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 + 1504 word(s) 1504 2021-08-17 05:20:11 |
2 format correct + 245 word(s) 1749 2021-09-22 11:26:55 |

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.
Ruvo, M. Recent Applications of Retro-Inverso Peptides. Encyclopedia. Available online: https://encyclopedia.pub/entry/14374 (accessed on 29 March 2024).
Ruvo M. Recent Applications of Retro-Inverso Peptides. Encyclopedia. Available at: https://encyclopedia.pub/entry/14374. Accessed March 29, 2024.
Ruvo, Menotti. "Recent Applications of Retro-Inverso Peptides" Encyclopedia, https://encyclopedia.pub/entry/14374 (accessed March 29, 2024).
Ruvo, M. (2021, September 21). Recent Applications of Retro-Inverso Peptides. In Encyclopedia. https://encyclopedia.pub/entry/14374
Ruvo, Menotti. "Recent Applications of Retro-Inverso Peptides." Encyclopedia. Web. 21 September, 2021.
Recent Applications of Retro-Inverso Peptides
Edit

Retro-inverso peptides possess reversed sequences and chirality compared to the parent molecules maintaining at the same time an identical array of side chains and in some cases similar structure. The inverted chirality renders them less prone to degradation by endogenous proteases conferring enhanced half-lives and an increased potential as new drugs. However, given their general incapability to adopt the 3D structure of the parent peptides their application should be careful evaluated and investigated case by case.

retro-inverso peptides anticancer peptides drug delivery peptide antigens IAPP antimicrobial peptides

1. Topology, Structural Characteristics

As therapeutic agents, peptides have fascinating properties, such as very high specificity and binding affinity, generally low toxicity, and low risk of drug interactions [1]. Moreover, due to the high diversity which is sequence- and structure-dependent, they can be designed as potential drugs to target almost any disease. On the other hand, natural peptides, due to their low size and high sensitivity to most proteases, are quickly excreted or anyhow degraded, resulting in poor biodistribution, bioavailability, and rapid clearance [2], and thus limited therapeutic potential [2]. To increase peptide half-life, stability and bioavailability, many approaches have been proposed including PEGylation, backbone modifications, cyclization, side chain stapling, and lipidation [3][4]. Among these, modification of the backbone is one of the most invasive approaches, as it may profoundly affect the conformation of peptides, especially when it involves alteration of the residue’s isomerization.

All amino acids (except glycine) possess chiral centers and occur in nature almost exclusively as L-enantiomers in proteins and natural peptides. D-amino acid-containing peptides are also found in nature, mainly in some frog species and bacteria, as a result of post-translational modifications [5].

The hydrogen bonds between the CO acceptors and NH donors generate a network of highly stabilizing interactions in peptides arranged as α-helices and β-sheets. If the network is removed, the stability of the 3D structure will be severely compromised in the retro-inverso mimetics, largely affecting their activity [6]. From a topological point of view, the RI analogues of larger peptides [7][8] could adopt conformations similar to that of the parent peptide when the full-length protein or part of it predominantly contains structural elements whose energy in the Ramachandran map is not drastically changed during the conversion like in β-sheets and γ-turns. In this case, they are likely to be stabilized by similar side-chain-to-side-chain interactions. This observation is consistent with results reported in literature [6][9][10][11]. As an example, Peggion and coworkers in 2009 [12] proposed a structure-function relationship study on a mimetic peptide of the Parathyroid hormone (PTH) spanning residues 1-11 (PTH(1-11)). This peptide is a ligand of the PTH type-I receptor and was studied through the synthesis and characterization of all-D PTH retro-inverso analogues. The retro-inverso RI-PTH(1-11) analogues showed a reduced biological activity compared to the parent peptide, because of the absence of the α-helical structure which could be induced by introducing an Aib residues on the N-terminal position [13].

The main advantages of retro-inverso peptides as potential biotherapeutics are the improved stability in vitro and in vivo and the novelty in terms of intellectual property as compared to the parent peptides. The interest around this subclass of molecules is driven by their potential use in a vast area of applications here reviewed, including diagnostics, cancer therapeutics, neurodegenerative diseases, and new antibiotics (as antimicrobial peptides). Some potential applications are here reported.

2. Anticancer Applications—Diagnostic

An example of the use of RI analogues is provided by Calvanese et al. [14] who designed a series of cyclic peptides embedding the retro-inverso (RI) version of the consensus sequences RPL/LPPR, corresponding to peptides capable of preventing VEGF binding to VEGFR1 [15] and VEGFR2 [16], and to specifically inhibit human endothelial cells (EC) proliferation in vitro [16]. Direct binding experiments of the peptides to VEGFR1 and VEGFR2 identified a peptide that bound both receptors with a K D in the low micromolar range but with a significant selectivity for VEGFR1 respect to VEGFR2 thus showing a potential as VEGFR1-selective diagnostic probe.

Other examples are reported below.

The peptide VAP ( Table 1 ) was shown to have high binding affinity in vitro to GRP78 protein, which is overexpressed in gliomas, glioma stem cells, vasculogenic mimicry, and neovasculature [17]. The prediction of binding for the analogue RI-VAP to GRP78 was similar to that of the parent peptide and, in addition, remarkable tumor accumulation was observed experimentally by imaging in vivo. RI-VAP-modified paclitaxel-loaded polymeric micelles had better anti-tumor efficacy compared to free taxol, to paclitaxel-loaded simple micelles, and to micelles modified with parent peptide.

Table 1. Names, peptide sequences reported in the review and their applications.
Name Sequence 1 Application Ref
Anticancer Applications—Diagnostic
VEGF-P3(CYC) I76TMQ79CG92IHQGQHPKIRMI80CE93MSF96 * Inhibition angiogenesis [18][19]
D(LPR) D(Leu-Pro-Arg) Inhibition retinal angiogenesis; Diagnostic [15][14][20]
SP5 PRPSPKMGVSVS * Drug delivery [21][22]
uPAR88–92 SRSRY * Maintaining chemotactic activity and triggers directed cell migration and angiogenesis [23][24]
RI-3 Ac-D(Tyr- Arg-Aib-Arg)- NH2 Prevent extracellular invasion by tumor cells [25]
D(RGD) D(Asp-Gly-Arg) Diagnostic [26][27][28]
VS SWFSRHRYSPFAVS * Glioblastoma multiforme (GBM) [29][30]
VAP SNTRVAP * Gliomas, glioma stem cells, vasculogenic mimicry and neovasculature [17]
WSW SYPGWSW * Glioma cells and tumor neovasculature [31]
BK RPPGFSPFR * Glioma cells [32]
FP21 YTRDLVYGDPARPGIQGTGTF * Ovarian cancer [33][34][35]
T7 HAIYPRH * Drug delivery [36]
Applications in Immunology
TG19320 (rty)4K2KG IgG binding [37][38]
VSVp RGYVYQGL * antigen surface of hepatitis B virus [39]
OVAp SIINFEKL * antigen surface of hepatitis B virus [39]
PS1 HQLDPAFGANSTNPD * antigen surface of hepatitis B virus [39]
HAI HAIYPRH * Crossing BBB [40]
THR THRPPMWSPVWP * Crossing BBB [40]
InsB:9–23 HLVEALYLVCGERGG * Analogue of diabetogenic islet peptide—prevents T-cell activation in humanized model mice [41][42]
Application in Neurodegenerative Diseases
Amytrap WKGEWTGR * Blocking the oligomerization and aggregation of Aβ1–42 [43][44][45][46][47]
IAPP11–20 RLANFLVHSS * Strong inhibitory effects on amylin aggregation in T2DM [48]
β-syn36–45 GVLYVGSKT * Reduction of amyloid fibril and oligomer formation [49]
Application in Antimicrobial Antibiotics
RI1018 rrwirvavilrv Preventing formation of Biofilm [50]
RI-JK6 rivwvrirrwqv Preventing formation of Biofilm [50]
RI-73 lwGvwrrvidwlr Damaging the bacterial membrane [51]
BMAP-28 GGLRSLGRKILRAWKKYGPIIVPIIRIG * Broad antimicrobial activities [52]
1 The sequences reported are those of the parent peptides * (L-residues), unless otherwise indicated, like reporting D residues as lower-case letters or adding a “D” before the sequence.

Follicle-stimulating hormone receptor (FSHR) expression is limited to the reproductive system [53][54] and might be targeted to deliver drugs against ovarian cancer with high selectivity and specificity. In particular, nanoparticles carrying the RI variant of the peptide FP21 showed to bind FSHR ( Table 1 ). They were thus used as an ovarian cancer targeted delivery system [33][34][35], showing improved biostability compared to the parent peptide, with no degradation even after 12 h incubation with proteolytic enzymes. The data obtained on the RI peptide encouraged further developments and optimizations of the molecule for treating ovarian cancers expressing FSHR [33].

3. Conclusions and Future Perspectives

In the field of peptidomimetics, retro-inversion has been largely explored to improve peptide stability while retaining the parent molecule’s activity. Changing the order of the amino acids and their configuration has been also a mean of introducing novelty and to overcome existing intellectual property claims [55]. The first examples of their use were reported by M. Goodman in the mid-1970s [56], who was interested in the study of stereochemical and conformational properties of retro–inverso (RI) amide bonds in linear peptides. Interesting examples were next reported by Merrifield with studies on the CAMEL peptide [57][58], which was a chimeric peptide derived from the merging of two AMP. Despite the amazing results reported in literature, the application of retro-inversion to generate peptidomimetics is still rare or however uncommon.

Beyond these basic rules applicable to short peptides or other specific examples, the reasons for the frequent failure of RI isomerization of longer molecules are still largely unclear, and definite instructions for possibly improving the success rate are unresolved. The reversal of the peptide backbone and the shift of the H-bond network it is involved into is a major alteration of the fine equilibrium of the forces that supports the conformation of a peptide having an organized 3D structure. Therefore, as for the folding of a natural molecule, the lack of one such important puzzle piece prevents the correct assembling of the structure although the side chains may potentially have access to the same conformational space of the parent molecule. We can thus conclude that the design of a successful retro-inverso analogue of a folded peptide has the same complications as for the de novo design of a new protein or peptide and one should thus proceed following the rules, still not well understood and codified, of protein folding, exploiting and using the geometrical and structural features of amino acids in D configuration. For example, the RI isomerization and structure reconstruction of the MDM2/MDMX peptide inhibitor stingin, which adopts an N-terminal loop and a C-terminal α-helix, lead to an isomer that partially retained binding (3.0–3.4 kcal/mol reduction) and showed a decreased ability to prevent the interaction with p53 [59]. These conformation and energy issues have been often discouraging because of the frequent loss of biological activity observed in larger molecules showing well-defined tridimensional organizations. Merrifield indeed soon observed that the efficiency of peptide retro-inversion was not only related to inversion of its chirality but to the global change of the 3D conformation [57]. These observations have been indirectly confirmed showing that retro-inverso analogues of unstructured peptides more often maintain or even increase the activity compared to the parent peptide [60].

On the other hands, peptides that assemble into β-sheets adopting extended conformations establish a large and well-organized network of interactions, mostly H-bonds, with the adjacent molecules. Also, the side chains are well packed each other. In this case, despite the strong backbone interactions, retro-inverso analogues have more chance to be successful if the registry of H-bonds and of side chain-to-side chain interactions is corrected to account for the inverted amide bonds. The molecular dynamic simulations of amyloid fibrils in AD [61] or amylin in T2D [62] indeed showed that the interactions of both side chains and backbone of RI peptides were re-aligned establishing different patterns of contacts and hydrogen bonding. Also, the twist of the RI analogue β-sheets was similar and the complex had only slightly lower stability compared to the parent peptides.

Computational approaches might be of great help and might open a new season in this field as suggested by Robson [63]. Despite their many limitations, we believe their use still has a place in the design of drugs based on bioactive peptides. This belief stems from the simplicity of the design, from the rapidity in making synthetic peptides and from the immediate benefits resulting when the molecules maintain their activity. Therefore, this review would be an incentive to continue working with these types of molecules, also to further investigate the conformational and topological space they need to occupy to fully mimic bioactive peptides with complex structure.

References

  1. Lee, A.C.; Harris, J.L.; Khanna, K.K.; Hong, J.H. A Comprehensive Review on Current Advances in Peptide Drug Development and Design. Int. J. Mol. Sci. 2019, 20, 2383.
  2. Bruno, B.J.; Miller, G.D.; Lim, C.S. Basics and recent advances in peptide and protein drug delivery. Ther. Deliv. 2013, 4, 1443–1467.
  3. Corbi-Verge, C.; Garton, M.; Nim, S.; Kim, P.M. Strategies to Develop Inhibitors of Motif-Mediated Protein-Protein Interactions as Drug Leads. Annu. Rev. Pharmacol. Toxicol. 2017, 57, 39–60.
  4. Fosgerau, K.; Hoffmann, T. Peptide therapeutics: Current status and future directions. Drug Discov. Today 2015, 20, 122–128.
  5. Kreil, G. D-amino acids in animal peptides. Annu. Rev. Biochem. 1997, 66, 337–345.
  6. Rai, J. Peptide and protein mimetics by retro and retroinverso analogs. Chem. Biol. Drug Des. 2019, 93, 724–736.
  7. Brady, L.; Dodson, G. Drug design. Reflections on a peptide. Nature 1994, 368, 692–693.
  8. Jameson, B.A.; McDonnell, J.M.; Marini, J.C.; Korngold, R. A rationally designed CD4 analogue inhibits experimental allergic encephalomyelitis. Nature 1994, 368, 744–746.
  9. Ruvo, M.; Fassina, G. End-group modified retro-inverso isomers of tripeptide oxytocin analogues: Binding to neurophysin II and enhancement of its self-association properties. Int. J. Pept. Protein Res. 1995, 45, 356–365.
  10. Taylor, E.M.; Otero, D.A.; Banks, W.A.; O’Brien, J.S. Retro-inverso prosaptide peptides retain bioactivity, are stable In vivo, and are blood-brain barrier permeable. J. Pharmacol. Exp. Ther. 2000, 295, 190–194.
  11. Banerjee, A.; Raghothama, S.R.; Karle, I.L.; Balaram, P. Ambidextrous molecules: Cylindrical peptide structures formed by fusing left- and right-handed helices. Biopolymers 1996, 39, 279–285.
  12. Caporale, A.; Biondi, B.; Schievano, E.; Wittelsberger, A.; Mammi, S.; Peggion, E. Structure-function relationship studies of PTH(1-11) analogues containing D-amino acids. Eur. J. Pharmacol. 2009, 611, 1–7.
  13. Crisma, M.; Bisson, W.; Formaggio, F.; Broxterman, Q.B.; Toniolo, C. Factors governing 3(10)-helix vs alpha-helix formation in peptides: Percentage of C(alpha)-tetrasubstituted alpha-amino acid residues and sequence dependence. Biopolymers 2002, 64, 236–245.
  14. Calvanese, L.; Caporale, A.; Foca, G.; Iaccarino, E.; Sandomenico, A.; Doti, N.; Apicella, I.; Incisivo, G.M.; De Falco, S.; Falcigno, L.; et al. Targeting VEGF receptors with non-neutralizing cyclopeptides for imaging applications. Amino Acids 2018, 50, 321–329.
  15. Giordano, R.J.; Cardo-Vila, M.; Salameh, A.; Anobom, C.D.; Zeitlin, B.D.; Hawked, D.H.; Valente, A.P.; Almeida, F.C.L.; Nor, J.E.; Sidman, R.L.; et al. From combinatorial peptide selection to drug prototype (I): Targeting the vascular endothelial growth factor receptor pathway. Proc. Natl. Acad. Sci. USA 2010, 107, 5112–5117.
  16. Binetruy-Tournaire, R.; Demangel, C.; Malavaud, B.; Vassy, R.; Rouyre, S.; Kraemer, M.; Plouet, J.; Derbin, C.; Perret, G.; Mazie, J.C. Identification of a peptide blocking vascular endothelial growth factor (VEGF)-mediated angiogenesis. Embo. J. 2000, 19, 1525–1533.
  17. Ran, D.N.; Mao, J.N.; Shen, Q.; Xie, C.; Zhan, C.Y.; Wang, R.F.; Lu, W.Y. GRP78 enabled micelle-based glioma targeted drug delivery. J. Control. Release 2017, 255, 120–131.
  18. Vicari, D.; Foy, K.C.; Liotta, E.M.; Kaumaya, P.T. Engineered conformation-dependent VEGF peptide mimics are effective in inhibiting VEGF signaling pathways. J. Biol. Chem. 2011, 286, 13612–13625.
  19. Foy, K.C.; Liu, Z.; Phillips, G.; Miller, M.; Kaumaya, P.T. Combination treatment with HER-2 and VEGF peptide mimics induces potent anti-tumor and anti-angiogenic responses in vitro and in vivo. J. Biol. Chem. 2011, 286, 13626–13637.
  20. Rezazadeh, F.; Sadeghzadeh, N.; Abedi, S.M.; Abediankenari, S. Tc-99m labeled (D)(LPR): A novel retro-inverso peptide for VEGF receptor-1 targeted tumor imaging. Nucl. Med. Biol. 2018, 62–63, 54–62.
  21. Lee, T.Y.; Lin, C.T.; Kuo, S.Y.; Chang, D.K.; Wu, H.C. Peptide-mediated targeting to tumor blood vessels of lung cancer for drug delivery. Cancer Res. 2007, 67, 10958–10965.
  22. Li, Y.; Lei, Y.; Wagner, E.; Xie, C.; Lu, W.Y.; Zhu, J.H.; Shen, J.; Wang, J.; Liu, M. Potent Retro-Inverso D-Peptide for Simultaneous Targeting of Angiogenic Blood Vasculature and Tumor Cells. Bioconjugate Chem. 2013, 24, 133–143.
  23. Bifulco, K.; Longanesi-Cattani, I.; Gala, M.; Di Carluccio, G.; Masucci, M.T.; Pavone, V.; Lista, L.; Arra, C.; Stoppelli, M.P.; Carriero, M.V. The soluble form of urokinase receptor promotes angiogenesis through its Ser(88)-Arg-Ser-Arg-Tyr(92) chemotactic sequence. J. Thromb. Haemost. 2010, 8, 2789–2799.
  24. Resnati, M.; Pallavicini, I.; Wang, J.M.; Oppenheim, J.; Serhan, C.N.; Romano, M.; Blasi, F. The fibrinolytic receptor for urokinase activates the G protein-coupled chemotactic receptor FPRL1/LXA4R. Proc. Natl. Acad. Sci. USA 2002, 99, 1359–1364.
  25. Carriero, M.V.; Bifulco, K.; Ingangi, V.; Costantini, S.; Botti, G.; Ragone, C.; Minopoli, M.; Motti, M.L.; Rea, D.; Scognamiglio, G.; et al. Retro-inverso Urokinase Receptor Antagonists for the Treatment of Metastatic Sarcomas. Sci. Rep. 2017, 7, 1–17.
  26. Karimi, H.; Sadeghzadeh, N.; Abediankenari, S.; Rezazadeh, F.; Hallajian, F. Radiochemical Evaluation and In Vitro Assessment of the Targeting Ability of a Novel Tc-99m-HYNIC-RGD for U87MG Human Brain Cancer Cells. Curr. Radiopharm. 2017, 10, 139–144.
  27. Torabizadeh, S.A.; Abedi, S.M.; Noaparast, Z.; Hosseinimehr, S.J. Comparative assessment of a Tc-99m labeled H1299.2-HYNIC peptide bearing two different co-ligands for tumor-targeted imaging. Bioorgan. Med. Chem. 2017, 25, 2583–2592.
  28. Liu, Y.Y.; Mei, L.; Yu, Q.W.; Zhang, Q.Y.; Gao, H.L.; Zhang, Z.R.; He, Q. Integrin alpha(v)beta(3) targeting activity study of different retro-inverso sequences of RGD and their potentiality in the designing of tumor targeting peptides. Amino Acids 2015, 47, 2533–2539.
  29. Ren, Y.C.; Zhan, C.Y.; Gao, J.; Zhang, M.F.; Wei, X.L.; Ying, M.; Liu, Z.N.; Lu, W.Y. A D-Peptide Ligand of Integrins for Simultaneously Targeting Angiogenic Blood Vasculature and Glioma Cells. Mol. Pharmaceut 2018, 15, 592–601.
  30. van Ommeren, R.; Staudt, M.D.; Xu, H.; Hebb, M.O. Advances in HSP27 and HSP90-targeting strategies for glioblastoma. J. Neuro Oncol. 2016, 127, 209–219.
  31. Ran, D.N.; Mao, J.N.; Zhan, C.Y.; Xie, C.; Ruan, H.T.; Ying, M.; Zhou, J.F.; Lu, W.L.; Lu, W.Y. D-Retroenantiomer of Quorum-Sensing Peptide-Modified Polymeric Micelles for Brain Tumor-Targeted Drug Delivery. ACS Appl. Mater. Inter. 2017, 9, 25672–25682.
  32. Xie, Z.X.; Shen, Q.; Xie, C.; Lu, W.Y.; Peng, C.M.; Wei, X.L.; Li, X.; Su, B.X.; Gao, C.L.; Liu, M. Retro-inverso bradykinin opens the door of blood-brain tumor barrier for nanocarriers in glioma treatment. Cancer Lett. 2015, 369, 144–151.
  33. Zhang, M.Y.; Zhang, M.X.; Wang, J.; Cai, Q.Q.; Zhao, R.; Yu, Y.; Tai, H.Y.; Zhang, X.Y.; Xu, C.J. Retro-inverso follicle-stimulating hormone peptide-mediated polyethylenimine complexes for targeted ovarian cancer gene therapy. Drug Deliv. 2018, 25, 995–1003.
  34. Zhang, M.X.; Hong, S.S.; Cai, Q.Q.; Zhang, M.; Chen, J.; Zhang, X.Y.; Xu, C.J. Transcriptional control of the MUC16 promoter facilitates follicle-stimulating hormone peptide-conjugated shRNA nanoparticle-mediated inhibition of ovarian carcinoma in vivo. Drug Deliv. 2018, 25, 797–806.
  35. Hong, S.S.; Zhang, M.X.; Zhang, M.; Yu, Y.; Chen, J.; Zhang, X.Y.; Xu, C.J. Follicle-stimulating hormone peptide-conjugated nanoparticles for targeted shRNA delivery lead to effective gro-alpha silencing and antitumor activity against ovarian cancer. Drug Deliv. 2018, 25, 576–584.
  36. Tang, J.J.; Wang, Q.T.; Yu, Q.W.; Qiu, Y.; Mei, L.; Wan, D.D.; Wang, X.H.; Li, M.; He, Q. A stabilized retro-inverso peptide ligand of transferrin receptor for enhanced liposome-based hepatocellular carcinoma-targeted drug delivery. Acta Biomater. 2019, 83, 379–389.
  37. Moiani, D.; Salvalaglio, M.; Cavallotti, C.; Bujacz, A.; Redzynia, I.; Bujacz, G.; Dinon, F.; Pengo, P.; Fassina, G. Structural characterization of a Protein A mimetic peptide dendrimer bound to human IgG. J. Phys. Chem. B 2009, 113, 16268–16275.
  38. Marino, M.; Ruvo, M.; De Falco, S.; Fassina, G. Prevention of systemic lupus erythematosus in MRL/lpr mice by administration of an immunoglobulin-binding peptide. Nat. Biotechnol. 2000, 18, 735–739.
  39. Nair, D.T.; Kaur, K.J.; Singh, K.; Mukherjee, P.; Rajagopal, D.; George, A.; Bal, V.; Rath, S.; Rao, K.V.S.; Salunke, D.M. Mimicry of native peptide antigens by the corresponding retro-inverso analogs is dependent on their intrinsic structure and interaction propensities. J. Immunol. 2003, 170, 1362–1373.
  40. Lee, J.H.; Engler, J.A.; Collawn, J.F.; Moore, B.A. Receptor mediated uptake of peptides that bind the human transferrin receptor. Eur. J. Biochem. 2001, 268, 2004–2012.
  41. Lee, K.H.; Wucherpfennig, K.W.; Wiley, D.C. Structure of a human insulin peptide-HLA-DQ8 complex and susceptibility to type 1 diabetes. Nat. Immunol. 2001, 2, 501–507.
  42. Nakayama, M.; Abiru, N.; Moriyama, H.; Babaya, N.; Liu, E.; Miao, D.; Yu, L.; Wegmann, D.R.; Hutton, J.C.; Elliott, J.F.; et al. Prime role for an insulin epitope in the development of type 1 diabetes in NOD mice. Nature 2005, 435, 220–223.
  43. Gandbhir, O.; Sundaram, P. Pre-Clinical Safety and Efficacy Evaluation of Amytrap, a Novel Therapeutic to Treat Alzheimer’s Disease. J. Alzheimers Dis. Rep. 2019, 3, 77–94.
  44. Parthsarathy, V.; McClean, P.L.; Holscher, C.; Taylor, M.; Tinker, C.; Jones, G.; Kolosov, O.; Salvati, E.; Gregori, M.; Masserini, M.; et al. A novel retro-inverso peptide inhibitor reduces amyloid deposition, oxidation and inflammation and stimulates neurogenesis in the APPswe/PS1DeltaE9 mouse model of Alzheimer’s disease. PLoS ONE 2013, 8, e54769.
  45. Gregori, M.; Taylor, M.; Salvati, E.; Re, F.; Mancini, S.; Balducci, C.; Forloni, G.; Zambelli, V.; Sesana, S.; Michael, M.; et al. Retro-inverso peptide inhibitor nanoparticles as potent inhibitors of aggregation of the Alzheimer’s Abeta peptide. Nanomedicine 2017, 13, 723–732.
  46. Morris, O.; Gregory, J.; Kadirvel, M.; Henderson, F.; Blykers, A.; McMahon, A.; Taylor, M.; Allsop, D.; Allan, S.; Grigg, J.; et al. Development & automation of a novel [(18)F]F prosthetic group, 2-[(18)F]-fluoro-3-pyridinecarboxaldehyde, and its application to an amino(oxy)-functionalised Abeta peptide. Appl. Radiat. Isot. 2016, 116, 120–127.
  47. Stark, T.; Lieblein, T.; Pohland, M.; Kalden, E.; Freund, P.; Zangl, R.; Grewal, R.; Heilemann, M.; Eckert, G.P.; Morgner, N.; et al. Peptidomimetics That Inhibit and Partially Reverse the Aggregation of Aβ1–42. Biochemistry 2017, 56, 4840–4849.
  48. Obasse, I.; Taylor, M.; Fullwood, N.J.; Allsop, D. Development of proteolytically stable N-methylated peptide inhibitors of aggregation of the amylin peptide implicated in type 2 diabetes. Interface Focus 2017, 7, 20160127.
  49. Windisch, M.; Hutter-Paier, B.; Schreiner, E.; Wronski, R. Beta-Synuclein-derived peptides with neuroprotective activity: An alternative treatment of neurodegenerative disorders? J. Mol. NeuroSci 2004, 24, 155–165.
  50. de la Fuente-Nunez, C.; Reffuveille, F.; Haney, E.F.; Straus, S.K.; Hancock, R.E. Broad-spectrum anti-biofilm peptide that targets a cellular stress response. PLoS Pathog 2014, 10, e1004152.
  51. Kumar, P.; Pletzer, D.; Haney, E.F.; Rahanjam, N.; Cheng, J.T.J.; Yue, M.; Aljehani, W.; Hancock, R.E.W.; Kizhakkedathu, J.N.; Straus, S.K. Aurein-Derived Antimicrobial Peptides Formulated with Pegylated Phospholipid Micelles to Target Methicillin-Resistant Staphylococcus aureus Skin Infections. ACS Infect. Dis. 2019, 5, 443–453.
  52. Lynn, M.A.; Kindrachuk, J.; Marr, A.K.; Jenssen, H.; Pante, N.; Elliott, M.R.; Napper, S.; Hancock, R.E.; McMaster, W.R. Effect of BMAP-28 antimicrobial peptides on Leishmania major promastigote and amastigote growth: Role of leishmanolysin in parasite survival. PLoS Negl. Trop. Dis. 2011, 5, e1141.
  53. Papadimitriou, K.; Kountourakis, P.; Kottorou, A.E.; Antonacopoulou, A.G.; Rolfo, C.; Peeters, M.; Kalofonos, H.P. Follicle-Stimulating Hormone Receptor (FSHR): A Promising Tool in Oncology? Mol. Diagn. Ther. 2016, 20, 523–530.
  54. Perales-Puchalt, A.; Svoronos, N.; Rutkowski, M.R.; Allegrezza, M.J.; Tesone, A.J.; Payne, K.K.; Wickramasinghe, J.; Nguyen, J.M.; O’Brien, S.W.; Gumireddy, K.; et al. Follicle-Stimulating Hormone Receptor Is Expressed by Most Ovarian Cancer Subtypes and Is a Safe and Effective Immunotherapeutic Target. Clin. Cancer Res. 2017, 23, 441–453.
  55. Fassina, G.; Verdoliva, A.; Ruvo, M. Antigenic Peptides. U.S. Patent US-5932692-A, 3 August 1999.
  56. Goodman, M.; Chorev, M. On the concept of linear modified retro-peptide structures. Acc. Chem. Res. 1979, 12, 1–7.
  57. Merrifield, R.B.; Juvvadi, P.; Andreu, D.; Ubach, J.; Boman, A.; Boman, H.G. Retro and retroenantio analogs of cecropin-melittin hybrids. Proc. Natl. Acad. Sci. USA 1995, 92, 3449–3453.
  58. Andreu, D.; Ubach, J.; Boman, A.; Wahlin, B.; Wade, D.; Merrifield, R.B.; Boman, H.G. Shortened cecropin A-melittin hybrids. Significant size reduction retains potent antibiotic activity. FEBS Lett. 1992, 296, 190–194.
  59. Li, C.; Zhan, C.; Zhao, L.; Chen, X.; Lu, W.Y.; Lu, W. Functional consequences of retro-inverso isomerization of a miniature protein inhibitor of the p53-MDM2 interaction. Bioorg. Med. Chem. 2013, 21, 4045–4050.
  60. Garton, M.; Nim, S.; Stone, T.A.; Wang, K.E.; Deber, C.M.; Kim, P.M. Method to generate highly stable D-amino acid analogs of bioactive helical peptides using a mirror image of the entire PDB. Proc. Natl. Acad Sci. USA 2018, 115, 1505–1510.
  61. Xi, W.; Hansmann, U.H.E. The effect of retro-inverse D-amino acid Abeta-peptides on Abeta-fibril formation. J. Chem. Phys. 2019, 150, 095101.
  62. Pandey, P.; Nguyen, N.; Hansmann, U.H.E. d-Retro Inverso Amylin and the Stability of Amylin Fibrils. J. Chem. Theory Comput. 2020, 16, 5358–5368.
  63. Robson, B. Computers and viral diseases. Preliminary bioinformatics studies on the design of a synthetic vaccine and a preventative peptidomimetic antagonist against the SARS-CoV-2 (2019-nCoV, COVID-19) coronavirus. Comput. Biol. Med. 2020, 119, 103670.
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: 431
Revisions: 2 times (View History)
Update Date: 22 Sep 2021
1000/1000