Spontaneous In Vitro and In Vivo Interaction of (-)-Oleocanthal with Glycine in Biological Fluids: Novel Pharmacokinetic Markers - PubMed (original) (raw)

. 2021 Jan 5;4(1):179-192.

doi: 10.1021/acsptsci.0c00166. eCollection 2021 Feb 12.

Aimilia Rigakou 2, Andrew Brannen 1, Mohammed H Qusa 3, Niki Tasiakou 2, Panagiotis Diamantakos 2, Miranda N Reed 1 4, Peter Panizzi 1, Melissa D Boersma 5, Eleni Melliou 2, Khalid A El Sayed 3, Prokopios Magiatis 2, Amal Kaddoumi 1 4

Affiliations

Spontaneous In Vitro and In Vivo Interaction of (-)-Oleocanthal with Glycine in Biological Fluids: Novel Pharmacokinetic Markers

Lucy I Darakjian et al. ACS Pharmacol Transl Sci. 2021.

Abstract

Since the first discovery of its ibuprofen-like anti-inflammatory activity in 2005, the olive phenolic (-)-oleocanthal gained great scientific interest and popularity due to its reported health benefits. (-)-Oleocanthal is a monophenolic secoiridoid exclusively occurring in extra-virgin olive oil (EVOO). While several groups have investigated oleocanthal pharmacokinetics (PK) and disposition, none was able to detect oleocanthal in biological fluids or identify its PK profile that is essential for translational research studies. Besides, oleocanthal could not be detected following its addition to any fluid containing amino acids or proteins such as plasma or culture media, which could be attributed to its unique structure with two highly reactive aldehyde groups. Here, we demonstrate that oleocanthal spontaneously reacts with amino acids, with high preferential reactivity to glycine compared to other amino acids or proteins, affording two products: an unusual glycine derivative with a tetrahydropyridinium skeleton that is named oleoglycine, and our collective data supported the plausible formation of tyrosol acetate as the second product. Extensive studies were performed to validate and confirm oleocanthal reactivity, which were followed by PK disposition studies in mice, as well as cell culture transport studies to determine the ability of the formed derivatives to cross physiological barriers such as the blood-brain barrier. To the best of our knowledge, we are showing for the first time that (-)-oleocanthal is biochemically transformed to novel products in amino acids/glycine-containing fluids, which were successfully monitored in vitro and in vivo, creating a completely new perspective to understand the well-documented bioactivities of oleocanthal in humans.

© 2021 American Chemical Society.

PubMed Disclaimer

Conflict of interest statement

The authors declare the following competing financial interest(s): Amal Kaddoumi and Khalid El Sayed are Chief Scientific Officers without compensation in the Shreveport, Louisiana-based Oleolive, LLC. All other authors declare no conflict of interest.

Figures

Figure 1

Figure 1

Chemical structures of oleocanthal (1), oleocanthadiol (2), oleoglycine (3a,b) and tyrosol acetate (4).

Figure 2

Figure 2

Typical chromatograms obtained from spiked oleocanthal in culture medium and glycine solution. (A) Oleocanthal (OC) free culture medium. (B) Culture medium spiked with oleocanthal 25 μg/mL and kept at RT for 2 h, then analyzed. (C) Culture medium spiked with oleocanthal 25 μg/mL and heated for 2 h at 60 °C. (D) 1 mM glycine solution spiked with 100 μg/mL oleocanthal at RT. (E) 1 mM glycine solution spiked with 100 μg/mL oleocanthal and heated for 2 h at 60 °C. (F) Effect of temperature and time on oleoglycine and tyrosol acetate formation. Peaks of OC, oleocanthadiol (OCdiol), oleoglycine (OG), and tyrosol acetate (TA) are shown.

Figure 3

Figure 3

1H NMR spectra of oleocanthal (1) and oleocanthadiol (2) equilibrium in D2O (green spectrum) and the reaction mixture between oleocanthal and glycine (5× excess) after 10 min in D2O (red spectrum). The reaction between oleocanthal with glycine to afford α/β-oleoglycine (3a,b) is also shown.

Figure 4

Figure 4

Selected 1H 1H-COSY (bold), 1H 13C-HMBC (black arrows), and 1H15N-HMBC (green arrows) correlations for compounds 3a and 3b.

Figure 5

Figure 5

MS/MS fragmentation spectra provided by LC/Q-TOF analysis of oleoglycine.

Figure 6

Figure 6

15N-MS/MS fragmentation spectra provided by LC/Q-TOF analysis of 15N-tagged oleoglycine.

Figure 7

Figure 7

Tyrosol acetate (4) spectroscopic confirmation. (A) High-resolution ESI MS/MS fragmentation showing synthetic tyrosol acetate. (B) High-resolution ESI MS/MS from mouse plasma spiked with oleocanthal. (C) 1H NMR spectrum of tyrosol acetate in CDCl3. (D) 1H 13C HMBC spectrum of tyrosol acetate showing the key (red-circled) HMBC correlations.

Figure 8

Figure 8

Typical chromatograms obtained from oleocanthal transport study across bEnd3 cells monolayer. (A) Sample obtained from donor compartment after 1 h of adding culture medium containing 25 μg/mL oleocanthal and directly analyzed. (B) Same sample described for (A) but exposed to heat for 2 h at 60 °C. (C) Sample obtained from receiver compartment after 1 h of adding culture medium containing 25 μg/mL oleocanthal to the donor compartment and directly analyzed. (D) Same sample described for (C) but exposed to heat for 2 h at 60 °C. (E) Sample obtained from receiver compartment after 1 h of adding oleocanthal-free culture medium to the donor compartment and exposed to heat for 2 h at 60 °C. (F) Apparent permeability (_P_app, cm/s) of oleoglycine and tyrosol acetate as a function of oleocanthal concentration.

Figure 9

Figure 9

Typical chromatograms obtained from mouse plasma and brain homogenate following IV administration of oleocanthal (30 mg/kg, n = 3 mice/time point). (A) Plasma sample after 15 min of oleocanthal dosing and directly analyzed. (B) Same sample described for (A) but exposed to heat for 2 h at 60 °C. (C) Brain homogenate sample obtained before oleocanthal administration exposed to heat for 2 h at 60 °C. (D) Brain homogenate sample after 15 min of oleocanthal dosing and direct analysis. (E) Same sample described for (D) but exposed to heat for 2 h at 60 °C. (F) Mean mouse plasma concentration vs time profiles of oleoglycine (red curves) and tyrosol acetate (black curves) obtained directly without heat and after exposure to heat for 2 h at 60 °C.

Similar articles

Cited by

References

    1. Franconi F.; Campesi I.; Romani A. (2020) Is extra virgin olive oil an ally for women’s and men’s cardiovascular health?. Cardiovasc. Ther. 2020, 6719301.10.1155/2020/6719301. - DOI - PMC - PubMed
    1. Wongwarawipat T.; Papageorgiou N.; Bertsias D.; Siasos G.; Tousoulis D. (2017) Olive oil-related anti-inflammatory effects on atherosclerosis: potential clinical implications. Endocr., Metab. Immune Disord.: Drug Targets 18, 51–62. 10.2174/1871530317666171116103618. - DOI - PubMed
    1. Al Rihani S. B.; Darakjian L. I.; Kaddoumi A. (2019) Oleocanthal-rich extra-virgin olive oil restores the blood-brain barrier function through NLRP3 inflammasome inhibition simultaneously with autophagy induction in TgSwDI mice. ACS Chem. Neurosci. 10, 3543–3554. 10.1021/acschemneuro.9b00175. - DOI - PMC - PubMed
    1. Batarseh Y. S.; Kaddoumi A. (2018) Oleocanthal-rich extra-virgin olive oil enhances donepezil effect by reducing amyloid-β load and related toxicity in a mouse model of Alzheimer’s disease. J. Nutr. Biochem. 55, 113–123. 10.1016/j.jnutbio.2017.12.006. - DOI - PMC - PubMed
    1. Qosa H.; Batarseh Y. S.; Mohyeldin M. M.; El Sayed K. A.; Keller J. N.; Kaddoumi A. (2015) Oleocanthal enhances amyloid-β clearance from the brains of TgSwDI mice and in vitro across a human blood-brain barrier model. ACS Chem. Neurosci. 6, 1849–1859. 10.1021/acschemneuro.5b00190. - DOI - PMC - PubMed

LinkOut - more resources