Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2934 word(s) 2934 2021-02-19 07:41:23 |
2 format correct Meta information modification 2934 2021-02-25 10:36:51 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Ferreras, J.M. Ribosome-Inactivating Proteins. Encyclopedia. Available online: https://encyclopedia.pub/entry/7476 (accessed on 27 July 2024).
Ferreras JM. Ribosome-Inactivating Proteins. Encyclopedia. Available at: https://encyclopedia.pub/entry/7476. Accessed July 27, 2024.
Ferreras, José Miguel. "Ribosome-Inactivating Proteins" Encyclopedia, https://encyclopedia.pub/entry/7476 (accessed July 27, 2024).
Ferreras, J.M. (2021, February 22). Ribosome-Inactivating Proteins. In Encyclopedia. https://encyclopedia.pub/entry/7476
Ferreras, José Miguel. "Ribosome-Inactivating Proteins." Encyclopedia. Web. 22 February, 2021.
Ribosome-Inactivating Proteins
Edit

Ribosome-inactivating proteins (RIPs) are rRNA N-glycosylases from plants (EC 3.2.2.22) that inactivate ribosomes thus inhibiting protein synthesis.

adenine polynucleotide glycosylase antiviral therapy human virus immunotoxin

1. Introduction

One of the main efforts of virologists and molecular biologists is the search for antivirals that can help in the fight against viruses causing diseases in animals and especially in humans. Strategies are also being searched to tackle the challenge of plant viruses causing significant crop losses. This has led to the discovery of a number of antivirals with different chemical nature or proteins with different enzymatic activities [1][2]. The search for more effective and safer antivirals continues to be a field of intense investigation and plants are one of the most used sources, since they have evolved a variety of protein-based defense mechanisms to tackle viral infections [3]. Regarding ribosome-inactivating proteins (RIPs), it is worth noting the fact that one of the first RIPs to be purified was PAP (pokeweed antiviral protein) and although many RIPs have been purified as protein synthesis inhibitors, many others have been isolated as powerful antivirals. For many years, RIPs have been studied as potent inhibitors of protein synthesis that can be used for the construction of immunotoxins [4]. Since linked to a monoclonal antibody or a protein that specifically binds to a receptor, they can be used to specifically kill tumor cells [5][4]. RIPs have initially been studied as a family of proteins widely distributed among angiosperms although they have also been found in other taxons [6][7]. They irreversibly inactivate ribosomes inhibiting protein synthesis and thus causing cell death [6][7]. The first RIPs to be isolated, the extremely potent toxins ricin and abrin, were purified at the end of the nineteenth century and it was believed that their red cell agglutinating activity was the reason for the toxic effect [8][9]. In the early 1970s, it was reported that abrin, ricin, and PAP strongly inhibited protein synthesis in a cell-free rabbit reticulocyte system [8][9][10]; and Barbieri and Stirpe classified these and other related proteins as type 1 RIPs (a single polypeptide chain, such as PAP) and type 2 RIPs (two chains, an A chain similar to type 1 RIPs, and a B chain that possesses lectin activity, such as abrin and ricin) [4]. The enzymatic activity of ricin was discovered by Endo and colleagues, that is, RIPs are considered as 28S rRNA N-glycosylases (EC 3.2.2.22) that cleave the N-glycosidic bond between the adenine No. 4324 and its ribose in the 60S subunit of rat ribosomes [11] or the equivalent one in sensitive ribosomes from other organisms [12]. This adenine is located in the sarcin-ricin loop (SRL) that is crucial for anchoring the elongation factors EFG and EF2 on the ribosome during mRNA-tRNA translocation in prokaryotes and eukaryotes, respectively. This loop is also the target of ribotoxins such as α-sarcin, enzymes with rRNA endonuclease activity (EC 3.1.27.10) [13]. However, some RIPs are also able to remove more than an adenine from the rRNA [14] and many of them are able to deadenylate not only rRNA but also other polynucleotide substrates such as DNA, poly(A), mRNA, tRNA, and viral RNA [15], and because of this, the name of adenine polynucleotide glycosylase (or polynucleotide: adenosine glycosidase) was proposed for RIPs [15]. Additionally, other activities have been reported for RIPs, just as shown in Table 1.

Table 1. Proposed activities and other biological properties of ribosome-inactivating proteins (RIPs).

A convincing picture of the role played by these proteins in plants is not yet available. They seem to play different roles in different species, so antiviral, antifungal, plant defense, storage, programmed senescence, antifeedant, stress protection, and development regulation roles have been proposed for RIPs [7].

2. Activity on Animal (Human) Viruses

Global health threats such as the emergence of human viruses resistant to commonly used antiviral drugs, has prompted the study of RIPs as possible tools for fighting these agents. Antiviral activity of RIPs against different animal viruses has been reported (Table 2).

Table 2. RIPs active against animal viruses. RIPs with antiviral activity, the families and species from which they have been obtained and the viruses in which this activity has been demonstrated are shown.

Species and RIP Virus
POACEAE  
Zea mays L.  
Maize RIP HIV, SHIV
EUPHORBIACEAE  
Ricinus communis L.  
Ricin A chain HIV
Suregada multiflora (A.Juss.) Baill. (=Gelonium multiflorum A.Juss.)
Gelonin HIV, HPV, HSV, PICV,
GAP31 HIV
CUCURBITACEAE  
Trichosanthes kirilowii Maxim  
Trichosanthin (TCS) HBV, HIV, HSV
TAP29 HIV
Trichobitacin HIV
Momordica charantia L.  
Momordin (M. charantia inhibitor) HPV, HSV
Alpha-momorcharin (α-MMC) HBV, HIV, HSV
Beta-momorcharin HIV
Momordica antiviral protein (MAP30) DENV-2, HHV8, HBV, HIV, HSV
Momordica balsamina L.  
Balsamin HIV
Luffa cylindrica (L.) M.Roem.  
Luffin HIV
Bryonia cretica subsp. dioica (Jacq.) Tutin (=Bryonia dioica Jacq.)
Bryodin HIV
CARYOPHYLLACEAE  
Saponaria officinalis L.  
Saporin HIV
Dianthus caryophyllus L.  
Dianthin 32 (DAP32) HIV, HPV, HSV
Dianthin 30 (DAP30) HIV
Agrostemma githago L.  
Agrostin HIV
PHYTOLACCACEAE  
Phytolacca americana L.  
PAP (PAPI) CHIKV, FLUV, HBV, HIV, HPV,
  HSV, HTLV, JEV, LCMV
PAPII HIV
PAPIII HIV
PAP-S HSV, HPV, HBV

Virus name abbreviations: CHIKV (chikungunya virus), DENV (dengue virus), FLUV (human influenza virus), HBV (hepatitis B virus), HHV (human gammaherpesvirus), HIV (human immunodeficiency virus), HPV (human poliovirus), HSV (herpes simplex virus), HTLV (human T-cell leukemia virus), JEV (Japanese encephalitis virus), LCMV (lymphocytic choriomeningitis virus), PICV (Pichinde virus), SHIV (simian–human immunodeficiency virus).

RIPs with antiviral activity belong to the main types of RIPs found in angiosperms [7]: monocot type 1 RIPs (Poaceae), dicot type 1 RIPs (Euphorbiaceae, Caryophyllaceae, Phytolaccaceae), type 2 RIPs (ricin, Euphorbiaceae), and type 1 RIPs derived from type 2 RIPs (Cucurbitaceae); which suggests that all these proteins could have, to a greater or lesser extent, antiviral activity and that their main biological role could be precisely the defense of the plant against viruses. However, researchers have focused on the study of proteins obtained from species of the families Phytolaccaceae, Cucurbitaceae, Caryophyllaceae, and Euphorbiaceae; and the most studied RIPs are pokeweed antiviral protein (PAP), trichosanthin (TCS) and Momordica antiviral protein (MAP30), which have been the subject of recent reviews [10][35][36][38][58]. It is noteworthy that RIPs have shown to be active against viruses of very different nature: double-stranded (ds) DNA viruses (hepatitis B virus, HBV; human gammaherpesvirus, HHV; human poliovirus, HPV; herpes simplex virus, HSV), retroviruses (human immunodeficiency virus, HIV; human T-cell leukemia virus, HTLV; simian–human immunodeficiency virus, SHIV), positive-sense single-stranded (ss) RNA viruses (Japanese encephalitis virus, JEV; dengue virus, DENV; chikungunya virus, CHIKV), and negative-sense (ss) RNA viruses (human influenza virus, FLUV; lymphocytic choriomeningitis virus, LCMV; Pichinde virus, PICV). Most of the viruses studied are enveloped viruses that infect humans, with the exceptions of the simian–human immunodeficiency virus (SHIV), the Pichinde virus (PICV), and the non-enveloped human poliovirus. This virus was the first in which activity against an animal virus was reported [59]. Results obtained with HEp-2 cells infected with human poliovirus or herpes simplex virus (HSV) showed that gelonin, momordin, dianthin 32, and PAP-S impaired viral replication by inhibiting protein synthesis in virus-infected cells, in which presumably they enter more easily than in uninfected cells [30], suggesting that antiviral activity could be a general property of RIPs.

2.1. Activity on Human Immunodeficiency Virus

The most studied virus is the human immunodeficiency virus (HIV). The lack of effective antivirals against this virus and its rapid spread around the world prompted studies on the activity of RIPs against this virus since 1989 [60]. At least 20 RIPs have shown activity against HIV (Table 2). Thus, several RIPs obtained from Euphorbiaceae and Caryophylaceae, but mostly from Cucurbitaceae and Phytolocaceae, inhibit the replication of HIV in vitro [35]. It has also been reported that maize RIP transiently reduces viral load in SHIV infected Chinese rhesus macaques [27]. The results obtained with RIPs promoted their use in clinical trials [61]. Although the development of specific HIV antivirals such as reverse-transcriptase and protease inhibitors have directed AIDS therapy to other treatments, these studies demonstrated the potential of RIPs for the treatment of virus-related diseases.

2.2. Activity on Herpes Simplex Virus

Another virus that has been targeted by RIPs is the herpes simplex virus (HSV). Currently, there is no treatment that completely eliminates HSV infection from the body, because once the virus enters an organism, it remains dormant until reactivated. This has encouraged researchers to study RIPs as candidates for HSV therapy. Gelonin, trichosanthin, dianthin 32, PAP, PAP-S, and several RIPs obtained from Momordica charantia have shown anti-HSV activity in vitro (Table 2).

2.3. Activity on Other Animal Viruses

Exposure of HepG2.2.15 cells to MAP30 [44], PAP-S [56], α-momorcharin [41], and an eukaryotic expression plasmid encoding PAP [56] inhibits the production of hepatitis B virus (HBV). Additionally, an extract from Radix Trichosanthis had a stronger inhibitive effect on expression of HBsAg and HBeAg in HepG2.2.15, and trichosanthin has been proposed as the main component of the aqueous extract responsible for the anti-hepatitis B viral effect [62].

On the other hand, it has also been reported that PAP inhibits replication of human T-cell leukemia (HTLV), human influenza, chikungunya (CHIKV), Japanese encephalitis (JEV), and lymphocytic choriomeningitis (LCMV) viruses, gelonin inhibits Pichinde virus replication, and MAP30 inhibits human gammaherpesvirus 8 (HHV8) and dengue virus [10][31][35][42][52][53][54][55].

2.4. Citotoxicity of RIPs

An important aspect to consider when working with antivirals is their cytotoxicity. In this sense, type 1 RIPs and type 2 RIPs can be distinguished. Type 1 RIPs consist of a polypeptide chain with rRNA N-glycosylase activity, while type 2 RIPs are constituted by two chains linked by a disulfide bond: The A chain (active) is equivalent to a type 1 RIP and the B chain (binding) is a lectin able to bind to membrane glycoproteins and glycolipids allowing endocytosis of RIP by cells. This is why RIPs such as ricin and abrin are extremely toxic showing IC50 (concentration that inhibits protein synthesis by 50%) values of 0.67–8 pM in cell cultures [63]. There are type 2 RIPs such as those from Sambucus which are much less toxic to cultured cells with IC50 values of 27–64 nM [64]. Type 1 RIPs are much less toxic and have highly variable IC50 values (0.2–10 μM) [63]. Due to the high cytotoxicity of type 2 RIPs, only type 1 RIPs or the ricin A-chain (which has a cytotoxicity similar to that of type 1 RIPs) [63] have been used as antiviral agents.

A good antiviral should display a substantial difference between the antiviral concentration and the cytotoxic concentration. Due to the diverse toxicities of type 1 RIPs, there are also differences in this regard, but the most commonly used proteins such as PAP, MAP30, or trichosanthin always show a substantial difference between toxic concentrations for cells (3–30 μM) [63][65][66] and concentrations that have antiviral activity (around 30 nM) [35].

Finally, it should be noted that some bacterial and fungal enzymes targeting the sarcin-ricin loop have also been reported to possess antiviral activity [2][67][68][69][70][71][72][73].

Therefore, RIPs have awakened over many years, and continue to do so, a keen interest as tools to fight viruses that cause diseases in humans. In fact, recently saporin and RTAM-PAP1 (a chimera constructed with ricin A-chain and PAP) have been proposed as candidates for therapy of COVID-19 [74][75].

3. Activity against Plant Viruses

To date, 39 RIPs have been described that display some type of activity against plant viruses (Table 3).

Table 3. RIPs active against plant viruses. RIPs with antiviral activity, the families and species from which they have been obtained and the viruses in which this activity has been demonstrated are shown.

 

 

These RIPs have been found in 26 plant species belonging to one family of monocotyledons and ten families of dicotyledons, that are distributed throughout the phylogenetic tree of angiosperms in a similar way to the RIP-containing plants [7], thus suggesting that most RIPs could be active against plant viruses. As a matter of fact, only two type 2 RIPs from Sambucus nigra (SNAI and SNLRP) have been reported to fail to protect transgenic plants against viral infection [76].

Despite the fact that these antiviral proteins are distributed in a great variety of families, most of them (thirty one) belong to the orders Caryophyllales and Lamiales (families Caryophyllaceae, Amaranthaceae, Phytolaccaceae, Nyctaginaceae, Basellaceae, Lamiaceae), which are RIPs with well-defined structural and phylogenetic characteristics [7].

RIPs seem to be active against a wide range of viruses (Table 3), all of them belonging to different families of positive-sense single-stranded (ss) RNA viruses. The exception is the geminivirus ACMV (African cassava mosaic virus), which contains a single-stranded circular DNA genome. They seem to protect all kinds of plants and, although the most commonly used plant for testing has been Nicotiana tabacum L., RIPs have also shown ability to protect other species of the genus Nicotiana (N. benthamiana Domin and N. glutinosa L.) as well as other species commonly used in research or crops such as Brassica rapa L. (=B. parachinensis L.H.Bailey) (choy sum), Cyamopsis tetragonoloba (L.) Taub. (guar), Crotalaria juncea L. (sunn hemp). Phaseolus vulgaris L. (common bean), Momordica charantia L. (bitter melon), Beta vulgaris L. (sugar beet), Cucurbita pepo L. (squash), Solanum tuberosum L. (potato), Carica papaya L. (papaya), Chenopodium quinoa Willd. (quinoa), or Lycopersicon esculentum Mill. (tomato).

It is difficult to compare the antiviral activity of the different RIPs because different criteria have been used to evaluate their antiviral capacity. In some cases, the putative antiviral character is based on their N-glycosylase activity on the virus genome [105]; all RIPs are able to release adenines from any kind of RNA or DNA, including viral genomes [4]. This adenine polynucleotide glycosylase activity has been detected by electrophoresis [87], or HPLC [103][105]. In many cases, the test has involved applying a RIP solution on the leaf surface of the plant together with the virus and comparing the result with the control that does not contain RIP. In some cases, the virus is applied simultaneously [86][92][113] and in others, sometime after the application of the RIP [90][115]. The evaluation of antiviral activity has been done by counting the number of lesions [88][93], the time of onset of symptoms [77][79], the number of infected plants [105], or the severity of the infection symptoms [78][115]. Virus levels have also been estimated by ELISA [99], Western blotting analysis [81], RT-PCR analysis [101], quantitative real-time PCR analysis [81][82], electron microscopy [92], or by determining the infection capacity of an extract from the infected plant [92]. Another approach has been the construction of virus-resistant transgenic plants [80][102]. The virus has been inoculated mechanically or by aphids [102] and the resistance has been determined by one of the methods listed above.

Other studies link RIPs to the defense of plants against viruses, especially studies of induction of RIPs through signaling compounds such as salicylic acid, hydrogen peroxide, or jasmonic acid, which are involved in the systemic acquired resistance (SAR) of plants against viruses and other pathogens. Thus, it has been reported that artichoke mottled crinkle virus (AMCV), salicylic acid, and hydrogen peroxide induce the expression of BE27 in both treated and untreated leaves of sugar beet plant [86][117]. On the other hand, it has been reported that alpha-momorcharin induces the generation of salicylic acid, jasmonic acid, and reactive oxygen species, which improve tobacco mosaic virus (TMV) tolerance [118]. Additionally, alpha-momorcharin induces the expression of the N gene [118], which encodes the N protein that recognizes the TMV replicase fragment and triggers signal transduction cascades, initiating a hypersensitive response (HR) and inhibiting the spread of TMV [118]. Other RIPs in which some type of elicitor activity has been reported are pokeweed antiviral protein II (PAPII) [104], CIP-29 [111], and CA-SRI [113][115]. By contrast, the antiviral activity of SNAI’ [116], IRIP and IRAb [77], and nigrin b [76] is not accompanied by an induction of pathogenesis-related proteins. All this suggests that some, but not all RIPs, could be part of the SAR or/and HR to defend the plant against viral infections.

References

  1. De Clercq, E. Looking Back in 2009 at the Dawning of Antiviral Therapy Now 50 Years Ago: An Historical Perspective. In Advances in Virus Research; Maramorosch, K., Shatkin, A.J., Murphy, F.A., Eds.; Elsevier Academic Press Inc: San Diego, CA, USA, 2009; Volume 73, pp. 1–53.
  2. Ng, T.; Cheung, R.C.F.; Wong, J.H.; Chan, W.-Y. Proteins, peptides, polysaccharides, and nucleotides with inhibitory activity on human immunodeficiency virus and its enzymes. Appl. Microbiol. Biotechnol. 2015, 99, 10399–10414, doi:10.1007/s00253-015-6997-z.
  3. Musidlak, O.; Nawrot, R.; Goździcka-Józefiak, A. Which Plant Proteins Are Involved in Antiviral Defense? Review on In Vivo and In Vitro Activities of Selected Plant Proteins against Viruses. Int. J. Mol. Sci. 2017, 18, 2300, doi:10.3390/ijms18112300.
  4. Bolognesi, A.; Bortolotti, M.; Maiello, S.; Battelli, M.G.; Polito, L. Ribosome-Inactivating Proteins from Plants: A Historical Overview. Molecules 2016, 21, 1627, doi:10.3390/molecules21121627.
  5. Ferreras, J.M.; Citores, L.; Iglesias, R.; Jiménez, P.; Girbés, T. Use of Ribosome-Inactivating Proteins from Sambucus for the Construction of Immunotoxins and Conjugates for Cancer Therapy. Toxins 2011, 3, 420–441, doi:10.3390/toxins3050420.
  6. Stirpe, F. Ribosome-inactivating proteins: From toxins to useful proteins. Toxicon 2013, 67, 12–16, doi:10.1016/j.toxicon.2013.02.005.
  7. Di Maro, A.; Citores, L.; Russo, R.; Iglesias, R.; Ferreras, J.M. Sequence comparison and phylogenetic analysis by the Maximum Likelihood method of ribosome-inactivating proteins from angiosperms. Plant Mol. Biol. 2014, 85, 575–588, doi:10.1007/s11103-014-0204-y.
  8. Olsnes, S. The history of ricin, abrin and related toxins. Toxicon 2004, 44, 361–370, doi:10.1016/j.toxicon.2004.05.003.
  9. Polito, L.; Bortolotti, M.; Battelli, M.G.; Calafato, G.; Bolognesi, A. Ricin: An Ancient Story for a Timeless Plant Toxin. Toxins 2019, 11, 324, doi:10.3390/toxins11060324.
  10. Domashevskiy, A.V.; Goss, D.J. Pokeweed Antiviral Protein, a Ribosome Inactivating Protein: Activity, Inhibition and Prospects. Toxins 2015, 7, 274–298, doi:10.3390/toxins7020274.
  11. 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, doi:10.1016/s0021-9258(18)68367-x.
  12. Iglesias, R.; Citores, L.; Ferreras, J.M. Ribosomal RNA N-glycosylase Activity Assay of Ribosome-inactivating Proteins. Bio-Protocol 2017, 7, e2180, doi:10.21769/bioprotoc.2180.
  13. 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, doi:10.1021/acschembio.9b00291.
  14. Barbieri, L.; Ferreras, J.M.; Barraco, A.; Ricci, P.; Stirpe, F. Some ribosome-inactivating proteins depurinate ribosomal RNA at multiple sites. Biochem. J. 1992, 286, 1–4, doi:10.1042/bj2860001.
  15. 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, doi:10.1093/nar/25.3.518.
  16. Ogasawara, T.; Sawasaki, T.; Morishita, R.; Ozawa, A.; Madin, K.; Endo, Y. A new class of enzyme acting on damaged ribosomes: Ribosomal RNA apurinic site specific lyase found in wheat germ. EMBO J. 1999, 18, 6522–6531, doi:10.1093/emboj/18.22.6522.
  17. Choudhary, N.L.; Yadav, O.P.; Lodha, M.L. Ribonuclease, deoxyribonuclease, and antiviral activity of Escherichia coli-expressed Bougainvillea xbuttiana antiviral protein-1. Biochemistry 2008, 73, 273–277, doi:10.1134/s000629790803005x.
  18. Chen, H.; Wang, Y.; Yan, M.; Yu, M.; Yao, Q. 5’-AMP Phosphatase activity on trichosanthin and other single chain ribosome inactivating proteins. Chin. Biochem. J. 1996, 12, 125–130.
  19. Li, X.; Chen, W.-F.; Liu, W.-Y.; Wang, G.-H. Large-Scale Preparation of Two New Ribosome-Inactivating Proteins—Cinnamomin and Camphorin from the Seeds of Cinnamomum camphora. Protein Expr. Purif. 1997, 10, 27–31, doi:10.1006/prep.1996.0706.
  20. Helmy, M.; Lombard, S.; Piéroni, G. Ricin RCA60: Evidence of Its Phospholipase Activity. Biochem. Biophys. Res. Commun. 1999, 258, 252–255, doi:10.1006/bbrc.1999.0618.
  21. Shih, N.; McDonald, K.A.; Jackman, A.P.; Girbés, T.; Iglesias, R. Bifunctional plant defence enzymes with chitinase and ribosome inactivating activities from Trichosanthes kirilowii cell cultures. Plant Sci. 1997, 130, 145–150, doi:10.1016/s0168-9452(97)00229-x.
  22. Iglesias, R.; Citores, L.; Di Maro, A.; Ferreras, J.M. Biological activities of the antiviral protein BE27 from sugar beet (Beta vulgaris L.). Planta 2014, 241, 421–433, doi:10.1007/s00425-014-2191-2.
  23. Polito, L.; Bortolotti, M.; Pedrazzi, M.; Mercatelli, D.; Battelli, M.G.; Bolognesi, A. Apoptosis and necroptosis induced by stenodactylin in neuroblastoma cells can be completely prevented through caspase inhibition plus catalase or necrostatin-1. Phytomedicine 2016, 23, 32–41, doi:10.1016/j.phymed.2015.11.006.
  24. Panda, P.K.; Behera, B.; Meher, B.R.; Das, D.N.; Mukhopadhyay, S.; Sinha, N.; Naik, P.P.; Roy, B.; Das, J.; Paul, S.; et al. Abrus Agglutinin, a type II ribosome inactivating protein inhibits Akt/PH domain to induce endoplasmic reticulum stress mediated autophagy-dependent cell death. Mol. Carcinog. 2016, 56, 389–401, doi:10.1002/mc.22502.
  25. Rustgi, S.; Pollmann, S.; Buhr, F.; Springer, A.; Reinbothe, C.; Von Wettstein, D.; Reinbothe, S. JIP60-mediated, jasmonate- and senescence-induced molecular switch in translation toward stress and defense protein synthesis. Proc. Natl. Acad. Sci. USA 2014, 111, 14181–14186, doi:10.1073/pnas.1415690111.
  26. Przydacz, M.; Jones, R.; Pennington, H.G.; Belmans, G.; Bruderer, M.; Greenhill, R.; Salter, T.; Wellham, P.A.; Cota, E.; Spanu, P.D. Mode of Action of the Catalytic Site in the N-Terminal Ribosome-Inactivating Domain of JIP60. Plant Physiol. 2020, 183, 385–398, doi:10.1104/pp.19.01029.
  27. Wang, R.-R.; Au, K.-Y.; Zheng, H.-Y.; Gao, L.-M.; Zhang, X.; Luo, R.-H.; Law, S.K.-Y.; Mak, A.N.-S.; Wong, K.-B.; Zhang, M.-X.; et al. The Recombinant Maize Ribosome-Inactivating Protein Transiently Reduces Viral Load in SHIV89.6 Infected Chinese Rhesus Macaques. Toxins 2015, 7, 156–169, doi:10.3390/toxins7010156.
  28. Law, S.K.-Y.; Wang, R.-R.; Mak, A.N.-S.; Wong, K.-B.; Zheng, Y.-T.; Shaw, P.C. A switch-on mechanism to activate maize ribosome-inactivating protein for targeting HIV-infected cells. Nucleic Acids Res. 2010, 38, 6803–6812, doi:10.1093/nar/gkq551.
  29. Au, K.-Y.; Wang, R.-R.; Wong, Y.-T.; Wong, K.-B.; Zheng, Y.-T.; Shaw, P.C. Engineering a switch-on peptide to ricin A chain for increasing its specificity towards HIV-infected cells. Biochim. Biophys. Acta Gen. Subj. 2014, 1840, 958–963, doi:10.1016/j.bbagen.2013.11.005.
  30. Foà-Tomasi, L.; Campadelli-Fiume, G.; Barbieri, L.; Stirpe, F. Effect of ribosome-inactivating proteins on virus-infected cells. Inhibition of virus multiplication and of protein synthesis. Arch. Virol. 1982, 71, 323–332, doi:10.1007/bf01315062.
  31. Barnett, B.B.; Burns, N.J.; Park, K.J.; Dawson, M.I.; Kende, M.; Sidwell, R.W. Antiviral immunotoxins: antibody-mediated delivery of gelonin inhibits Pichinde virus replication in vitro. Antivir. Res. 1991, 15, 125–138, doi:10.1016/0166-3542(91)90030-u.
  32. Au, T.K.; Collins, R.A.; Lam, T.L.; Ng, T.B.; Fong, W.P.; Wan, D. The plant ribosome inactivating proteins luffin and saporin are potent inhibitors of HIV-1 integrase. FEBS Lett. 2000, 471, 169–172, doi:10.1016/s0014-5793(00)01389-2.
  33. Li, H.-G.; Huang, P.L.; Zhang, D.; Sun, Y.; Chen, H.-C.; Zhang, J.; Huang, P.L.; Kong, X.-P.; Lee-Huang, S. A new activity of anti-HIV and anti-tumor protein GAP31: DNA adenosine glycosidase – Structural and modeling insight into its functions. Biochem. Biophys. Res. Commun. 2010, 391, 340–345, doi:10.1016/j.bbrc.2009.11.060.
  34. Lee-Huang, S.; Kung, H.-F.; Huang, P.L.; Huang, P.L.; Li, B.-Q.; Huang, P.; Huang, H.I.; Chen, H.-C. A new class of anti-HIV agents: GAP31, DAPs 30 and 32. FEBS Lett. 1991, 291, 139–144, doi:10.1016/0014-5793(91)81122-o.
  35. Kaur, I.; Gupta, R.C.; Puri, M. Ribosome inactivating proteins from plants inhibiting viruses. Virol. Sin. 2011, 26, 357–365, doi:10.1007/s12250-011-3223-8.
  36. Fang, E.F.; Ng, T.B.; Shaw, P.C.; Wong, R.N.S. Recent Progress in Medicinal Investigations on Trichosanthin and other Ribo-some Inactivating Proteins from the Plant Genus Trichosanthes. Curr. Med. Chem. 2011, 18, 4410–4417.
  37. He, D.-X.; Tam, S.-C. Trichosanthin affects HSV-1 replication in Hep-2 cells. Biochem. Biophys. Res. Commun. 2010, 402, 670–675, doi:10.1016/j.bbrc.2010.10.080.
  38. Shi, W.-W.; Wong, K.-B.; Shaw, P.C. Structural and Functional Investigation and Pharmacological Mechanism of Trichosanthin, a Type 1 Ribosome-Inactivating Protein. Toxins 2018, 10, 335, doi:10.3390/toxins10080335.
  39. Zheng, Y.T.; Ben, K.L.; Jin, S.W. Anti-HIV-1 activity of trichobitacin, a novel ribosome-inactivating protein. Acta Pharmacol. Sin. 2000, 21, 179–182.
  40. Meng, Y.; Liu, S.; Li, J.; Zhao, X. Preparation of an antitumor and antivirus agent: Chemical modification of α-MMC and MAP30 from Momordica Charantia L. with covalent conjugation of polyethyelene glycol. Int. J. Nanomed. 2012, 7, 3133–3142, doi:10.2147/ijn.s30631.
  41. Yao, X.; Li, J.; Deng, N.; Wang, S.; Meng, Y.; Shen, F. Immunoaffinity purification of α-momorcharin from bitter melon seeds (Momordica charantia). J. Sep. Sci. 2011, 34, 3092–3098, doi:10.1002/jssc.201100235.
  42. Rothan, H.A.; Bahrani, H.; Mohamed, Z.; Rahman, N.A.; Yusof, R. Fusion of Protegrin-1 and Plectasin to MAP30 Shows Significant Inhibition Activity against Dengue Virus Replication. PLoS ONE 2014, 9, e94561, doi:10.1371/journal.pone.0094561.
  43. Sun, Y.; Huang, P.L.; Li, J.J.; Huang, Y.Q.; Zhang, L.; Huang, P.L.; Lee-Huang, S. Anti-HIV Agent MAP30 Modulates the Expression Profile of Viral and Cellular Genes for Proliferation and Apoptosis in AIDS-Related Lymphoma Cells Infected with Kaposi’s Sarcoma-Associated Virus. Biochem. Biophys. Res. Commun. 2001, 287, 983–994, doi:10.1006/bbrc.2001.5689.
  44. Fan, J.M.; Zhang, Q.; Xu, J.; Zhu, S.; Ke, T.; Gao, D.F.; Xu, Y.B. Inhibition on Hepatitis B virus in vitro of recombinant MAP30 from bitter melon. Mol. Biol. Rep. 2007, 36, 381–388, doi:10.1007/s11033-007-9191-2.
  45. Lee-Huang, S.; Huang, P.L.; Bourinbaiar, A.S.; Chen, H.C.; Kung, H.F. Inhibition of the integrase of human immunodeficiency virus (HIV) type 1 by anti-HIV plant proteins MAP30 and GAP31. Proc. Natl. Acad. Sci. USA 1995, 92, 8818–8822, doi:10.1073/pnas.92.19.8818.
  46. Bourinbaiar, A.S.; Lee-Huang, S. The Activity of Plant-Derived Antiretroviral Proteins MAP30 and GAP31 against Herpes Simplex Virus Infectionin Vitro. Biochem. Biophys. Res. Commun. 1996, 219, 923–929, doi:10.1006/bbrc.1996.0334.
  47. Kaur, I.; Puri, M.; Ahmed, Z.; Blanchet, F.P.; Mangeat, B.; Piguet, V. Inhibition of HIV-1 Replication by Balsamin, a Ribosome Inactivating Protein of Momordica balsamina. PLoS ONE 2013, 8, e73780, doi:10.1371/journal.pone.0073780.
  48. Wachinger, M.; Samtleben, R.; Gerhauser, C.; Wagner, H.; Erfle, V. Bryodin, a single-chain ribosome-inactivating protein, selectively inhibits the growth of HIV-1-infected cells and reduces HIV-1 production. Res. Exp. Med. 1993, 193, 1–12, doi:10.1007/bf02576205.
  49. Leitman, E.M.; Palmer, C.D.; Buus, S.; Chen, F.; Riddell, L.; Sims, S.; Klenerman, P.; Saez-Cirion, A.; Walker, B.D.; Hess, P.R.; et al. Saporin-conjugated tetramers identify efficacious anti-HIV CD8+ T-cell specificities. PLoS ONE 2017, 12, e0184496, doi:10.1371/journal.pone.0184496.
  50. Yadav, S.K.; Batra, J.K. Mechanism of Anti-HIV Activity of Ribosome Inactivating Protein, Saporin. Protein Pept. Lett. 2015, 22, 497–503, doi:10.2174/0929866522666150428120701.
  51. Vivanco, J.M.; Tumer, N.E. Translation Inhibition of Capped and Uncapped Viral RNAs Mediated by Ribosome-Inactivating Proteins. Phytopathology 2003, 93, 588–595, doi:10.1094/phyto.2003.93.5.588.
  52. Mansouri, S.; Choudhary, G.; Sarzala, P.M.; Ratner, L.; Hudak, K.A. Suppression of Human T-cell Leukemia Virus I Gene Expression by Pokeweed Antiviral Protein. J. Biol. Chem. 2009, 284, 31453–31462, doi:10.1074/jbc.m109.046235.
  53. Rothan, H.A.; Bahrani, H.; Shankar, E.M.; Rahman, N.A.; Yusof, R. Inhibitory effects of a peptide-fusion protein (Latarcin–PAP1–Thanatin) against chikungunya virus. Antivir. Res. 2014, 108, 173–180, doi:10.1016/j.antiviral.2014.05.019.
  54. Ishag, H.Z.A.; Li, C.; Huang, L.; Sun, M.-X.; Ni, B.; Guo, C.-X.; Mao, X. Inhibition of Japanese encephalitis virus infection in vitro and in vivo by pokeweed antiviral protein. Virus Res. 2013, 171, 89–96, doi:10.1016/j.virusres.2012.10.032.
  55. Uckun, F.M.; Rustamova, L.; Vassilev, A.O.; Tibbles, H.E.; Petkevich, A.S. CNS activity of Pokeweed Anti-viral Protein (PAP) in mice infected with Lymphocytic Choriomeningitis Virus (LCMV). BMC Infect. Dis. 2005, 5, 9, doi:10.1186/1471-2334-5-9.
  56. He, Y.-W.; Guo, C.-X.; Pan, Y.-F.; Peng, C.; Weng, Z.-H. Inhibition of hepatitis B virus replication by pokeweed antiviral protein in vitro. World J. Gastroenterol. 2008, 14, 1592–1597, doi:10.3748/wjg.14.1592.
  57. Rajamohan, F.; Venkatachalam, T.K.; Irvin, J.D.; Uckun, F.M. Pokeweed Antiviral Protein Isoforms PAP-I, PAP-II, and PAP-III Depurinate RNA of Human Immunodeficiency Virus (HIV)-1. Biochem. Biophys. Res. Commun. 1999, 260, 453–458, doi:10.1006/bbrc.1999.0922.
  58. Di, R.; Tumer, N.E. Pokeweed Antiviral Protein: Its Cytotoxicity Mechanism and Applications in Plant Disease Resistance. Toxins 2015, 7, 755–772, doi:10.3390/toxins7030755.
  59. Ussery, M.A.; Irvin, J.D.; Hardesty, B. Inhibition of Poliovirus Replication by A Plant Antiviral Peptide. Ann. N. Y. Acad. Sci. 1977, 284, 431–440, doi:10.1111/j.1749-6632.1977.tb21979.x.
  60. McGrath, M.S.; Hwang, K.M.; Caldwell, S.E.; Gaston, I.; Luk, K.C.; Wu, P.; Ng, V.L.; Crowe, S.; Daniels, J.; Marsh, J. GLQ223: An inhibitor of human immunodeficiency virus replication in acutely and chronically infected cells of lymphocyte and mononuclear phagocyte lineage. Proc. Natl. Acad. Sci. USA 1989, 86, 2844–2848, doi:10.1073/pnas.86.8.2844.
  61. Byers, V.S.; Levin, A.S.; Waites, L.A.; Starrett, B.A.; Mayer, R.A.; Clegg, J.A.; Price, M.R.; Robins, R.A.; Delaney, M.; Baldwin, R.W. A phase I/II study of trichosanthin treatment of HIV disease. AIDS 1990, 4, 1189–1196, doi:10.1097/00002030-199012000-00002.
  62. Wen, D.; Wang, J.; Yan, H.; Chen, J.; Xia, K.; Liu, J.; Zhang, A. Effect of Radix Trichosanthis and Trichosanthin on Hepatitis B Virus in HepG2.2.15 Cells. J. Nanosci. Nanotechnol. 2015, 15, 2094–2098, doi:10.1166/jnn.2015.9271.
  63. Barbieri, L.; Battelli, M.G.; Stirpe, F. Ribosome-inactivating proteins from plants. Biochim. Biophys. Acta Rev. Biomembr. 1993, 1154, 237–282, doi:10.1016/0304-4157(93)90002-6.
  64. Citores, L.; Iglesias, R.; Ferreras, J.M. Ribosome Inactivating Proteins from Plants: Biological Properties and their Use in Experimental Therapy. In Antitumor Potential and Other Emerging Medicinal Properties of Natural Compounds; Fang, E.F., Ng, T.B., Eds.; Springer: Dordrecht, The Netherlands. 2013; pp. 127–143.
  65. Lv, Q.; Yang, X.-Z.; Fu, L.-Y.; Lu, Y.-T.; Lu, Y.-H.; Zhao, J.; Wang, F.-J. Recombinant expression and purification of a MAP30-cell penetrating peptide fusion protein with higher anti-tumor bioactivity. Protein Expr. Purif. 2015, 111, 9–17, doi:10.1016/j.pep.2015.03.008.
  66. Lin, B.; Yang, X.-Z.; Cao, X.-W.; Zhang, T.-Z.; Wang, F.-J.; Zhao, J. A novel trichosanthin fusion protein with increased cytotoxicity to tumor cells. Biotechnol. Lett. 2016, 39, 71–78, doi:10.1007/s10529-016-2222-0.
  67. Ferens, W.A.; Hovde, C.J. Antiviral Activity of Shiga Toxin 1: Suppression of Bovine Leukemia Virus-Related Spontaneous Lymphocyte Proliferation. Infect. Immun. 2000, 68, 4462–4469, doi:10.1128/iai.68.8.4462-4469.2000.
  68. Shi, P.L.; Binnington, B.; Sakac, D.; Katsman, Y.; Ramkumar, S.; Gariépy, J.; Kim, M.; Branch, D.R.; Lingwood, C. Verotoxin A Subunit Protects Lymphocytes and T Cell Lines against X4 HIV Infection in Vitro. Toxins 2012, 4, 1517–1534, doi:10.3390/toxins4121517.
  69. Yadav, S.K.; Batra, J.K. Ribotoxin restrictocin manifests anti-HIV-1 activity through its specific ribonuclease activity. Int. J. Biol. Macromol. 2015, 76, 58–62, doi:10.1016/j.ijbiomac.2015.01.062.
  70. Wong, J.H.; Wang, H.X.; Ng, T.B. Marmorin, a new ribosome inactivating protein with antiproliferative and HIV-1 reverse transcriptase inhibitory activities from the mushroom Hypsizigus marmoreus. Appl. Microbiol. Biotechnol. 2008, 81, 669–674, doi:10.1007/s00253-008-1639-3.
  71. Wang, H.; Ng, T.B. Isolation and characterization of velutin, a novel low-molecular-weight ribosome-inactivating protein from winter mushroom (Flammulina velutipes) fruiting bodies. Life Sci. 2001, 68, 2151–2158, doi:10.1016/s0024-3205(01)01023-2.
  72. Lam, S.; Ng, T. First Simultaneous Isolation of a Ribosome Inactivating Protein and an Antifungal Protein from a Mushroom (Lyophyllum shimeji) Together with Evidence for Synergism of their Antifungal Effects. Arch. Biochem. Biophys. 2001, 393, 271–280, doi:10.1006/abbi.2001.2506.
  73. Yao, Q.Z.; Yu, M.M.; Ooi, L.S.; Ng, T.B.; Chang, S.T.; Sun, S.S.; Ooi, V.E. Isolation and Characterization of a Type 1 Ribosome-Inactivating Protein from Fruiting Bodies of the Edible Mushroom (Volvariella volvacea). J. Agric. Food Chem. 1998, 46, 788–792, doi:10.1021/jf970551h.
  74. Arslan, I.; Akgul, H.; Kara, M. Saporin, a Polynucleotide–Adenosine Nucleosidase, May Be an Efficacious Therapeutic Agent for SARS-CoV-2 Infection. SLAS Discov. Adv. Life Sci. 2020, 26, 330-335. doi:10.1177/2472555220970911.
  75. Hassan, Y.; Ogg, S.; Ge, H. Novel Binding Mechanisms of Fusion Broad Range Anti-Infective Protein Ricin A Chain Mutant-Pokeweed Antiviral Protein 1 (RTAM-PAP1) against SARS-CoV-2 Key Proteins in Silico. Toxins 2020, 12, 602, doi:10.3390/toxins12090602.
  76. Vandenbussche, F.; Desmyter, S.; Ciani, M.; Proost, P.; Peumans, W.J.; Van Damme, E.J.M. Analysis of the in planta antiviral activity of elderberry ribosome-inactivating proteins. Eur. J. Biochem. 2004, 271, 1508–1515, doi:10.1111/j.1432-1033.2004.04059.x.
  77. Vandenbussche, F.; Peumans, W.J.; Desmyter, S.; Proost, P.; Ciani, M.; Van Damme, E.J.M. The type-1 and type-2 ribosome-inactivating proteins from Iris confer transgenic tobacco plants local but not systemic protection against viruses. Planta 2004, 220, 211–221, doi:10.1007/s00425-004-1334-2.
  78. 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. 2007, 54, 115–123, doi:10.1007/s10725-007-9234-7.
  79. Lam, Y.-H.; Wong, Y.-S.; Wang, B.; Wong, R.N.-S.; Yeung, H.-W.; Shaw, P.-C. Use of trichosanthin to reduce infection by turnip mosaic virus. Plant Sci. 1996, 114, 111–117, doi:10.1016/0168-9452(95)04310-1.
  80. Krishnan, R.; McDonald, K.A.; Dandekar, A.M.; Jackman, A.P.; Falk, B. Expression of recombinant trichosanthin, a ribosome-inactivating protein, in transgenic tobacco. J. Biotechnol. 2002, 97, 69–88, doi:10.1016/s0168-1656(02)00058-5.
  81. Zhu, F.; Zhang, P.; Meng, Y.-F.; Xu, F.; Zhang, D.-W.; Cheng, J.; Lin, H.-H.; Xi, D. 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 2012, 237, 77–88, doi:10.1007/s00425-012-1746-3.
  82. Yang, T.; Meng, Y.; Chen, L.-J.; Lin, H.; Xi, D.-H. The Roles of Alpha-Momorcharin and Jasmonic Acid in Modulating the Response of Momordica charantia to Cucumber Mosaic Virus. Front. Microbiol. 2016, 7, 1796, doi:10.3389/fmicb.2016.01796.
  83. Ruan, X.-L.; Liu, L.-F.; Li, H. Transgenic tobacco plants with ribosome inactivating protein gene cassin from Cassia occidentalis and their resistance to tobacco mosaic virus. J. Plant Physiol. Mol. Biol. 2007, 33, 517–523.
  84. Hong, Y.; Saunders, K.; Hartley, M.R.; Stanley, J. Resistance to Geminivirus Infection by Virus-Induced Expression of Dianthin in Transgenic Plants. Virology 1996, 220, 119–127, doi:10.1006/viro.1996.0292.
  85. Stirpe, F.; Williams, D.G.; Onyon, L.J.; Legg, R.F.; Stevens, W.A. Dianthins, ribosome-damaging proteins with anti-viral properties from Dianthus caryophyllus L. (carnation). Biochem. J. 1981, 195, 399–405, doi:10.1042/bj1950399.
  86. Iglesias, R.; Pérez, Y.; de Torre, C.; Ferreras, J.M.; Antolín, P.; Jiménez, P.; Ángeles Rojo, M.; Méndez, E.; Girbés, 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, doi:10.1093/jxb/eri164.
  87. 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, doi:10.1016/j.bbagen.2016.03.011.
  88. Roy, S.; Sadhana, P.; Begum, M.; Kumar, S.; Lodha, M.; Kapoor, H. Purification, characterization and cloning of antiviral/ribosome inactivating protein from Amaranthus tricolor leaves. Phytochemistry 2006, 67, 1865–1873, doi:10.1016/j.phytochem.2006.06.011.
  89. Kwon, S.-Y.; An, C.S.; Liu, J.R.; Paek, K.-H. A Ribosome–Inactivating Protein fromAmaranthus viridis. Biosci. Biotechnol. Biochem. 1997, 61, 1613–1614, doi:10.1271/bbb.61.1613.
  90. Gholizadeh, A. Purification of a ribosome-inactivating protein with antioxidation and root developer potencies from Celosia plumosa. Physiol. Mol. Biol. Plants 2018, 25, 243–251, doi:10.1007/s12298-018-0577-5.
  91. Baranwal, V.K.; Tumer, N.E.; Kapoor, H.C. Depurination of ribosomal RNA and inhibition of viral RNA translation by an antiviral protein of Celosia cristata. Indian J. Exp. Biol. 2002, 40, 1195–1197.
  92. Balasubrahmanyam, A.; Baranwal, V.K.; Lodha, M.; Varma, A.; Kapoor, H. Purification and properties of growth stage-dependent antiviral proteins from the leaves of Celosia cristata. Plant Sci. 2000, 154, 13–21, doi:10.1016/s0168-9452(99)00192-2.
  93. Begam, M.; Kumar, S.; Roy, S.; Campanella, J.J.; Kapoor, H. Molecular cloning and functional identification of a ribosome inactivating/antiviral protein from leaves of post-flowering stage of Celosia cristata and its expression in E. coli. Phytochemistry 2006, 67, 2441–2449, doi:10.1016/j.phytochem.2006.08.015.
  94. Dutt, S.; Narwal, S.; Kapoor, H.C.; Lodha, M.L. Isolation and Characterization of Two Protein Isoforms with Antiviral Activity from Chenopodium album L Leaves. J. Plant Biochem. Biotechnol. 2003, 12, 117–122, doi:10.1007/bf03263171.
  95. Park, J.-S.; Hwang, D.-J.; Lee, S.-M.; Kim, Y.-T.; Choi, S.-B.; Cho, K.-J. Ribosome-inactivating activity and cDNA cloning of antiviral protein isoforms of Chenopodium album. Mol. Cells 2004, 17, 73–80.
  96. Torky, Z.A. Isolation and characterization of antiviral protein from Salsola longifolia leaves expressing polynucleotide adenosine glycoside activity. TOJSAT 2012, 2, 52–58.
  97. Straub, P.; Adam, G.; Mundry, K.-W. Isolation and Characterization of a Virus Inhibitor from Spinach (Spinacia oleracea L.). J. Phytopathol. 1986, 115, 357–367, doi:10.1111/j.1439-0434.1986.tb04349.x.
  98. Kawade, K.; Ishizaki, T.; Masuda, K. Differential expression of ribosome-inactivating protein genes during somatic embryogenesis in spinach (Spinacia oleracea). Physiol. Plant. 2008, 134, 270–281, doi:10.1111/j.1399-3054.2008.01129.x.
  99. Moon, Y.H.; Song, S.K.; Choi, K.W.; Lee, J.S. Expression of a cDNA encoding Phytolacca insularis antiviral protein confers virus resistance on transgenic potato plants. Mol. Cells 1997, 7, 807–815.
  100. Bulgari, D.; Landi, N.; Ragucci, S.; Faoro, F.; Di Maro, A. Antiviral Activity of PD-L1 and PD-L4, Type 1 Ribosome Inactivating Proteins from Leaves of Phytolacca dioica L. in the Pathosystem Phaseolus vulgaris–Tobacco Necrosis Virus (TNV). Toxins 2020, 12, 524, doi:10.3390/toxins12080524.
  101. Sipahioglu, H.M.; Kaya, I.; Usta, M.; Ünal, M.; Ozcan, D.; Özer, M.; Güller, A.; Pallás, V. Pokeweed (Phytolacca americana L.) antiviral protein inhibits Zucchini yellow mosaic virus infection in a dose-dependent manner in squash plants. Turk. J. Agric. For. 2017, 41, 256–262, doi:10.3906/tar-1612-30.
  102. Lodge, J.K.; Kaniewski, W.K.; Tumer, N.E. Broad-spectrum virus resistance in transgenic plants expressing pokeweed antiviral protein. Proc. Natl. Acad. Sci. USA 1993, 90, 7089–7093, doi:10.1073/pnas.90.15.7089.
  103. Domashevskiy, A.V.; Williams, S.; Kluge, C.; Cheng, S.-Y. Plant Translation Initiation Complex eIFiso4F Directs Pokeweed Antiviral Protein to Selectively Depurinate Uncapped Tobacco Etch Virus RNA. Biochemistry 2017, 56, 5980–5990, doi:10.1021/acs.biochem.7b00598.
  104. Wang, P.; 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, doi:10.1023/a:1006084925016.
  105. Bolognesi, A.; Polito, L.; Olivieri, F.; Valbonesi, P.; Barbieri, L.; Battelli, M.G.; Carusi, M.V.; Benvenuto, E.; Blanco, F.D.V.; Di Maro, A.; et al. New ribosome-inactivating proteins with polynucleotide:adenosine glycosidase and antiviral activities from Basella rubra L. and Bougainvillea spectabilis Willd. Planta 1997, 203, 422–429, doi:10.1007/s004250050209.
  106. Srivastava, S.; Verma, H.N.; Srivastava, A.; Prasad, V. BDP-30, a systemic resistance inducer from Boerhaavia diffusa L., suppresses TMV infection, and displays homology with ribosome-inactivating proteins. J. Biosci. 2015, 40, 125–135, doi:10.1007/s12038-014-9494-0.
  107. Bolognesi, A.; Polito, L.; Lubelli, C.; Barbieri, L.; Parente, A.; Stirpe, F. Ribosome-inactivating and Adenine Polynucleotide Glycosylase Activities in Mirabilis jalapa L. Tissues. J. Biol. Chem. 2002, 277, 13709–13716, doi:10.1074/jbc.m111514200.
  108. Güller, A.; Sipahioğlu, H.M.; Usta, M.; Durak, E.D. Antiviral and Antifungal Activity of Biologically Active Recombinant Bouganin Protein from Bougainvillea spectabilis Willd. J. Agric. Sci. 2018, 24, 227–237, doi:10.15832/ankutbd.446442.
  109. Narwal, S.; Balasubrahmanyam, A.; Lodha, M.L.; Kapoor, H.C. Purification and properties of antiviral proteins from the leaves of Bougainvillea xbuttiana. Indian J. Biochem. Biophys. 2001, 38, 342–347.
  110. Narwal, S.; Balasubrahmanyam, A.; Sadhna, P.; Kapoor, H.; Lodha, M.L. A systemic resistance inducing antiviral protein with N-glycosidase activity from Bougainvillea xbuttiana leaves. Indian J. Exp. Biol. 2001, 39, 600–603.
  111. Olivieri, F.; Prasad, V.; Valbonesi, P.; Srivastava, S.; Ghosal-Chowdhury, P.; Barbieri, L.; Bolognesi, A.; Stirpe, F. A systemic antiviral resistance-inducing protein isolated from Clerodendrum inerme Gaertn. is a polynucleotide:adenosine glycosidase (ribosome-inactivating protein). FEBS Lett. 1996, 396, 132–134, doi:10.1016/0014-5793(96)01089-7.
  112. Prasad, V.; Mishra, S.K.; Srivastava, S.; Srivastava, A. A virus inhibitory protein isolated from Cyamopsis tetragonoloba (L.) Taub. upon induction of systemic antiviral resistance shares partial amino acid sequence homology with a lectin. Plant Cell Rep. 2014, 33, 1467–1478, doi:10.1007/s00299-014-1630-7.
  113. Verma, H.; Srivastava, S.; Kumar, D. Induction of Systemic Resistance in Plants Against Viruses by a Basic Protein from Clerodendrum aculeatum Leaves. Phytopathology 1996, 86, 485–492, doi:10.1094/phyto-86-485.
  114. Verma, H.N.; Tewari, K.K.; Kumar, D.; Tuteja, N. Cloning and characterisation of a gene encoding an antiviral protein from Clerodendrum aculeatum L. Plant Mol. Biol. 1997, 33, 745–751, doi:10.1023/A:1005716103632.
  115. Srivastava, A.; Trivedi, S.; Krishna, S.K.; Verma, H.N.; Prasad, V. Suppression of Papaya ringspot virus infection in Carica papaya with CAP-34, a systemic antiviral resistance inducing protein from Clerodendrum aculeatum. Eur. J. Plant Pathol. 2008, 123, 241–246, doi:10.1007/s10658-008-9358-2.
  116. Chen, Y.; Peumans, W.J.; Van Damme, E.J.M. The Sambucus nigratype-2 ribosome-inactivating protein SNA-I’ exhibits in planta antiviral activity in transgenic tobacco. FEBS Lett. 2002, 516, 27–30, doi:10.1016/s0014-5793(02)02455-9.
  117. Iglesias, R.; Pérez, Y.; Citores, L.; Ferreras, J.M.; Méndez, E.; Girbés, T. Elicitor-dependent expression of the ribosome-inactivating protein beetin is developmentally regulated. J. Exp. Bot. 2008, 59, 1215–1223, doi:10.1093/jxb/ern030.
  118. Yang, T.; Zhu, L.-S.; Meng, Y.; Lv, R.; Zhou, Z.; Zhu, L.; Lin, H.-H.; Xi, D. Alpha-momorcharin enhances Tobacco mosaic virus resistance in tobacco NN by manipulating jasmonic acid-salicylic acid crosstalk. J. Plant Physiol. 2018, 223, 116–126, doi:10.1016/j.jplph.2017.04.011.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 554
Revisions: 2 times (View History)
Update Date: 25 Feb 2021
1000/1000
Video Production Service