Next Article in Journal
Toxicity of Cry- and Vip3Aa-Class Proteins and Their Interactions against Spodoptera frugiperda (Lepidoptera: Noctuidae)
Previous Article in Journal
Botulinum Toxin Injections for Psychiatric Disorders: A Systematic Review of the Clinical Trial Landscape
Previous Article in Special Issue
Edodin: A New Type of Toxin from Shiitake Mushroom (Lentinula edodes) That Inactivates Mammalian Ribosomes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Antifungal Activity of Ribosome-Inactivating Proteins

by
Rosario Iglesias
1,†,
Lucía Citores
1,†,
Claudia C. Gay
2 and
José M. Ferreras
1,*
1
Department of Biochemistry and Molecular Biology and Physiology, Faculty of Sciences, University of Valladolid, E-47011 Valladolid, Spain
2
Laboratory of Protein Research, Institute of Basic and Applied Chemistry of Northeast Argentina (UNNE-CONICET), Faculty of Exact and Natural Sciences and Surveying, Av. Libertad 5470, Corrientes 3400, Argentina
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Toxins 2024, 16(4), 192; https://doi.org/10.3390/toxins16040192
Submission received: 14 March 2024 / Revised: 4 April 2024 / Accepted: 12 April 2024 / Published: 15 April 2024
(This article belongs to the Special Issue Biological Activities of Ribosome Inactivating Proteins II)

Abstract

:
The control of crop diseases caused by fungi remains a major problem and there is a need to find effective fungicides that are environmentally friendly. Plants are an excellent source for this purpose because they have developed defense mechanisms to cope with fungal infections. Among the plant proteins that play a role in defense are ribosome-inactivating proteins (RIPs), enzymes obtained mainly from angiosperms that, in addition to inactivating ribosomes, have been studied as antiviral, fungicidal, and insecticidal proteins. In this review, we summarize and discuss the potential use of RIPs (and other proteins with similar activity) as antifungal agents, with special emphasis on RIP/fungus specificity, possible mechanisms of antifungal action, and the use of RIP genes to obtain fungus-resistant transgenic plants. It also highlights the fact that these proteins also have antiviral and insecticidal activity, which makes them very versatile tools for crop protection.
Key Contribution: Ribosome-inactivating proteins (RIPs) can be very useful in the fight against fungal crop diseases.

1. Introduction

Diseases caused by plant pathogens are continuously increasing, causing severe losses in agricultural production, as disease prevalence can reach 70–80% of the total plant population and yields can decrease in some cases up to 80–98% [1]. The main phytopathogens are viruses, bacteria, and fungi [1]. Fungi are responsible for 80% of plant diseases [2] and fungal epidemics have had significant social and economic repercussions throughout history and today [3]. To control these diseases, chemical-based fungicides are used, which are very effective but bring with them problems such as environmental contamination, development of resistance, and residual toxicity [2,4,5]. Therefore, the fight against fungal diseases remains a major challenge and there is a need to find effective fungicides that are environmentally friendly. In this context, the search for more effective and safer fungicides continues to be a field of intense research. Plants are one of the most widely used sources, as they have developed various protein-based defense mechanisms to cope with fungal infections. However, the control of crop diseases using this type of fungicides has drawbacks, such as the instability of many of these agents in the field [6] and their high cost of production [2]. A solution to these drawbacks may be the use of transgenic plants carrying genes that code for antifungal proteins. Antifungal proteins include chitinases, glucanases, thaumatin-like proteins, thionins, cyclophilin-like proteins, lectins, ribonucleases, deoxyribonucleases, peroxidases, protease inhibitors, and ribosome-inactivating proteins [7,8]. Ribosome-inactivating proteins (RIPs) may be an excellent choice, as they have been attributed a defense role in plants and have been shown to have great potential against viruses, fungi, and insects [9,10,11,12].
In this review, we summarize and discuss the potential use of ribosome-inactivating proteins (RIPs) for agricultural applications from a bioengineering and biotechnology perspective, with special emphasis on RIP/fungus specificity, possible mechanisms of antifungal action, and the use of RIP genes to obtain transgenic plants resistant to fungi.

2. Ribosome-Inactivating Proteins

Ribosome-inactivating proteins (RIPs) are a group of proteins that inactivate ribosomes, leading to irreversible inhibition of protein synthesis (with IC50 values, i.e., concentration inhibiting protein synthesis by 50%, for animal cell-free systems in the range of 0.015 to 3.5 nM) and, consequently, cell death [13,14,15,16,17]. RIPs have been classified according to their structure into type 1 RIPs, consisting of a polypeptide chain with enzymatic activity, and type 2 RIPs, made up of two polypeptide chains, an A chain with enzymatic activity and a B chain with lectin activity that can bind to cell surface receptors facilitating RIP entry [13]. In addition, a third class of RIPs, termed type 3 RIPs, has been recognized, which includes a few members, such as jasmonate-induced protein (JIP60) and maize b-32 protein, which are activated by proteolysis [13,14,16,17]. Some type 2 RIPs such as ricin and abrin are extremely toxic (with IC50 values for cell cultures between 0.3 and 8 pM), while others have low toxicity, because the binding of the B-chain to oligosaccharides on the cell surface is less efficient and because, once internalized, the RIP follows a different intracellular pathway than ricin [16]. The toxicity of type 1 RIPs is lower (with IC50 values for cell cultures between 2 and 34 µM) since they lack the lectin part and are therefore unable to bind to cells as type 2 RIPs do. The structure, activity, and mode of action of RIPs have been studied over the last decades, but their biological function has not been demonstrated, although there is a broad consensus that these proteins play an important role in the defense of plants against viruses, fungi, and insects [9,12,18].
Due to their diverse activities, RIPs, alone or as part of conjugates, are good candidates for developing selective antiviral and anticancer agents [12,19,20,21,22]. Conjugates consist of a targeting moiety, such as an antibody, lectin, or growth factor, linked to a toxic moiety. RIPs have been used as the toxic part in several conjugates that have been tested in experimental therapies against various malignancies. In agriculture, RIPs have been shown to increase resistance against viruses, fungi, and insects in transgenic plants [9,12,21].
RIPs are present in a large number of angiosperm plants, both monocotyledonous and dicotyledonous, although in some plant families it is more common to find RIPs than in others, thus there are families such as the Poaceae, Euphorbiaceae, Cucurbitaceae, Caryophyllaceae, Amaranthaceae, and Phytolacaceae, where several species have been found with RIPs, and other families where they have never been found [13,14,23]. Some bacteria possess toxins with rRNA N-glycosylase activity whose sequence is homologous to plant RIPs [24]. The best known are Shiga toxin and related proteins consisting of an A chain with N-glycosylase activity and a B subunit, which is a pentamer that binds to specific glycolipids of the plasma membrane facilitating their endocytosis [25].
Although RIPs were initially studied as inhibitors of mammalian ribosomes, they can also inactivate ribosomes from other animals [26] and fungi [27,28,29,30], and, in some cases, ribosomes from bacteria [9] and plants [31]. The ability to inactivate plant ribosomes is, as we will discuss later, of particular significance in the defense against pathogens. Table 1 identifies the RIPs that inactivate or do not inactivate ribosomes of various plant species.
The inhibitory activity of RIPs on plant ribosomes is very diverse. It appears that RIPs from Poaceae do not inhibit protein synthesis in plants, or, if they do (OsRIP1 in the germ system of wheat or tritin-L in wheat and tobacco), it is at very high concentrations. Neither type 2 RIPs from Abrus precatorius L. or Viscum album L., nor RIPs from the genus Sambucus, whether type 1 or 2, inhibit protein synthesis in plants. Type 2 RIPs from Ricinus comunis L. (ricin and RCA) inhibit protein synthesis in different plant germ systems but do so at very high concentrations. The case of cucurbits is inconclusive because only bryodin has been tested in the germ-derived system of Cucumis sativus L. However, type 1 RIPs from the Phytolaccaceae, Amaranthaceae, and Caryophyllaceae families inhibit protein synthesis in practically all systems in which they have been tested in the nM range, a concentration which, given the reported yield for purification of many RIPs, could easily be achieved in vivo. The reason for not inactivating their ribosomes is that RIPs of dicots, like those of monocots of the family Asparagaceae, have a leader peptide, are synthesized in the endoplasmic reticulum and exported to the apoplast, thus avoiding contact with ribosomes [32]. In this regard, it should be noted that the RIPs from Phytolacca americana L. and Phytolacca dodecandra L’Hér. have been tested on their ribosomes, which are sensitive to the toxins [33].
Table 1. Sensitivity of plant ribosomes to ribosome-inactivating proteins. The families and species of both the RIP source (rows) and the ribosome source (columns) are indicated.
Table 1. Sensitivity of plant ribosomes to ribosome-inactivating proteins. The families and species of both the RIP source (rows) and the ribosome source (columns) are indicated.
RIPIC50 * (nM)References
CucurbitaceaeBrassicaceaEuphorbiaceaeFabaceaePhytolaccaceaePoaceaeSolanaceae
AMARANTHACEAE
Beta vulgaris L.
BE27 Yes ** (VS) [30,34]
ASPARAGACEAE
Agave tequilana F.A.C.Weber
Mayahuelin 10.43 (TA) [35]
Asparagus officinalis L.
Asparin 11333 (CM) No * (VS) [36,37]
Muscari armeniacum H.J.Veitch
Musarmins 1-2-3 No * (VS) No * (TA) [38]
CARYOPHYLLACEAE
Dianthus caryophyllus L.
Dianthin 30No * (CM) [37]
Dianthin 32 Yes ** (NT)[39]
Saponaria officinalis L.
Saporin-L126.7 (CM)
17.3 (CS)
0.99 (VS) 40 (TA) [37,40]
Saporin-L231 (CS) 20.91 (VS) 23.7 (TA) [40]
Saporin-R11105 (CS) 0.22 (VS) 582 (TA) [40]
Saporin-R255.7 (CS) 0.97 (VS) 3.1 (TA) [40]
Saporin-R3959 (CS) 0.02 (VS) 176 (TA) [40]
Saporin-S510 (CM)
0.03 (CS)
0.34–0.48 (VS) 772 (TA) [31,37,40,41]
Saporin-S63606 (CS) 0.31 (VS) 32 (TA) [40]
Silene glaucifolia Lag. (=Petrocoptis glaucifolia Boiss.)
Petroglaucin 1219 (CS) 49 (VS) 30 (TA) [42]
Petroglaucin 227–29 (CS) 0.2–6 (VS) 127 (VL) 30–73 (TA) [36,42,43]
Silene laxipruinosa Mayol and Rosselló (=Petrocoptis grandiflora Rothm.)
Petrograndin186 (CS) 5 (VS) 100(TA) [42]
CUCURBITACEAE
Bryonia dioica Sessé and Moc.
BryodinNo * (CS) [41]
EUPHORBIACEAE
Ricinus communis L.
Ricin1473 (CL) 1470 (RC)923 (PS)1700 (PA)1313 (TA) 980 (HV)Yes ** (NT)[33,39,44]
RCANo * (CM)
3767 (CL)
8167 (RC)No * (VS) 7500 (PS) 19,333 (TA) 7567 (HV) [44,45]
FABACEAE
Abrus precatorius L.
APANo * (CM) No * (VS) No * (TA) [45]
PHYTOLACCACEAE
Phytolacca americana L
PAP (PAP I) 1.1–2.9 (PA)0.3 (TA)Yes ** (NT)[33,39]
PAP II 3.9 (PA) [33]
PAP-S54–500 (CS) 0.26–0.38 (VS)6.7 (PA)4.5 (TA) [31,33,36,41,46]
Phytolacca dioica L.
PD L4-S2, Dioicin 2 Yes ** (VS) [30]
Phytolacca dodecandra L’Hér.
Dodecandrin 0.8–3.1 (PD)0.2 (TA) [33]
POACEAE
Hordeum vulgare L.
Barley RIP 30 No ** (NT)[39]
Oryza sativa L.
OsRIP1 1500 (TA) [47]
Triticum aestivum L.
Tritin (Tritin-S) No ** (AT) No ** (LJ) No ** (TA)No ** (NT)[39,48]
Tritin-L Yes ** (AT) Yes ** (LJ) Yes ** (TA)Yes ** (NT)[48]
Zea mays L.
pro-RIP, αβ RIP No ** (ZM) [49]
SANTALACEAE
Viscum album L.
VAANo * (CM) No * (VS) No * (TA) [45]
VIBURNACEAE
Sambucus ebulus L.
Ebulin fNo * (CS) No * (VS) No * (TA) [50]
Ebulin r1–r2No * (CM) No * (VS) No * (TA) [50]
α-β-γ-EbulitinNo * (CM) No * (VS) No * (TA) [50]
Sambucus nigra L.
Nigrin bNo * (CS) No * (VS) No * (TA) [50]
Nigrin fNo * (CS) No * (VS) No * (TA) [50]
basic Nigrin b No * (TA) [50]
Nigritin f1–f2No * (CS) No * (VS) No * (TA) [50]
The table lists both protein synthesis inhibition assays (IC50 is indicated, i.e., concentration inhibiting protein synthesis by 50%) * and N-glycosylase activity assays ** on ribosomes of the following species: Arabidopsis thaliana (L.) Heynh (AT); Citrullus lanatus (Thunb.) Matsum. and Nakai (CL); Cucumis melo L. (CM); Cucumis sativus L. (CS); Hordeum vulgare L. (HV); Lotus japonicus (Regel) K. Larsen (LJ); Nicotiana tabacum L. (NT); Phytolacca americana L. (PA); Phytolacca dodecandra L’Hér. (PD); Pisum sativum L. (PS); Ricinus communis L. (RC); Triticum aestivum L. (TA); Vicia lens (L.) Coss. and Germ. (VL); Vicia sativa L. (VA); Zea mays L. (ZM).

3. Mechanism of Ribosome Inactivation by RIPs

The mechanism of ribosome inactivation by RIPs has been known since 1987 [51,52]. RIPs are 28S rRNA N-glycosylases (EC 3.2.2.22) that catalyze the hydrolysis of the N-glycosidic bond of adenosine 4324 in the sarcin-ricin loop (SRL) of the large RNA of the 60S subunit of rat ribosomes or the equivalent in sensitive ribosomes of other organisms [30]. The SRL is part of the GTPase-associated center (GAC), which is the landing platform for translational GTPases (trGTPases) such as the prokaryote elongation factors EF-Tu and EF-G, and their eukaryote counterparts eEF1A and eEF2 [53]. The GAC consists of the SRL and the ribosomal stalk. The ribosomal stalk consists of a base made up of two ribosomal proteins and the lateral elements that are made up of several copies of proteins. The SRL and the base of the ribosomal stalk are conserved in prokaryotes and eukaryotes, while the lateral proteins are not conserved and are precisely the docking points of trGTPases and RIPs [54,55]. This may be the basis for the different specificity of RIPs for ribosomes of different species. The removal of adenine from the SRL, which is essential for the binding of trGTPases and even appears to be involved in the catalysis process, irreversibly inactivates ribosomes and has been reported to prevent eEF2 binding and GTP hydrolysis in eukaryotes [54].
A similar effect is caused by ribotoxins, such as α-sarcin, which are a group of extracellular ribonucleases that show cytotoxic activity towards animal cells [56,57]. These proteins are highly specific rRNA endonucleases (EC 4.6.1.23) that catalyze the hydrolysis of the phosphodiester bond between guanosine 4325 and adenosine 4326 in the SRL of rat 28S rRNA [58] (or the equivalent phosphodiester bond in ribosomes of other organisms), which prevents the binding of elongation factors [59]. Ribotoxins are produced by a few species of ascomycetes, mostly from the genus Aspergillus [57]. Due to the translation inhibitory and apoptotic activities of ribotoxins, they have been used as components of immunotoxins [57].
Finally, it should be noted that some RIPs can also remove more than one adenine from rRNA and many of them can catalyze the deadenylation not only of rRNA but also of other polynucleotide substrates such as DNA, poly(A), mRNA, tRNA, and viral RNA, and because of this, the name adenine polynucleotide glycosylase (or polynucleotide: adenosine glycosidase) was proposed for RIPs [12,60]. In addition, other activities have been reported for RIPs that could play a role in their possible function as defense elements [12,17].

4. RIP-like Proteins and Ribotoxin-like Proteins

There are plant proteins that have rRNA N-glycosylase activity but do not show homology or structural similarity to type 1 RIPs [61], and have been classified under various names (e.g., “small RIPs”, “small RIP 1 candidates”, or “RIP-Like Proteins”). These proteins could therefore exhibit all or at least some of the biological properties of RIPs and could also be used as crop defense tools. However, it should be noted that, although they inhibit protein synthesis, the N-glycosylase assay has not been performed on all of them, so some may have a different enzymatic activity. In fungi, some proteins with rRNA N-glycosylase activity have also been found without homology or structural similarity to plant RIPs [62,63], some of which have shown antifungal activity.
Ribotoxins have been found exclusively in ascomycetes; however, recently, proteins with the same activity have been found in basidiomycetes, and because they are not homologous with ascomycete ribotoxins they have been named Ribotoxin-Like Proteins (RLPs) [64]. As will be discussed later, some of them have shown antifungal activity.

5. Endocytosis of Ribosome-Inactivating Proteins

A very important question is how RIPs enter cells to carry out their enzymatic activity on ribosomes, since cytotoxic activity depends more on their ability to access ribosomes than on their catalytic power [16,28]. RIPs must interact with the cell membrane, and, following initial internalization, they are transported within the cell to the particular membrane where toxin translocation to the cytosol occurs. In addition, to enter fungal cells, RIPs must pass through the fungal cell wall. The internalization routes of RIPs have been studied in animal cells, where it has been observed that they can follow different internalization routes. In this respect, the most studied is ricin, a type 2 RIP that is highly toxic to animal cells. At picomolar concentrations, it binds to plasma membrane glycoproteins and is internalized into the cell [16,65]. Some protein molecules are recycled back to the plasma membrane, others undergo degradation in lysosomes, and a small number are transported first to the Golgi network and then to the endoplasmic reticulum. In the endoplasmic reticulum, the disulfide bridge is reduced, and the A-chain is translocated to the cytosol via the endoplasmic reticulum-associated degradation (ERAD) pathway. Finally, in the cytosol, the A-chain inactivates ribosomes, leading to cell death. Nigrin b and other type 2 RIPs from species of the genus Sambucus, which are thousands of times less toxic than ricin, can bind to plasma membrane glycoproteins other than ricin and internalize into the cell. All protein molecules are recycled back to the plasma membrane or transported to lysosomes for degradation. However, at a much higher extracellular concentration (40,000-fold), saturation of the endosome with nigrin b can lead to spontaneous release of nigrin b into the cytosol, causing inactivation of the ribosomes [16]. Type 1 RIPs such as saporin, trichosanthin, and curcins enter by binding to receptors of the LDLR (low-density lipoprotein receptor) family [66,67] and also follow different routes to those of ricin, most of them being located in endosomal compartments, which causes them to only reach the ribosomes at much higher concentrations than those required for ricin [66,68,69].
The internalization routes of RIPs have not been studied in fungi, but, given that they also possess endocytic mechanisms similar to those of animal cells [70,71], it would be expected that the routes of access to ribosomes would be similar to those discovered in animal cells. In addition, the cell wall represents an important barrier to the passage of macromolecules [72], which, given the diversity of the composition and structure of the cell walls of different fungi [73], could be an important element explaining, in part, the different sensitivities of fungi to different RIPs.

6. Inhibition of Fungal Growth by RIPs

At least 34 fungal species are sensitive to some RIPs or RIP-Like Proteins (Table 2). This includes a wide variety of species belonging to various families of basidiomycetes and ascomycetes. Furthermore, three species of fungus-like organisms are also sensitive to RIPs; thus, it has been reported that ME1 and ME2 from the roots of Mirabilis expansa (Ruiz and Pav.) Standl. slightly inhibit the growth of Globisporangium irregulare (Buisman) Uzuhashi, Tojo and Kakish, and Phytophthora drechsleri Tucker [27], and that the RIP isolated from the sarcocarp of Cucurbita moschata Duchesne strongly inhibits the growth of Phytophthora infestans (Mont.) de Bary [74]. Therefore, growth inhibition by the different RIPs could cover a spectrum that practically encompasses the entire fungal kingdom. The fungi on which RIPs have been tested are plant pathogens, although some of them, such as those of the genus Aspergillus, may be opportunistic pathogens in humans [75].
The antifungal activity of at least 20 RIPs obtained from 17 different species has been demonstrated (Table 3 and Table 4). This includes type 1 RIPs from two Poaceae (barley and maize) and several dicots and a type 2 RIP from the dicot Sambucus nigra L.
The antifungal activity of RIPs has been demonstrated both in vitro assays (Table 3) and, as discussed later, in transgenic plants (Table 4). Different types of in vitro assays have been used, thus, the RIP 30 from barley has been tested on microtiter plates and by different types of assay on agar plates. Sensitivity seems to be higher on microtiter plates than on agar plates [76]. Thus, inhibition of Trichoderma reesei E. G. Simmons growth was seen with a concentration of 11 μg/mL of RIP 30 on microtitre plates, whereas discs impregnated with 15 times more concentration are needed to obtain the same result [77]. In any case, RIP 30 was shown to inhibit the growth of several fungi: Rhizoctonia solani J.G. Kühn [76], T. reesei [76,77], Fusarium sporotrichioides Sherb. [76], and Botrytis cinerea Pers. [76]. However, about 16 fungal species have shown resistance to this RIP when tested on agar plates, including Phycomyces blakesleeanus Burgeff, Alternaria alternariae (Cooke) Woudenb. and Crous, and Neurospora crassa Shear and B.O. Dodge [77].
Table 2. Fungi that have been described as sensitive to RIPs or RIP-Like Proteins.
Table 2. Fungi that have been described as sensitive to RIPs or RIP-Like Proteins.
OrderFamilySpeciesReferences
Division Basidiomycota
CLASS AGARICOMYCETES
CantharellalesCeratobasidiaceaeRhizoctonia solani J.G. Kühn[28,76,78,79,80,81,82,83,84,85,86,87]
PolyporalesPolyporaceaeGanoderma boninense Pat.[88]
AgaricalesAgaricaceaeCoprinus comatus (O.F. Müll.) Pers.[89,90]
Division Ascomycota
CLASS LEOTIOMYCETES
HelotialesErysiphaceaeBlumeria graminis (DC.) Speer[91]
SclerotiniaceaeBotrytis cinerea Pers.[76,87,92]
Clarireedia homoeocarpa (F.T. Benn.) L.A. Beirn, B.B. Clarke, C. Salgado and J.A. Crouch[93]
Sclerotinia sclerotiorum (Lib.) de Bary[94]
CLASS SORDARIOMYCETES
AmphisphaerialesPestalotiopsidaceaePestalotia sp.[95]
DiaporthalesCryphonectriaceaeCryphonectria parasitica (Murrill) M.E. Barr[96]
ValsaceaeCytospora sp. *[95]
GlomerellalesPlectosphaerellaceaeVerticillium dahliae Kleb.[27]
HypocrealesHypocreaceaeTrichoderma reesei E.G. Simmons[27,95]
Trichoderma harzianum Rifai[27]
NectriaceaeFusarium culmorum (Wm.G. Sm.) Sacc.[97]
Fusarium fujikuroi Nirenberg[98,99]
Fusarium graminearum Schwabe[94]
Fusarium oxysporum Schltdl.[27,90,95,100,101,102]
F. proliferatum (Matsush.) Nirenberg ex Gerlach and Nirenberg[27]
Fusarium sporotrichioides Sherb.[76]
MagnaporthalesPyriculariaceaePyricularia grisea Cooke ex Sacc.[103]
Pyricularia oryzae Cavara[104,105]
SordarialesSordariaceaeNeurospora crassa Shear and B.O. Dodge **[77]
XylarialesHyponectriaceaePhysalospora pyricola Nose[89,90]
CLASS EUROTIOMYCETES
EurotialesAspergillaceaeAspergillus flavus Link[94,106]
Aspergillus nidulans (Eidam) G. Winter[106]
Aspergillus niger Tiegh.[94,107]
Penicillium digitatum (Pers.) Sacc.[29,30,108,109]
CLASS DOTHIDEOMYCETES
PleosporalesCorynesporascaceaeCorynespora cassiicola (Berk. and M.A. Curtis) C.T. Wei[110]
DidymellaceaeDidymella arachidicola (Khokhr.) Tomilin[90,101]
Phoma sp.[95]
PleosporaceaeAlternaria alternata (Fr.) Keissl.[92]
Alternaria brassicae (Berk.) Sacc.[111]
Alternaria solani Sorauer[27,102]
Cochliobolus heterostrophus (Drechsler) Drechsler[94]
* Sensitive only to RIP-Like Proteins; ** Mutant os-1. Synonyms used in the cited articles: Blumeria graminis (DC.) Speer (=Erysiphe graminis DC.); Fusarium fujikuroi Nirenberg (=Fusarium verticillioides (Sacc.) Nirenberg); Clarireedia homoeocarpa (F.T. Benn.) L.A. Beirn, B.B. Clarke, C. Salgado and J.A. Crouch (=Sclerotinia homoeocarpa F.T. Benn.); Pyricularia grisea Cooke ex Sacc. (=Magnaporthe grisea (T.T. Hebert) M.E. Barr); Cytospora sp. (Cytospora canker); Globisporangium irregulare (Buisman) Uzuhashi, Tojo and Kakish. (=Pythium irregulare Buisman); Fusarium oxysporum Schltdl. (=Fusarium oxysporum var. solani Raillo); Cochliobolus heterostrophus (Drechsler) Drechsler (=Bipolaris maydis (Y. Nisik. and C. Miyake) Shoemaker); Aspergillus flavus Link (=Aspergillus oryzae (Ahlb.) Cohn); Cucumis melo L. (=Luffa cylindrica M.Roem.); Didymella arachidicola (Khokhr.) Tomilin (=Mycosphaerella arachidicola Khokhr.).
Table 3. Ribosome-inactivating proteins that inhibit fungal growth in vitro. The RIPs, the families and species from which they have been obtained, and the fungi in which this activity has been demonstrated are shown.
Table 3. Ribosome-inactivating proteins that inhibit fungal growth in vitro. The RIPs, the families and species from which they have been obtained, and the fungi in which this activity has been demonstrated are shown.
Species and RIPFungiRef.
POACEAE
Hordeum vulgare L.
Barley RIP30Botrytis cinerea, Fusarium sporotrichioides, Neurospora crassa *, Rhizoctonia solani, Trichoderma reesei[76,77]
Zea mays L.
Maize b-32 (MOD1)Aspergillus flavus, A. nidulans, R. solani[80,106]
AMARANTHACEAE
Salsola soda L.
Sodin 5Penicillium digitatum[109]
Chenopodium quinoa Willd.
QuinoinCryphonectria parasitica, P. digitatum[96,109]
Beta vulgaris L.
BE27P. digitatum[29,108]
PHYTOLACCACEAE
Phytolacca dioica L.
Dioicin 2P. digitatum[30]
PD-S2P. digitatum[30]
Phytolacca heterotepala H.Walter
PhRIP IB. cinerea[92]
NYCTAGINACEAE
Mirabilis expansa (Ruiz and Pav.) Standl.
ME1 and ME2Alternaria solani, Fusarium oxysporum, F. proliferatum, Globisporangium irregulare, Phytophthora drechsleri, R. solani, Trichoderma harzianum, T. reesei, Verticillium dahliae[27,28]
CUCURBITACEAE
Momordica charantia L.
Alpha-momorcharin (α-MMC)A. flavus, A. niger, Cochliobolus heterostrophus, Fusarium graminearum, F. oxysporum, F. solani, Pyricularia oryzae, Sclerotinia sclerotiorum[94,100,105]
Momordica balsamina L.
MbRIP-1Aspergillus niger[107]
Benincasa hispida Cogn.
HispinCoprinus comatus, Didymella arachidicola, F. oxysporum, Physalospora pyricola[90]
SOLANACEAE
Nicotiana tabacum L.
TRIPC. heterostrophus, Cytospora sp., F. oxysporum, Pestalotia sp., Phoma sp., T. reesei[95]
ARECACEAE
Elaeis guineensis Jacq.
EgRIP-1a and EgRIP-1bGanoderma boninense[88]
VIBURNACEAE
Sambucus ebulus L.
PebulinA. solani, F. oxysporum[102]
* Mutant os-1.
Notably, the protoplast-forming mutant os-1 of N. crassa was sensitive to RIP 30, indicating that, at least in this organism, the presence of an intact cell wall protects against the antifungal activity of the RIP [77]. N. crassa ribosomes were approximately 10 times more sensitive to inactivation than ribosomes from ascites cells [77], indicating that it is not the sensitivity of the ribosomes to RIP that determines their toxicity but rather their ability to reach the ribosomes. The inhibition of T. reesei growth by RIP 30 was enhanced in the presence of both chitinase and barley β-1,3-glucanase, whereas with F. sporotrichioides it was only enhanced in the presence of chitinase [76]. This synergistic inhibition suggests that inhibition by RIP 30 is enhanced when hyphal cell walls are permeabilized by the action of these hydrolases.
Table 4. Transgenic fungus-resistant plants bearing RIP genes.
Table 4. Transgenic fungus-resistant plants bearing RIP genes.
RIPHostPathogenRef.
Barley RIP30Nicotiana tabacum L.Rhizoctonia solani[78,79]
Triticum aestivum L.Blumeria graminis[91]
Solanum tuberosum L.R. solani[86,112]
Brassica juncea (L.) Czern.Alternaria brassicae[111]
Vigna mungo (L.) HepperCorynespora cassiicola[110]
Maize b-32N. tabacumR. solani[80]
T. aestivum LFusarium culmorum[97]
Zea mays L.Fusarium fujikuroi[98]
MOD1Oryza sativa L.R. solani[84]
Z. maysF. fujikuroi[99]
PAP (PAP I)N. tabacumR. solani[81,83]
PAPIIN. tabacumR. solani[82]
Agrostis stolonifera L.Clarireedia homoeocarpa[93]
PhRIP IN. tabacumBotrytis cinerea, Alternaria alternata[92]
S. tuberosumB. cinerea, R. solani[87]
TCSO. sativaPyricularia oryzae[104]
α-MMCO. sativaPyricularia grisea[103]
Curcin 2N. tabacumR. solani[85]
Aspergillus flavus Link (which is not an aggressive pathogen of maize, but has a great economic impact due to its production of aflatoxin) and Aspergillus nidulans (Eidam) G. Winter were sensitive to MOD1 (RIP1), i.e., an engineered form of maize RIP b-32 (proRIP1) that does not require proteolytic activation [106]. MOD1 not only affected the growth of the fungi but also altered their morphology. The growth inhibition was concentration-dependent, being evident at 200 μg/mL, and above. R. solani was more sensitive to RIP b-32; from 0.6 µg/mL, the growth inhibition was shown in a microtiter plate assay [80].
In dicotyledons, type 1 RIPs with antifungal activity have been found in species of the families Amaranthaceae, Phytolaccaceae, Nyctaginaceae, Cucurbitaceae, Solanaceae, and Arecaceae (Table 3).
Penicillium digitatum (Pers.) Sacc. was very sensitive to different type 1 RIPs: sodin 5 [109], quinoin [109], BE27 [29,108], dioicin 2 [30], and PD-S2 [30]. The one that exerted the greatest effect was BE27 since the growth inhibition was evident at 0.6 µg/mL [29]. The other RIPs mentioned inhibited fungal growth from concentrations of 5–10 μg/mL; however, PD-L4 did not inhibit fungal growth at 30 μg/mL [30]. Notably, like sodin 5, quinoin, BE27, diocin 2, and PD-S2, PD-L4 also have N-glycosylase activity on yeast ribosomes [30,109]. BE27, diocin 2, PD-S2, and PD-L4 have also been reported to be active against P. digitatum ribosomes [30].
The fact that PD-L4 does not show antifungal activity [30] despite showing high homology with PD-S2 [113] suggests that entry into cells may be the limiting step for the fungicidal capacity of RIPs, and it has been suggested that the amphipathicity of the carboxyl-terminal domain could play a relevant role in the different degrees of toxicity of RIPs towards fungi [30]. In fact, BE27, which is the most toxic to P. digitatum, is the one with the highest degree of amphipathicity in the carboxyl-terminal domain and has been shown to be able to internalize and depurinate fungal ribosomes [29]. This is in agreement with the studies reported with RIPs obtained from M. expansa. ME1 and ME2 are two RIPs, obtained from the root of M. expansa, which showed both rRNA N-glycosylase activity on yeast ribosomes and antifungal activity in an agar plate assay against various fungi (Table 3), but were shown to be inactive against others [27]. Park et al. [28] compared the activity of three RIPs (ricin A-chain, saporin-S6, and ME from M. expansa) on fungal ribosomes and their antifungal activity. Ricin A-chain and saporin-S6 were much more active on ribosomes of Alternaria solani Sorauer, R. solani, T. reesei, and Candida albicans (C.P. Robin) Berkhout than the RIP of M. expansa; however, this was the only one able to inhibit the growth of R. solani because it was the only one able to enter the fungus.
Quinoin also induced a slight inhibition of the growth of Cryphonectria parasitica (Murrill) M.E. Barr [96]. Another RIP from Phytolacaceae, PhRIP I, was able to inhibit the germination of B. cinerea spores [92].
Alpha-momorcharin (α-MMC), a RIP from Momordica charantia L. seeds, inhibited sporulation [94] or mycelial growth [100,105] of a wide variety of pathogenic fungi (Table 3); however, it was ineffective in inhibiting the growth of C. albicans [100], supporting the hypothesis that antifungal activity also depends on the fungus studied. Interestingly, α-MMC, in addition to antifungal activity, also has antibacterial [100] and antiviral [12] activity, making it an excellent tool for crop protection against a wide variety of pathogens. Two other RIPs obtained from cucurbits, MbRIP-1 [107] and hispin [90], have antifungal activity. The latter, despite inhibiting the growth of several fungi (Table 3), was shown to be ineffective against B. cinerea [90].
The other type 1 RIPs that have shown antifungal activity are TRIP from tobacco [95] and two isoforms, EgRIP-1a and EgRIP-1b, obtained from oil palm (Elaeis guineensis Jacq.) [88]. TRIP showed rRNA N-glycosylase activity against yeast and T. resei ribosomes and presented growth inhibitory activity against T. resei and other fungi in agar plate assays [95]. The activity was different in different fungal species and, in some cases, was ineffective. Partially purified oil palm RIPs showed rRNA N-glycosylase activity on yeast ribosomes and inhibited the growth of Ganoderma boninense Pat., an oil palm pathogen causing basal stem rot (BSR) [88]. The only type 2 RIP reported to have antifungal activity is pebulin, a recombinant protein from Sambucus ebulus L. This protein was able to completely inhibit the germination of A. solani and Fusarium oxysporum Schltdl. spores at a concentration of 5 µg/mL [102].
In addition, peptides of around 10 kDa obtained from the cucurbits Cucumis melo L. and Benincasa hispida Cogn. have been reported as RIPs with antifungal activity [89,101]. On the other hand, the ribotoxin α-sarcin [108] and the RLPs ageritin [114,115] and eryngitins 3 and 4 [116] have also shown antifungal activity, indicating that, although never used for that purpose, these proteins could also be tools to defend crops against fungal diseases.

7. Mechanisms of Antifungal Activity

Several mechanisms have been proposed for the antifungal action of RIPs (Figure 1).
Since many RIPs inactivate plant ribosomes (Table 1), it has been proposed that they could be part of a “suicide mechanism” [32,33,117]. The RIPs that have been reported to have antifungal activity are, except pebulin, type 1 RIPs (Table 3 and Table 4). RIPs from Poaceae do not have a leader peptide [118] and are localized in the cytoplasm [47,49]. However, it seems that these RIPs do not affect the ribosomes of the same plant (Table 1). Some are synthesized as precursors and subsequently undergo processing, but this does not appear to significantly increase enzymatic activity against their own plant ribosomes [49]. Many type 1 RIPs from dicots are potent inhibitors of protein synthesis in plants (Table 1), but, having leader peptides [118], they are synthesized in the endoplasmic reticulum, and are located in the apoplast, the space between the plasma membrane and the cell wall [34,117,119], thus avoiding contact with ribosomes. Therefore, it has been assumed that pathogen infection would alter the permeability of the host cell membrane, allowing RIPs access to ribosomes and leading to the arrest of protein synthesis and cell death. This would prevent the spread of the pathogen throughout the rest of the plant [32,33,117]. In addition, RIP expression could be increased by the presence of the pathogen, since infection causes the release of pathogen-associated molecular patterns (PAMPs) that are recognized by pattern recognition receptors (PRRs) and damage-associated molecular patterns (DAMPs) that are recognized by wall-associated kinases (WAKs) leading to the increase of defense signal molecules such as hydrogen peroxide, salicylic acid, or jasmonic acid [120]. Hydrogen peroxide and salicylic acid have been shown to increase BE27 expression [29,34], jasmonic acid has been shown to increase the expression of α-momorcharin (α-MMC) [121], and methyl jasmonate and salicylic acid have been shown to increase curcin-L expression [122]. These are type 1 RIPs that are expressed in Beta vulgaris leaves, in different tissues of M. charantia, and in Jatropha curcas L. leaves, respectively.
A second mechanism Involves a direct effect on pathogen ribosomes. Many, but not all, RIPs inhibit the growth of various fungi in vitro (Table 3). Such inhibition appears to be related to the ability of the RIP to reach ribosomes by traversing the fungal cell wall and membrane [28,29]. Chitinases and glucanases can degrade the fungal cell wall and favor RIP entry, as these enzymes have been shown to enhance the antifungal capacity of some RIPs [76].
The third proposed mechanism involves the generation of signaling molecules that defend the plant from attack by fungi and other pathogens [120,123]. Not all RIPs generate the same signals and different results have been obtained depending on the RIP and the plant studied. Thus, it has been reported that α-MMC, in Nicotiana benthamiana Domin plants sprayed with a solution of the RIP, up-regulates the expression of genes related to the scavenging of reactive oxygen species (ROS), modulating ROS homeostasis, and some defense-related genes responsive to salicylic acid [94,124], and that in Nicotiana tabacum L. plants it induces an increase in both jasmonic acid and salicylic acid [125]. In contrast, the same RIP sprayed on M. charantia plants increases jasmonic acid biosynthesis and ROS induction without a relevant increase in salicylic acid [121]. PAP and PAPII (two type 1 RIPs, obtained from spring and early summer leaves of P. amaricana, respectively) generate a signal leading to overexpression of pathogenesis-related proteins in the absence of increased salicylic acid levels, making transgenic tobacco plants resistant to virus and fungal infection [81,82,83,126].
The relationship between the enzymatic activity of RIPs and their ability to induce the production of signaling molecules in plants has not been studied. In animals, ricin, α-sarcin, and Shiga toxin, as a consequence of their enzymatic action on the sarcin-ricin loop (SRL), activate signaling pathways through the mitogen-activated protein kinases (MAPKs) p38 and JNK [127]. Deoxynivalenol (DON) and T-2 toxin (both trichothecene mycotoxins) inhibit protein synthesis and induce ERK1/2 and p38 MAPK activation in several cell lines, followed by increased cytokine production [128]. This ribosome-mediated MAPK activation is termed “ribotoxic stress response” [128]. In Arabidopsis thaliana (L.) Heynh., DON, and T-2 toxin induce the expression of MPK3 and MPK6, which are implicated as positive regulators of the hypersensitive response through ethylene and ROS signaling [128]. It is therefore possible that the generation of signaling compounds by plants is a response to the ribotoxic stress produced by RIPs.
In conclusion, RIPs could exert their antifungal action through various mechanisms. Probably, depending on the RIP and the pathogen, one mechanism could predominate over the others or the effect could be a combination of several of them.

8. Transgenic Plants Resistant to Fungal Infection

Genetic engineering has proved to be an excellent method for obtaining fungus-resistant plants [120,129,130]. In this way, plants expressing genes that protect the plant from fungal infections have been obtained. Using this strategy, transgenic plants have been designed that carry the gene for a RIP and are resistant to pathogenic fungi that cause disease (Table 4). The most commonly used model is the tobacco plant (N. tabacum), but plants have also been obtained from some important crops such as wheat (Triticum aestivum L.), potato (Solanum tuberosum L.), Indian mustard (Brassica juncea (L.) Czern.), black gram (Vigna mungo (L.) Hepper), maize (Zea mays L.), rice (Oryza sativa L.), or creeping bentgrass (Agrostis stolonifera L.), widely used as turf (Table 4).
The most commonly used method to obtain these transgenic plants is Agrobacterium tumefaciens-mediated transformation, although direct methods such as biolistics [91,93,97,104] and polyethylene glycol-mediated transfer [98] have also been used. Different RIPs have been constitutively expressed under the control of a strong promoter such as the cauliflower mosaic virus promoter (CaMV 35S) (Figure 2), since there is a strong correlation between the level of RIP expression and the level of resistance against the fungi [85]. Genes conferring resistance to kanamycin and neomycin, hygromycin, or glufosinate (Figure 2), also controlled by strong promoters, are used as selection marker genes.
Although cases have been reported in which the transgenic plants show a normal phenotype [85,87,91,104], or at most a slightly smaller size [97], in other cases the constitutively expressed RIPs were toxic to the plants, probably due to their ability to inactivate host plant ribosomes [81,83,93]. This major drawback has been overcome by using the RIP gene with the sequence that directs it to the apoplast [85], by introducing mutations that reduce RIP toxicity without affecting its antifungal activity [81,83,93], or by using RIP genes that are less toxic to plants [82]. The latter may depend on the RIP and the host, e.g., PAPII has been reported to be toxic to creeping bentgrass [93], but not to tobacco [82]. This is also true for the RIP-pathogen relationship, so it has been reported that transgenic plants resistant to one fungus are not resistant to others [84,112]. Another strategy is to use inducible promoters that respond to the damage caused by the fungus in the plant (Figure 2) so that RIP is only expressed when the plant is attacked by the fungus without affecting its development [78,80,87,92].
To enhance the antifungal activity of RIPs, genes encoding chitinases have also been introduced [79,84,86,99,110,111], which as we have seen exert a synergistic effect with RIPs, and even the lectin WGA [99] (Figure 2). In addition, some transgenic plants carrying RIPs are also resistant to viruses and insects [9,12], which adds even more interest to this type of strategy. For example, tobacco plants carrying the PAP II gene are resistant to the fungus R. solani and the viruses TMV and PVX [82]; those carrying curcin 2 are resistant to R. solani and TMV [85]; and maize plants carrying maize ribosome-inactivating protein (MRIP), tobacco hornworm chitinase (THWC), and wheat germ agglutinin (WGA) are resistant to the fungus Fusarium fujikuroi Nirenberg and the insects Spodoptera frugiperda Walker and Helicoverpa zea Boddie [99].

9. Conclusions

The use of ribosome-inactivating proteins (RIPs) is a promising alternative to chemical-based fungicides, which cause problems such as environmental contamination, resistant development, and residual toxicity. The fact that RIPs also possess antiviral and insecticidal activities makes them an ideal tool for disease and pest control in crops.
In view of the published results, it seems that the most efficient way to use these proteins would be to construct transgenic plants carrying genes for RIPs and genes for other defense proteins with which they show a synergistic effect.
However, in order to use these proteins effectively, further studies are still needed to shed light on the toxicity of the different RIPs to the host plants, the efficacy of each RIP on the fungi causing the diseases to be controlled, the synergistic effect with other fungicidal agents, as well as the mechanisms of antifungal action.

Author Contributions

Conceptualization, R.I. and J.M.F.; writing—original draft preparation, L.C. and J.M.F.; writing—review and editing, L.C., R.I. and C.C.G.; funding acquisition, J.M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Consejería de Educación (Junta de Castilla y León) to the GIR ProtIBio, grant number VA033G19. C.C.G. was supported by a fellowship from Programa Estancias Cortas Postdoctorales 2023, Fundación Carolina-Universidad Nacional del Nordeste.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Nazarov, P.; Baleev, D.; Ivanova, M.; Sokolova, L.; Karakozova, M. Infectious Plant Diseases: Etiology, Current Status, Problems and Prospects in Plant Protection. Acta Nat. 2020, 12, 46–59. [Google Scholar] [CrossRef] [PubMed]
  2. Shuping, D.S.S.; Eloff, J.N. The use of plants to protect plants and food against fungal pathogens. Afr. J. Tradit. Complement. Altern. Med. 2017, 14, 120–127. [Google Scholar] [CrossRef] [PubMed]
  3. Doehlemann, G.; Ökmen, B.; Zhu, W.; Sharon, A. Plant Pathogenic Fungi. Microbiol. Spectr. 2017, 5, 701–726. [Google Scholar] [CrossRef] [PubMed]
  4. Yoon, M.; Cha, B.; Kim, J. Recent Trends in Studies on Botanical Fungicides in Agriculture. Plant Pathol. J. 2013, 29, 1–9. [Google Scholar] [CrossRef] [PubMed]
  5. Chiu, T.; Poucet, T.; Li, Y. The potential of plant proteins as antifungal agents for agricultural applications. Synth. Syst. Biotechnol. 2022, 7, 1075–1083. [Google Scholar] [CrossRef]
  6. Meng, J.; Zhang, X.; Han, X.; Fan, B. Application and Development of Biocontrol Agents in China. Pathogens 2022, 11, 1120. [Google Scholar] [CrossRef]
  7. Ng, T. Antifungal proteins and peptides of leguminous and non-leguminous origins. Peptides 2004, 25, 1215–1222. [Google Scholar] [CrossRef] [PubMed]
  8. Wong, J.; Ng, T.; Cheung, R.; Ye, X.; Wang, H.; Lam, S.; Lin, P.; Chan, Y.; Fang, E.; Ngai, P.; et al. Proteins with antifungal properties and other medicinal applications from plants and mushrooms. Appl. Microbiol. Biotechnol. 2010, 87, 1221–1235. [Google Scholar] [CrossRef] [PubMed]
  9. Zhu, F.; Zhou, Y.K.; Ji, Z.L.; Chen, X.R. The Plant Ribosome-Inactivating Proteins Play Important Roles in Defense against Pathogens and Insect Pest Attacks. Front. Plant Sci. 2018, 9, 146. [Google Scholar] [CrossRef]
  10. Stirpe, F. Ribosome-inactivating proteins. Toxicon 2004, 44, 371–383. [Google Scholar] [CrossRef]
  11. Bertholdo-Vargas, L.; Martins, J.; Bordin, D.; Salvador, M.; Schafer, A.; de Barros, N.; Barbieri, L.; Stirpe, F.; Carlini, C. Type 1 ribosome-inactivating proteins-Entomotoxic, oxidative and genotoxic action on Anticarsia gemmatalis (Hubner) and Spodoptera frugiperda (JE Smith) (Lepidoptera: Noctuidae). J. Insect Physiol. 2009, 55, 51–58. [Google Scholar] [CrossRef] [PubMed]
  12. Citores, L.; Iglesias, R.; Ferreras, J.M. Antiviral Activity of Ribosome-Inactivating Proteins. Toxins 2021, 13, 80. [Google Scholar] [CrossRef] [PubMed]
  13. Stirpe, F. Ribosome-inactivating proteins: From toxins to useful proteins. Toxicon 2013, 67, 12–16. [Google Scholar] [CrossRef] [PubMed]
  14. Barbieri, L.; Battelli, M.G.; Stirpe, F. Ribosome-Inactivating Proteins from Plants. Biochim. Biophys. Acta 1993, 1154, 237–282. [Google Scholar] [CrossRef] [PubMed]
  15. Bolognesi, A.; Bortolotti, M.; Maiello, S.; Battelli, M.G.; Polito, L. Ribosome-Inactivating Proteins from Plants: A Historical Overview. Molecules 2016, 21, 19. [Google Scholar] [CrossRef]
  16. Stirpe, F.; Battelli, M. Ribosome-inactivating proteins: Progress and problems. Cell. Mol. Life Sci. 2006, 63, 1850–1866. [Google Scholar] [CrossRef]
  17. Peumans, W.; Hao, Q.; Van Damme, E. Ribosome-inactivating proteins from plants: More than RNA N-glycosidases? FASEB J. 2001, 15, 1493–1506. [Google Scholar] [CrossRef]
  18. Wang, P.; Tumer, N.; Maramorosch, K.; Murphy, F.; Shatkin, A. Virus resistance mediated by ribosome inactivating proteins. Adv. Virus Res. 2000, 55, 325–355. [Google Scholar] [CrossRef]
  19. Bolognesi, A.; Polito, L. Immunotoxins and other conjugates: Pre-clinical studies. Mini-Rev. Med. Chem. 2004, 4, 563–583. [Google Scholar] [CrossRef]
  20. Pastan, I.; Hassan, R.; FitzGerald, D.; Kreitman, R. Immunotoxin treatment of cancer. Annu. Rev. Med. 2007, 58, 221–237. [Google Scholar] [CrossRef]
  21. Becker, N.; Benhar, I. Antibody Based Immunotoxins for the Treatment of Cancer. Antibodies 2012, 1, 39–69. [Google Scholar] [CrossRef]
  22. Polito, L.; Djemil, A.; Bortolotti, M. Plant Toxin-Based Immunotoxins for Cancer Therapy: A Short Overview. Biomedicines 2016, 4, 12. [Google Scholar] [CrossRef]
  23. Dougherty, K.; Hudak, K. Phylogeny and domain architecture of plant ribosome inactivating proteins. Phytochemistry 2022, 202, 113337. [Google Scholar] [CrossRef]
  24. Liu, J.; Wen, D.; Song, X.; Su, P.; Lou, J.; Yao, D.; Zhang, C. Evolution and natural selection of ribosome-inactivating proteins in bacteria, fungi, and plants. Int. J. Biol. Macromol. 2023, 248, 125929. [Google Scholar] [CrossRef] [PubMed]
  25. Sandvig, K.; Bergan, J.; Dyve, A.; Skotland, T.; Torgersen, M. Endocytosis and retrograde transport of Shiga toxin. Toxicon 2010, 56, 1181–1185. [Google Scholar] [CrossRef] [PubMed]
  26. Cenini, P.; Carnicelli, D.; Stirpe, F. Effect of Plant Ribosome-Inactivating Proteins on Ribosomes from Various Metazoan Species. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 1990, 96, 581–584. [Google Scholar] [CrossRef]
  27. Vivanco, J.M.; Savary, B.J.; Flores, H.E. Characterization of two novel type I ribosome-inactivating proteins from the storage roots of the Andean crop Mirabilis expansa. Plant Physiol. 1999, 119, 1447–1456. [Google Scholar] [CrossRef]
  28. Park, S.W.; Stevens, N.M.; Vivanco, J.M. Enzymatic specificity of three ribosome-inactivating proteins against fungal ribosomes, and correlation with antifungal activity. Planta 2002, 216, 227–234. [Google Scholar] [CrossRef] [PubMed]
  29. Citores, L.; Iglesias, R.; Gay, C.; Miguel Ferreras, J. Antifungal activity of the ribosome-inactivating protein BE27 from sugar beet (Beta vulgaris L.) against the green mould Penicillium digitatum. Mol. Plant Pathol. 2016, 17, 261–271. [Google Scholar] [CrossRef]
  30. Iglesias, R.; Citores, L.; Ragucci, S.; Russo, R.; Di Maro, A.; Ferreras, J.M. Biological and antipathogenic activities of ribosome-inactivating proteins from Phytolacca dioica L. Biochim. Biophys. Acta Gen. Subj. 2016, 1860, 1256–1264. [Google Scholar] [CrossRef]
  31. Iglesias, R.; Arias, F.; Rojo, M.; Escarmis, C.; Ferreras, J.; Girbes, T. Molecular Action of the Type 1 Ribosome-Inactivating Protein Saporin 5 on Vicia Sativa Ribosomes. FEBS Lett. 1993, 325, 291–294. [Google Scholar] [CrossRef] [PubMed]
  32. Park, S.W.; Vepachedu, R.; Sharma, N.; Vivanco, J.M. Ribosome-inactivating proteins in plant biology. Planta 2004, 219, 1093–1096. [Google Scholar] [CrossRef] [PubMed]
  33. Bonness, M.S.; Ready, M.P.; Irvin, J.D.; Mabry, T.J. Pokeweed Antiviral Protein Inactivates Pokeweed Ribosomes-Implications for the Antiviral Mechanism. Plant J. 1994, 5, 173–183. [Google Scholar] [CrossRef] [PubMed]
  34. Iglesias, R.; Perez, Y.; de Torre, C.; Ferreras, J.; Antolin, P.; Jimenez, P.; Rojo, M.; Mendez, E.; Girbes, T. Molecular characterization and systemic induction of single-chain ribosome-inactivating proteins (RIPs) in sugar beet (Beta vulgaris) leaves. J. Exp. Bot. 2005, 56, 1675–1684. [Google Scholar] [CrossRef] [PubMed]
  35. Lledías, F.; Gutiérrez, J.; Martínez-Hernández, A.; García-Mendoza, A.; Sosa, E.; Hernández-Bermúdez, F.; Dinkova, T.; Reyes, S.; Cassab, G.; Nieto-Sotelo, J. Mayahuelin, a Type I Ribosome Inactivating Protein: Characterization, Evolution, and Utilization in Phylogenetic Analyses of Agave. Front. Plant Sci. 2020, 11, 573. [Google Scholar] [CrossRef] [PubMed]
  36. Arias, F.; Rojo, M.; Iglesias, R.; Ferreras, J.; Girbes, T. Vicia sativa L. “Run-off” and Purified Ribosomes: Polyphenylalanine Synthesis and Molecular Action of Ribosome-Inactivating Proteins. J. Exp. Bot. 1993, 44, 1297–1304. [Google Scholar] [CrossRef]
  37. Rojo, M.; Arias, F.; Ferreras, J.; Iglesias, R.; Munoz, R.; Girbes, T. Development of a Cell-free Translation System from Cucumis melo: Preparation, Optimization and Evaluation of Sensitivity to some Translational Inhibitors. Plant Sci. 1993, 90, 127–134. [Google Scholar] [CrossRef]
  38. Arias, F.; Antolin, P.; de Torre, C.; Barriuso, B.; Iglesias, R.; Rojo, M.; Ferreras, J.; Benvenuto, E.; Mendez, E.; Girbes, T. Musarmins: Three single-chain ribosome-inactivating protein isoforms from bulbs of Muscari armeniacum L. and Miller. Int. J. Biochem. Cell Biol. 2003, 35, 61–78. [Google Scholar] [CrossRef]
  39. Taylor, S.; Massiah, A.; Lomonossoff, G.; Roberts, L.M.; Lord, J.M.; Hartley, M. Correlation between the Activities of 5 Ribosome-Inactivating Proteins in Depurination of Tobacco Ribosomes and Inhibition of Tobacco Mosaic Virus Infection. Plant J. 1994, 5, 827–835. [Google Scholar] [CrossRef]
  40. Ferreras, J.; Barbieri, L.; Girbes, T.; Battelli, M.; Rojo, M.; Arias, F.; Rocher, M.; Soriano, F.; Mendez, E.; Stirpe, F. Distribution and Properties of Major Ribosome-Inactivating Proteins (28 S rRNA N-Glycosidases) of the Plant Saponaria officinalis L. (Caryophyllaceae). Biochim. Biophys. Acta 1993, 1216, 31–42. [Google Scholar] [CrossRef]
  41. Rojo, M.; Arias, F.; Iglesias, R.; Ferreras, J.; Munoz, R.; Girbes, T. A Cucumis Sativus Cell-free Translation System: Preparation, Optimization and Sensitivity to some Antibiotics and Ribosome-Inactivating Proteins. Physiol. Plant. 1993, 88, 549–556. [Google Scholar] [CrossRef] [PubMed]
  42. Arias, F.; Rojo, M.; Ferreras, J.; Iglesias, R.; Munoz, R.; Soriano, F.; Mendez, E.; Barbieri, L.; Girbes, T. Isolation and Characterization of Two New N-Glycosidase Type 1 Ribosome-Inactivating Proteins, Unrelated in Amino Acid Sequence, from Petrocoptis Species. Planta 1994, 194, 487–491. [Google Scholar] [CrossRef] [PubMed]
  43. Arias, F.; Rojo, M.; Ferreras, J.; Iglesias, R.; Munoz, R.; Rocher, A.; Mendez, E.; Barbieri, L.; Girbes, T. Isolation and Partial Characterization of a New Ribosome-Inactivating Protein from Petrocoptis glaucifolia (Lag.) Boiss. Planta 1992, 186, 532–540. [Google Scholar] [CrossRef] [PubMed]
  44. Harley, S.M.; Beevers, H. Ricin Inhibition of In Vitro Protein Synthesis by Plant Ribosomes. Proc. Natl. Acad. Sci. USA 1982, 79, 5935–5938. [Google Scholar] [CrossRef] [PubMed]
  45. Citores, L.; Ferreras, J.; Iglesias, R.; Carbajales, M.; Arias, F.; Jimenez, P.; Rojo, M.; Girbes, T. Molecular mechanism of inhibition of mammalian protein synthesis by some four-chain agglutinins. Proposal of an extended classification of plant ribosome-inactivating proteins (rRNA N-glycosidases). FEBS Lett. 1993, 329, 59–62. [Google Scholar] [CrossRef] [PubMed]
  46. Arias, F.; Rojo, M.; Ferreras, J.; Iglesias, R.; Munoz, R.; Girbes, T. Preparation and Optimization of a Cell-free Translation System from Vicia sativa Germ Lacking Ribosome-Inactivating Protein Activity. J. Exp. Bot. 1992, 43, 729–737. [Google Scholar] [CrossRef]
  47. De Zaeytijd, J.; Rouge, P.; Smagghe, G.; Van Damme, E.J.M. Structure and Activity of a Cytosolic Ribosome-Inactivating Protein from Rice. Toxins 2019, 11, 325. [Google Scholar] [CrossRef] [PubMed]
  48. Massiah, A.J.; Hartley, M.R. Wheat Ribosome-Inactivating Proteins—Seed and Leaf Forms with Different Specificities and Cofactor Requirements. Planta 1995, 197, 633–640. [Google Scholar] [CrossRef] [PubMed]
  49. Hey, T.; Hartley, M.; Walsh, T. Maize ribosome-inactivating protein (b-32)-homologs in related species, effects on maize ribosomes, and modulation of activity by pro-peptide deletions. Plant Physiol. 1995, 107, 1323–1332. [Google Scholar] [CrossRef]
  50. Ferreras, J.; Citores, L.; Martinez de Benito, F.; Arias, F.; Rojo, M.; Munoz, R.; Iglesias, R.; Girbes, T. Ribosome-inactivating proteins and lectins from Sambucus. Curr. Top. Phytochem. 2000, 3, 113–128. [Google Scholar]
  51. Endo, Y.; Mitsui, K.; Motizuki, M.; Tsurugi, K. The Mechanism of Action of Ricin and Related Toxic Lectins on Eukaryotic Ribosomes—The Site and the Characteristics of the Modification in 28-S Ribosomal-RNA Caused by the Toxins. J. Biol. Chem. 1987, 262, 5908–5912. [Google Scholar] [CrossRef] [PubMed]
  52. Endo, Y.; Tsurugi, K. The RNA N-glycosidase activity of ricin A-chain. The characteristics of the enzymatic activity of ricin A-chain with ribosomes and with rRNA. J. Biol. Chem. 1988, 263, 8735–8739. [Google Scholar] [CrossRef]
  53. Maracci, C.; Rodnina, M. Translational GTPases. Biopolymers 2016, 105, 463–475. [Google Scholar] [CrossRef] [PubMed]
  54. Grela, P.; Szajwaj, M.; Horbowicz-Drozdzal, P.; Tchorzewski, M. How Ricin Damages the Ribosome. Toxins 2019, 11, 241. [Google Scholar] [CrossRef] [PubMed]
  55. Choi, A.; Wong, E.; Lee, K.; Wong, K. Structures of Eukaryotic Ribosomal Stalk Proteins and Its Complex with Trichosanthin, and Their Implications in Recruiting Ribosome-Inactivating Proteins to the Ribosomes. Toxins 2015, 7, 638–647. [Google Scholar] [CrossRef]
  56. Lacadena, J.; Alvarez-García, E.; Carreras-Sangrà, N.; Herrero-Galán, E.; Alegre-Cebollada, J.; García-Ortega, L.; Oñaderra, M.; Gavilanes, J.; del Pozo, A. Fungal ribotoxins: Molecular dissection of a family of natural killers. FEMS Microbiol. Rev. 2007, 31, 212–237. [Google Scholar] [CrossRef]
  57. Olombrada, M.; Lazaro-Gorines, R.; Lopez-Rodriguez, J.; Martinez-del-Pozo, A.; Onaderra, M.; Maestro-Lopez, M.; Lacadena, J.; Gavilanes, J.; Garcia-Ortega, L. Fungal Ribotoxins: A Review of Potential Biotechnological Applications. Toxins 2017, 9, 71. [Google Scholar] [CrossRef]
  58. Endo, Y.; Wool, I.G. The site of action of alpha-sarcin on eukaryotic ribosomes—The sequence at the alpha-sarcin cleavage site in 28 S-ribosomal ribonucleic-acid. J. Biol. Chem. 1982, 257, 9054–9060. [Google Scholar] [CrossRef] [PubMed]
  59. García-Ortega, L.; Alvarez-García, E.; Gavilanes, J.; Martínez-del-Pozo, A.; Joseph, S. Cleavage of the sarcin-ricin loop of 23S rRNA differentially affects EF-G and EF-Tu binding. Nucleic Acids Res. 2010, 38, 4108–4119. [Google Scholar] [CrossRef]
  60. Barbieri, L.; Valbonesi, P.; Bonora, E.; Gorini, P.; Bolognesi, A.; Stirpe, F. Polynucleotide:adenosine glycosidase activity of ribosome-inactivating proteins: Effect on DNA, RNA and poly(A). Nucleic Acids Res. 1997, 25, 518–522. [Google Scholar] [CrossRef]
  61. Schrot, J.; Weng, A.; Melzig, M. Ribosome-Inactivating and Related Proteins. Toxins 2015, 7, 1556–1615. [Google Scholar] [CrossRef] [PubMed]
  62. Landi, N.; Hussain, H.Z.F.; Pedone, P.V.; Ragucci, S.; Di Maro, A. Ribotoxic Proteins, Known as Inhibitors of Protein Synthesis, from Mushrooms and Other Fungi According to Endo’s Fragment Detection. Toxins 2022, 14, 403. [Google Scholar] [CrossRef] [PubMed]
  63. Citores, L.; Ragucci, S.; Russo, R.; Gay, C.; Chambery, A.; Di Maro, A.; Iglesias, R.; Ferreras, J. Structural and functional characterization of the cytotoxic protein ledodin, an atypical ribosome-inactivating protein from shiitake mushroom (Lentinula edodes). Protein Sci. 2023, 32, e4621. [Google Scholar] [CrossRef] [PubMed]
  64. Ragucci, S.; Landi, N.; Russo, R.; Valletta, M.; Pedone, P.; Chambery, A.; Di Maro, A. Ageritin from Pioppino Mushroom: The Prototype of Ribotoxin-Like Proteins, a Novel Family of Specific Ribonucleases in Edible Mushrooms. Toxins 2021, 13, 263. [Google Scholar] [CrossRef] [PubMed]
  65. Spooner, R.; Lord, J. Ricin Trafficking in Cells. Toxins 2015, 7, 49–65. [Google Scholar] [CrossRef] [PubMed]
  66. Fabbrini, M.; Katayama, M.; Nakase, I.; Vago, R. Plant Ribosome-Inactivating Proteins: Progesses, Challenges and Biotechnological Applications (and a Few Digressions). Toxins 2017, 9, 314. [Google Scholar] [CrossRef]
  67. Qin, S.; Wang, X.; Han, P.; Lai, Z.; Ren, Y.; Ma, R.; Cheng, C.; Wang, T.; Xu, Y. LRP1-Mediated Endocytosis May Be the Main Reason for the Difference in Cytotoxicity of Curcin and Curcin C on U2OS Osteosarcoma Cells. Toxins 2022, 14, 771. [Google Scholar] [CrossRef] [PubMed]
  68. Bolognesi, A.; Polito, L.; Scicchitano, V.; Orrico, C.; Pasquinelli, G.; Musiani, S.; Santi, S.; Riccio, M.; Bortolotti, M.; Battelli, M. Endocytosis and intracellular localisation of type 1 ribosome-inactivating protein saporin-S6. J. Biol. Regul. Homeost. Agents 2012, 26, 97–109. [Google Scholar] [PubMed]
  69. Walsh, M.; Dodd, J.; Hautbergue, G. Ribosome-inactivating proteins Potent poisons and molecular tools. Virulence 2013, 4, 774–784. [Google Scholar] [CrossRef]
  70. Goode, B.; Eskin, J.; Wendland, B. Actin and Endocytosis in Budding Yeast. Genetics 2015, 199, 315–358. [Google Scholar] [CrossRef]
  71. Commer, B.; Shaw, B.D. Current views on endocytosis in filamentous fungi. Mycology 2020, 12, 1–9. [Google Scholar] [CrossRef]
  72. Ebrahimi, H.; Siavoshi, F.; Jazayeri, M.; Sarrafnejad, A.; Saniee, P.; Mobini, M. Physicochemical properties of intact fungal cell wall determine vesicles release and nanoparticles internalization. Heliyon 2023, 9, e13834. [Google Scholar] [CrossRef] [PubMed]
  73. Gow, N.; Latge, J.; Munro, C. The Fungal Cell Wall: Structure, Biosynthesis, and Function. Microbiol. Spectr. 2017, 5, 1–25. [Google Scholar] [CrossRef] [PubMed]
  74. Barbieri, L.; Polito, L.; Bolognesi, A.; Ciani, M.; Pelosi, E.; Farini, V.; Jha, A.; Sharma, N.; Vivanco, J.; Chambery, A.; et al. Ribosome-inactivating proteins in edible plants and purification and characterization of a new ribosome-inactivating protein from Cucurbita moschata. Biochim. Biophys. Acta Gen. Subj. 2006, 1760, 783–792. [Google Scholar] [CrossRef] [PubMed]
  75. Hohl, T.; Feldmesser, M. Aspergillus fumigatus: Principles of pathogenesis and host defense. Eukaryot. Cell 2007, 6, 1953–1963. [Google Scholar] [CrossRef] [PubMed]
  76. Leah, R.; Tommerup, H.; Svendsen, I.; Mundy, J. Biochemical and molecular characterization of 3 barley seed proteins with antifungal properties. J. Biol. Chem. 1991, 266, 1564–1573. [Google Scholar] [CrossRef] [PubMed]
  77. Roberts, W.; Selitrennikoff, C. Isolation and partial characterization of 2 antifungal proteins from barley. Biochim. Biophys. Acta 1986, 880, 161–170. [Google Scholar] [CrossRef] [PubMed]
  78. Logemann, J.; Jach, G.; Tommerup, H.; Mundy, J.; Schell, J. Expression of a barley ribosome-inactivating protein leads to increased fungal protection in transgenic tobacco plants. Bio-Technology 1992, 10, 305–308. [Google Scholar] [CrossRef]
  79. Jach, G.; Gornhardt, B.; Mundy, J.; Logemann, J.; Pinsdorf, P.; Leah, R.; Schell, J.; Maas, C. Enhanced quantitative resistance against fungal disease by combinatorial expression of different barley antifungal proteins in transgenic tobacco. Plant J. 1995, 8, 97–109. [Google Scholar] [CrossRef]
  80. Maddaloni, M.; Forlani, F.; Balmas, V.; Donini, G.; Stasse, L.; Corazza, L.; Motto, M. Tolerance to the fungal pathogen Rhizoctonia solani AG4 of transgenic tobacco expressing the maize ribosome-inactivating protein b-32. Transgenic Res. 1997, 6, 393–402. [Google Scholar] [CrossRef]
  81. Zoubenko, O.; Uckun, F.; Hur, Y.; Chet, I.; Tumer, N. Plant resistance to fungal infection induced by nontoxic pokeweed antiviral protein mutants. Nat. Biotechnol. 1997, 15, 992–996. [Google Scholar] [CrossRef] [PubMed]
  82. Wang, P.G.; Zoubenko, O.; Tumer, N.E. Reduced toxicity and broad spectrum resistance to viral and fungal infection in transgenic plants expressing pokeweed antiviral protein II. Plant Mol. Biol. 1998, 38, 957–964. [Google Scholar] [CrossRef]
  83. Zoubenko, O.; Hudak, K.; Tumer, N.E. A non-toxic pokeweed antiviral protein mutant inhibits pathogen infection via a novel salicylic acid-independent pathway. Plant Mol. Biol. 2000, 44, 219–229. [Google Scholar] [CrossRef]
  84. Kim, J.; Jang, I.; Wu, R.; Zu, W.; Boston, R.; Lee, Y.; Alm, P.; Nahm, B. Co-expression of a modified maize ribosome-inactivating protein and a rice basic chitinase gene in transgenic rice plants confers enhanced resistance to sheath blight. Transgenic Res. 2003, 12, 475–484. [Google Scholar] [CrossRef] [PubMed]
  85. Huang, M.-X.; Hou, P.; Wei, Q.; Xu, Y.; Chen, F. A ribosome-inactivating protein (curcin 2) induced from Jatropha curcas can reduce viral and fungal infection in transgenic tobacco. Plant Growth Regul. 2008, 54, 115–123. [Google Scholar] [CrossRef]
  86. M’hamdi, M.; Chikh-Rouhou, H.; Boughalleb, N.; de Galarreta, J. Enhanced resistance to Rhizoctonia solani by combined expression of chitinase and Ribosome Inactivating Protein in transgenic potatoes (Solanum tuberosum L.). Span. J. Agric. Res. 2012, 10, 778–785. [Google Scholar] [CrossRef]
  87. Gonzales-Salazar, R.; Cecere, B.; Ruocco, M.; Rao, R.; Corrado, G. A comparison between constitutive and inducible transgenic expression of the PhRIP I gene for broad-spectrum resistance against phytopathogens in potato. Biotechnol. Lett. 2017, 39, 1049–1058. [Google Scholar] [CrossRef]
  88. Sargolzaei, M.; Ho, C.; Wong, M. Characterization of novel type I ribosome-inactivating proteins isolated from oil palm (Elaeis guineensis) inoculated with Ganoderma boninense, the causal agent of basal stem rot. Physiol. Mol. Plant Pathol. 2016, 94, 53–61. [Google Scholar] [CrossRef]
  89. Ng, T.; Parkash, A.; Tso, W. Purification and characterization of α- and β-benincasins, arginine/glutamate-rich peptides with translation-inhibiting activity from wax gourd seeds. Peptides 2003, 24, 11–16. [Google Scholar] [CrossRef]
  90. Ng, T.; Parkash, A. Hispin, a novel ribosome inactivating protein with antifungal activity from hairy melon seeds. Protein Expr. Purif. 2002, 26, 211–217. [Google Scholar] [CrossRef]
  91. Bieri, S.; Potrykus, I.; Fütterer, J. Expression of active barley seed ribosome-inactivating protein in transgenic wheat. Theor. Appl. Genet. 2000, 100, 755–763. [Google Scholar] [CrossRef]
  92. Corrado, G.; Bovi, P.; Ciliento, R.; Gaudio, L.; Di Maro, A.; Aceto, S.; Lorito, M.; Rao, R. Inducible expression of a Phytolacca heterotepala ribosome-inactivating protein leads to enhanced resistance against major fungal pathogens in tobacco. Phytopathology 2005, 95, 206–215. [Google Scholar] [CrossRef]
  93. Dai, W.; Bonos, S.; Guo, Z.; Meyer, W.; Day, P.; Belanger, F. Expression of pokeweed antiviral proteins in creeping bentgrass. Plant Cell Rep. 2003, 21, 497–502. [Google Scholar] [CrossRef] [PubMed]
  94. Zhu, F.; Zhang, P.; Meng, Y.F.; Xu, F.; Zhang, D.W.; Cheng, J.; Lin, H.H.; Xi, D.H. Alpha-momorcharin, a RIP produced by bitter melon, enhances defense response in tobacco plants against diverse plant viruses and shows antifungal activity in vitro. Planta 2013, 237, 77–88. [Google Scholar] [CrossRef]
  95. Sharma, N.; Park, S.; Vepachedu, R.; Barbieri, L.; Ciani, M.; Stirpe, F.; Savary, B.; Vivanco, J. Isolation and characterization of an RIP (ribosome-inactivating protein)-like protein from tobacco with dual enzymatic activity. Plant Physiol. 2004, 134, 171–181. [Google Scholar] [CrossRef]
  96. Ragucci, S.; Bulgari, D.; Landi, N.; Russo, R.; Clemente, A.; Valletta, M.; Chambery, A.; Gobbi, E.; Faoro, F.; Di Maro, A. The Structural Characterization and Antipathogenic Activities of Quinoin, a Type 1 Ribosome-Inactivating Protein from Quinoa Seeds. Int. J. Mol. Sci. 2021, 22, 8964. [Google Scholar] [CrossRef] [PubMed]
  97. Balconi, C.; Lanzanova, C.; Conti, E.; Triulzi, T.; Forlani, F.; Cattaneo, M.; Lupotto, E. Fusarium head blight evaluation in wheat transgenic plants expressing the maize b-32 antifungal gene. Eur. J. Plant Pathol. 2007, 117, 129–140. [Google Scholar] [CrossRef]
  98. Lanzanova, C.; Giuffrida, M.; Motto, M.; Baro, C.; Donn, G.; Hartings, H.; Lupotto, E.; Careri, M.; Elviri, L.; Balconi, C. The Zea mays b-32 ribosome-inactivating protein efficiently inhibits growth of Fusarium verticillioides on leaf pieces in vitro. Eur. J. Plant Pathol. 2009, 124, 471–482. [Google Scholar] [CrossRef]
  99. Dowd, P.; Johnson, E.; Price, N. Enhanced Pest Resistance of Maize Leaves Expressing Monocot Crop Plant-Derived Ribosome-Inactivating Protein and Agglutinin. J. Agric. Food Chem. 2012, 60, 10768–10775. [Google Scholar] [CrossRef]
  100. Wang, S.; Zhang, Y.; Liu, H.; He, Y.; Yan, J.; Wu, Z.; Ding, Y. Molecular cloning and functional analysis of a recombinant ribosome-inactivating protein (alpha-momorcharin) from Momordica charantia. Appl. Microbiol. Biotechnol. 2012, 96, 939–950. [Google Scholar] [CrossRef]
  101. Parkash, A.; Ng, T.; Tso, W. Isolation and characterization of luffacylin, a ribosome inactivating peptide with anti-fungal activity from sponge gourd (Luffa cylindrica) seeds. Peptides 2002, 23, 1019–1024. [Google Scholar] [CrossRef] [PubMed]
  102. Rezaei-Moshaei, M.; Dehestani, A.; Bandehagh, A.; Pakdin-Parizi, A.; Golkar, M.; Heidari-Japelaghi, R. Recombinant pebulin protein, a type 2 ribosome-inactivating protein isolated from dwarf elder (Sambucus ebulus L.) shows anticancer and antifungal activities in vitro. Int. J. Biol. Macromol. 2021, 174, 352–361. [Google Scholar] [CrossRef] [PubMed]
  103. Qian, Q.; Huang, L.; Yi, R.; Wang, S.; Ding, Y. Enhanced resistance to blast fungus in rice (Oryza sativa L.) by expressing the ribosome-inactivating protein alpha-momorcharin. Plant Sci. 2014, 217, 1–7. [Google Scholar] [CrossRef] [PubMed]
  104. Yuan, H.; Ming, X.; Wang, L.; Hu, P.; An, C.; Chen, Z. Expression of a gene encoding trichosanthin in transgenic rice plants enhances resistance to fungus blast disease. Plant Cell Rep. 2002, 20, 992–998. [Google Scholar] [CrossRef]
  105. Nguyen, T.; Dang, D. Molecular cloning and isolation of a recombinant alpha-Momorcharin in E. coli against Pyricularia oryzae. Sci. Technol. Dev. J. 2023, 26, 2665–2671. [Google Scholar] [CrossRef]
  106. Nielsen, K.; Payne, G.; Boston, R. Maize ribosome-inactivating protein inhibits normal development of Aspergillus nidulans and Aspergillus flavus. Mol. Plant Microbe Interact. 2001, 14, 164–172. [Google Scholar] [CrossRef]
  107. Kushwaha, G.; Pandey, N.; Sinha, M.; Singh, S.; Kaur, P.; Sharma, S.; Singh, T. Crystal structures of a type-1 ribosome inactivating protein from Momordica balsamina in the bound and unbound states. Biochim. Biophys. Acta Proteins Proteom. 2012, 1824, 679–691. [Google Scholar] [CrossRef] [PubMed]
  108. Citores, L.; Valletta, M.; Singh, V.; Pedone, P.; Iglesias, R.; Ferreras, J.; Chambery, A.; Russo, R. Deciphering Molecular Determinants Underlying Penicillium digitatum’s Response to Biological and Chemical Antifungal Agents by Tandem Mass Tag (TMT)-Based High-Resolution LC-MS/MS. Int. J. Mol. Sci. 2022, 23, 680. [Google Scholar] [CrossRef]
  109. Landi, N.; Ragucci, S.; Citores, L.; Clemente, A.; Hussain, H.; Iglesias, R.; Ferreras, J.; Di Maro, A. Isolation, Characterization and Biological Action of Type-1 Ribosome-Inactivating Proteins from Tissues of Salsola soda L. Toxins 2022, 14, 566. [Google Scholar] [CrossRef]
  110. Chopra, R.; Saini, R. Transformation of Blackgram (Vigna mungo (L.) Hepper) by Barley Chitinase and Ribosome-Inactivating Protein Genes Towards Improving Resistance to Corynespora Leaf Spot Fungal Disease. Appl. Biochem. Biotechnol. 2014, 174, 2791–2800. [Google Scholar] [CrossRef]
  111. Chhikara, S.; Chaudhury, D.; Dhankher, O.; Jaiwal, P. Combined expression of a barley class II chitinase and type I ribosome inactivating protein in transgenic Brassica juncea provides protection against Alternaria brassicae. Plant Cell Tissue Organ Cult. 2012, 108, 83–89. [Google Scholar] [CrossRef]
  112. M’hamdi, M.; Chikh-Rouhou, H.; Boughalleb, N.; de Galarreta, J. Ribosome Inactivating Protein of barley enhanced resistance to Rhizoctonia solani in transgenic potato cultivar ‘Desiree’ in greenhouse conditions. Biotechnol. Agron. Soc. Environ. 2013, 17, 20–26. [Google Scholar]
  113. Parente, A.; Chambery, A.; Di Maro, A.; Russo, R.; Severino, V. Ribosome-inactivating Proteins from Phytolaccaceae. In Ribosome-Inactivating Proteins: Ricin and Related Proteins; Stirpe, F., Lappi, D.A., Eds.; Wiley: Ames, IA, USA, 2014; pp. 28–43. [Google Scholar] [CrossRef]
  114. Citores, L.; Ragucci, S.; Ferreras, J.M.; Di Maro, A.; Iglesias, R. Ageritin, a Ribotoxin from Poplar Mushroom (Agrocybe aegerita) with Defensive and Antiproliferative Activities. ACS Chem. Biol. 2019, 14, 1319–1327. [Google Scholar] [CrossRef] [PubMed]
  115. Ragucci, S.; Castaldi, S.; Landi, N.; Isticato, R.; Di Maro, A. Antifungal Activity of Ageritin, a Ribotoxin-like Protein from Cyclocybe aegerita Edible Mushroom, against Phytopathogenic Fungi. Toxins 2023, 15, 578. [Google Scholar] [CrossRef] [PubMed]
  116. Ragucci, S.; Landi, N.; Citores, L.; Iglesias, R.; Russo, R.; Clemente, A.; Saviano, M.; Pedone, P.; Chambery, A.; Ferreras, J.; et al. The Biological Action and Structural Characterization of Eryngitin 3 and 4, Ribotoxin-like Proteins from Pleurotus eryngii Fruiting Bodies. Int. J. Mol. Sci. 2023, 24, 14435. [Google Scholar] [CrossRef] [PubMed]
  117. Ready, M.; Brown, D.; Robertus, J. Extracellular localization of Pokeweed Antiviral Protein. Proc. Natl. Acad. Sci. USA 1986, 83, 5053–5056. [Google Scholar] [CrossRef] [PubMed]
  118. De Zaeytijd, J.; Van Damme, E. Extensive Evolution of Cereal Ribosome-Inactivating Proteins Translates into Unique Structural Features, Activation Mechanisms, and Physiological Roles. Toxins 2017, 9, 123. [Google Scholar] [CrossRef]
  119. Marshall, R.; D’Avila, F.; Di Cola, A.; Traini, R.; Spanò, L.; Fabbrini, M.; Ceriotti, A. Signal peptide-regulated toxicity of a plant ribosome-inactivating protein during cell stress. Plant J. 2011, 65, 218–229. [Google Scholar] [CrossRef]
  120. Kaur, S.; Samota, M.; Choudhary, M.; Pandey, A.; Sharma, A.; Thakur, J. How do plants defend themselves against pathogens-Biochemical mechanisms and genetic interventions. Physiol. Mol. Biol. Plants 2022, 28, 485–504. [Google Scholar] [CrossRef]
  121. Yang, T.; Meng, Y.; Chen, L.J.; Lin, H.-H.; Xi, D.-H. The Roles of Alpha-Momorcharinand Jasmonic Acid in Modulating the Response of Momordica charantia to Cucumber Mosaic Virus. Front. Microbiol. 2016, 7, 1796. [Google Scholar] [CrossRef]
  122. Qin, X.; Shao, C.; Hou, P.; Gao, J.; Lei, N.; Jiang, L.; Ye, S.; Gou, C.; Luo, S.; Zheng, X.; et al. Different functions and expression profiles of curcin and curcin-L in Jatropha curcas L. Z. Naturforsch. C J. Biosci. 2010, 65, 355–362. [Google Scholar] [CrossRef] [PubMed]
  123. Pieterse, C.; Van der Does, D.; Zamioudis, C.; Leon-Reyes, A.; Van Wees, S.; Schekman, R. Hormonal Modulation of Plant Immunity. Annu. Rev. Cell Dev. Biol. 2012, 28, 489–521. [Google Scholar] [CrossRef] [PubMed]
  124. Zhu, F.; Zhu, P.X.; Xu, F.; Che, Y.P.; Ma, Y.M.; Ji, Z.L. Alpha-momorcharin enhances Nicotiana benthamiana resistance to tobacco mosaic virus infection through modulation of reactive oxygen species. Mol. Plant Pathol. 2020, 21, 1212–1226. [Google Scholar] [CrossRef] [PubMed]
  125. Yang, T.; Zhu, L.-S.; Meng, Y.; Lv, R.; Zhou, Z.; Zhu, L.; Lin, H.-h.; Xi, D.-h. Alpha-momorcharin enhances Tobacco mosaic virus resistance in tobacco(NN) by manipulating jasmonic acid-salicylic acid crosstalk. J. Plant Physiol. 2018, 223, 116–126. [Google Scholar] [CrossRef] [PubMed]
  126. Smirnov, S.; Shulaev, V.; Tumer, N.E. Expression of pokeweed antiviral protein in transgenic plants induces virus resistance in grafted wild-type plants independently of salicylic acid accumulation and pathogenesis-related protein synthesis. Plant Physiol. 1997, 114, 1113–1121. [Google Scholar] [CrossRef] [PubMed]
  127. Vind, A.C.; Genzor, A.V.; Bekker-Jensen, S. Ribosomal stress-surveillance: Three pathways is a magic number. Nucleic Acids Res. 2020, 48, 10648–10661. [Google Scholar] [CrossRef] [PubMed]
  128. Arunachalam, C.; Doohan, F.M. Trichothecene toxicity in eukaryotes: Cellular and molecular mechanisms in plants and animals. Toxicol. Lett. 2013, 217, 149–158. [Google Scholar] [CrossRef] [PubMed]
  129. Moosa, A.; Farzand, A.; Sahi, S.; Khan, S. Transgenic expression of antifungal pathogenesis-related proteins against phytopathogenic fungi-15 years of success. Isr. J. Plant Sci. 2018, 65, 38–54. [Google Scholar] [CrossRef]
  130. Collinge, D.; Sarrocco, S. Transgenic approaches for plant disease control: Status and prospects 2021. Plant Pathol. 2022, 71, 207–225. [Google Scholar] [CrossRef]
Figure 1. Proposed mechanisms for the antifungal activity of RIPs. The infection causes the release of PAMPs that are recognized by PRRs and DAMPs which, in turn, are recognized by WAKs, leading to an increase in signal molecules, inducing RIP expression. Type 1 RIPs from dicots are synthesized in the endoplasmic reticulum and are localized in the apoplast. Infection by pathogens can alter the permeability of the host cell membrane, allowing RIP to enter the cytoplasm and inactivate ribosomes, leading to cell death, which prevents the spread of the pathogen. RIP can also pass through the cell wall and membrane of the fungus, inactivating its ribosomes and causing its death. In the case of cytosolic RIPs (Poaceae), these may be released as a consequence of fungal damage to the cell membrane. Chitinases and glucanases can degrade the fungal cell wall and favor RIP entry. RIP can also trigger fungal defense signaling pathways. The activation of these pathways could be a consequence of ribotoxic stress caused by RIPs.
Figure 1. Proposed mechanisms for the antifungal activity of RIPs. The infection causes the release of PAMPs that are recognized by PRRs and DAMPs which, in turn, are recognized by WAKs, leading to an increase in signal molecules, inducing RIP expression. Type 1 RIPs from dicots are synthesized in the endoplasmic reticulum and are localized in the apoplast. Infection by pathogens can alter the permeability of the host cell membrane, allowing RIP to enter the cytoplasm and inactivate ribosomes, leading to cell death, which prevents the spread of the pathogen. RIP can also pass through the cell wall and membrane of the fungus, inactivating its ribosomes and causing its death. In the case of cytosolic RIPs (Poaceae), these may be released as a consequence of fungal damage to the cell membrane. Chitinases and glucanases can degrade the fungal cell wall and favor RIP entry. RIP can also trigger fungal defense signaling pathways. The activation of these pathways could be a consequence of ribotoxic stress caused by RIPs.
Toxins 16 00192 g001
Figure 2. Strategies used for the construction of the T-DNA region of binary vectors for plant transformation with A. tumefaciens. Vectors with constitutive promoters, inducible promoters, and vectors expressing RIPs and chitinases or wheat germ agglutinin (WGA) have been designed. LB: left border, RB: right border, NOS: nopaline synthase, rbcS: rice rbcS gene, 35S: cauliflower mosaic virus (CMV), CFDV: coconut foliar decay virus, Ubi: ubiquitin, BAR: bar gene (resistance to glufosinate), NPT II: neomycin phosphotransferase II (resistance to kanamycin and neomycin), hph: hygromycin phosphotransferase gene (hygromycin resistance), Act1: rice actin 1, pwun1: promoter of the potato wun1 gene (wound-inducible), PGIP: bean polygalacturonase gene I promoter, PinII: 3’ region of the potato proteinase inhibitor II gene, CHI: chitinase, WGA: wheat germ agglutinin, HBT: HBT promoter (of the C4PPDK gene). The most used elements are written in bold.
Figure 2. Strategies used for the construction of the T-DNA region of binary vectors for plant transformation with A. tumefaciens. Vectors with constitutive promoters, inducible promoters, and vectors expressing RIPs and chitinases or wheat germ agglutinin (WGA) have been designed. LB: left border, RB: right border, NOS: nopaline synthase, rbcS: rice rbcS gene, 35S: cauliflower mosaic virus (CMV), CFDV: coconut foliar decay virus, Ubi: ubiquitin, BAR: bar gene (resistance to glufosinate), NPT II: neomycin phosphotransferase II (resistance to kanamycin and neomycin), hph: hygromycin phosphotransferase gene (hygromycin resistance), Act1: rice actin 1, pwun1: promoter of the potato wun1 gene (wound-inducible), PGIP: bean polygalacturonase gene I promoter, PinII: 3’ region of the potato proteinase inhibitor II gene, CHI: chitinase, WGA: wheat germ agglutinin, HBT: HBT promoter (of the C4PPDK gene). The most used elements are written in bold.
Toxins 16 00192 g002
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

Iglesias, R.; Citores, L.; Gay, C.C.; Ferreras, J.M. Antifungal Activity of Ribosome-Inactivating Proteins. Toxins 2024, 16, 192. https://doi.org/10.3390/toxins16040192

AMA Style

Iglesias R, Citores L, Gay CC, Ferreras JM. Antifungal Activity of Ribosome-Inactivating Proteins. Toxins. 2024; 16(4):192. https://doi.org/10.3390/toxins16040192

Chicago/Turabian Style

Iglesias, Rosario, Lucía Citores, Claudia C. Gay, and José M. Ferreras. 2024. "Antifungal Activity of Ribosome-Inactivating Proteins" Toxins 16, no. 4: 192. https://doi.org/10.3390/toxins16040192

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