Next Article in Journal
Recovery of Bioactive Compounds from Exhausted Olive Pomace
Previous Article in Journal
Chromatographic Analysis of Iodoacetate Encapsulated in Liposomes and Biochemical Assessment of Its Toxic Effect during Course Application, Cancer Research Institute Tomsk NRMC, 2020
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Ultra-Short Cyclo-Peptides as Bio-Inspired Therapeutics: Proline-Based 2,5-Diketopiperazines (DKP) †

Faculty of Chemistry, Institute of General & Inorganic Chemistry, Lodz University of Technology, Zeromskiego 116, 90-924 Lodz, Poland
*
Author to whom correspondence should be addressed.
Presented at the 1st International Electronic Conference on Biomolecules: Natural and Bio-Inspired Therapeutics for Human Diseases, 1–13 December 2020; Available online: https://iecbm2020.sciforum.net/.
Proceedings 2021, 79(1), 10; https://doi.org/10.3390/IECBM2020-08804
Published: 1 December 2020

Abstract

:
2,5-diketopiperazines (DKP) are the simplest cyclo-peptides in nature, which could play a key role in the origin of life. They are ubiquitous in microorganisms, higher species, and in food and beverages. These dipeptides have been known since the beginning of the 20th century, but they have only recently been gaining interest due to diverse, noteworthy bioactivities, such as, but not limited to, anticancer, antiviral, antioxidant, and neuroprotective properties. DKPs have relevance in quorum sensing, cell–cell signaling, or as drug delivery systems. They have less toxicity, increased cell permeability, and binding affinity. Proline-containing DKPs have an extra-rigid conformation and are more resistant to degradation by enzymes. They represent an attractive subclass of cyclo-peptides with a high potential in future therapies.

1. Introduction

The main aim of this mini-review is to provide an overview of the scientific findings on diversity and the enormous biological potential of the simplest natural cyclo-peptides in light of the newest literature, structural databases, and patents. Further, we have emphasized the attractiveness of their pharmacokinetic profile in relation to the future innovative effective therapeutical and bio-control agents.

2. Cyclo-Peptides: General Considerations

Recently, short peptides have been enjoyed greater and greater significant interest as a unique class of bio-molecules filling a therapeutic niche between small chemical drugs and macro-molecular agents with diverse well-known limitations. Oligopeptides are a primary source of bio-molecules, which are components of proteins participating in bio-processes. Therefore, they have significant advantages, such as diverse bioactivities, high selectivity due to specific interactions of peptides with targets, low toxicity, because they do not accumulate in the organs, and amino acids are degradation products.
On the other hand, shortcomings of the peptides, such as poor oral absorption, low stability in vivo, high conformational freedom, or low cell permeability, can not be neglected [1].
Nevertheless, cyclic peptides have specific structural features resulting in a superior pharmacological profile [1,2]. Amino and carboxyl terminals are linked together with a peptide bond, forming a circular chain, which results in rigid conformation. This ”head to tail” cyclization leads to increased stability against proteolysis and better bioavailability. Besides, cyclo-peptides have less cytotoxicity, higher bioactivity, specificity and efficacy, increased cell permeability, or binding affinity [3].

3. Diketopiperazines

Cyclic dipeptides containing a 2,5-diketopiperazine backbone, known also as cyclodipeptides, 2,5-diketopiperazines (DKPs), piperazine-2,5-diones, 2,5-dioxopiperazines, or dipeptide anhydrides, represent a unique class of compounds with extra advantages in drug development [4]. The DKP scaffold consists of a six-membered ring with, or without, various substituents orientated in a defined way. It provides three-dimensionality, increased rigidity, chiral nature, enables the control of the substituent’s stereochemistry, stimulation of the pharmacophoric peptide groups, promotes the intermolecular H-bonding interactions with bio-targets via corresponding sites of donors and acceptors [5]. The rigid DKP core enables the preferential conformation of peptide to be mimicked, and allows the dual behavior of amino acids to be activated in an either constrained or flexible manner [6].
It should be mentioned that even though 2,5-DKPs are the most popular in nature and therapies [4,7], 2,3-DKPs and 2,6-DKPs are possible as well—see Scheme 1 [8].

3.1. Historical Background and Occurence: Origin of Life

In nature, imperative fundamental functions of amino acids are indisputable. The cyclic dipeptides probably play a key role in the origin of life in the context of chemical evolution. They could have relevance in the process of catalyse chiral selection and act as precursors in the formation of the peptide, which is considered an essential condition of the beginning of life [9].
2,5-diketopiperazines (DKP) are the simplest naturally occurring cyclo-peptides. They are biosynthesized from amino acids, which are catalyzed through two enzyme families, both nonribosomal peptide synthetase and cyclodipeptide synthase enzymes, resulting in the formation of the two peptide bonds [10,11,12,13,14]. The DKP skeleton is ubiquitous in various environments, either in microorganisms, bacteria such as Bacillus subtilis, Streptomyces, Pseudomonas aeruginosa, or Lactobacillus plantarum [15,16,17], marine and terrestrial fungi [18] as Aspergillus flavus or Alternaria alternata and Penicillium, respectively [19,20], or higher species, such as marine sponges such as Dysidea herbacea and fragilis [21], but also proteobacteria Alcaligenes faecalis, algae, lichens, gorgonians, tunicates, plants, or animals venoms. Notably, they have been found in the central nervous system, gastrointestinal tract, and blood of humans [22]. Additionally, they are present in food and beverages [4,23,24]. DKPs can be chemical byproducts, e.g., in Pu-erh tea, cocoa, dried bonito, roasted coffee, sake, beer, cheese, casein, chicken extract, or stewed beef, giving a special metallic bitter taste [25,26,27,28], but also in drugs because of intramolecular cyclization of the dipeptidyl moiety in active peptide-based substances. The latter is a common phenomenon in different therapeutics, in aminopenicillin, amoxicillin, ACE inhibitors [29,30,31], etc. Notably, proteins and peptides can be cyclized to DKPs by heating [32].
Surprisingly, these natural dipeptides have been known since the beginning of the 20th century [33,34] but they have been neglected for a long time. Only recently, they have been enjoyed greater and greater significant attention, and their biological profile is investigated in detail.

3.2. Properties and Possibilities

The simple biomolecules containing the bis-lactam core of DKP have a wide spectrum of biological activities, inter alia anticancer, T-cell mediated immunity, antiviral, nootropic and neuroprotective in neurodegenerative diseases (Alzheimer’s or Parkinson’s diseases and amyotrophic lateral sclerosis), cytotoxic, immunosuppressive, antibacterial, antifungal, antimutagenic, anti-inflammatory, antihyperglycemic, antiarrhythmic, antimalarial, antiparasitic, anthelmintic, insecticidal, antifouling, anti-prion, vasorelaxant or metabolic regulatory [5,24,34,35,36,37]. They have relevance in quorum sensing, improving the classical theory of quorum sensing, unique communication manner between bacteria and gene regulation systems [38]. DKP ring plays a key role in fighting oxidative stress [39]. Moreover, they play role in ion-transport, cell–cell signaling, and a high affinity to many receptors and enzymes [4]. They are useful in delivery systems of drugs, which have low permeability to crossing the blood–brain barrier [34]. Additionally, cell-penetrating peptides containing DKP have better properties in relation to anti-cancer drugs. They provide high cell membrane penetration or transport cargo into the cell [3]. The DKPs can be easily synthesized through conventional procedures providing an attractive scaffold in new drug design due to its main simplicity and marvelous structural diversity [40,41]. They are an excellent model for theoretical studies in the context of the constrained structural scaffold with a relevant pharmacophore [4,5,42]. Furthermore, they are used in the synthesis of many natural products, alkaloids [23]. DKP framework is present in culture broths fermented with lactic acid bacteria. Therefore, they provide an eco-friendly approach for food and feed preservation [43].

3.3. DKP-Based Drugs

Interestingly, many relatively new drugs, such as tadalafil, phosphodiesterase-5 inhibitor for the treatment of erectile dysfunction [44], retosiban, as an oxytocin antagonist for preterm labor [45], epelsiban, as an oxytocin antagonist in premature ejaculation in men [4], the vascular disrupting and tubulin-depolymerizing pinabulin, on the base of the marine fungal halimide, potential therapeutical agent in lung cancer [46,47], as well as other anticancer natural agents as ambewelamides, phenylahistin, dehydrophenylahistin [48], verticillin A [49], but also Aplaviroc for HIV [50], antiviral and immunosuppressive activities (gliotoxin and sirodesmin PL), antimicrobial pulcherimin, antibacterial albonoursin, brevianamide S or bicyclomycin, avrainvillamide [51], anti-inflammatory agents, e.g., FR106969 and FR900452 [6], and many others, contain DKP core.

3.4. Cyclo-Dipeptides Containing Proline: Towards Effective Therapies

Currently, the growing attention in terms of cyclic dipeptides containing proline moiety is noticeable. Among the amino acids, proline is unique due to its specific structure and a lot of biological properties [52]. Both L and D-proline-based DKPs, widespread in nature, are an interesting sub-family of cyclo-dipeptides [53]. It could be mentioned that they are more predominant in heated and fermented foods than another type of DKPs, even at the level of ~90%. Pro-based DKPs are more easily generated in comparison to other DKPs and can be considered as significant components of flavor and bioactivity [28]. The structural complexity and bioactivity of them are highly impressive [54], inter alia in the stimulation of hematopoiesis, bacterial, viral, and fungal infections, food intake inhibition, control the activity of many receptors, as markers in the protein pyrolysates [55] etc. [28,36,56]. They have been potential as cytotoxic and anticancer agents, in treating renal inflammation [57] or cardio-metabolic disorders [58]. In the proline-containing cyclo-dipeptides, the 2,5-DKP a six-membered piperazine nucleus is fused to pyrrolidine ring, which leads to prominent bio-properties, such as extra rigid heterocyclic structure, and consequently, inter alia greater resistance to degradation by enzymes. Proline- and hydroxyproline-based DKPs can form novel essential quorum sensing inhibitors, which are involved in intracellular communication [38]. They are also promising candidates in neurodegeneration prevention, e.g., in the treatment of Alzheimer disease [59], in flavor response [28], etc.
General mechanism of DKP formation was decribed by Gomes et al. [60], while mechanism of proline-based DKPs formation is proposed by Otsuka et al. [9,28].
The diverse mechanisms of DKP bio-actions peptides as well as their targets have not been precisely known and understood yet. Nevertheless, the growing number of scientific reports in this topic are observable. As an example, neuroprotective action of diketopiperazine-(proline) based agents via different mechanisms are discussed by Cornachia et al. [34], while antibacterial, antifungal activities are considered by Zhao et al. [18]. They proved that presence of specific substituents and other modifications have relevance for endowing bio-action. Propositions of mechanisms via inter alia different inhibitions were summarized.

3.5. Databases Survey

The survey of the Research Collaboratory for Structural Bioinformatics Protein Data Bank (RCSB PDB) [61] revealed 42 bio-complexes related to DKP moiety, including proline-based one, signed by reference codes: 6SSG, 6SSF, 6SSE, 6SSD [62], 6F0B, 6F0C [63], 41CT, 41Q7, 41PS, 41PW, 41Q9 [64], 5YL4 [65], 6EZ3, 5MLQ, 5OCD, 5MLP [66], 4Q24 [67], 4E0T, 4E0U [68], 3S7T, 3OQJ, 3OQH, 3OQI [69], 3OQV [70], 3NC6, 3NC7, 3NC3, 3NC5 [71], 3N1A [72], 2X9Q [73], 3G5H, 3G5F [74], 1W1T, 1W1P, 1W1V, 1W1Y [75], 6VXV, 6VZB, 6WOS, 6VZA [76], 1O6I [77]. The latter represents proline-based cyclic-dipeptide as chitinase inhibitor with chemotherapeutic potential against fungi, insects and protozoan/nematodal parasites—see Figure 1. In the previous cases, insight into the mechanism of action of biomacromolecules, especially a new class of small proteins, cyclodipeptide sythases, the molecular bases of the interactions with DKP ring towards the design of more effective diverse therapetical agents are discussed. Interestingly, only one complex, 1QZR [78], contains 2,6-DKP core.
On the other hand, the thorough analysis of the Cambridge Structure Database (CSD) [79,80] leads to 256 entries of 2,5-DKPs, 52 2,3-DKPs and 5 hits of 2,6-DKPs (see Scheme, and Supplementary Materials).
It is noteworthy that CSD collects a huge structural knowledge on potential peptide-based ligands that can be applicable at the macromolecular level. Small-molecular crystal structures, especially peptides and their derivatives, have a natural synergy with proteins. The rational design of modern effective ligands should be based not only on the 3D-structure of macromolecular target but also on potential ligands. Groom and Cole [81] said that the designers should try to “understand and exploit what small-molecule crystal structures tell them; it is just a matter of listening”.

4. Conclusions and Future Prospects

Taking all the above into account, we can conclude that DKPs in general, and proline-based DKPs peptides especially, offer a highly functionalized natural arsenal and huge potential as biological tools for either better understanding bio-mechanisms or future more effective therapies. These the simplest either natural or unnatural cyclo-peptides possessing economically beneficial biological properties are valuable molecular scaffolds in synthetic biology and protein engineering.

Supplementary Materials

The following are available online at www.mdpi.com/2504-3900/83/1/10/s1, Table S1: Crystal structures containing DKP moiety, retrieved from the CSD.

Author Contributions

Conceptualization, J.B.; software, J.B.; writing—original draft preparation, J.B. writing—review and editing, J.B. and W.M.W.; supervision, W.M.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Joo, S.H. Cyclic peptides as therapeutic agents and biochemical tools. Biomol. Ther. 2012, 20, 19–26. [Google Scholar] [CrossRef] [PubMed]
  2. Gang, D.; Kim, D.W.; Park, H.S. Cyclic peptides: Promising scaffolds for biopharmaceuticals. Genes 2018, 9, 557. [Google Scholar] [CrossRef] [PubMed]
  3. Feni, L.; Jutten, L.; Parente, S.; Piarulli, U.; Neundorf, I.; Dia, D. Cell-penetrating peptides containing 2,5-DKP scaffolds as shuttles for anti-cancer drugs: Conformational studies and biological activities. Chem. Commun. 2020, 56, 5685–5688. [Google Scholar] [CrossRef] [PubMed]
  4. Borthwick, A.D. 2,5-Diketopiperazines: Synthesis, reactions, medicinal chemistry, and bioactive natural products. Chem. Rev. 2012, 112, 3641–3716. [Google Scholar] [CrossRef]
  5. Martins, M.B.; Carvalho, I. Diketopiperazines: Biological activity and synthesis. Tetrahedron 2007, 63, 9923–9932. [Google Scholar] [CrossRef]
  6. Giessen, T.W.; Marahiel, M.A. Rational and combinatorial tailoring of bioactive cyclic dipeptides. Front. Microbiol. 2015, 6, 785. [Google Scholar] [CrossRef]
  7. Ma, Y.M.; Liang, X.A.; Kong, Y.; Jia, B. Structural diversity and biological activities of indole diketopiperazine alkaloids from fungi. J. Agric. Food Chem. 2016, 64, 6659–6671. [Google Scholar] [CrossRef]
  8. Sano, S.; Nakao, M. Chemistry of 2,5-diketopiperazine and its bis-lactim ether: A brief review. Heterocycles 2015, 91, 1349–1375. [Google Scholar] [CrossRef]
  9. Ying, J.; Lin, R.; Xu, P.; Wu, Y.; Zhao, Y. Prebiotic formation of cyclic dipeptides under potentially early Earth conditions Sci. Rep. 2018, 8, 936. [Google Scholar]
  10. Gondry, M.; Sauguet, L.; Belin, P.; Thai, R.; Amouroux, R.; Tellier, C.; Tuphile, K.; Jacquet, M.; Braud, S.; Courcon, M.; et al. Cyclodipeptide synthases are a family of tRNA-dependent peptide bond-forming enzymes. Nat. Chem. Biol. 2009, 5, 414–420. [Google Scholar] [CrossRef]
  11. García-Estrada, C.; Ullán, R.; Albillos, S.; Fernández-Bodega, M.; Durek, P.; Vondöhren, H.; Martín, J. A single cluster of coregulated genes encodes the biosynthesis of the mycotoxins roquefortine C and meleagrin in Penicillium chrysogenum. Chem. Biol. 2011, 18, 1499–1512. [Google Scholar] [CrossRef] [PubMed]
  12. Borgman, P.; Lopez, R.D.; Lane, A.L. The expanding spectrum of DKP natural product biosynthetic pathways containing cyclodipeptide synthases. Org. Biomol. Chem. 2019, 17, 2305–2314. [Google Scholar] [CrossRef] [PubMed]
  13. Moutiez, M.; Belin, P.; Gondry, M. Aminoacyl-tRNA-Utilizing Enzymes in Natural Product Biosynthesis. Chem. Rev. 2017, 117, 5578–5618. [Google Scholar] [CrossRef] [PubMed]
  14. Payne, J.A.; Schoppet, M.; Hansen, M.H.; Cryle, M.J. Diversity of nature's assembly lines – recent discoveries in non-ribosomal peptide synthesis. Mol. BioSyst. 2016, 13, 9–22. [Google Scholar] [CrossRef]
  15. Elkahoui, S.; Abdel Rahim, H.; Tabbene, O.; Shaaban, M.; Limam, F.; Laatsch, H. Cyclo-(His,Leu): A new microbial diketopiperazine from a terrestrial Bacillus subtilis strain B38. Nat. Prod. Res. 2013, 27, 108–116. [Google Scholar] [CrossRef] [PubMed]
  16. Holden, M.T.; Ram Chhabra, S.; de Nys, R.; Stead, P.; Bainton, N.J.; Hill, P.J.; Manefield, M.; Kumar, N.; Labatte, M.; England, D.; et al. Quorum-sensing cross talk: Isolation and chemical characterization of cyclic dipeptides from Pseudomonas aeruginosa and other Gram-negative bacteria. Mol. Microbiol. 1999, 33, 1254–1266. [Google Scholar] [CrossRef]
  17. Ström, K.; Sjögren, M.; Broberg, A.; Schnürer, J. Lactobacillus plantarum MiLAB 393 produces the antifungal cyclic dipeptides cyclo(L-Phe-L-Pro) and cyclo(L-Phe-trans-4-OH-L-Pro) and 3-phenyllactic acid. Appl. Environ. Microbiol. 2002, 68, 4322–4327. [Google Scholar] [CrossRef]
  18. Zhao, P.; Xue, Y.; Li, J.; Li, X.; Zu, X.; Zhao, Z.; Quan, C.; Gao, W.; Feng, S. Non-lipopeptide fungi-derived peptide antibiotics developed since 2000. Biotechnol. Lett. 2019, 41, 651–673. [Google Scholar] [CrossRef]
  19. Lin, A.; Fang, Y.; Zhu, T.; Gu, Q.; Zhu, W. A new diketopiperazine alkaloid isolated from an algicolous Aspergillus flavus strain. Pharmazie 2008, 63, 323–325. [Google Scholar]
  20. Stierle, A.C.; Cardellina, J.H.; Strobel, G.A. Maculosin, a host-specific phytotoxin for spotted knapweed from Alternaria alternate. Proc. Natl. Acad. Sci. USA 1988, 85, 8008–8011. [Google Scholar] [CrossRef]
  21. Kazlauskas, R.; Murphy, P.T.; Wells, P.J. A diketopiperazine derived from trichloroleucine from the sponge Dysidea herbacea. Tetrahedron Lett. 1978, 49, 4945–4948. [Google Scholar] [CrossRef]
  22. Prasad, C. Cyclo(His-Pro): Its distribution, origin and function in the human. Neurosci. Biobehav. Rev. 1988, 12, 19–22. [Google Scholar] [CrossRef]
  23. Huang, R.M.; Yi, X.X.; Zhou, Y.; Su, X.; Pengs, Y.; Gao, C.H. An Update on 2,5-Diketopiperazines from Marine Organisms. Mar. Drugs 2014, 12, 6213–6235. [Google Scholar] [CrossRef] [PubMed]
  24. Harizani, M.; Katsini, E.; Georgantea, P.; Roussis, V.; Ioannou, E. New chlorinated 2,5DKPs from marine-derived bacteria isolated from sediments of the eastern Mediterranean sea. Molecules 2020, 25, 1509. [Google Scholar] [CrossRef] [PubMed]
  25. Ginz, M.; Engelhardt, U.H. Identification of proline-based diketopiperazines in roasted coffee. J. Agric. Food Chem. 2000, 48, 3528–3532. [Google Scholar] [CrossRef]
  26. Chen, M.Z.; Dewis, M.L.; Kraut, K.; Merritt, D.; Reiber, L.; Trinnaman, L.; Da Costa, N.C. 2,5-Diketopiperazines (cyclic dipeptides) in beef: Identification, synthesis, and sensory evaluation. J. Food Sci. 2009, 74, C100–C105. [Google Scholar] [CrossRef]
  27. Gautschi, M.; Schmid, J.P.; Peppard, T.L.; Ryan, T.P.; Tuorto, R.M.; Yang, X. Chemical characterization of diketopiperazines in beer. J. Agric. Food Chem. 1997, 45, 3183–3189. [Google Scholar] [CrossRef]
  28. Otsuka, Y.; Arita, H.; Sakaji, M.; Yamamoto, K.; Kashiwagi, T.; Shimamura, T.; Ukeda, H. Investigation of the formation mechanism of proline-containing cyclic dipeptide from the linear peptide. Biosci. Biotechnol. Biochem. 2019, 12, 2355–2363. [Google Scholar] [CrossRef]
  29. Bojarska, J.; Remko, L.; Maniukiewicz, W.; Sieron, M. An orthorhombic polymorph of a cyclization product of perindopril. Acta Crystallogr. C 2013, 69, 630–633. [Google Scholar] [CrossRef]
  30. Bojarska, J.; Maniukiewicz, W.; Główka, M.L.; Siero’n, L.; Remko, M. Crystal structure of perindopril cyclization product. J. Chil. Chem. Soc. 2013, 58, 1530–1532. [Google Scholar] [CrossRef]
  31. Remko, M.; Bojarska, J.; Jezko, L.; Olczak, A.; Maniukiewicz, W. Molecular structure of antihypertensive drug perindopril, its active metabolite perindoprilat and impurity F. J. Mol. Struct. 2013, 1036, 292–297. [Google Scholar] [CrossRef]
  32. Yue, L.; Fangfang, L.; Yanyan, Z.; Li, X.; Zhou, Z.; Liu, C.; Zhang, W.; Tang, M. DFT study onreaction mechanisms of cyclic dipeptide generation. Struct. Chem. 2016, 27, 1165–1173. [Google Scholar]
  33. Corey, R. Diketopiperazine. J. Am. Chem. Soc. 1938, 60, 1598. [Google Scholar] [CrossRef]
  34. Cornacchia, C.; Cacciatore, I.; Baldassarre, L.; Mollica, A.; Feliciani, F.; Pinnen, F. Mini Reviews in Medicinal Chemistry. BenthamScience 2021, 21. [Google Scholar]
  35. Prasad, C. Bioactive cyclic dipeptides. Peptides 1995, 16, 151–164. [Google Scholar] [CrossRef]
  36. Wang, Y.; Wang, P.; Ma, H.; Zhu, W. Developments around the bioactive diketopiperazines: A patent review. Expert Opin. Ther. Patnets 2013, 23, 1415–1433. [Google Scholar] [CrossRef]
  37. Mishra, A.K.; Choi, J.; Choi, S.J.; Baek, K.H. Cyclodipeptides: An overview of their biosynthesis and biological activity. Molecules 2017, 22, 1796. [Google Scholar] [CrossRef]
  38. Sun, S.J.; Liu, Y.C.; Weng, C.H.; Sun, S.W.; Li, F.; Li, H.; Zhu, H. Cyclic dipeptides mediating quorum sensing and their bilogical effects in Hypsizygus Marmoreus. Biomolecules 2020, 10, 298. [Google Scholar] [CrossRef]
  39. Zhang, L.; Yu, H.; Zhao, X. Neuroprotective e_ects of salidroside against beta-amyloid-induced oxidative stress in SH-SY5Y human neuroblastoma cells. Neurochem. Int. 2010, 57, 547–555. [Google Scholar] [CrossRef]
  40. Fischer, P.M. Diketopiperazines in peptide and combinatorial chemistry. J. Peptide Sci. 2003, 9, 9–35. [Google Scholar] [CrossRef]
  41. Dubois, P.; Correia, I.; Le Chevalier, F.; Dubois, S.; Jacques, I.; Canu, N.; Moutiez, M.; Thai, R.; Gondry, M.; Lequin, O.; et al. Reprogramming Escherichia coli for the production of prenylated indole diketopiperazine alkaloids. Sci. Rep. 2019, 9, 9208–9220. [Google Scholar] [CrossRef] [PubMed]
  42. Ressurreição, A.S.M.; Delatouche, R.; Gennari, C.; Piarulli, U. Bifunctional 2,5-diketopiperazines as rigid three-dimensional scaffolds in receptors and peptidomimetics. Eur. J. Org. Chem. 2011, 2, 217–228. [Google Scholar] [CrossRef]
  43. Crowley, S.; Mahony, J.; van Sinderen, D. Current perspectives on antifungal lactic acid bacteria as natural bio-preservatives. Trends Food Sci. Technol. 2013, 33, 93–109. [Google Scholar] [CrossRef]
  44. Sung, B.J.; Hwang, K.Y.; Jeon, Y.H.; Lee, J.I.; Heo, Y.S.; Kim, J.H.; Moon, J.; Yoon, J.M.; Hyun, Y.L.; Kim, E.; et al. Structure of the catalytic domain of human phosphodiesterase 5 with bound drug molecules. Nature 2003, 425, 98–102. [Google Scholar] [CrossRef] [PubMed]
  45. Liddle, J.; Allen, M.J.; Borthwick, A.D.; Brooks, D.P.; Davies, D.E.; Edwards, R.M.; Exall, A.M.; Hamlett, C.; Irving, W.R.; Mason, A.M.; et al. The discovery of GSK221149A: A potent and selective oxytocin antagonist. Bioorg. Med. Chem. Lett. 2008, 18, 90–94. [Google Scholar] [CrossRef]
  46. Gomes, N.G.M.; Lefranc, F.; Kijjoa, A.; Kiss, R. Can some marine-derived fungal metabolites become actual anticancer agents? Mar. Drugs 2015, 13, 3950–3991. [Google Scholar] [CrossRef]
  47. Mohanlal, R.W.; Lloyd, K.; Huang, L. Plinabulin, a novel small molecule clinical stage IO agent with anti-cancer activity, to prevent chemo–induced neutropenia and immune related AEs. J. Clin. Oncol. 2018, 36, 126. [Google Scholar] [CrossRef]
  48. Kanzaki, H.; Yanagisawa, S.; Kanoh, K.; Nitoda, T. A novel potent cell cycle inhibitor dehydrophenylahisti enzymatic synthesis and inhibitory activity towards eaurchin embryo. J. Antibiot. 2002, 55, 1042–1047. [Google Scholar] [CrossRef]
  49. Gu, B.; He, S.; Yan, X.; Zhang, L. Tentative biosynthetic pathways of some microbial diketopiperazines. Appl. Microbiol. Biotechnol. 2013, 97, 8439–8453. [Google Scholar] [CrossRef]
  50. Maeda, K.; Nakata, H.; Koh, Y.; Miyakawa, T.; Ogata, H.; Takaoka, Y. Spiro diketopiperazine-based CCR5 inhibitor whichpreservesCC-chemokine/CCR5 interactions and exerts potent activitya gainst R5 human immuno deficiency virus type 1 in vitro. J. Virol. 2004, 78, 8654–8662. [Google Scholar] [CrossRef]
  51. Sugie, Y.; Hirai, H.; Inagaki, T.; Ishiguro, M.; Kim, Y.J.; Kojima, Y. A new antibiotic CJ-17,665 from Aspergillusochraceus. J. Antibiot. 2001, 54, 911–916. [Google Scholar] [CrossRef] [PubMed]
  52. Bojarska, J.; Remko, M.; Breza, M.; Madura, I.D.; Kaczmarek, K.; Zabrocki, J.; Wolf, W.M. A supramolecular approach to structure-based design with a focus on synthon hierarchy in ornithine-derived ligands: Review, synthesis, experimental and in silico studies. Molecules 2020, 25, 1135. [Google Scholar] [CrossRef] [PubMed]
  53. Trost, B.M.; Stiles, D.T. Total synthesis of spirotryprostatin B via diastereoselective prenylation. Org. Lett. 2007, 9, 2763–2766. [Google Scholar] [CrossRef]
  54. Cui, C.; Kakeya, H.; Osada, H. Novel mammalian cell cycle inhibitors, cyclotroprostatins A–D, produced by Aspergillus fumigatus, which inhibit mammalian cell cycle at G2/M phase. Tetrahedron 1997, 53, 59–72. [Google Scholar] [CrossRef]
  55. Fabbri, D.; Adamiano, A.; Falini, G.; De Marco, R.; Mancini, I. Analytical pyrolysis of dipeptides containing proline and amino acids with polar side chains. Novel 2,5-diketopiperazine markers in the pyrolysates of proteins. J. Anal. Appl. Pyrolysis. 2012, 95, 145–155. [Google Scholar] [CrossRef]
  56. Xiang, W.X.; Liu, Q.; Li, X.M.; Lu, C.H.; Shen, Y.M. Four pairs of proline-containing cyclic dipeptides from Nocardiopsis sp. HT88, an endophytic bacterium of Mallotus nudiflorus L. Nat. Prod. Res. 2020, 34, 2219–2224. [Google Scholar] [CrossRef] [PubMed]
  57. Begum Ahil, S.; Hira, K.; Shaik, A.B.; Pal, P.P.; Kulkarni, O.P.; Araya, H.; Fujimoto, Y. L-Proline-based-cyclic dipeptides from Pseudomonas sp. (ABS-36) inhibit pro-inflammatory cytokines and alleviate crystal-induced renal injury in mice. Int. Immunopharmacol. 2019, 73, 395–404. [Google Scholar] [CrossRef]
  58. Li, F.; Liu, K.; Gray, C.; Harris, P.; Reynolds, C.M.; Vickers, M.H.; Guan. Cyclic glycine-proline normalizes systolic blood pressure in high-fat diet-induced obese male rats. J. Nutr. Metab. Cardiovasc. Dis. 2020, 30, 339–346. [Google Scholar]
  59. Turkez, H.; Cacciatore, I.; Arslan, M.E.; Fornasari, E.; Marinelli, L.; Di Stefano, A.; Mardinoglu, A. Histydyl-proline diketopiperazine isomers as multipotent anti-Alzheimer drug candidates. Biomolecules 2020, 10, 737–756. [Google Scholar] [CrossRef]
  60. Gomes, P.; Vale, R.; Moreira, R. Cyclization-activated prodrugs. Molecules 2007, 12, 2484–2506. [Google Scholar] [CrossRef]
  61. Burley, S.K.; Berman, H.M. RCSB Protein Data Bank: Biological macromolecular structures enabling research and education in fundamental biology, biomedicine, biotechnology and energy. Nucleic Acids Res. 2019, 47, D464–D474. [Google Scholar] [CrossRef] [PubMed]
  62. Gao, S.; Liu, H.; de Crecy-Lagard, V.; Zhu, W.; Richards, N.G.J.; Naismith, J.H. PMP-diketopiperazine adducts form at the active site of a PLP dependent enzyme involved in formycin biosynthesis. Chem. Commun. 2019, 55, 14502–14505. [Google Scholar] [CrossRef] [PubMed]
  63. Alkhalaf, L.M.; Barry, S.M.; Rea, D.; Gallo, A.; Griffiths, D.; Lewandowski, J.R.; Fulop, V.; Challis, G.L. Binding of Distinct Substrate Conformations Enables Hydroxylation of Remote Sites in Thaxtomin D by Cytochrome P450 TxtC. J. Am. Chem. Soc. 2019, 141, 216–222. [Google Scholar] [CrossRef]
  64. Fonvielle, M.; Le Du, M.H.; Lequin, O.; Jacquet, M.; Thai, R.; Dubois, S.; Gondry, M.; Belin, P. Substrate and Reaction Specificity of Mycobacterium tuberculosis Cytochrome P450 CYP121: Insights from biochemical studies and crystal structures. J. Biol. Chem. 2013, 288, 17347–17359. [Google Scholar] [CrossRef]
  65. Fu, Z.Y.; Li, W.B. Design, synthesis and biological activity evaluation of plinabulin derivatives based on co-crystal structure. Bioorganic & Medicinal Chemistry. 2018, 26, 2061–2072. [Google Scholar]
  66. Bourgeois, G.; Seguin, J.; Babin, M.; Belin, P.; Moutiez, M.; Mechulan, Y.; Gondry, M.; Schmitt, E. Structural basis for partition of the cyclodipeptide synthases into two subfamilies. J. Struct. Biol. 2019, 203, 17–26. [Google Scholar] [CrossRef] [PubMed]
  67. Moutiez, M.; Schmitt, E.; Seguin, J.; Thai, R.; Favry, E.; Belin, P.; Mechulan, Y.; Gondry, M. Unravelling the mechanism of non-ribosomal peptide synthesis by cyclodipeptide synthases. Nat. Commun. 2014, 5, 5141–5141. [Google Scholar] [CrossRef]
  68. Schuller, J.M.; Zocher, G.; Liebhold, M.; Xie, X.; Stahl, M.; Li, S.M.; Stehle, T. Structure and catalytic mechanism of a cyclic dipeptide prenyltransferase with broad substrate promiscuity. J. Mol. Biol. 2012, 422, 87–99. [Google Scholar] [CrossRef]
  69. Bonnefond, L.; Arai, T.; Sakaguchi, Y.; Suzuki, T.; Ishitani, R.; Nureki, O. Structural basis for nonribosomal peptide synthesis by an aminoacyl-tRNA synthetase paralog. Proc. Natl. Acad. Sci. USA 2011, 108, 3912–3917. [Google Scholar] [CrossRef]
  70. Sauguet, L.; Moutiez, M.; Li, Y.; Belin, P.; Seguin, J.; Le Du, M.H.; Thai, R.; Masson, C.; Fonvielle, M.; Pernodet, J.L.; et al. Cyclodipeptide synthases, a family of class-I aminoacyl-tRNA synthetase-like enzymes involved in non-ribosomal peptide synthesis. Nucleic Acids. Res. 2011, 39, 4475–4489. [Google Scholar] [CrossRef]
  71. Cryle, M.J.; Bell, S.G.; Schlichting, I. Structural and biochemical characterization of the cytochrome P450 CypX (CYP134A1) from Bacillus subtilis: A cyclo-L-leucyl-L-leucyl dipeptide oxidase. Biochemistry 2010, 49, 7282–7296. [Google Scholar] [CrossRef] [PubMed]
  72. Hsieh, Y.C.; Wu, Y.J.; Chiang, Y.Y.; Kuo, C.Y.; Shrestha, K.L.; Chao, C.F.; Huang, Y.C.; Chuankhayan, P.; Wu, W.G.; Li, Y.K.; et al. Crystal structures of bacillus cereus NCTU2 chitinase complexes with chitooligomers reveal novel substrate binding for catalysis: A chitinase without chitin-binding and insertion domains. J. Biol. Chem. 2010, 285, 31603–31615. [Google Scholar] [CrossRef]
  73. Vetting, M.W.; Hegde, S.S.; Blanchard, J.S. The Structure and Mechanism of the Mycobacterium Tuberculosis Cyclodityrosine Synthetase. Nat. Chem. Biol. 2010, 6, 797. [Google Scholar] [CrossRef] [PubMed]
  74. Belin, P.; Le Du, M.H.; Fielding, A.; Lequin, O.; Jacquet, M.; Charbonnier, J.B.; Lecoq, A.; Thai, R.; Courcon, M.; Masson, C.; et al. Identification and structural basis of the reaction catalyzed by CYP121, an essential cytochrome P450 in Mycobacterium tuberculosis. Proc. Natl. Acad. Sci. USA 2009, 106, 7426–7431. [Google Scholar] [CrossRef] [PubMed]
  75. Houston, D.R.; Synstad, B.; Ejsink, V.G.H.; Stark, M.J.; Eggleston, I.; Van Aalten, D.M.F. Structure-Based Exploration of Cyclic Dipeptide Chitinase Inhibitors. J. Med. Chem. 2004, 47, 5713. [Google Scholar] [CrossRef] [PubMed]
  76. Sun, C.; Luo, Z.; Zhang, W.; Deng, Z.; Mobli, M.; Kobe, B.; Jia, X.; Qu, X. Molecular Basis of Regio- and Stereo-Specificity in Biosynthesis of Bacterial Heterodimeric Diketopiperazines. Nature Commun. 2020, 11, 6251–6261. [Google Scholar] [CrossRef] [PubMed]
  77. Houston, D.R.; Eggleston, I.; Synstad, B.; Ejsink, V.G.; van Aalten, D.M. The cyclic dipeptide CI-4 [cyclo-(l-Arg-d-Pro)] inhibits family 18 chitinases by structural mimicry of a reaction intermediate. Biochem. J. 2002, 368, 23–27. [Google Scholar] [CrossRef]
  78. Classen, S.; Olland, S.; Berger, J.M. Structure of the topoisomerase II ATPase region and its mechanism of inhibition by the chemotherapeutic agent ICRF-187. Proc. Natl. Acad. Sci. USA 2003, 100, 10629–10634. [Google Scholar] [CrossRef]
  79. Allen, F.H. The Cambridge Structural Database: A quarter of a million crystal structures and rising. Acta Cryst. B 2002, 58, 380–388. [Google Scholar] [CrossRef]
  80. Cole, J.C.; Wiggin, S.; Stanzione, F. New insights and innovation from a million crystal structures in the Cambridge Structural Database. Struc. Dyn. 2019, 6, 1–6. [Google Scholar] [CrossRef]
  81. Groom, C.R.; Cole, J.C. The use of small-molecule structures to complement protein-ligand crystal structures in drug discovery. Acta Cryst. D 2017, 73, 240–245. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Structure of 2,5-DKP (on the left), 2,3-DKP (in the middle) and 2,6-DKP (on the right).
Scheme 1. Structure of 2,5-DKP (on the left), 2,3-DKP (in the middle) and 2,6-DKP (on the right).
Proceedings 79 00010 sch001
Figure 1. Bio-complex containing proline-based 2,5-DKP moiety, 1O6I.pdb [76].
Figure 1. Bio-complex containing proline-based 2,5-DKP moiety, 1O6I.pdb [76].
Proceedings 79 00010 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bojarska, J.; Wolf, W.M. Ultra-Short Cyclo-Peptides as Bio-Inspired Therapeutics: Proline-Based 2,5-Diketopiperazines (DKP). Proceedings 2021, 79, 10. https://doi.org/10.3390/IECBM2020-08804

AMA Style

Bojarska J, Wolf WM. Ultra-Short Cyclo-Peptides as Bio-Inspired Therapeutics: Proline-Based 2,5-Diketopiperazines (DKP). Proceedings. 2021; 79(1):10. https://doi.org/10.3390/IECBM2020-08804

Chicago/Turabian Style

Bojarska, Joanna, and Wojciech M. Wolf. 2021. "Ultra-Short Cyclo-Peptides as Bio-Inspired Therapeutics: Proline-Based 2,5-Diketopiperazines (DKP)" Proceedings 79, no. 1: 10. https://doi.org/10.3390/IECBM2020-08804

Article Metrics

Back to TopTop