PR-3 has been reported at the gene transcription or protein accumulation levels. The expression level of PR-1 significantly increased in Vietnamese snowbell under waterlogging stress [193]. Similarly, PR-2 and PR-9 expressions were upregulated in soybean and maize under waterlogging stress [178]. Transcriptomics analysis revealed that transcript levels of PR-3 increased in the leaves and roots of waterlogged wheat, specifically involving class IV and class VIII chitinases [194]. In addition, PR-12 gene expression increased in onion leaves when plants were subjected to waterlogging stress [195].
Overall, among the various PR families, PR-1 is the most frequently reported PR family involved in abiotic stress responses. Studies on changes in the gene expression levels of various PR families under stress are more numerous than studies at the protein level. Among the different abiotic stressors influencing PR levels, drought stress has been the most extensively studied.

4. Regulation of PR Proteins Involved in Abiotic Stress Responses

PR proteins, as key components of the plant defense system, play crucial roles in mediating responses to both biotic and abiotic stresses. While the above sections describe significant alterations in PR gene expression and protein abundance under abiotic stress conditions, the signaling transduction pathways by which abiotic stress signals regulate PR genes or proteins remain poorly understood. Notably, abiotic stress signaling pathways may utilize components of biotic stress signaling pathways; therefore, this section first briefly introduces the signaling pathways regulating PR genes under biotic stress (Figure 2).
Extensive research has revealed that pathogen signals activate PR gene expression primarily through hormone signaling pathways such as SA and JA. However, PR proteins are not regulated by a single hormone alone, but function through complex crosstalk involving multiple stress hormones and their signaling components—such as SA (via NPR1), JA (via MYC2), and ethylene (via ethylene response factor)—alongside synergistic or antagonistic interactions with growth hormones like gibberellin and cytokinin. This dynamic network balances defense and growth, reflecting the multidimensional complexity of the plant stress responses rather than a simple summation of individual pathways [159]. In this process, transcription factors, including the bZIP family factor TGA, the NAC family factor ATAF2, and the AP2/ERF family factor EREBP1, play important roles in the transcriptional activation of PR genes. The SA signaling pathway relies on the coactivator NPR1 interacting with the bZIP transcription factor TGA. This interaction promotes TGA binding to the LS7 cis-element in the PR-1 promoter, thereby activating PR-1 transcription and mediating system-acquired resistance (SAR) [196]. The expression of the Arabidopsis NAC transcription factor ATAF2 is induced by mechanical wounding, SA, and MeJA. ATAF2 binds to the promoters of PR-1, PR-4, PR-12, and other PR genes, repressing their transcription [197]. In rice, the MAPK cascade kinase BWMK1 integrates signals from fungal elicitors, H2O2, SA, and JA. BWMK1 translocates to the nucleus via its C-terminal domain, phosphorylates the AP2/ERF transcription factor OsEREBP1, thereby enhancing its binding affinity to the GCC-box cis-element (AGCCGCC) in PR genes promoters, and coordinately activates the expression of PR-1, PR-2, PR-5, and other PR genes, conferring broad-spectrum resistance to fungi and bacterial pathogens [198].
Under abiotic stress conditions, the transcriptional regulatory mechanisms governing plant PR proteins are not fully elucidated. Promoter analysis indicates that the promoter regions of various plant PR genes contain multiple cis-acting elements associated with both biotic and abiotic stress responses, harboring binding sites for transcription factors such as WRKY, HD-Zip, AP2/ERF, and MYB. In soybean, predicted transcription factor family binding sites in the PR-1 gene promoter include members of the WRKY, GATA, HD-Zip, ERF, and C2H2 families [101]. Similarly, the rice PR-1 gene promoter is predicted to harbor binding sites for WRKY, bZIP, AP2/ERF, bHLH, MYB, TCP, SBP, and B3 transcription factors [199].
Experimental evidence supports the involvement of several transcription factor families, including C2H2, NAC, WRKY, and MYB, in mediating the regulation of PR genes expression by abiotic stress signals such as drought, low temperatures, and heavy metal exposure. Under drought conditions, the Arabidopsis transcription factor Di19 (Drought-induced), a C2H2 zinc finger protein, activates the expression of PR-1, PR-2, and PR-5 genes by binding to the conserved sequence TACA (A/G)T within their promoters [10]. The Arabidopsis Di19 gene family encodes seven hydrophilic proteins, with Di19 being one of them, containing two atypical C2H2-type zinc finger domains and a nuclear localization signal (NLS). A drought-induced cotton R2R3-MYB transcription factor, GhMYB36, binds to the PR-1 promoter and upregulates PR-1 expression [200].
Under cold stress conditions, the Arabidopsis NAC transcription factor NTL6 directly binds to and activates the expression of PR-1, PR-2, and PR-5 genes [11]. NTL6 belongs to membrane-bound transcription factors containing an N-terminal NAC domain (DNA-binding domain) and a C-terminal transmembrane motif. A cold-stress-responsive Arabidopsis ERF/AP2 transcription factor, DREB1, induces PR-2 gene transcription by binding to the DRE/CRT element (GCCGAC) within its promoter [201]. Under low-temperature and osmotic stress, A. mongolicus transcription factor AmWRKY14 directly binds to the W-box cis-element of the PR-12 gene promoter and activates its expression [202]. Under cold, drought, and wounding stress, the jojoba transcription factor ScWRKY41 promotes PR-5 expression by binding to W-box elements in its promoter region [203]. Similarly, under low-temperature stress, ScMYC2, a bHLH family member in jojoba, regulates PR-3 expression by binding to the E-box cis-element in its promoter [204].
Under heavy metal stress, the rice transcription factor OsNAC300, a NAC transcription factor, directly binds to the CATGTG motif in the promoter region of PR-10 genes and activates their transcription [205].
Compared with well-characterized regulatory pathways involved in PR gene activation under biotic stress, the regulatory mechanisms underlying PR gene regulation under abiotic stress remain less defined. However, some experimental evidence exists. For instance, within the bHLH transcription factor family, it was found that low temperature increases the JA content in leaves. JA acts as a signaling molecule, initiating a signal transduction pathway where the transcription factor MYC2 binds to specific cis-elements in the promoter regions of target genes to regulate their expression [204].

5. Biological Functions of PR Proteins in Abiotic Stress Responses

Previous sections summarized the induction of PR genes expression under abiotic stress conditions and the transcriptional regulatory mechanism involved. How do PR proteins function under abiotic stress conditions? Given that different PR families possess distinct conserved domains, PR proteins within a single PR family typically exhibit similar molecular activities. Below, we detail the specific biological functions of PR-1 to PR-17 in the context of abiotic stress responses.
The overexpression of PR-1 enhanced tolerance to drought and salt stress in transgenic Arabidopsis [9] and cold tolerance in transgenic chrysanthemum [206]. Several mechanisms have been proposed through which PR-1 may help plants to tolerate abiotic stress: (1) PR-1 helps to maintain plasma membrane stability. The CAP superfamily domain in PR-1 proteins may regulate membrane stability under stress conditions by binding sterols. (2) PR-1 forms chimeric receptor proteins: The fusion of PR-1 with receptor-like protein kinases (RLKs) potentially generates novel proteins that perceive extracellular signals, undergo conformational changes, and activate intracellular phosphorylation cascades in response to extracellular cues [207]. Additionally, PR-1 may indirectly confer tolerance by modulating antioxidant enzyme activity [206] and regulating stomatal aperture [10].
Transgenic Arabidopsis plants overexpressing PR-2 (β-1,3-glucanase) showed increased aluminum tolerance [208], while other studies suggest that PR-2 accumulation may contribute to enhancing low-temperature tolerance in citrus [209]. PR-2 may play a part in abiotic stress responses in the following ways: (1) PR-2 helps to protect thylakoid membranes. Tobacco PR-2 protects spinach thylakoids from freeze-thaw induced damage [20]. (2) PR-2 proteins function as ice-binding proteins (IBPs). Winter rye PR-2 specifically binds ice crystals, inhibiting their growth and recrystallization, thereby conferring freezing tolerance [19]. (3) PR-2 proteins hydrolyze carbohydrates. Winter rye PR-2 hydrolyzes β-glucan polymers, generating small soluble carbohydrates that increase cellular solute concentration and depress freezing points. (4) PR-2 proteins form synergistic complexes. PR-2 complexes with other antifreeze proteins (e.g., chitinases, PR-5) exhibit significantly higher activity in modifying carbohydrates and controlling ice crystal morphology than individual PR-2 [19].
PR-3, PR-4, PR-8, and PR-11 belong to the chitinase family. Chitinase overexpression enhanced tolerance to low temperature and osmotic stress in transgenic Arabidopsis and jojoba [204], as well as to salt and heavy metal stress in tobacco [188]. These proteins may contribute to abiotic stress tolerance through two primary mechanisms: (1) Chitinases indirectly promote plant tolerance via enhancing antioxidant enzyme activity. They reduce malondialdehyde (MDA) accumulation and effectively scavenge reactive oxygen species (ROS), thereby mitigating oxidative damage induced by stress like drought. (2) Chitinases participate in stress responses by promoting chitosan oligosaccharide (COS) production. Chitinases in the GH18 subfamily, dependent on the DxDxE motif, are proposed to generate COS, which participates in antioxidation and osmoregulation. Notably, higher COS polymerization degrees are correlated with stronger antioxidant activity in barley [210].
The overexpression of PR-5 (thaumatin-like proteins, TLPs) enhanced drought and salt tolerance in transgenic tobacco, rice, and Arabidopsis [132,181,211]. The first direct way in which PR-5 helps plants to tolerate stress is Na+ sequestration. Under high salinity, tobacco PR-5 binds to Na+ channels on the plasma membrane and coordinates Na+ ions, sequestering and compartmentalizing them into the apoplast and vacuoles to regulate osmotic potential [212]. The second direct way in which PR-5 helps plants to tolerate stress is by modulating cytoskeleton dynamics. Potato PR-5 stabilizes actin filaments (AFs) by direct binding and blocks Ca2+ channel activity, preventing cytosolic Ca2+ elevation and enhancing cold tolerance [213]. The third direct way in which PR-5 helps plants to tolerate stress is by protecting the plant cell membrane. Tobacco PR-5 interacts with membrane lipids and competitively binds lipid peroxidation products, thereby reducing oxidative damage under drought and salt stress [214]. Rice PR-5 acts as a molecular chaperone, interacting with membrane proteins to prevent their denaturation and membrane disintegration under drought conditions [215]. The fourth way in which PR-5 helps plants to tolerate stress is the fusion with kinases to form transmembrane receptors. The Arabidopsis chimeric receptor PR5-RLK contains an extracellular PR-5-like domain, transmembrane domain, and intracellular serine/threonine kinase domain, which are induced by drought conditions. Under stress conditions, PR5-RLK phosphorylates downstream transcription factors and ion channel proteins, inducing drought-responsive gene expression and stomatal closure [216,217]. In addition, PR-5 can indirectly participate in plant stress tolerance in three different ways, including increasing the thickness of secondary cell walls [120], promoting proline synthesis via the ABA signaling pathway [218,219], and enhancing enzyme activity such as antioxidant enzymes [211,215].
Studies suggest that several PR-9 proteins (peroxidases) may help to enhance plant heavy metal tolerance. Water lily utilizes PR-9 to produce phenolic polymers that capture cadmium (Cd), sequestering it as Ca–Cd crystals in specific glands on the leaf water surface [65]. Similarly, the overexpression of PR-13 (thionin) enhanced Cd tolerance in transgenic rice [82]. Under Cd stress, rice PR-13 localizes to the cell wall, significantly increasing Cd binding to the wall, thereby reducing Cd upward translocation and accumulation in shoots and straw [82].
Overexpression of PR-10 (ribonuclease-like proteins) genes has been shown to enhance salt and drought tolerance in rice and the halophyte Halostachys capsica [152,179]. PR-10 indirectly improves stress tolerance by increasing the activity of the plant antioxidant enzyme system and regulating osmoprotectant synthesis to effectively scavenge ROS, thereby protecting cells from oxidative damage [179].
Transgenic Arabidopsis overexpressing PR-14 (LTP) exhibited enhanced tolerance to salt and drought stress [220]. PR-14 interacts with cellular membranes, regulating membrane lipid fluidity and permeability. Under drought, salt, or low temperature, the hydrophobic cavity of PR-14 binds membrane lipids, stabilizing membrane structure and preventing lipid peroxidation. In addition, PR-14 also modulates membrane composition, participates in cuticle and suberin biosynthesis, and facilitates the transport of hydrophobic lipid precursors to the cell wall, thereby strengthening the epidermal barrier function [166].
The accumulation of PR-15 (oxalate oxidase) may be beneficial for plant tolerance to waterlogging, high-temperature, and high-salinity conditions. PR-15 functions through both OXO and SOD activities, and it catalyzes oxalate degradation to produce H2O2, which may participate in cell wall lignification, enhancing wall stability and helping plants to withstand high-temperature damage to cellular structures [221]. Moreover, H2O2 acts as a signaling molecule, activating defense mechanisms enabling plant adaptation to waterlogging stress [170]. Furthermore, the Mn2+/Mn3+-dependent SOD activity of moss PR-15 scavenges excess intracellular superoxide radicals, mitigating oxidative damage and protecting cells from high salinity [222].
Transgenic Arabidopsis overexpressing PR-16 (oxalate oxidase-like proteins) showed enhanced tolerance to drought and salt stress [92]. PR-16 in moss and soybean exhibits SOD activity [222], thereby protecting chloroplasts by scavenging ROS and maintaining ion homeostasis. Additionally, PR-16 enhances stress tolerance by regulating the accumulation of osmoprotectants and inducing stomatal closure to reduce water loss under drought and salt stress [92].
In summary, PR families contribute to abiotic stress tolerance through diverse mechanisms (Figure 3). The first one is directly enhancing membrane stability via binding lipids/membrane proteins (PR-1, PR-5, and PR-14). The second is directly binding metal ions to participate in heavy metal tolerance (PR-9 and PR-13). The third one is indirectly enhancing antioxidant capacity, and this mechanism involves PR-1, Chitinases (PR-3, 4, 8, 11), PR-5, PR-10, PR-15, and PR-16. The fourth one is indirectly reducing stomatal aperture (PR-1, PR-2, PR-5, and PR-16). The fifth way is indirectly regulating osmotic potential via soluble compounds accumulation, and this mechanism involves PR-2, Chitinases (PR-3, 4, 8, 11), PR-5, PR-10, and PR-16. The sixth way is indirectly increasing cell wall thickness, and this mechanism involves PR-5, PR-14, and PR-15.
Some PR families rely on their enzymatic activity to help plants to tolerate abiotic stress, including PR-2 (β-1,3-glucanase), PR-3, 4, 8, and 11 (chitinases), PR-9 (peroxidase), PR-10 (ribonuclease), PR-15 (oxalate oxidase), and PR-16 (oxalate oxidase-like protein) (Table 2). These PR families directly or indirectly participate in stress responses through catalytic reactions. While other PR families use their specific three-dimensional structures to help plants to tolerate abiotic stress, including PR-1, PR-5, PR-13, and PR-14, these PR families rely on the three-dimensional structure of proteins, such as binding sites and membrane interaction regions, to exert membrane stability, ion sequestration, and other functions.

6. Conclusions

6.1. Critical Knowledge Gaps

Research on PR proteins in the context of abiotic stress responses holds significant scientific importance and application value. A major critical knowledge gap lies in the lack of a clear distinction between the direct and indirect roles of PR proteins in abiotic stress responses, which has hindered their systematic study in this context. Evolutionarily conserved as core components of biotic stress defense, PR proteins are often induced under abiotic stress through secondary mechanisms—such as crosstalk with general stress signaling networks (e.g., SA-ABA interplay, ROS-mediated pathways) or overlap with shared transcriptional regulators (e.g., WRKY, ERF/AP2 factors)—rather than via dedicated abiotic stress-specific pathways. This indirect involvement, coupled with historical research focus on their biotic defense functions, has limited in-depth exploration of their abiotic roles. Further, the mechanistic basis for why PR proteins are induced under abiotic stresses remains poorly synthesized. While transgenic overexpression of PR genes can enhance abiotic tolerance, this is often attributed to pleiotropic effects, rather than direct functional roles. Additionally, the ecological relevance of such induction—for instance, whether it primarily serves to preempt pathogen colonization under abiotic stress-weakened states, or reflects genuine adaptation to abiotic stress—lacks systematic analysis. Finally, current studies largely report PR expression patterns under abiotic stresses without integrating these observations into a unifying framework. This gap limits our understanding of whether PR proteins act as “stress integrators” (linking biotic and abiotic responses) or merely as byproducts of overlapping stress pathways. Although the mechanisms by which specific PR protein families confer abiotic stress tolerance to plants are partially understood, critical knowledge gaps persist regarding the molecular mechanisms underlying the involvement of most PR families. For instance, it has been clarified that PR-16, which has SOD activity, can scavenge ROS in plants and enhance their antioxidant capacity. In contrast, the mechanisms by which PR-1, PR-3, PR-4, and other families enhance plant antioxidant enzyme activity remain largely unknown. Similarly, the molecular basis of stomatal aperture regulation by PR-1, PR-2, and PR-5 has yet to be elucidated.
Although upstream transcription factors and signaling pathways regulating specific PR genes are known for some individual PR proteins, the precise molecular pathways governing the regulation of most PR genes by abiotic stress signals remain largely obscure. In particular, the coupling mechanisms between upstream transcription factors (e.g., WRKY and NAC) and abiotic stress signals are not fully revealed, particularly concerning the synergistic regulatory networks involving hormone pathways (SA, JA, and ABA) and PR genes. Moreover, current research on the impact of post-translational modifications (PTMs) (e.g., phosphorylation and glycosylation) on PR proteins activity and stability is rather limited. For example, the structural basis by which phosphorylation enhances the ribonuclease activity of PR-10 remains unelucidated.
Systematic studies on the functional specificity and cross-regulatory mechanisms of PR proteins under combined abiotic stresses (e.g., drought + heat, salinity + heavy metals) are still lacking, and how they balance stress tolerance with growth and development requires deeper investigation. The mechanisms underlying functional redundancy and functional divergence within PR families are unclear. For instance, the distinct roles and potential synergistic interactions of different classes of chitinases belonging to PR-3, PR-4, PR-8, and PR-11 families in abiotic stress responses need further delineation.
The evolutionary trajectories and functional innovations of PR families in extremophile plants have not been systematically investigated, and the mining of stress-resistance gene resources remains largely confined to model plants. In addition, functional studies on PR-like proteins in lower plants (e.g., bryophytes and algae) are virtually non-existent, leaving the evolutionary process of their primordial defense mechanisms adapting to abiotic stresses largely unexplored.

6.2. Prospects for Study on PR Proteins

Bioinformatics approaches should be employed to conduct systematic evolutionary analyses of individual PR families across the plant kingdom. This will identify the full repertoire of PR families from lower to higher plants, clarify their evolutionary paths, define the structural changes occurring at specific evolutionary junctures, and elucidate the acquisition of novel functions. Special focus should be placed on understanding how PR proteins transitioned from classical defense roles against biotic stresses to functions in abiotic stress tolerance. In addition, the exploitation of PR gene resources in extremophile plants should be strengthened. PR homologous genes in desert plants and halophytes should be systematically screened, and the association between their structural features and abiotic stress tolerance should be analyzed.
Multi-dimensional technologies should be combined with spatial omics to reveal dynamic expression patterns of PR proteins in specific tissues upon abiotic stress treatment. Structural biology approaches and AI prediction should be leveraged, with cryo-electron microscopy and machine learning employed to resolve high-resolution three-dimensional structures of PR proteins and design functionally enhanced mutants. Post-translational modification landscapes of PR proteins should be deciphered using phosphoproteomics and ubiquitinomics analyses to elucidate regulatory effects on PR protein stability and interaction networks under stress conditions.
The increasing frequency of extreme weather events due to global climate change should be addressed through transgenic approaches utilizing PR proteins to develop crops tolerant to both biotic and abiotic stresses. Gene editing and synthetic biology should be prioritized to investigate multi-gene synergistic regulation. In particular, CRISPR-Cas9 technology should be employed for the targeted editing of multiple PR family genes to overcome functional redundancy while enhancing broad-spectrum stress resistance. The design and application of stress-specific promoters should be developed to drive high-efficiency expression of PR genes under defined stress conditions. Concurrently, crop quality traits should be optimized through PR gene editing to achieve synergistic improvement of stress tolerance and agronomic yield.
The yield performance and stress tolerance stability of PR transgenic crops should be evaluated under field conditions in realistic agricultural settings (e.g., drought and saline–alkaline soils), and phenotype–genotype association databases should be established. Furthermore, the impact of PR overexpression on soil microbiomes and ecosystem health should be assessed to ensure the environmental safety of transgenic crops. Environmental remediation applications should be explored by exploiting the pollutant-degrading capacity of PR-9 peroxidases, with phytoremediation technologies developed for soils contaminated with heavy metals and organic pollutants. Innovative nanocarrier-mediated delivery systems should be investigated through the design of PR protein nanoparticles for foliar application, enabling rapid activation of plant stress responses while reducing reliance on genetic modification. Finally, artificial intelligence should be utilized to construct predictive models of PR protein functioning, thereby accelerating intelligent molecular breeding for stress-tolerant crops.

Author Contributions

Conceptualization, F.G.; formal analysis, Y.Z.; writing—original draft preparation, Y.Z. and F.G.; writing—review and editing, Y.Z. and F.G.; funding acquisition, F.G. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Natural Science Foundation of China (Grant number 31670335).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

We would like to express our gratitude to Pedro Garcia Caparros for his assistance in improving the language quality of this review. During the preparation of this manuscript, the authors used DeepSeek-R1 (via the official DeepSeek web interface) to improve language quality, including grammar, spelling, punctuation, and fluency. The authors have reviewed and edited the output and take full responsibility for the content of this publication.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sels, J.; Mathys, J.; De Coninck, B.M.A.; Cammue, B.P.A.; De Bolle, M.F.C. Plant pathogenesis-related (PR) proteins: A focus on PR peptides. Plant Physiol. Biochem. 2008, 46, 941–950. [Google Scholar] [CrossRef]
  2. Van Loon, L.C.; Van Kammen, A. Polyacrylamide disc electrophoresis of the soluble leaf proteins from Nicotiana tabacum var. “Samsun” and “Samsun NN”. II. Changes in protein constitution after infection with tobacco mosaic virus. Virology 1970, 40, 190–211. [Google Scholar] [CrossRef] [PubMed]
  3. Li, X.-Y.; Gao, L.; Zhang, W.-H.; Liu, J.-K.; Zhang, Y.-J.; Wang, H.-Y.; Liu, D.-Q. Characteristic expression of wheat PR5 gene in response to infection by the leaf rust pathogen, Puccinia triticina. J. Plant Interact. 2015, 10, 132–141. [Google Scholar] [CrossRef]
  4. Van Loon, L.C.; Pierpoint, W.S.; Boller, T.; Conejero, V. Recommendations for naming plant pathogenesis-related proteins. Plant Mol. Biol. Rep. 1994, 12, 245–264. [Google Scholar] [CrossRef]
  5. Gordon-Weeks, R.; Sugars, J.; Antoniw, J.; White, R. Accumulation of a novel PR1 protein in Nicotiana langsdorfii leaves in response to virus infection or treatment with salicylic acid. Physiol. Mol. Plant Pathol. 1997, 50, 263–273. [Google Scholar] [CrossRef]
  6. Niki, T.; Mitsuhara, I.; Seo, S.; Ohtsubo, N.; Ohashi, Y. Antagonistic Effect of Salicylic Acid and Jasmonic Acid on the Expression of Pathogenesis-Related (PR) Protein Genes in Wounded Mature Tobacco Leaves. Plant Cell Physiol. 1998, 39, 500–507. [Google Scholar] [CrossRef]
  7. Fernández, C.; Szyperski, T.; Bruyère, T.; Ramage, P.; Mösinger, E.; Wüthrich, K. NMR solution structure of the pathogenesis-related protein P14a. J. Mol. Biol. 1997, 266, 576–593. [Google Scholar] [CrossRef]
  8. Sun, T.; Yan, N.; Liu, Q.; Bai, T.; Gao, H.; Chen, J. Re-Examination Characterization and Screening of Stripe Rust Resistance Gene of Wheat TaPR1 Gene Family Based on the Transcriptome in Xinchun 32. Int. J. Mol. Sci. 2025, 26, 640. [Google Scholar] [CrossRef]
  9. Kothari, K.S.; Dansana, P.K.; Giri, J.; Tyagi, A.K. Rice stress associated protein 1 (OsSAP1) interacts with aminotransferase (OsAMTR1) and pathogenesis-related 1a protein (OsSCP) and regulates abiotic stress responses. Front. Plant Sci. 2016, 7, 1057. [Google Scholar] [CrossRef]
  10. Liu, W.-X.; Zhang, F.-C.; Zhang, W.-Z.; Song, L.-F.; Wu, W.-H.; Chen, Y.-F. Arabidopsis Di19 functions as a transcription factor and modulates PR1, PR2, and PR5 expression in response to drought stress. Mol. Plant 2013, 6, 1487–1502. [Google Scholar] [CrossRef]
  11. Seo, P.J.; Kim, M.J.; Park, J.Y.; Kim, S.Y.; Jeon, J.; Lee, Y.H.; Kim, J.; Park, C.M. Cold activation of a plasma membrane-tethered NAC transcription factor induces a pathogen resistance response in Arabidopsis. Plant J. 2010, 61, 661–671. [Google Scholar] [CrossRef] [PubMed]
  12. Akbudak, M.A.; Yildiz, S.; Filiz, E. Pathogenesis related protein-1 (PR-1) genes in tomato (Solanum lycopersicum L.): Bioinformatics analyses and expression profiles in response to drought stress. Genomics 2020, 112, 4089–4099. [Google Scholar] [CrossRef]
  13. Roy Choudhury, S.; Roy, S.; Singh, S.K.; Sengupta, D.N. Molecular characterization and differential expression of β-1, 3-glucanase during ripening in banana fruit in response to ethylene, auxin, ABA, wounding, cold and light–dark cycles. Plant Cell Rep. 2010, 29, 813–828. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, Y.; Liu, L.; Yang, S.; Liu, G.; Zeng, Q.; Liu, Y. Molecular characterization and functional analysis of a pathogenesis-related β-1, 3-glucanase gene in spruce (Picea asperata). Eur. J. Plant Pathol. 2022, 164, 177–192. [Google Scholar] [CrossRef]
  15. Yoshikawa, M.; Keen, N.T.; Wang, M.-C. A Receptor on Soybean Membranes for a Fungal Elicitor of Phytoalexin Accumulation 1. Plant Physiol. 1983, 73, 497–506. [Google Scholar] [CrossRef] [PubMed]
  16. Levy, A.; Erlanger, M.; Rosenthal, M.; Epel, B.L. A plasmodesmata-associated β-1,3-glucanase in Arabidopsis. Plant J. 2007, 49, 669–682. [Google Scholar] [CrossRef] [PubMed]
  17. Wu, S.-W.; Kumar, R.; Iswanto, A.B.B.; Kim, J.Y. Callose balancing at plasmodesmata. J. Exp. Bot. 2018, 69, 5325–5339. [Google Scholar] [CrossRef] [PubMed]
  18. Parveen, S.; Mazumder, M.; Bhattacharya, A.; Mukhopadhyay, S.; Saha, U.; Mukherjee, A.; Mondal, B.; Debnath, A.J.; Das, S.; Sikdar, S. Identification of anther-specific genes from sesame and functional assessment of the upstream region of a tapetum-specific β-1, 3-glucanase gene. Plant Mol. Biol. Rep. 2018, 36, 149–161. [Google Scholar] [CrossRef]
  19. Yaish, M.W.; Doxey, A.C.; McConkey, B.J.; Moffatt, B.A.; Griffith, M. Cold-active winter rye glucanases with ice-binding capacity. Plant Physiol. 2006, 141, 1459–1472. [Google Scholar] [CrossRef] [PubMed]
  20. Hincha, D.K.; Meins Jr, F.; Schmitt, J.M. β-1, 3-glucanase is cryoprotective in vitro and is accumulated in leaves during cold acclimation. Plant Physiol. 1997, 114, 1077–1083. [Google Scholar] [CrossRef]
  21. Su, Y.-C.; Wang, Z.-Q.; Liu, F.; Li, Z.; Peng, Q.; Guo, J.-L.; Xu, L.-P.; Que, Y.-X. Isolation and Characterization of ScGluD 2, a New Sugarcane β-1, 3-Glucanase D Family Gene Induced by Sporisorium scitamineum, ABA, H2O2, NaCl, and CdCl2 Stresses. Front. Plant Sci. 2016, 7, 1348. [Google Scholar] [CrossRef]
  22. Kumar, M.; Brar, A.; Yadav, M.; Chawade, A.; Vivekanand, V.; Pareek, N. Chitinases—Potential candidates for enhanced plant resistance towards fungal pathogens. Agriculture 2018, 8, 88. [Google Scholar] [CrossRef]
  23. Hong, J.K.; Hwang, B.K. Induction by pathogen, salt and drought of a basic class II chitinase mRNA and its in situ localization in pepper (Capsicum annuum). Physiol. Plant. 2002, 114, 549–558. [Google Scholar] [CrossRef] [PubMed]
  24. Bravo, J.M.; Campo, S.; Murillo, I.; Coca, M.; San Segundo, B. Fungus-and wound-induced accumulation of mRNA containing a class II chitinase of the pathogenesis-related protein 4 (PR-4) family of maize. Plant Mol. Biol. 2003, 52, 745–759. [Google Scholar] [CrossRef] [PubMed]
  25. Beintema, J.J. Structural features of plant chitinases and chitin-binding proteins. FEBS Lett. 1994, 350, 159–163. [Google Scholar] [CrossRef]
  26. Ohnuma, T.; Numata, T.; Osawa, T.; Mizuhara, M.; Lampela, O.; Juffer, A.H.; Skriver, K.; Fukamizo, T. A class V chitinase from Arabidopsis thaliana: Gene responses, enzymatic properties, and crystallographic analysis. Planta 2011, 234, 123–137. [Google Scholar] [CrossRef]
  27. Gomez, L.; Allona, I.; Casado, R.; Aragoncillo, C. Seed chitinases. Seed Sci. Res. 2002, 12, 217–230. [Google Scholar] [CrossRef]
  28. Maia, L.B.L.; Pereira, H.D.M.; Garratt, R.C.; Brandão-Neto, J.; Henrique-Silva, F.; Toyama, D.; Dias, R.O.; Bachega, J.F.R.; Peixoto, J.V.; Silva-Filho, M.C. Structural and Evolutionary Analyses of PR-4 SUGARWINs Points to a Different Pattern of Protein Function. Front. Plant Sci. 2021, 12, 734248. [Google Scholar] [CrossRef] [PubMed]
  29. Sasaki, C.; Yokoyama, A.; Itoh, Y.; Hashimoto, M.; Watanabe, T.; Fukamizo, T. Comparative study of the reaction mechanism of eamily 18 chitinases from plants and microbes. J. Biochem. 2002, 131, 557–564. [Google Scholar] [CrossRef] [PubMed]
  30. Nagpure, A.; Choudhary, B.; Gupta, R.K. Chitinases: In agriculture and human healthcare. Crit. Rev. Biotechnol. 2014, 34, 215–232. [Google Scholar] [CrossRef] [PubMed]
  31. Chandra, S.; Dutta, A.K.; Chandrashekara, K.N.; Acharya, K. In silico characterization, homology modeling of Camellia sinensis chitinase and its evolutionary analyses with other plant chitinases. Proc. Natl. Acad. Sci. India Sect. B Biol. Sci. 2017, 87, 685–695. [Google Scholar] [CrossRef]
  32. Šindelářová, M.; Šindelář, L. Isolation of pathogenesis-related proteins from TMV-infected tobacco and their influence on infectivity of TMV. Plant Prot. Sci. 2005, 41, 52–57. [Google Scholar] [CrossRef]
  33. Tapia, G.; Morales-Quintana, L.; Inostroza, L.; Acuna, H. Molecular characterisation of Ltchi7, a gene encoding a Class III endochitinase induced by drought stress in Lotus spp. Plant Biol. 2011, 13, 69–77. [Google Scholar] [CrossRef]
  34. Janská, A.; Maršík, P.; Zelenková, S.; Ovesná, J. Cold stress and acclimation–what is important for metabolic adjustment? Plant Biol. 2010, 12, 395–405. [Google Scholar] [CrossRef]
  35. Mészáros, P.; Rybanský, Ľ.; Spieß, N.; Socha, P.; Kuna, R.; Libantová, J.; Moravčíková, J.; Piršelová, B.; Hauptvogel, P.; Matušíková, I. Plant chitinase responses to different metal-type stresses reveal specificity. Plant Cell Rep. 2014, 33, 1789–1799. [Google Scholar] [CrossRef] [PubMed]
  36. Lu, H.-C.; Lin, J.-H.; Chua, A.C.N.; Chung, T.-Y.; Tsai, I.C.; Tzen, J.T.C.; Chou, W.-M. Cloning and expression of pathogenesis-related protein 4 from jelly fig (Ficus awkeotsang Makino) achenes associated with ribonuclease, chitinase and anti-fungal activities. Plant Physiol. Biochem. 2012, 56, 1–13. [Google Scholar] [CrossRef] [PubMed]
  37. Edens, L.; Heslinga, L.; Klok, R.; Ledeboer, A.M.; Maat, J.; Toonen, M.Y.; Visser, C.; Verrips, C.T. Cloning of cDNA encoding the sweet-tasting plant protein thaumatin and its expression in Escherichia coli. Gene 1982, 18, 1–12. [Google Scholar] [CrossRef]
  38. Kauffmann, S.; Legrand, M.; Fritig, B. Isolation and characterization of six pathogenesis-related (PR) proteins of Samsun NN tobacco. Plant Mol. Biol. 1990, 14, 381–390. [Google Scholar] [CrossRef]
  39. Brandazza, A.; Angeli, S.; Tegoni, M.; Cambillau, C.; Pelosi, P. Plant stress proteins of the thaumatin-like family discovered in animals. FEBS Lett. 2004, 572, 3–7. [Google Scholar] [CrossRef]
  40. Piggott, N.; Ekramoddoullah, A.K.; Liu, J.-J.; Yu, X. Gene cloning of a thaumatin-like (PR-5) protein of western white pine (Pinus monticola D. Don) and expression studies of members of the PR-5 group. Physiol. Mol. Plant Pathol. 2004, 64, 1–8. [Google Scholar] [CrossRef]
  41. Koiwa, H.; Kato, H.; Nakatsu, T.; Oda, J.i.; Yamada, Y.; Sato, F. Crystal structure of tobacco PR-5d protein at 1.8 Å resolution reveals a conserved acidic cleft structure in antifungal thaumatin-like proteins11Edited by R. Huber. J. Mol. Biol. 1999, 286, 1137–1145. [Google Scholar] [CrossRef] [PubMed]
  42. Li, S.; Zhang, Y.; Xin, X.; Ding, C.; Lv, F.; Mo, W.; Xia, Y.; Wang, S.; Cai, J.; Sun, L.; et al. The Osmotin-Like Protein Gene PdOLP1 Is Involved in Secondary Cell Wall Biosynthesis during Wood Formation in Poplar. Int. J. Mol. Sci. 2020, 21, 3993. [Google Scholar] [CrossRef]
  43. Singh, S.; Tripathi, R.K.; Lemaux, P.G.; Buchanan, B.B.; Singh, J. Redox-dependent interaction between thaumatin-like protein and β-glucan influences malting quality of barley. Proc. Natl. Acad. Sci. USA 2017, 114, 7725–7730. [Google Scholar] [CrossRef]
  44. Kobayashi, K.; Fukuda, M.; Igarashi, D.; Sunaoshi, M. Cytokinin-binding proteins from tobacco callus share homology with osmotin-like protein and an endochitinase. Plant Cell Physiol. 2000, 41, 148–157. [Google Scholar] [CrossRef] [PubMed]
  45. Stintzi, A.; Heitz, T.; Kauffmann, S.; Legrand, M.; Fritig, B. Identification of a basic pathogenesis-related, thaumatin-like protein of virus-infected tobacco as osmotin. Physiol. Mol. Plant Pathol. 1991, 38, 137–146. [Google Scholar] [CrossRef]
  46. Singh, N.K.; Bracker, C.A.; Hasegawa, P.M.; Handa, A.K.; Buckel, S.; Hermodson, M.A.; Pfankoch, E.; Regnier, F.E.; Bressan, R.A. Characterization of osmotin: A thaumatin-like protein associated with osmotic adaptation in plant cells. Plant Physiol. 1987, 85, 529–536. [Google Scholar] [CrossRef] [PubMed]
  47. Breiteneder, H.; Radauer, C. A classification of plant food allergens. J. Allergy Clin. Immunol. 2004, 113, 821–831. [Google Scholar] [CrossRef]
  48. Myagmarjav, D.; Sukweenadhi, J.; Kim, Y.-J.; Jang, M.-G.; Rahimi, S.; Silva, J.; Choi, J.-Y.; Mohanan, P.; Kwon, W.-S.; Kim, C. Molecular characterization and expression analysis of pathogenesis related protein 6 from Panax ginseng. Russ. J. Genet. 2017, 53, 1211–1220. [Google Scholar] [CrossRef]
  49. Doares, S.H.; Syrovets, T.; Weiler, E.W.; Ryan, C.A. Oligogalacturonides and chitosan activate plant defensive genes through the octadecanoid pathway. P. Natl. A. Sci. India. B 1995, 92, 4095–4098. [Google Scholar] [CrossRef] [PubMed]
  50. Terras, F.R.; Schoofs, H.M.; Thevissen, K.; Osborn, R.W.; Vanderleyden, J.; Cammue, B.P.; Broekaert, W.F. Synergistic enhancement of the antifungal activity of wheat and barley thionins by radish and oilseed rape 2S albumins and by barley trypsin inhibitors. Plant Physiol. 1993, 103, 1311–1319. [Google Scholar] [CrossRef]
  51. Koiwa, H.; Bressan, R.A.; Hasegawa, P.M. Regulation of protease inhibitors and plant defense. Trends Plant Sci. 1997, 2, 379–384. [Google Scholar] [CrossRef]
  52. Umemoto, N.; Kakitani, M.; Iwamatsu, A.; Yoshikawa, M.; Yamaoka, N.; Ishida, I. The structure and function of a soybean β-glucan-elicitor-binding protein. PNAS 1997, 94, 1029–1034. [Google Scholar] [CrossRef] [PubMed]
  53. Lavrova, V.; Matveeva, E.; Zinovieva, S. Expression of genes, encoded defense proteins, in potato plants infected with the cyst-forming nematode Globodera rostochiensis (Wollenweber 1923) Behrens, 1975 and modulation of their activity during short-term exposure to low temperatures. Biol. Bull. 2017, 44, 128–136. [Google Scholar] [CrossRef]
  54. Giacometti, R.; Barneto, J.; Barriga, L.G.; Sardoy, P.M.; Balestrasse, K.; Andrade, A.M.; Pagano, E.A.; Alemano, S.G.; Zavala, J.A. Early perception of stink bug damage in developing seeds of field-grown soybean induces chemical defences and reduces bug attack. Pest Manag. Sci. 2016, 72, 1585–1594. [Google Scholar] [CrossRef]
  55. Vera, P.; Conejero, V. The induction and accumulation of the pathogenesis-related P69 proteinase in tomato during citrus exocortis viroid infection and in response to chemical treatments. Physiol. Mol. Plant Pathol. 1989, 34, 323–334. [Google Scholar] [CrossRef]
  56. Tornero, P.; Conejero, V.; Vera, P. Primary structure and expression of a pathogen-induced protease (PR-P69) in tomato plants: Similarity of functional domains to subtilisin-like endoproteases. PNAS 1996, 93, 6332–6337. [Google Scholar] [CrossRef] [PubMed]
  57. Siezen, R.J.; de Vos, W.M.; Leunissen, J.A.; Dijkstra, B.W. Homology modelling and protein engineering strategy of subtilases, the family of subtilisin-like serine proteinases. Protein Eng. Des. Sel. 1991, 4, 719–737. [Google Scholar] [CrossRef]
  58. Ahmad, A.; Shafique, S.; Shafique, S. Molecular basis of antifungal resistance in tomato varieties. Pak. J. Agr. Sci. 2014, 51, 683–687. [Google Scholar]
  59. Campos, M.A.; Rosa, D.D.; Teixeira, J.É.C.; Targon, M.L.P.; Souza, A.A.; Paiva, L.V.; Stach-Machado, D.R.; Machado, M.A. PR gene families of citrus: Their organ specific-biotic and abiotic inducible expression profiles based on ESTs approach. Genet. Mol. Biol. 2007, 30, 917–930. [Google Scholar] [CrossRef]
  60. Baboulène, L.; Silvestre, J.; Pinelli, E.; Morard, P. Effect of Ca Deficiency on Growth and Leaf Acid Soluble Proteins of Tomato. J. Plant Nutr. 2007, 30, 497–515. [Google Scholar] [CrossRef]
  61. Irigoyen, M.L.; Garceau, D.C.; Bohorquez-Chaux, A. Genome-wide analyses of cassava Pathogenesis-related (PR) gene families reveal core. BMC Genom. 2020, 21, 1. [Google Scholar] [CrossRef]
  62. Vera, P.; Conejero, V. Effect of Ethephon on Protein Degradation and the Accumulation of Pathogenesis-Related’(PR) Proteins in Tomato Leaf Discs. Plant Physiol. 1990, 92, 227–233. [Google Scholar] [CrossRef] [PubMed]
  63. Barceló, A.R.; Ros, L.V.G.; Carrasco, A.E.J. Looking for syringyl peroxidases. Trends Plant Sci. 2007, 12, 486–491. [Google Scholar] [CrossRef] [PubMed]
  64. Nawrot, R.; Barylski, J.; Nowicki, G.; Broniarczyk, J.; Buchwald, W.; Goździcka-Józefiak, A. Plant antimicrobial peptides. Folia Microbiol. 2014, 59, 181–196. [Google Scholar] [CrossRef] [PubMed]
  65. Lavid, N.; Schwartz, A.; Yarden, O.; Tel-Or, E. The involvement of polyphenols and peroxidase activities in heavy-metal accumulation by epidermal glands of the waterlily (Nymphaeaceae). Planta 2001, 212, 323–331. [Google Scholar] [CrossRef] [PubMed]
  66. Park, S.Y.; Ryu, S.H.; Kwon, S.Y.; Lee, H.S.; Kim, J.; Kwak, S.S. Differential expression of six novel peroxidase cDNAs from cell cultures of sweetpotato in response to stress. Mol. Gen. Genom. 2003, 269, 542–552. [Google Scholar] [CrossRef]
  67. Liu, J.-J.; Ekramoddoullah, A.K. The family 10 of plant pathogenesis-related proteins: Their structure, regulation, and function in response to biotic and abiotic stresses. Physiol. Mol. Plant Pathol. 2006, 68, 3–13. [Google Scholar] [CrossRef]
  68. Biesiadka, J.; Bujacz, G.; Sikorski, M.M.; Jaskolski, M. Crystal structures of two homologous pathogenesis-related proteins from yellow lupine. J. Mol. Biol. 2002, 319, 1223–1234. [Google Scholar] [CrossRef] [PubMed]
  69. Park, C.J.; Kim, K.J.; Shin, R.; Park, J.M.; Shin, Y.C.; Paek, K.H. Pathogenesis-related protein 10 isolated from hot pepper functions as a ribonuclease in an antiviral pathway. Plant J. 2004, 37, 186–198. [Google Scholar] [CrossRef] [PubMed]
  70. Xie, Y.-R.; Chen, Z.-Y.; Brown, R.L.; Bhatnagar, D. Expression and functional characterization of two pathogenesis-related protein 10 genes from Zea mays. J. Plant Physiol. 2010, 167, 121–130. [Google Scholar] [CrossRef] [PubMed]
  71. Pühringer, H.; Moll, D.; Hoffmann-Sommergruber, K.; Watillon, B.; Katinger, H.; da Câmara Machado, M.L. The promoter of an apple Ypr10 gene, encoding the major allergen Mal d 1, is stress-and pathogen-inducible. Plant Sci. 2000, 152, 35–50. [Google Scholar] [CrossRef]
  72. Gonneau, M.; Pagant, S.; Brun, F.; Laloue, M. Photoaffinity labelling with the cytokinin agonist azido-CPPU of a 34 kDa peptide of the intracellular pathogenesis-related protein family in the moss Physcomitrella patens. Plant Mol. Biol. 2001, 46, 539–548. [Google Scholar] [CrossRef] [PubMed]
  73. Odintsova, T.I.; Slezina, M.P.; Istomina, E.A. Defensins of Grasses: A Systematic Review. Biomolecules 2020, 10, 1029. [Google Scholar] [CrossRef] [PubMed]
  74. Jha, S.; Chattoo, B.B. Expression of a plant defensin in rice confers resistance to fungal phytopathogens. Transgenic Res. 2010, 19, 373–384. [Google Scholar] [CrossRef] [PubMed]
  75. Yount, N.Y.; Yeaman, M.R. Peptide antimicrobials: Cell wall as a bacterial target. Ann. N. Y. Acad. Sci. 2013, 1277, 127–138. [Google Scholar] [CrossRef]
  76. Koike, M.; Okamoto, T.; Tsuda, S.; Imai, R. A novel plant defensin-like gene of winter wheat is specifically induced during cold acclimation. Biochem. Biophys. Res. Commun. 2002, 298, 46–53. [Google Scholar] [CrossRef]
  77. Do, H.M.; Lee, S.C.; Jung, H.W.; Sohn, K.H.; Hwang, B.K. Differential expression and in situ localization of a pepper defensin (CADEF1) gene in response to pathogen infection, abiotic elicitors and environmental stresses in Capsicum annuum. Plant Sci. 2004, 166, 1297–1305. [Google Scholar] [CrossRef]
  78. Mirouze, M.; Sels, J.; Richard, O.; Czernic, P.; Loubet, S.; Jacquier, A.; François, I.E.; Cammue, B.P.; Lebrun, M.; Berthomieu, P. A putative novel role for plant defensins: A defensin from the zinc hyper-accumulating plant, Arabidopsis halleri, confers zinc tolerance. Plant J. 2006, 47, 329–342. [Google Scholar] [CrossRef]
  79. Bohlmann, H.; Broekaert, W. The role of thionins in plant protection. Crit. Rev. Plant Sci. 1994, 13, 1–16. [Google Scholar] [CrossRef]
  80. Chan, Y.-L.; Prasad, V.; Sanjaya; Chen, K.H.; Liu, P.C.; Chan, M.-T.; Cheng, C.-P. Transgenic tomato plants expressing an Arabidopsis thionin (Thi2. 1) driven by fruit-inactive promoter battle against phytopathogenic attack. Planta 2005, 221, 386–393. [Google Scholar] [CrossRef]
  81. Leybourne, D.J.; Valentine, T.A.; Binnie, K.; Taylor, A.; Karley, A.J.; Bos, J.I.B. Drought stress increases the expression of barley defence genes with negative consequences for infesting cereal aphids. J. Exp. Bot. 2022, 73, 2238–2250. [Google Scholar] [CrossRef] [PubMed]
  82. Liu, X.; Gong, X.; Zhou, D.; Jiang, Q.; Liang, Y.; Ye, R.; Zhang, S.; Wang, Y.; Tang, X.; Li, F.; et al. Plant Defensin-Dissimilar Thionin OsThi9 Alleviates Cadmium Toxicity in Rice Plants and Reduces Cadmium Accumulation in Rice Grains. J. Agric. Food Chem. 2023, 71, 8367–8380. [Google Scholar] [CrossRef] [PubMed]
  83. García-Olmedo, F.; Molina, A.; Alamillo, J.M.; Rodríguez-Palenzuéla, P. Plant defense peptides. Pept. Sci. 1998, 47, 479–491. [Google Scholar] [CrossRef]
  84. Wang, S.-Y.; Wu, J.-H.; Ng, T.; Ye, X.-Y.; Rao, P.-F. A non-specific lipid transfer protein with antifungal and antibacterial activities from the mung bean. Peptides 2004, 25, 1235–1242. [Google Scholar] [CrossRef] [PubMed]
  85. Xue, Y.; Zhang, C.; Shan, R.; Li, X.; Tseke Inkabanga, A.; Li, L.; Jiang, H.; Chai, Y. Genome-wide identification and expression analysis of nsLTP gene family in rapeseed (Brassica napus) reveals their critical roles in biotic and abiotic stress responses. Int. J. Mol. Sci. 2022, 23, 8372. [Google Scholar] [CrossRef] [PubMed]
  86. Govindan, G.; Sandhiya, K.R.; Alphonse, V.; Somasundram, S. Role of Germin-Like Proteins (GLPs) in Biotic and Abiotic Stress Responses in Major Crops: A Review on Plant Defense Mechanisms and Stress Tolerance. Plant Mol. Biol. Rep. 2024, 42, 450–468. [Google Scholar] [CrossRef]
  87. Islam, M.M.; El-Sappah, A.H.; Ali, H.M.; Zandi, P.; Huang, Q.; Soaud, S.A.; Alazizi, E.M.Y.; Wafa, H.A.; Hossain, M.A.; Liang, Y. Pathogenesis-related proteins (PRs) countering environmental stress in plants: A review. S. Afr. J. Bot. 2023, 160, 414–427. [Google Scholar] [CrossRef]
  88. Folkerts, O. Overexpression of a gene encoding H2O2-generating oxalate oxidase evoke defense responses in sunflower. Plant Physiol. 2003, 133, 170181. [Google Scholar]
  89. Mittler, R. Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci. 2002, 7, 405–410. [Google Scholar] [CrossRef]
  90. Krishna, G.; Singh, B.K.; Kim, E.K.; Morya, V.K.; Ramteke, P.W. Progress in genetic engineering of peanut (Arachis hypogaea L.)—A review. Plant Biotechnol. J. 2015, 13, 147–162. [Google Scholar] [CrossRef]
  91. Druka, A.; Kudrna, D.; Kannangara, C.G.; von Wettstein, D.; Kleinhofs, A. Physical and genetic mapping of barley (Hordeum vulgare) germin-like cDNAs. Proc. Natl. Acad. Sci. USA 2002, 99, 850–855. [Google Scholar] [CrossRef] [PubMed]
  92. Li, Y.; Zhang, D.; Li, W.; Mallano, A.I.; Zhang, Y.; Wang, T.; Lu, M.; Qin, Z.; Li, W. Expression study of soybean germin-like gene family reveals a role of GLP7 gene in various abiotic stress tolerances. Can. J. Plant Sci. 2016, 96, 296–304. [Google Scholar] [CrossRef]
  93. Christensen, A.B.; Cho, B.H.; Naesby, M.; Gregersen, P.L.; Brandt, J.; Madriz-Ordenana, K.; Collinge, D.B.; Thordal-Christensen, H. The molecular characterization of two barley proteins establishes the novel PR-17 family of pathogenesis-related proteins. Mol. Plant Pathol. 2002, 3, 135–144. [Google Scholar] [CrossRef]
  94. Schweizer, P.; Pokorny, J.; Abderhalden, O.; Dudler, R. A transient assay system for the functional assessment of defense-related genes in wheat. Mol. Plant-Microbe Interact. 1999, 12, 647–654. [Google Scholar] [CrossRef]
  95. Benhamou, N.; Bélanger, R.R. Benzothiadiazole-Mediated Induced Resistance to Fusarium oxysporum f. sp.radicis-lycopersici in Tomato. Plant Physiol. 1998, 118, 1203–1212. [Google Scholar] [CrossRef]
  96. Luo, X.; Tian, T.; Feng, L.; Yang, X.; Li, L.; Tan, X.; Wu, W.; Li, Z.; Treves, H.; Serneels, F. Pathogenesis-related protein 1 suppresses oomycete pathogen by targeting against AMPK kinase complex. J. Adv. Res. 2023, 43, 13–26. [Google Scholar] [CrossRef] [PubMed]
  97. Han, Z.; Xiong, D.; Schneiter, R.; Tian, C. The function of plant PR1 and other members of the CAP protein superfamily in plant-pathogen interactions. Mol. Plant Pathol. 2023, 24, 651–668. [Google Scholar] [CrossRef] [PubMed]
  98. Breen, S.; Williams, S.J.; Outram, M.; Kobe, B.; Solomon, P.S. Emerging Insights into the Functions of Pathogenesis-Related Protein 1. Trends Plant Sci. 2017, 22, 871–879. [Google Scholar] [CrossRef] [PubMed]
  99. Guo, J.; Bai, Y.; Wei, Y.; Dong, Y.; Zeng, H.; Reiter, R.J.; Shi, H. Fine-tuning of pathogenesis-related protein 1 (PR1) activity by the melatonin biosynthetic enzyme ASMT2 in defense response to cassava bacterial blight. J. Pineal Res. 2022, 72, e12784. [Google Scholar] [CrossRef]
  100. Niderman, T.; Genetet, I.; Bruyère, T.; Gees, R.; Stintzi, A.; Legrand, M.; Fritig, B.; Mösinger, E. Pathogenesis-related PR-1 proteins are antifungal. Isolation and characterization of three 14-kilodalton proteins of tomato and of a basic PR-1 of tobacco with inhibitory activity against Phytophthora infestans. Plant Physiol. 1995, 108, 17–27. [Google Scholar] [CrossRef] [PubMed]
  101. Almeida-Silva, F.; Venancio, T.M. Pathogenesis-related protein 1 (PR-1) genes in soybean: Genome-wide identification, structural analysis and expression profiling under multiple biotic and abiotic stresses. Gene 2022, 809, 146013. [Google Scholar] [CrossRef] [PubMed]
  102. Goyal, R.K.; Fatima, T.; Topuz, M.; Bernadec, A.; Sicher, R.; Handa, A.K.; Mattoo, A.K. Pathogenesis-Related Protein 1b1 (PR1b1) Is a Major Tomato Fruit Protein Responsive to Chilling Temperature and Upregulated in High Polyamine Transgenic Genotypes. Front. Plant Sci. 2016, 7, 901. [Google Scholar] [CrossRef] [PubMed]
  103. Zribi, I.; Ghorbel, M.; Haddaji, N.; Besbes, M.; Brini, F. Genome-wide identification and expression profiling of pathogenesis-related protein 1 (PR-1) genes in durum wheat (Triticum durum Desf.). Plants 2023, 12, 1998. [Google Scholar] [CrossRef] [PubMed]
  104. Sarowar, S.; Kim, Y.J.; Kim, E.N.; Kim, K.D.; Hwang, B.K.; Islam, R.; Shin, J.S. Overexpression of a pepper basic pathogenesis-related protein 1 gene in tobacco plants enhances resistance to heavy metal and pathogen stresses. Plant Cell Rep. 2005, 24, 216–224. [Google Scholar] [CrossRef] [PubMed]
  105. Green, R.; Fluhr, R. UV-B-induced PR-1 accumulation is mediated by active oxygen species. Plant Cell 1995, 7, 203–212. [Google Scholar] [CrossRef]
  106. Leah, R.; Tommerup, H.; Svendsen, I.; Mundy, J. Biochemical and molecular characterization of three barley seed proteins with antifungal properties. J. Biol. Chem. 1991, 266, 1564–1573. [Google Scholar] [CrossRef] [PubMed]
  107. Pei, Y.; Xue, Q.; Zhang, Z.; Shu, P.; Deng, H.; Bouzayen, M.; Hong, Y.; Liu, M. β-1, 3-GLUCANASE10 regulates tomato development and disease resistance by modulating callose deposition. Plant Physiol. 2023, 192, 2785–2802. [Google Scholar] [CrossRef]
  108. Nawrath, C.; Métraux, J.-P. Salicylic acid induction–deficient mutants of Arabidopsis express PR-2 and PR-5 and accumulate high levels of camalexin after pathogen inoculation. Plant Cell 1999, 11, 1393–1404. [Google Scholar] [CrossRef]
  109. Sunpapao, A.; Pornsuriya, C. Overexpression of β-1, 3-glucanase gene in response to Phytophthora palmivora infection in leaves of Hevea brasiliensis clones. Wal. J. Sci. Technol. 2016, 13, 35–43. [Google Scholar]
  110. Li, Y.; Jiao, M.; Li, Y.; Zhong, Y.; Li, X.; Chen, Z.; Chen, S.; Wang, J. Penicillium chrysogenum polypeptide extract protects tobacco plants from tobacco mosaic virus infection through modulation of ABA biosynthesis and callose priming. J. Exp. Bot. 2021, 72, 3526–3539. [Google Scholar] [CrossRef]
  111. Rabie, M.; Aseel, D.G.; Younes, H.A.; Behiry, S.I.; Abdelkhalek, A. Transcriptional responses and secondary metabolites variation of tomato plant in response to tobacco mosaic virus infestation. Sci. Rep. 2024, 14, 19565. [Google Scholar] [CrossRef]
  112. Bozbuga, R. Commonalities of molecular response in tomato plants against parasitic nematodes. Biol. Bull. 2021, 48, S12–S21. [Google Scholar] [CrossRef]
  113. Wang, L.; Li, R.; Li, K.; Qu, Z.; Zhou, R.; Lu, G.; Li, P.; Li, G. Genome-wide identification of the grapevine β-1, 3-glucanase gene (VviBG) family and expression analysis under different stresses. BMC Plant Biol. 2024, 24, 911. [Google Scholar] [CrossRef] [PubMed]
  114. Sock, J.; Rohringer, R.; Kang, Z. Extracellular beta-1,3-Glucanases in Stem Rust-Affected and Abiotically Stressed Wheat Leaves: Immunocytochemical Localization of the Enzyme and Detection of Multiple Forms in Gels by Activity Staining with Dye-Labeled Laminarin. Plant Physiol. 1990, 94, 1376–1389. [Google Scholar] [CrossRef] [PubMed]
  115. Surplus, S.; Jordan, B.; Murphy, A.; Carr, J.; Thomas, B.; Mackerness, S.H. Ultraviolet-B-induced responses in Arabidopsis thaliana: Role of salicylic acid and reactive oxygen species in the regulation of transcripts encoding photosynthetic and acidic pathogenesis-related proteins. Plant Cell Environ. 1998, 21, 685–694. [Google Scholar] [CrossRef]
  116. Barre, A.; Damme, E.; Simplicien, M.; Benoist, H.; Rouge, P. Are Dietary Lectins Relevant Allergens in Plant Food Allergy? Foods 2020, 9, 1724. [Google Scholar] [CrossRef]
  117. Schlumbaum, A.; Mauch, F.; Vögeli, U.; Boller, T. Plant chitinases are potent inhibitors of fungal growth. Nature 1986, 324, 365–367. [Google Scholar] [CrossRef]
  118. Chandrashekar, N.; Ali, S.; Grover, A. Exploring expression patterns of PR-1, PR-2, PR-3, and PR-12 like genes in Arabidopsis thaliana upon Alternaria brassicae inoculation. 3 Biotech 2018, 8, 230. [Google Scholar] [CrossRef]
  119. Omar, A.Z.; Hamdy, E.; Hamed, E.A.; Hafez, E.; Abdelkhalek, A. The curative activity of some arylidene dihydropyrimidine hydrazone against Tobacco mosaic virus infestation. J. Saudi Chem. Soc. 2022, 26, 101504. [Google Scholar] [CrossRef]
  120. Kumar, S.A.; Kumari, P.H.; Jawahar, G.; Prashanth, S.; Suravajhala, P.; Katam, R.; Sivan, P.; Rao, K.S.; Kirti, P.B.; Kishor, P.B.K. Beyond just being foot soldiers—osmotin like protein (OLP) and chitinase (Chi11) genes act as sentinels to confront salt, drought, and fungal stress tolerance in tomato. Environ. Exp. Bot. 2016, 132, 53–65. [Google Scholar] [CrossRef]
  121. Yun, H.-K.; Yi, S.-Y.; Yu, S.-H.; Choi, D. Cloning of a Pathogenesis-Related Protein-1 Gene from Nicotians glutinosa L. and Its Salicylic Acid-Independent Induction by Copper and β-Aminobutyric Acid. J. Plant Physiol. 1999, 154, 327–333. [Google Scholar] [CrossRef]
  122. Guevara-Morato, M.A.; García de Lacoba, M.; García-Luque, I.; Serra, M.T. Characterization of a pathogenesis-related protein 4 (PR-4) induced in Capsicum chinense L3 plants with dual RNase and DNase activities. J. Exp. Bot. 2010, 61, 3259–3271. [Google Scholar] [CrossRef]
  123. Chouhan, R.; Ahmed, S.; Gandhi, S.G. Over-expression of PR proteins with chitinase activity in transgenic plants for alleviation of fungal pathogenesis. J. Plant Pathol. 2023, 105, 69–81. [Google Scholar] [CrossRef]
  124. Wang, N.; Xiao, B.; Xiong, L. Identification of a cluster of PR4-like genes involved in stress responses in rice. J. Plant Physiol. 2011, 168, 2212–2224. [Google Scholar] [CrossRef]
  125. Kim, Y.; Lee, H.; Jang, M.; Kwon, W.; Kim, S.; Yang, D. Cloning and characterization of pathogenesis-related protein 4 gene from Panax ginseng. Russ. J. Plant Physl. 2014, 61, 664–671. [Google Scholar] [CrossRef]
  126. Bashir, M.A.; Silvestri, C.; Ahmad, T.; Hafiz, I.A.; Abbasi, N.A.; Manzoor, A.; Cristofori, V.; Rugini, E. Osmotin: A Cationic Protein Leads to Improve Biotic and Abiotic Stress Tolerance in Plants. Plants 2020, 9, 992. [Google Scholar] [CrossRef]
  127. de Jesus-Pires, C.; Ferreira-Neto, J.R.C.; Pacifico Bezerra-Neto, J.; Kido, E.A.; de Oliveira Silva, R.L.; Pandolfi, V.; Wanderley-Nogueira, A.C.; Binneck, E.; da Costa, A.F.; Pio-Ribeiro, G.; et al. Plant Thaumatin-like Proteins: Function, Evolution and Biotechnological Applications. Curr. Protein Pept. Sci. 2020, 21, 36–51. [Google Scholar] [CrossRef] [PubMed]
  128. Ullah, A.; Hussain, A.; Shaban, M.; Khan, A.H.; Alariqi, M.; Gul, S.; Jun, Z.; Lin, S.; Li, J.; Jin, S. Osmotin: A plant defense tool against biotic and abiotic stresses. Plant Physiol. Biochem. 2018, 123, 149–159. [Google Scholar] [CrossRef] [PubMed]
  129. Ruiz-Medrano, R.; Jimenez-Moraila, B.; Herrera-Estrella, L.; Rivera-Bustamante, R.F. Nucleotide sequence of an osmotin-like cDNA induced in tomato during viroid infection. Plant Mol. Biol. 1992, 20, 1199–1202. [Google Scholar] [CrossRef] [PubMed]
  130. Aseel, D.G.; Abdelkhalek, A.; Alotibi, F.O.; Samy, M.A.; Al-Askar, A.A.; Arishi, A.A.; Hafez, E.E. Foliar application of nanoclay promotes potato (Solanum tuberosum L.) growth and induces systemic resistance against potato virus Y. Viruses 2022, 14, 2151. [Google Scholar] [CrossRef]
  131. Viktorova, J.; Krasny, L.; Kamlar, M.; Novakova, M.; Mackova, M.; Macek, T. Osmotin, a pathogenesis-related protein. Curr. Protein Pept. Sci. 2012, 13, 672–681. [Google Scholar] [CrossRef]
  132. Rajam, M.; Chandola, N.; Saiprasad Goud, P.; Singh, D.; Kashyap, V.; Choudhary, M.; Sihachakr, D. Thaumatin gene confers resistance to fungal pathogens as well as tolerance to abiotic stresses in transgenic tobacco plants. Biol. Plant. 2007, 51, 135–141. [Google Scholar] [CrossRef]
  133. Singh, V.; Hallan, V.; Pati, P.K. Withania somnifera osmotin (WsOsm) confers stress tolerance in tobacco and establishes novel interactions with the defensin protein (WsDF). Physiol. Plant. 2024, 176, e14513. [Google Scholar] [CrossRef] [PubMed]
  134. Patade, V.Y.; Khatri, D.; Kumari, M.; Grover, A.; Mohan Gupta, S.; Ahmed, Z. Cold tolerance in Osmotin transgenic tomato (Solanum lycopersicum L.) is associated with modulation in transcript abundance of stress responsive genes. SpringerPlus 2013, 2, 117. [Google Scholar] [CrossRef] [PubMed]
  135. Zhang, X.; Zhou, Q.; Wang, X.; Cai, J.; Dai, T.; Cao, W.; Jiang, D. Physiological and transcriptional analyses of induced post-anthesis thermo-tolerance by heat-shock pretreatment on germinating seeds of winter wheat. Environ. Exp. Bot. 2016, 131, 181–189. [Google Scholar] [CrossRef]
  136. Djemal, R.; Bahloul, O.; Khoudi, H. A novel thaumatin-like protein from durum wheat, TdPR-5, is homologous to known plant allergens, responsive to stress exposure, and confers multiple-abiotic stress tolerances to transgenic yeast. Plant Gene 2022, 31, 100360. [Google Scholar] [CrossRef]
  137. Yarullina, L.; Cherepanova, E.A.; Burkhanova, G.F.; Sorokan, A.V.; Zaikina, E.A.; Tsvetkov, V.O.; Mardanshin, I.S.; Fatkullin, I.Y.; Kalatskaja, J.N.; Yalouskaya, N.A.; et al. Stimulation of the Defense Mechanisms of Potatoes to a Late Blight Causative Agent When Treated with Bacillus subtilis Bacteria and Chitosan Composites with Hydroxycinnamic Acids. Microorganisms 2023, 11, 1993. [Google Scholar] [CrossRef]
  138. Sabnam, N.; Hussain, A.; Saha, P. The secret password: Cell death-inducing proteins in filamentous phytopathogens—As versatile tools to develop disease-resistant crops. Microb. Pathog. 2023, 183, 106276. [Google Scholar] [CrossRef] [PubMed]
  139. Abdelkhalek, A.; Al-Askar, A.A.; Alsubaie, M.M.; Behiry, S.I. First report of protective activity of Paronychia argentea extract against tobacco mosaic virus infection. Plants 2021, 10, 2435. [Google Scholar] [CrossRef] [PubMed]
  140. Bozbuga, R. Molecular analysis of nematode-responsive defence genes CRF1, WRKY45, and PR7 in Solanum lycopersicum tissues during the infection of plant-parasitic nematode species of the genus Meloidogyne. Genome 2022, 65, 265–275. [Google Scholar] [CrossRef] [PubMed]
  141. Derikvand, F.; Bazgir, E.; Darvishnia, M.; Mirzaei Najafgholi, H. Evaluation the activity of peroxidase and catalase enzymes and the expression level of PR1 and PR8 genes in apple fruit following brown rot (Monilinia laxa) disease. Plant Genet. Res. 2023, 10, 29–42. [Google Scholar]
  142. Fagerstedt, K.V.; Kukkola, E.M.; Koistinen, V.V.; Takahashi, J.; Marjamaa, K. Cell wall lignin is polymerised by class III secretable plant peroxidases in Norway spruce. J. Integr. Plant Biol. 2010, 52, 186–194. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, X.; El Hadrami, A.; Adam, L.R.; Daayf, F. Genes encoding pathogenesis-related proteins PR-2, PR-3 and PR-9, are differentially regulated in potato leaves inoculated with isolates from US-1 and US-8 genotypes of Phytophthora infestans (Mont.) de Bary. Physiol. Mol. Plant Pathol. 2005, 67, 49–56. [Google Scholar] [CrossRef]
  144. Lambais, M.R. In silico differential display of defense-related expressed sequence tags from sugarcane tissues infected with diazotrophic endophytes. Genet. Mol. Biol. 2001, 24, 103–111. [Google Scholar] [CrossRef]
  145. Carrillo, M.G.C.; Martin, F.; Variar, M.; Bhatt, J.C.; Perez-Quintero, A.L.; Leung, H.; Leach, J.E.; Vera Cruz, C.M. Accumulating candidate genes for broad-spectrum resistance to rice blast in a drought-tolerant rice cultivar. Sci. Rep. 2021, 11, 21502. [Google Scholar] [CrossRef]
  146. Morris, J.S.; Caldo, K.M.P.; Liang, S.; Facchini, P.J. PR10/Bet v1-like Proteins as Novel Contributors to Plant Biochemical Diversity. ChemBioChem 2021, 22, 264–287. [Google Scholar] [CrossRef]
  147. Lopes, N.D.S.; Santos, A.S.; de Novais, D.P.S.; Pirovani, C.P.; Micheli, F. Pathogenesis-related protein 10 in resistance to biotic stress: Progress in elucidating functions, regulation and modes of action. Front. Plant Sci. 2023, 14, 1193873. [Google Scholar] [CrossRef] [PubMed]
  148. Barreto, L.R.; Barreto, T.; Melo, S.; Pungartnik, C.; Brendel, M. Sensitivity of Yeast Mutants Deficient in Mitochondrial or Vacuolar ABC Transporters to Pathogenesis-Related Protein TcPR-10 of Theobroma cacao. Biology 2018, 7, 35. [Google Scholar] [CrossRef] [PubMed]
  149. Regner, A.; Szepannek, N.; Wiederstein, M.; Fakhimahmadi, A.; Paciosis, L.F.; Blokhuis, B.R.; Redegeld, F.A.; Hofstetter, G.; Dvorak, Z.; Jensen-Jarolim, E.; et al. Binding to Iron Quercetin Complexes Increases the Antioxidant Capacity of the Major Birch Pollen Allergen Bet v 1 and Reduces Its Allergenicity. Antioxidants 2022, 12, 42. [Google Scholar] [CrossRef]
  150. Besbes, F.; Franz-Oberdorf, K.; Schwab, W. Phosphorylation-dependent ribonuclease activity of Fra a 1 proteins. J. Plant Physiol. 2019, 233, 1–11. [Google Scholar] [CrossRef]
  151. Wang, T.; Xie, M.; Hou, S.; Ma, J.; Lin, Y.; Chen, S.; Li, D.; Yang, G. Walnut PR10/Bet v1-like proteins interact with chitinase in response to anthracnose stress. J. Evol. Biol. 2025, 38, 391–403. [Google Scholar] [CrossRef] [PubMed]
  152. Cao, J.; Maitirouzi, A.; Feng, Y.; Zhang, H.; Heng, Y.; Zhang, J.; Wang, Y. Heterologous expression of Halostachys caspica pathogenesis-related protein 10 increases salt and drought resistance in transgenic Arabidopsis thaliana. Plant Mol. Biol. 2024, 115, 5. [Google Scholar] [CrossRef]
  153. Jomová, K.; Feszterová, M.; Morovič, M. Expression of pathogenesis-related protein genes and changes of superoxide dismutase activity induced by toxic elements in Lupinus luteus L. J. Microb. Biotec. Food 2011, 1, 437–445. [Google Scholar]
  154. Pinto, M.P.; Ricardo, C.P. Lupinus albus L. pathogenesis-related proteins that show similarity to PR-10 proteins. Plant Physiol. 1995, 109, 1345–1351. [Google Scholar] [CrossRef]
  155. Han, Z.; Schneiter, R. Dual functionality of pathogenesis-related proteins: Defensive role in plants versus immunosuppressive role in pathogens. Front. Plant Sci. 2024, 15, 1368467. [Google Scholar] [CrossRef]
  156. Heitz, T.; Segond, S.; Kauffmann, S.; Geoffroy, P.; Prasad, V.; Brunner, F.; Fritig, B.; Legrand, M. Molecular characterization of a novel tobacco pathogenesis-related (PR) protein: A new plant chitinase/lysozyme. Mol. Gen. Genet. 1994, 245, 246–254. [Google Scholar] [CrossRef]
  157. Zribi, I.; Ghorbel, M.; Brini, F. Pathogenesis Related Proteins (PRs): From Cellular Mechanisms to Plant Defense. Curr. Protein Pept. Sci. 2021, 22, 396–412. [Google Scholar] [CrossRef] [PubMed]
  158. Scala, E.; Villalta, D.; Meneguzzi, G.; Giani, M.; Asero, R. Storage molecules from tree nuts, seeds and legumes: Relationships and amino acid identity among homologue molecules. Eur. Ann. Allergy Clin. Immunol. 2018, 50, 148–155. [Google Scholar] [CrossRef] [PubMed]
  159. Kumar, P.; Pandey, S.; Pati, P.K. Interaction between pathogenesis-related (PR) proteins and phytohormone signaling pathways in conferring disease tolerance in plants. Physiol. Plant 2025, 177, e70174. [Google Scholar] [CrossRef] [PubMed]
  160. Ali, S.; Ganai, B.A.; Kamili, A.N.; Bhat, A.A.; Mir, Z.A.; Bhat, J.A.; Tyagi, A.; Islam, S.T.; Mushtaq, M.; Yadav, P. Pathogenesis-related proteins and peptides as promising tools for engineering plants with multiple stress tolerance. Microbiol. Res. 2018, 212, 29–37. [Google Scholar] [CrossRef] [PubMed]
  161. Hawamda, A.I.; Reichert, S.; Ali, M.A.; Nawaz, M.A.; Austerlitz, T.; Schekahn, P.; Abbas, A.; Tenhaken, R.; Bohlmann, H. Characterization of an Arabidopsis defensin-like gene conferring resistance against nematodes. Plants 2022, 11, 280. [Google Scholar] [CrossRef] [PubMed]
  162. Singiri, J.R.; Swetha, B.; Sikron-Persi, N.; Grafi, G. Differential response to single and combined salt and heat stresses: Impact on accumulation of proteins and metabolites in dead pericarps of Brassica juncea. Int. J. Mol. Sci. 2021, 22, 7076. [Google Scholar] [CrossRef]
  163. Srivastava, S.; Dashora, K.; Ameta, K.L.; Singh, N.P.; El-Enshasy, H.A.; Pagano, M.C.; Hesham, A.E.; Sharma, G.D.; Sharma, M.; Bhargava, A. Cysteine-rich antimicrobial peptides from plants: The future of antimicrobial therapy. Phytother. Res. 2021, 35, 256–277. [Google Scholar] [CrossRef]
  164. Rayapuram, C.; Wu, J.; Haas, C.; Baldwin, I.T. PR-13/Thionin but not PR-1 mediates bacterial resistance in Nicotiana attenuata in nature, and neither influences herbivore resistance. Mol. Plant-Microbe Interact. 2008, 21, 988–1000. [Google Scholar] [CrossRef] [PubMed]
  165. Gaudet, D.A.; Laroche, A.; Frick, M.; Huel, R.; Puchalski, B. Cold induced expression of plant defensin and lipid transfer protein transcripts in winter wheat. Physiol. Plant. 2003, 117, 195–205. [Google Scholar] [CrossRef]
  166. Finkina, E.I.; Melnikova, D.N.; Bogdanov, I.V.; Ovchinnikova, T.V. Plant Pathogenesis-Related Proteins PR-10 and PR-14 as Components of Innate Immunity System and Ubiquitous Allergens. Curr. Med. Chem. 2017, 24, 1772–1787. [Google Scholar] [CrossRef]
  167. Bao, Y.; Pu, Y.; Yu, X.; Gregory, B.D.; Srivastava, R.; Howell, S.H.; Bassham, D.C. IRE1B degrades RNAs encoding proteins that interfere with the induction of autophagy by ER stress in Arabidopsis thaliana. Autophagy 2018, 14, 1562–1573. [Google Scholar] [CrossRef] [PubMed]
  168. Hu, J.; Jiang, J.; Wang, N. Control of Citrus Huanglongbing via Trunk Injection of Plant Defense Activators and Antibiotics. Phytopathology 2018, 108, 186–195. [Google Scholar] [CrossRef] [PubMed]
  169. Deng, X.; Liu, Y.; Xu, X.; Liu, D.; Zhu, G.; Yan, X.; Wang, Z.; Yan, Y. Comparative Proteome Analysis of Wheat Flag Leaves and Developing Grains Under Water Deficit. Front. Plant Sci. 2018, 9, 425. [Google Scholar] [CrossRef] [PubMed]
  170. Woo, E.J.; Dunwell, J.M.; Goodenough, P.W.; Marvier, A.C.; Pickersgill, R.W. Germin is a manganese containing homohexamer with oxalate oxidase and superoxide dismutase activities. Nat. Struct. Mol. Biol. 2000, 7, 1036–1040. [Google Scholar] [CrossRef]
  171. Park, C.J.; An, J.M.; Shin, Y.C.; Kim, K.J.; Lee, B.J.; Paek, K.H. Molecular characterization of pepper germin-like protein as the novel PR-16 family of pathogenesis-related proteins isolated during the resistance response to viral and bacterial infection. Planta 2004, 219, 797–806. [Google Scholar] [CrossRef]
  172. Edreva, A. Pathogenesis-related Proteins: Research Progress in the last 15 Years. Gen. Appl. Plant Physiol. 2005, 35, 105–124. [Google Scholar]
  173. Breen, J.; Bellgard, M. Germin-like proteins (GLPs) in cereal genomes: Gene clustering and dynamic roles in plant defence. Funct. Integr. Genom. 2010, 10, 463–476. [Google Scholar] [CrossRef]
  174. Zaynab, M.; Peng, J.; Sharif, Y.; Fatima, M.; Albaqami, M.; Al-Yahyai, R.; Khan, K.A.; Alotaibi, S.S.; Alaraidh, I.A.; Shaikhaldein, H.O. Genome-wide identification and expression profiling of germin-like proteins reveal their role in regulating abiotic stress response in potato. Front. Plant Sci. 2022, 12, 831140. [Google Scholar]
  175. Okushima, Y.; Koizumi, N.; Kusano, T.; Sano, H. Secreted proteins of tobacco cultured BY2 cells: Identification of a new member of pathogenesis-related proteins. Plant Mol. Biol. 2000, 42, 479–488. [Google Scholar] [CrossRef]
  176. Zhang, H.-M.; Zhu, J.-H.; Gong, Z.-Z.; Zhu, J.-K. Abiotic stress responses in plants. Nat. Rev. Genet. 2022, 23, 104–119. [Google Scholar] [CrossRef] [PubMed]
  177. Wang, J.; Mao, X.; Wang, R.; Li, A.; Zhao, G.; Zhao, J.; Jing, R. Identification of wheat stress-responding genes and TaPR-1-1 function by screening a cDNA yeast library prepared following abiotic stress. Sci. Rep. 2019, 9, 141. [Google Scholar] [CrossRef]
  178. Kosova, K.; Vitamvas, P.; Urban, M.O.; Klima, M.; Roy, A.; Prasil, I.T. Biological Networks Underlying Abiotic Stress Tolerance in Temperate Crops--A Proteomic Perspective. Int. J. Mol. Sci. 2015, 16, 20913–20942. [Google Scholar] [CrossRef]
  179. Wu, J.; Kim, S.G.; Kang, K.Y.; Kim, J.G.; Park, S.R.; Gupta, R.; Kim, Y.H.; Wang, Y.; Kim, S.T. Overexpression of a Pathogenesis-Related Protein 10 Enhances Biotic and Abiotic Stress Tolerance in Rice. Plant Pathol. J. 2016, 32, 552–562. [Google Scholar] [CrossRef]
  180. Wu, X.; Gong, F.; Cao, D.; Hu, X.; Wang, W. Advances in crop proteomics: PTMs of proteins under abiotic stress. Proteomics 2016, 16, 847–865. [Google Scholar] [CrossRef] [PubMed]
  181. Misra, R.C.; Sandeep; Kamthan, M.; Kumar, S.; Ghosh, S. A thaumatin-like protein of Ocimum basilicum confers tolerance to fungal pathogen and abiotic stress in transgenic Arabidopsis. Sci. Rep. 2016, 6, 25340. [Google Scholar] [CrossRef] [PubMed]
  182. Zheng, L.; Liu, Q.; Wu, R.; Songbuerbatu; Zhu, M.; Dorjee, T.; Zhou, Y.; Gao, F. The alteration of proteins and metabolites in leaf apoplast and the related gene expression associated with the adaptation of Ammopiptanthus mongolicus to winter freezing stress. Int. J. Biol. Macromol. 2023, 240, 124479. [Google Scholar] [CrossRef] [PubMed]
  183. Archambault, A.; Strömvik, M.V. PR-10, defensin and cold dehydrin genes are among those over expressed in Oxytropis (Fabaceae) species adapted to the arctic. Funct. Integr. Genom. 2011, 11, 497–505. [Google Scholar] [CrossRef]
  184. Alhaithloul, H.A.S. Impact of combined heat and drought stress on the potential growth responses of the desert grass Artemisia sieberi alba: Relation to biochemical and molecular adaptation. Plants 2019, 8, 416. [Google Scholar] [CrossRef] [PubMed]
  185. Carbonell-Bejerano, P.; Santa María, E.; Torres-Pérez, R.; Royo, C.; Lijavetzky, D.; Bravo, G.; Aguirreolea, J.; Sánchez-Díaz, M.; Antolín, M.C.; Martínez-Zapater, J.M. Thermotolerance Responses in Ripening Berries of Vitis vinifera L. cv Muscat Hamburg. Plant Cell Physiol. 2013, 54, 1200–1216. [Google Scholar] [CrossRef]
  186. Domingo, G.; Locato, V.; Cimini, S.; Ciceri, L.; Marsoni, M.; De Gara, L.; Bracale, M.; Vannini, C. A comprehensive characterization and expression profiling of defensin family peptides in Arabidopsis thaliana with a focus on their abiotic stress-specific transcriptional modulation. Curr. Plant Biol. 2024, 39, 100376. [Google Scholar] [CrossRef]
  187. Pós, V.; Hunyadi-Gulyás, É.; Caiazzo, R.; Jócsák, I.; Medzihradszky, K.; Lukács, N. Induction of pathogenesis-related proteins in intercellular fluid by cadmium stress in barley (Hordeum vulgare L.)—A proteomic analysis. Acta Aliment. 2011, 40, 164–175. [Google Scholar] [CrossRef]
  188. Dana, M.d.l.M.; Pintor-Toro, J.A.; Cubero, B. Transgenic tobacco plants overexpressing chitinases of fungal origin show enhanced resistance to biotic and abiotic stress agents. Plant Physiol. 2006, 142, 722–730. [Google Scholar] [CrossRef]
  189. Cho, U.H.; Park, J.O. Mercury-induced oxidative stress in tomato seedlings. Plant Sci. 2000, 156, 1–9. [Google Scholar] [CrossRef]
  190. Utriainen, M.; Kokko, H.; Auriola, S.; Sarrazin, O.; Kärenlampi, S. PR-10 protein is induced by copper stress in roots and leaves of a Cu/Zn tolerant clone of birch, Betula pendula. Plant Cell Environ. 1998, 21, 821–828. [Google Scholar] [CrossRef]
  191. Berna, A.; Bernier, F. Regulation by biotic and abiotic stress of a wheat germin gene encoding oxalate oxidase, a H2O2-producing enzyme. Plant Mol. Biol. 1999, 39, 539–549. [Google Scholar] [CrossRef]
  192. Ota, E.; Nishimura, F.; Mori, M.; Tanaka, M.; Kanto, T.; Hosokawa, M.; Osakabe, M.; Satou, M.; Takeshita, M. Up-regulation of pathogenesis-related genes in strawberry leaves treated with powdery mildew-suppressing ultraviolet irradiation. Plant Pathol. 2021, 70, 1378–1387. [Google Scholar] [CrossRef]
  193. Chen, H.; Wu, Q.; Ni, M.; Chen, C.; Han, C.; Yu, F. Transcriptome Analysis of Endogenous Hormone Response Mechanism in Roots of Styrax tonkinensis Under Waterlogging. Front. Plant Sci. 2022, 13, 896850. [Google Scholar] [CrossRef]
  194. Cid, G.A.; Francioli, D.; Kolb, S.; Moya, Y.A.T.; Wirén, N.V.; Hajirezaei, M.R. Transcriptomic and metabolomic approaches elucidate the systemic response of wheat plants under waterlogging. J. Exp. Bot. 2022, 75, 1510–1529. [Google Scholar] [CrossRef] [PubMed]
  195. Gedam, P.A.; Khandagale, K.; Shirsat, D.; Thangasamy, A.; Kulkarni, O.; Kulkarni, A.; Patil, S.S.; Barvkar, V.T.; Mahajan, V.; Gupta, A.J.; et al. Elucidating the molecular responses to waterlogging stress in onion (Allium cepa L.) leaf by comparative transcriptome profiling. Front. Plant Sci. 2023, 14, 1150909. [Google Scholar] [CrossRef]
  196. Kesarwani, M.; Yoo, J.; Dong, X. Genetic interactions of TGA transcription factors in the regulation of pathogenesis-related genes and disease resistance in Arabidopsis. Plant Physiol. 2007, 144, 336–346. [Google Scholar] [CrossRef] [PubMed]
  197. Delessert, C.; Kazan, K.; Wilson, I.W.; Van Der Straeten, D.; Manners, J.; Dennis, E.S.; Dolferus, R. The transcription factor ATAF2 represses the expression of pathogenesis-related genes in Arabidopsis. Plant J. 2005, 43, 745–757. [Google Scholar] [CrossRef] [PubMed]
  198. Cheong, Y.H.; Moon, B.C.; Kim, J.K.; Kim, C.Y.; Kim, M.C.; Kim, I.H.; Park, C.Y.; Kim, J.C.; Park, B.O.; Koo, S.C.; et al. BWMK1, a rice mitogen-activated protein kinase, locates in the nucleus and mediates pathogenesis-related gene expression by activation of a transcription factor. Plant Physiol. 2003, 132, 1961–1972. [Google Scholar] [CrossRef] [PubMed]
  199. Kumari, D.; Prasad, B.D.; Dwivedi, P.; Sahni, S.; Kumar, M.; Alamri, S.; Adil, M.F.; Alakeel, K.A. Comprehensive analysis of transcription factor binding sites and expression profiling of rice pathogenesis related genes (OsPR1). Front. Plant Sci. 2024, 15, 1463147. [Google Scholar] [CrossRef] [PubMed]
  200. Liu, T.; Chen, T.; Kan, J.; Yao, Y.; Guo, D.; Yang, Y.; Ling, X.; Wang, J.; Zhang, B. The GhMYB36 transcription factor confers resistance to biotic and abiotic stress by enhancing PR1 gene expression in plants. Plant Biotechnol. J. 2022, 20, 722–735. [Google Scholar] [CrossRef]
  201. Ito, Y.; Katsura, K.; Maruyama, K.; Taji, T.; Kobayashi, M.; Seki, M.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Functional analysis of rice DREB1/CBF-type transcription factors involved in cold-responsive gene expression in transgenic rice. Plant Cell Physiol. 2006, 47, 141–153. [Google Scholar] [CrossRef] [PubMed]
  202. Liu, Q.; Yan, H.; Zhang, Z.; Zheng, L.; Zhou, Y.; Gao, F. AmDEF2. 7, a tandem duplicated defensin gene from Ammopiptanthus mongolicus, activated by AmWRKY14, enhances the tolerance of Arabidopsis to low temperature and osmotic stress. Environ. Exp. Bot. 2024, 227, 105956. [Google Scholar] [CrossRef]
  203. Zheng, L.; Li, B.; Zhou, Y.; Gao, F. A WRKY-regulated TLP gene mediates the response to cold, drought, and wound stress in jojoba. Ind. Crop Prod. 2024, 220, 119224. [Google Scholar] [CrossRef]
  204. Zheng, L.-M.; Li, B.-J.; Liu, Q.; Richardson, J.E.; Zhou, Y.-J.; Gao, F. Proteomic analysis of leaf apoplast reveals that a jasmonate-regulated CHIA participates in the response to cold and drought stress in jojoba. Hortic. Plant J. 2025, early view, online version. [Google Scholar] [CrossRef]
  205. Hu, S.; Shinwari, K.I.; Song, Y.; Xia, J.; Xu, H.; Du, B.; Luo, L.; Zheng, L. OsNAC300 positively regulates cadmium stress responses and tolerance in rice roots. Agronomy 2021, 11, 95. [Google Scholar] [CrossRef]
  206. Zhao, Y.; Gong, J.; Shi, R.; Wu, Z.; Liu, S.; Chen, S.; Tao, Y.; Li, S.; Tian, J. Application of proteomics in investigating the responses of plant to abiotic stresses. Planta 2025, 261, 128. [Google Scholar] [CrossRef]
  207. Lu, S.; Faris, J.D.; Edwards, M.C. Molecular cloning and characterization of two novel genes from hexaploid wheat that encode double PR-1 domains coupled with a receptor-like protein kinase. Mol. Genet. Genom. 2017, 292, 435–452. [Google Scholar] [CrossRef] [PubMed]
  208. Zhang, H.; Shi, W.L.; You, J.F.; Bian, M.D.; Qin, X.M.; Yu, H.; Liu, Q.; Ryan, P.R.; Yang, Z.M. Transgenic Arabidopsis thaliana plants expressing a β-1,3-glucanase from sweet sorghum (Sorghum bicolor L.) show reduced callose deposition and increased tolerance to aluminium toxicity. Plant Cell Environ. 2015, 38, 1178–1188. [Google Scholar] [CrossRef] [PubMed]
  209. Sanchez-Ballesta, M.T.; Gosalbes, M.J.; Rodrigo, M.J.; Granell, A.; Zacarias, L.; Lafuente, M.T. Characterization of a β-1,3-glucanase from citrus fruit as related to chilling-induced injury and ethylene production. Postharvest. Biol. Technol. 2006, 40, 133–140. [Google Scholar] [CrossRef]
  210. Wan, K.; Buitrago, S.; Cheng, B.; Zhang, W.; Pan, R. Analysis of chitinase gene family in barley and function study of HvChi22 involved in drought tolerance. Mol. Biol. Rep. 2024, 51, 731. [Google Scholar] [CrossRef]
  211. Mandal, A.B.; Biswas, A.; Mukherjee, P.; Mondal, R.; Mandal, C. Osmotin: A PR gene imparts tolerance to excess salt in indica rice. Adv. Life Sci. 2018, 8, 39–50. [Google Scholar]
  212. Husaini, A.M.; Abdin, M.Z. Overexpression of tobacco osmotin gene leads to salt stress tolerance in strawberry (Fragaria x ananassa Duch.) plants. Indian J. Biotechnol. 2008, 7, 465–471. [Google Scholar]
  213. Takemoto, D.; Furuse, K.; Doke, N.; Kawakita, K. Identification of chitinase and osmotin-like protein as actin-binding proteins in suspension-cultured potato cells. Plant Cell Physiol. 1997, 38, 441–448. [Google Scholar] [CrossRef] [PubMed]
  214. Abad, L.R.; D’Urzo, M.P.; Liu, D.; Narasimhan, M.L.; Reuveni, M.; Zhu, J.K.; Niu, X.; Singh, N.K.; Hasegawa, P.M.; Bressan, R.A. Antifungal activity of tobacco osmotin has specificity and involves plasma membrane permeabilization. Plant Sci. 1996, 118, 11–23. [Google Scholar] [CrossRef]
  215. Priya, M.; Dhanker, O.P.; Siddique, K.H.; HanumanthaRao, B.; Nair, R.M.; Pandey, S.; Singh, S.; Varshney, R.K.; Prasad, P.V.; Nayyar, H. Drought and heat stress-related proteins: An update about their functional relevance in imparting stress tolerance in agricultural crops. Theor. Appl. Genet. 2019, 132, 1607–1638. [Google Scholar] [CrossRef]
  216. Wang, X.; Zafian, P.; Choudhary, M.; Lawton, M. The PR5K receptor protein kinase from Arabidopsis thaliana is structurally related to a family of plant defense proteins. Proc. Natl. Acad. Sci. USA 1996, 93, 2598–2602. [Google Scholar] [CrossRef]
  217. Baek, D.; Kim, M.C.; Kumar, D.; Park, B.; Cheong, M.S.; Choi, W.; Park, H.C.; Chun, H.J.; Park, H.J.; Lee, S.Y.; et al. AtPR5K2, a PR5-Like Receptor Kinase, Modulates Plant Responses to Drought Stress by Phosphorylating Protein Phosphatase 2Cs. Front. Plant Sci. 2019, 10, 1146. [Google Scholar] [CrossRef] [PubMed]
  218. Barthakur, S.; Babu, V.; Bansa, K. Over-expression of osmotin induces proline accumulation and confers tolerance to osmotic stress in transgenic tobacco. J. Plant Biochem. Biotechnol. 2001, 10, 31–37. [Google Scholar] [CrossRef]
  219. Hmida-Sayari, A.; Gargouri-Bouzid, R.; Bidani, A.; Jaoua, L.; Savouré, A.; Jaoua, S. Overexpression of Δ1-pyrroline-5-carboxylate synthetase increases proline production and confers salt tolerance in transgenic potato plants. Plant Sci. 2005, 169, 746–752. [Google Scholar] [CrossRef]
  220. Jung, H.W.; Kim, K.D.; Hwang, B.K. Identification of pathogen-responsive regions in the promoter of a pepper lipid transfer protein gene (CALTPI) and the enhanced resistance of the CALTPI transgenic Arabidopsis against pathogen and environmental stresses. Planta 2005, 221, 361–373. [Google Scholar] [CrossRef]
  221. Cui, S.; Takeda, K.Y.; Wakatake, T.; Luo, J.; Tobimatsu, Y.; Yoshida, S. Striga hermonthica induces lignin deposition at the root tip to facilitate prehaustorium formation and obligate parasitism. Plant Commun. 2025, 6, 101294. [Google Scholar] [CrossRef] [PubMed]
  222. Yamahara, T.; Shiono, T.; Suzuki, T.; Tanaka, K.; Takio, S.; Sato, K.; Yamazaki, S.; Satoh, T. Isolation of a germin-like protein with manganese superoxide dismutase activity from cells of a moss, Barbula unguiculata. J. Biol. Chem. 1999, 274, 33274–33278. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Types of PR proteins induced by different abiotic stresses. Purple italic font represents the PR protein genes; orange font represents the PR proteins.
Figure 1. Types of PR proteins induced by different abiotic stresses. Purple italic font represents the PR protein genes; orange font represents the PR proteins.
Biomolecules 15 01103 g001
Figure 2. Transcription factors directly regulate PR genes under abiotic stress. The text within green boxes denotes transcription factor families that regulate the expression of PR genes under abiotic stress conditions. The text in blue boxes represents PR genes. Yellow boxes contain sequences of cis-regulatory elements in the promoter regions of PR genes that are bound by transcription factors. Arrows indicate that the transcription factor family can bind to the corresponding cis-regulatory element of the PR genes.
Figure 2. Transcription factors directly regulate PR genes under abiotic stress. The text within green boxes denotes transcription factor families that regulate the expression of PR genes under abiotic stress conditions. The text in blue boxes represents PR genes. Yellow boxes contain sequences of cis-regulatory elements in the promoter regions of PR genes that are bound by transcription factors. Arrows indicate that the transcription factor family can bind to the corresponding cis-regulatory element of the PR genes.
Biomolecules 15 01103 g002
Figure 3. Mechanism of PR proteins enhancing plant abiotic stress tolerance. PR-1, PR-5, and PR-14 enhance plant tolerance to abiotic stress by directly binding to membrane lipids or proteins, thereby increasing membrane stability. PR-9 and PR-13 mitigate heavy metal stress through direct binding to metal ions. PR-2 and PR-5 contribute to abiotic stress tolerance through the formation of fusion proteins with receptor kinases. PR-1, chitinases (PR-3, 4, 8, 11), PR-5, PR-10, PR-15, and PR-16 confer indirect tolerance by enhancing the plant’s antioxidant capacity; the specific mechanism of PR-1, chitinases (PR-3, 4, 8, and 11), PR-5, and PR-10′s actions remain currently unclear. PR-1, PR-2, PR-5, and PR-16 reduce stomatal aperture to indirectly alleviate abiotic stress. PR-2, chitinases (PR-3, 4, 8, and 11), PR-5, PR-10, and PR-16 modulate osmotic potential through the accumulation of soluble compounds. PR-5, PR-14, and PR-15 fortify cell walls to indirectly enhance stress tolerance. Abbreviations: PR, pathogenesis-related protein; ROS, reactive oxygen species.
Figure 3. Mechanism of PR proteins enhancing plant abiotic stress tolerance. PR-1, PR-5, and PR-14 enhance plant tolerance to abiotic stress by directly binding to membrane lipids or proteins, thereby increasing membrane stability. PR-9 and PR-13 mitigate heavy metal stress through direct binding to metal ions. PR-2 and PR-5 contribute to abiotic stress tolerance through the formation of fusion proteins with receptor kinases. PR-1, chitinases (PR-3, 4, 8, 11), PR-5, PR-10, PR-15, and PR-16 confer indirect tolerance by enhancing the plant’s antioxidant capacity; the specific mechanism of PR-1, chitinases (PR-3, 4, 8, and 11), PR-5, and PR-10′s actions remain currently unclear. PR-1, PR-2, PR-5, and PR-16 reduce stomatal aperture to indirectly alleviate abiotic stress. PR-2, chitinases (PR-3, 4, 8, and 11), PR-5, PR-10, and PR-16 modulate osmotic potential through the accumulation of soluble compounds. PR-5, PR-14, and PR-15 fortify cell walls to indirectly enhance stress tolerance. Abbreviations: PR, pathogenesis-related protein; ROS, reactive oxygen species.
Biomolecules 15 01103 g003
Table 1. Classification, molecular function, and stress response of PR proteins.
Table 1. Classification, molecular function, and stress response of PR proteins.
PropertyMolecular FunctionBiotic StressAbiotic Stress
PR-1AntifungalAntimicrobial activity [96]
Pathogen toxin degradation
Viral coat protein-receptor binding inhibition [8]
Recruitment of PR proteins (PR-5/PR-14) [97]
Sterol binding [98]
Bacteria [99]
Fungi [100]
Viruses [8]
Oomycetes
Insects
Nematodes [101]
Drought [12]
High salinity [9]
Low temperature [102]
High temperature [103]
Heavy metals [104]
UV radiation [105]
PR-2β-1,3-glucanaseMicrobial growth suppression [106]
Fungal cell wall degradation [14]
Callose deposition modulation [107]
Bacteria [108]
Fungi [14]
Oomycetes [109]
Viruses [110,111]
Nematodes [112]
Drought [18]
High salinity [21]
Low temperature [113]
Heavy metals [114]
UV radiation [115]
PR-3Endochitinase
(Classes I, II, IV, V, VI, VII)
Chitinase activity [116]
Fungal cell wall decomposition [117]
Fungi [118]
Viruses [119]
Nematodes [112]
Drought [59]
High salinity [120]
High temperature [36]
Heavy metals [121]
PR-4Endochitinase
(Classes I, II)
Spore germination inhibition
Hyphal growth suppression
Cell wall degradation assistance
Ribonuclease/DNase activity [28,122]
Fungi [123]
Insect and pests [28]
Drought [124]
High salinity [23]
Low temperature [125]
Heavy metals [24]
PR-5Thaumatin-like proteinMicrobial growth inhibition [106]
Fungal membrane permeabilization
Membrane potential dissipation [126]
Fruit ripening promotion [127]
Fungi [128]
Viruses [129,130]
Bacteria [131]
Nematodes [112]
Drought [132]
High salinity [133]
Osmotic stress [46]
Low temperature [134]
High temperature [135]
Heavy metals [136]
UV radiation [115]
PR-6Protease inhibitorPathogen protease inhibition
Host cell degradation prevention [137]
Bacteria,
Fungi [137]
Low temperature [53]
PR-7Alkaline endoproteaseFungal structural proteins degradation
Cell wall integrity impairment [138]
Fungi [58]
Bacteria [59]
Viruses [139]
Nematodes [140]
-
PR-8Class III chitinaseBacterial cell wall hydrolysis [123]Fungi [141]Drought [33]
PR-9PeroxidaseROS concentration modulation
Lignin-mediated cell wall reinforcement
Cytotoxic radical production [142]
Fungi [143]
Bacteria [144]
Viruses [64]
Drought [145]
High salinity [66]
Heavy metals [65]
PR-10Ribonuclease-like proteinsRibonuclease activity [70]
Small hydrophobic ligand binding [146]
Fungal/bacterial growth inhibition [147]
Secondary metabolite biosynthesis regulation [148]
Stress response participation
Iron chelation [149]
Growth/development promotion [150]
Viruses
Bacteria [151]
Fungi
Nematodes
Insects [147]
High salinity
Drought [152]
Heavy metals [153]
Low temperature [70]
UV radiation [154]
PR-11Class I chitinaseβ-1,4-chitin glycosidic bond hydrolysis
Hyphal structure disruption
Spore germination inhibition [155]
Pathogen defense enhancement [123]
Bacteria [59]
Viruses [156]
Drought [59]
PR-12Plant defensinAntimicrobial activity
Systemic defense potentiation [157]
Human IgE-mediated allergenicity [158]
Membrane pore formation
Pathogen protease inhibition [159]
Fungi [160]
Viruses [64]
Nematodes [161]
Drought [77]
High-salt/high-temperature stress [162]
Low temperature [76]
Heavy metals [78]
PR-13ThioninPhospholipid binding
Transmembrane pore formation
Ion leakage induction [163]
Damage-associated molecular pattern function [1]
Bacteria [164]
Fungi [80]
Viruses [64]
Drought [81]
High-salt/high-temperature stress [162]
Low temperature [165]
Heavy metals [82]
PR-14Lipid transfer proteinsLipid binding/transport
Membrane biosynthesis participation
Pathogen membrane disruption [166]
Lipid signaling modulation [166,167]
SAR network synergy [81]
Bacteria
Fungi [84]
Viruses [64]
Drought
High salinity
Low temperature [166]
PR-15Oxalate oxidaseROS-controlled antimicrobial activity [168]Fungi [88]
Bacteria [59]
Viruses
Insect and pests [90]
Drought [169]
Low temperature [91]
Heavy metals [86]
High temperature
High salinity
Waterlogging [170]
PR-16Oxalate oxidase-likeFungal toxin inhibition [171]
SOD-mediated superoxide scavenging
Defense gene activation [92,172]
H2O2-dependent cell wall fortification [173]
ABA sensitivity enhancement [92]
Pathogen wall hydrolysis [86]
Fungi [88]
Bacteria
Viruses [59]
Drought [59]
High salinity [174]
PR-17Antifungal and antiviralExtracellular protease activity [93]Viruses
Fungi [94]
Bacteria [59]
Drought [175]
Table 2. PR Families with enzymatic activity.
Table 2. PR Families with enzymatic activity.
PR FamilyEnzymatic Activity TypeBiological Functions
PR-2β-1,3-GlucanaseHydrolysis of β-1,3-glucans in fungal cell walls
Release of elicitors
Carbohydrate metabolism
Osmotic adjustment
PR-3, 4, 8, 11ChitinaseDegradation of chitin (fungal cell walls/insect exoskeletons)
Partial lysozyme activity
Antioxidant response
COS generation
PR-9PeroxidaseH2O2-dependent oxidative burst catalysis
Lignification promotion
ROS scavenging
Phenolic compound metabolism
PR-10RibonucleaseRNA degradation
Signal transduction
Phytohormone binding
Modulation of antioxidant enzyme activities
PR-15Oxalate oxidaseOxalate degradation with H2O2 generation
Cell wall lignification
ROS signaling
PR-16Oxalate oxidase-like proteinSOD activity
ROS scavenging
Ion homeostasis maintenance
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

Zhu, Y.; Gao, F. Involvement of Pathogenesis-Related Proteins and Their Roles in Abiotic Stress Responses in Plants. Biomolecules 2025, 15, 1103. https://doi.org/10.3390/biom15081103

AMA Style

Zhu Y, Gao F. Involvement of Pathogenesis-Related Proteins and Their Roles in Abiotic Stress Responses in Plants. Biomolecules. 2025; 15(8):1103. https://doi.org/10.3390/biom15081103

Chicago/Turabian Style

Zhu, Yilin, and Fei Gao. 2025. "Involvement of Pathogenesis-Related Proteins and Their Roles in Abiotic Stress Responses in Plants" Biomolecules 15, no. 8: 1103. https://doi.org/10.3390/biom15081103

APA Style

Zhu, Y., & Gao, F. (2025). Involvement of Pathogenesis-Related Proteins and Their Roles in Abiotic Stress Responses in Plants. Biomolecules, 15(8), 1103. https://doi.org/10.3390/biom15081103

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