Next Article in Journal
A Theoretical Analysis for Assessing the Variability of Secondary Model Thermal Inactivation Kinetic Parameters
Next Article in Special Issue
In Situ Raman Analysis of CO2—Assisted Drying of Fruit-Slices
Previous Article in Journal
Acknowledgement to Reviewers of Foods in 2016
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Extraction, Identification and Photo-Physical Characterization of Persimmon (Diospyros kaki L.) Carotenoids

by
Khalil Zaghdoudi
1,2,3,
Orleans Ngomo
1,4,
Régis Vanderesse
2,
Philippe Arnoux
1,
Bauyrzhan Myrzakhmetov
1,
Céline Frochot
1,* and
Yann Guiavarc’h
1
1
Laboratoire Réactions et Génie des Procédés, Université de Lorraine-CNRS, UMR 7274, 1 Rue Grandville, BP 20451, 54001 Nancy Cedex, France
2
Laboratoire de Chimie Physique Macromoléculaire, Université de Lorraine-CNRS, UMR 7375, 1 Rue Grandville, BP 20451, 54001 Nancy Cedex, France
3
Laboratoire d’Application de la Chimie aux Ressources et Substances Naturelles et à l’Environnement, Faculté des Sciences de Bizerte, Université de Carthage, Carthage 1054, Tunisie
4
Laboratoire des Matériaux et Chimie Industrielle Inorganique, ENSAI—University of Ngaoundere, Ngaoundere 454, Cameroun
*
Author to whom correspondence should be addressed.
Submission received: 24 September 2016 / Revised: 21 December 2016 / Accepted: 3 January 2017 / Published: 12 January 2017
(This article belongs to the Special Issue High Pressure Technologies in Food Processing)

Abstract

:
Carotenoid pigments were extracted and purified from persimmon fruits using accelerated solvent extraction (ASE). Eleven pigments were isolated and five of them were clearly identified as all-trans-violaxanthine, all-trans-lutein, all-trans-zeaxanthin all-trans-cryptoxanthin and all-trans-β-carotene. Absorption and fluorescence spectra were recorded. To evaluate the potential of 1O2 quenching of the purified carotenoids, we used a monocarboxylic porphyrin (P1COOH) as the photosensitizer to produce 1O2. The rate constants of singlet oxygen quenching (Kq) were determined by monitoring the near-infrared (1270 nm) luminescence of 1O2 produced by photosensitizer excitation. The lifetime of singlet oxygen was measured in the presence of increasing concentrations of carotenoids in hexane. Recorded Kq values show that all-trans-β-cryptoxanthin, all-trans-β-carotene, all-trans-lycopene and all-trans-zeaxanthin quench singlet oxygen in hexane efficiently (associated Kq values of 1.6 × 109, 1.3 × 109, 1.1 × 109 and 1.1 × 109 M−1·s−1, respectively). The efficiency of singlet oxygen quenching of β-cryptoxanthin can thus change the consideration that β-carotene and lycopene are the most efficient singlet oxygen quenchers acting as catalysts for deactivation of the harmful 1O2.

1. Introduction

Carotenoids are some of the most widespread lipidic soluble pigments in nature; more than 750 naturally occurring carotenoids have been identified to date, but only a few are commercially available in a pure form and are expensive. They are produced as secondary metabolites by all photosynthetic organisms (plants, fungi, yeasts and non-photosynthetic bacteria) and are also accumulated from human and animal diets [1,2]. Carotenoids contribute to both organoleptic and nutritional properties of food and food by-products providing red, yellow or orange colors for fruits and vegetables. They are used as food colorants in foods and animal products such as egg yolk, butter, crustaceans, trout, salmon and shrimp [3]. They also contribute to the pigmentation of birds’ plumage [4]. Generally, most carotenoids can be derived from C40 tetraterpenoids polyene backbone including a system of conjugated double bonds in which the π electrons are effectively delocalized over the entire chain; the skeleton can carry cyclic ends groups with substitution of functional groups containing oxygen [5]. Through their structure, carotenoids are classified into two main groups: (i) carotenes also called carotenoid hydrocarbons, which only contain carbon and hydrogen; and (ii) xanthophylls or oxygenated carotenoids that may contain different functional groups (epoxy, methoxy, hydroxy, carbonyl and carboxyl acid groups) [6,7]. The most characteristic feature of carotenoids is light absorption in the high-energy zone of the visible region (400–500 nm) by the long conjugated double bond system. Such a conjugated chain structure is crucial for carotenoid function especially in photosynthetic organisms as well as for photo-protection in all living organisms [8]. Carotenoids are involved in photosystem assembly, where light harvesting provides protection from excess light absorption through energy dissipation and radical detoxification [9].
Besides their role in photosynthesis and photo-protection, carotenoids are involved in the prevention of several human diseases and represent part of the antioxidant defense system in humans [10]. They act as cardioprotective agents and prevent acute myocardial infarction [11], prevent the development of reactive oxygen species-mediated disorders related to oxidative stress and can be responsible for reducing the low-density lipoprotein (LDL) cholesterol level [12]. Since two decades, several prospective studies have also shown a correlation between the consumption of fruits and vegetables rich in carotenoids and a decreased risk of many types of cancer [13]. Among these malignant diseases, prostate cancer [14], liver tumors, skin tumorigenesis and breast carcinogenesis [15] can be cited. Each carotenoid can be particularly involved—for example, canthaxanthin suppresses human colon cancer cells proliferation [16], while astaxanthin exhibits inhibitory activity in relation to cancer development in urinary bladder [17] and colorectum [18].
As previously written, one of the carotenoid functions is the protection against damage mediated by light, and carotenoids can act as quenchers to prevent the formation of 1O2 in biological systems. Moreover, in the presence of photosensitizers (PS) such as bacteriochlorophyll, chlorophyll or protoporphyrin IX, which absorb light and generate 1O2, they can quench the triplet state energy from the sensitizer or from 1O2 [19]. Foote et al. were the first to carry out a comprehensive study of 1O2 quenching mechanisms by β-carotene [20]. Due to these 1O2 quenching properties, the use of carotenoids for anticancer photodynamic therapy (PDT) has been developed, particularly to design “photomolecular beacons” (PMB), which are composed of an energy acceptor and an energy donor connected one to the other by a link (usually a peptide or nucleotide sensitive to an endogenous stimulus). This concept has been first developed by Zheng’s team [21,22] and is a part of our current studies [23]. Briefly, the energy acceptor (carotenoid) and the donor (PS) are maintained sufficiently close to each other to allow a transfer of energy, and the carotenoid can quench 1O2 generation through PS triplet-state energy transfer and/or 1O2 scavenging. If a specific enzyme cleaves the peptidic or nucleosidic link the photophysical properties of the photosensitizer is restored and PDT can be applied.
The aim of this study was to try to find unusual carotenoids, which would have a potential application in PDT as 1O2 quenchers. Eleven carotenoids were extracted and purified from persimmon fruit cultivated in Tunisia and known for its richness in carotenoid pigments, using accelerated solvent extraction (ASE) [24]. After separation, the structure of five of the carotenoids was determined and their potential evaluated as singlet oxygen quenchers in hexane.

2. Materials and Methods

2.1. Chemicals

All-trans-β-carotene (Type II synthetic, purity >95%), lutein from marigold, β-cryptoxanthin (purity >97%), zeaxanthin (purity >95%), lycopene from tomato (purity >90%), butylhydroxytoluene (BHT, purity ≥99%) and triethylamine (TEA, purity ≥99%) were obtained from Sigma-Aldrich (Shnelldorf, Germany). High-performance liquid chromatography (HPLC) organic solvents were of analytical grade: methanol (MeOH), absolute ethanol (EtOH), chloroform and hexane were from Carlo Erba (Val-de-Reuil, France), methyl-tert-butyl-ether (MTBE) was from Fisher Scientific (Loughborough, UK), petroleum ether (PE) was from VWR Prolabo (Fontenay-sous-Bois, France). Ultrapure water was obtained from a purified water system Arium® 611UV from Sartorius (Göttingen, Germany) with a resistivity of 18.2 MΩ*cm. Sodium chloride (NaCl, purity >99%) and potassium hydroxide (KOH, purity >99%) were obtained from VWR Prolabo (Fontenay-sous-Bois, France), and activated magnesium silicate Florisil® at 100–200 mesh (Sigma-Aldrich, Shnelldorf, Germany). Liquid Nitrogen was from Air Liquide (Nancy, France). The synthesis of 5-(4-carboxyphenyl)-10,15,20-triphenyl porphyrin (P1COOH) has been described elsewhere [25].

2.2. Fruit Materials

The present study focuses on 27 persimmon fruit (Diospyros kaki L.) cultivated in Tunisia (persimmon known under the appellation “Krima”); persimmon fruits were harvested at full maturity from three selected trees. The choice of persimmon as extraction matrix was built on earlier observations from previous investigation on carotenoid profiling from Prunus persica (peach), Prunus armeniaca (apricot) and Diospyros kaki (persimmon) fruits, largely produced in Tunisia [24], which shows that persimmon hole fruit is a rich source of carotenoids (xanthophylls and carotenes). Persimmon carotenoid profiles gave a large pigment range to assay; however, commercial carotenoids are still (i) rare; (ii) expensive; and (iii) with doubtful purity.

2.3. Sample Preparation and Characterization

Selected fruits were immediately washed with distillated water and placed in hermetical polystyrene boxes in which we added some liquid nitrogen in order to prevent oxidation reactions, before to be stored at −20 °C. Whole fruits (flesh and peel, 5 kg) were manually sliced into approximately 2 g pieces, frozen at −80 °C, and at once freeze-dried during 48 h with a pilot CryoTec freeze-dryer (Saint-Gély-du-Fesc, France) with a separated cold trap at −70 °C, and plate temperature set to −20 °C at a working pressure of 15 Pa. Freeze-dried fruit pieces were then placed in a desiccator containing P2O5 in order to prevent water sorption. After freeze-drying, fruit pieces were immediately grounded with a Moulinex coffee grinder (Moulinex, Ecully, France). A small amount of liquid nitrogen were added to prevent oxidation during grinding. Powders were stored in a desiccator containing P2O5 in order to prevent water sorption, and their residual water content was measured after 24 h at 105 °C. Particle size distribution of fruit powders was measured through dynamic light scattering with a Malvern Mastersizer 2000 apparatus (Malvern Instruments Ltd, Malvern, UK) in absolute ethanol.

2.4. Accelerated Solvent Extraction of Carotenoids

Carotenoid extraction experiments were carried out in a static mode with a Dionex ASE 350 extractor (Salt Lake City, UT, USA) using 12 stainless steel cells of 22 mL volume. Amber vials (60 mL) were used for extract collection. Common extraction parameters were as follows: (i) fruit powder (2 g) was loaded into the cell, and residual cell volume was partially filled with 2 mm glass beads; (ii) the cell was filled with solvent to a pressure of 1500 psi (103 bars); (iii) heat was applied for an initial heat-up time of 5 min; (iv) static extraction with all system valves closed was performed with a 5 min cycle time; (v) the cell was rinsed with 60% of the cell volume with extraction solvent; (vi) the solvent was purged from the cell with N2 gas for 60 s, and (vii) the system was depressurized. Extraction solvents, MeOH and tetrahydrofuran (THF), were degassed for 30 min by ultrasonic bath Sonoclean® Labo Moderne (Paris, France) before use. Florisil (2 g) was added as a dispersing agent for persimmon powder. Extraction parameters were set according to the common optimal conditions found in our previous study through an experimental design [24] as follows: temperature was set at 40 °C, 5 min static time extraction (one cycle), and binary extraction solvent (MeOH/THF, 20/80, v/v) under 103 bars.

2.5. Preparation of ASE Carotenoids Extracts for HPLC Semi-Preparative Purification

Pressurized liquid carotenoid extracts were prepared according to the method described in [24] and were conducted under limited light. At first, the ASE extract was homogenized with a vortex (Scientific Industries Inc., New York, NY, USA) at 13,500 rpm during 1 min before being filtered under vacuum through an AP 25 glass fiber filter disc of 2 µm porosity (Millipore, Darmstadt, Germany). Then, 10 mL of the filtrate was added to 10 mL of 10% (w/v) aqueous sodium chloride solution and carotenoids were transferred to a 10 mL petroleum ether phase containing 0.1% BHT (w/v) using a liquid–liquid extraction repeated at least three times in order to achieve a total transfer of carotenoids toward the petroleum ether phase. Organic phase was subsequently washed three times with 10 mL of ultrapure water in order to remove residual NaCl traces under neutral pH. Petroleum ether was then evaporated at 30 °C under nitrogen flow using a Turbo Vap® LV (Biotage AB, Uppsala, Sweden). The residue was rapidly dissolved in 5 mL of mobile phase (MeOH/MTBE, 1/1, v/v) containing 0.1% BHT (w/v) and filtered through 0.45 µm polyvinylidene fluoride (PVDF) syringe filters (Pall Life Sciences, Ann Arbor, PN, USA). Then, the filtered solution was mixed with 2.5 mL of MeOH containing 0.1% BHT (w/v) and saponified with 2.5 mL of 20% (w/v) methanolic KOH solution under nitrogen, in the dark, for 1 h, at room temperature. Finally, the saponified sample was injected for the semi preparative assay.

2.6. Semi Preparative HPLC Reversed Phase Carotenoid Purifications

Carotenoid separation was performed by reverse phase chromatography on a preparative Gilson HPLC GX-271 system consisting of a GX-271 Handler injector/collector equipped with a single 402 syringe pump, a 322 Semi-preparative HPLC pump and a 156 DAD detector at 450 nm from Gilson (Middleton, WI, USA) controlled by Trilution LCv2.1 software (Gilson, Middleton, WI, USA) We used a 250 length × 4.6 mm inside diameter S-5 µm Yamamura Chemical (YMC) C30 semi preparative column (Yamamura Chemical Europe GmbH, Schöttmannshof, Germany) coupled to a 10 × 4.0 mm inside diameter (ID) S-5µm guard cartridge semi preparative column (ImChem, Versailles, France). The purification assay was performed at 25 °C. The mobile phase consisted of a MeOH gradient containing 0.05 M ammonium acetate (A), and MTBE (B). Both solvents A and B contained 0.1% BHT (w/v) and 0.05% of triethylamine (TEA) (w/v). The flow rate was 4.0 mL/min and the injected volume was 3.0 mL. The gradient profile of the mobile phase was set as follows: a linear increasing gradient from 2% B to 20% B in 60 min; a linear increasing gradient from 20% B to 95% B in 20 min; a linear gradient at 95% B in 10 min, decreasing from 95% B to 2% B in 5 min; and a linear gradient at 2% B in 5 min. The column was equilibrated for 10 min in the starting conditions before each injection. Before each batch of HPLC semi-preparative purification, the stabilization time of the column was 30 min, and a blank (MeOH/MTBE, 1/1, v/v) was injected. The same pooled fraction of carotenoids was collected and then concentrated at 30 °C under nitrogen flow using a Turbo Vap® LV (Biotage AB, Uppsala, Sweden) in amber glass vials prior to characterization.

2.7. Identification of Unknown Carotenoids

In order to identify additional carotenoids and compare the fruits carotenoids profiles, liquid chromatography-mass spectrometry (LC-MS) runs were performed in positive atmospheric pressure chemical ionization (APCI) mode with a Shimadzu LC-MS-2020 liquid chromatograh-mass spectrometer (Shimadzu, Marne-la-Vallée, France). A scanning rate of 2143 u/s was used in the range 50–2000 amu. Nebulizing gas flow was fixed at 1.5 mL/min. Interface voltage and temperature were 4.5 kV and 250 °C, respectively. The gradient profile of the mobile phase was set as follows: linear increasing gradient from 5% B to 30% B in 30 min; linear increasing gradient from 30% B to 95% B in 20 min; and linear decreasing gradient from 95% B to 5% B in 10 min. The column was equilibrated for 5 min at the starting conditions before each injection.

2.8. Photophysical Properties

Absorption spectra were recorded on a PerkinElmer (Lambda EZ 210, Shelton, CT, USA) double beam UV-visible spectrophotometer. Fluorescence spectra were recorded on a Horiba Jobin Yvon Fluorolog-3 (FL3-222) spectrofluorometer equipped with a 450 W Xenon lamp, a thermostated cell compartment (25 °C), an R928 (Horiba, Longjumeau, France) UV-visible photomultiplier and a liquid nitrogen cooled InGaAs infrared detector (DSS-16A020L Electro-Optical System Inc., Phoenixville, PA, USA). The excitation spectrometer is a SPEX double grating monochromator (1200 grating/mm blazed 330 nm). The fluorescence was measured by the UV-visible detector through a SPEX double grating monochromator (600 grooves/mm blazed 1 nm). Singlet oxygen lifetime measurements of 5-(4-carboxyphenyl)-10,15,20-triphenyl porphyrin (P1COOH) with or without carotenoids were performed on a TEMPRO-01 spectrophotometer (Horiba Jobin Yvon, Palaiseau, France) composed of a pulsed diode excitation source SpectraLED-415 emitting at 415 nm, a cuvette compartment, a Seya–Namioka type emission monochromator (600–2000 nm) and a H10330-45 near-infrared photomultiplier tube with a thermoelectric cooler (Hamamatsu) as the detection device. The system was monitored by a single photon counting FluoroHub-B controller and the DataStation and DAS6 software (version, Horiba Jobin Yvon, City, US State abbrev if applicable, Country).
The Stern–Volmer equation was used to evaluate the constant quenching Kq of 1O2 by the different carotenoids τ0/τ = 1 + Kq(Car), where τ0 is the 1O2 lifetime after excitation of P1COOH in hexane without carotenoid, τ is the 1O2 lifetime after excitation of P1COOH in hexane with carotenoid, and (Car) is the carotenoid concentration. The concentration of P1COOH was kept constant at 2.25 × 10−6 mol/L, whereas the concentration of carotenoids varied between 0 and 1.5 × 10−3 mol/L. The ratio of (Car)/(P1COOH) was increasing from 0 to 670.

3. Results and Discussion

3.1. Particle Size Distribution of Persimmon Freeze-Dried Powder

Figure 1 shows the particle size distribution of persimmon freeze dried powder used for carotenoid extraction. All the particles present a diameter below 500 µm, which is currently recognized as appropriate for good ASE extractions [26]. The particle size of the freeze dried matrix is indeed a key factor in extraction of bioactive analytes from complex food matrices. The efficiency of the extraction depends on the analyte to be extracted, the nature of the sample matrix and the analyte location within the matrix [27]. Particle size, together with pressure and temperature in the ASE process, is a critical factor to enhance extraction efficiency. Reducing the particle size of the solid matrix increases the mass transfer surfaces with the extraction solvent. Reducing the particle size also facilitates solvent diffusion through the matrix pores and the partitioning of carotenoids between the matrix and the extraction fluid, thus leading to accelerated extraction.

3.2. Semi-Preparative HPLC Purification and Identification of Carotenoids in the ASE Extract

Analytical HPLC separation conditions of persimmon carotenoids were well described previously [24]. Figure 2 shows the ASE carotenoid extract chromatogram. Eleven carotenoids were detected due to their absorption at 450 nm and the retention times were given in Table 1. Their masses were calculated and are also found in Table 1. Five of the 11 purified carotenoids were identified using commercial standards and LC-MS (APCI positive mode), that is to say compounds 1, 2, 3, 6 and 9 as violaxanthin, lutein, zeaxanthin, β-cryptoxanthin and β-carotene, respectively [24].
As can be seen in Table 1, after fraction collections corresponding to each peak, eleven peaks were detected and identified as carotenoids based (i) on their retention time (as compared to commercial standards when available) and (ii) on their spectral characteristics after chromatographic solvent evaporation and solubilization in petroleum ether (λmax, fine spectral data as compared to other spectral data from the literature as described in Table 1 footnotes). In each case, m/z values were also determined from liquid chromatography coupled to mass spectrometry (LC-MS) analysis. The six first peaks identified as carotenoids logically correspond to polar carotenoids (xanthophylls). The higher the number of hydroxyl and/or epoxy groups, the shorter the retention time. While peak 7 is not clearly identified (mixture), logically, the last four peaks were identified as carotenes (no polar groups such as OH or epoxy groups). Cis-isomeric forms of carotenoids can easily be detected from the presence of a first λmax in the UV domain around 330 nm (therefore, cis-isomeric forms are characterized by 4 λmax instead of 3 λmax only for trans-isomeric forms). Each peak identification in Table 1 is explained in detail below.
  • All trans-violaxanthin (peak 1): [M + H]+ m/z, λmax, absence of cis-peak and a short retention time are in perfect agreement with the identification of all-trans-violaxanthin, which contains two hydroxyl and two epoxy groups. All-trans-violaxanthin had already been identified in persimmon [27,30]. However the % (III/II) observed here was 50 instead of 98, which may be due to the origin of the petroleum ether that we used. It should be noted that petroleum ether is not a well-defined molecule but a mixture of C5 and C6 hydrocarbons.
  • All trans-lutein (peak 2): [M + H]+ m/z, λmax, absence of cis-peak and a short retention time similar to that of the standard are in perfect agreement with the identification of all-trans-lutein, which contains two OH groups and should have a slightly longer retention time than violaxanthin. All-trans-lutein had already been identified in persimmon [30] However, the % (III/II) observed here was 25 instead of 60, which may be due to the origin of the petroleum ether that we used.
  • All trans-zeaxanthin (peak 3): the characteristics found fit the data perfectly from the literature [28]. The retention time is also in agreement with that of the standard.
  • Not identified (peak 4): The [M + H]+ m/z value of 601 indicates a xanthophyll with four oxygens. The fact that, despite these fours oxygens, the fact that this xanthophyll has a longer retention time than lutein, and even zeaxanthin, indicates that this is a cis-form of xanthophyll and indeed we observed a slight cis-peak at 330 nm and an associated % AB:AII of 14%.
  • Suspected 5,6-epoxy-α-carotene (peak 5): [M + H]+ m/z, λmax, the absence of a cis-peak, as well as the % III:II, all suggest the presence of 5,6-epoxy-α-carotene.
  • All-trans-β-cryptoxanthin (peak 6): the characteristics found fit the data from the literature [29]. The retention time is also in agreement with that of the standard. This xanthophyll is known to be the most prevalent carotenoid in persimmon.
  • Unidentified (peak 7): [M + H]+ m/z values of 553 and 537 indicate a mixture of a single-oxygenated xanthophyll and of one carotene, with retention times that do not allow separation by the preparative chromatography fraction collector. A very slight increase in absorbance is observed at 330 nm, which could indicate a cis-form. It is not possible to be more precise.
  • Cis-isomer of β-carotene, supposed position 13 (peak 8): the mass indicates that this is a carotene and the presence of a slight peak at 330 nm indicates a cis-form of carotene. λmax as well as the % (III:II) indicates that it is most likely a cis-isomer of β-carotene. However, the % AB:AII is lower than that found by De Rosso [29], but it should be noted that they used a different solvent. Therefore, this position 13 of cis-isomerization can only be presumed.
  • All-trans-β-carotene (peak 9): the characteristics found fit the data from the literature perfectly [28]. The retention time is also in agreement with that of the standard.
  • Cis-isomer of β-carotene, supposed position 9 (peak 10): the mass indicates that this is a carotene and the presence of a slight peak at 330 nm indicates a cis-form of carotene. λmax as well as the % (III:II) indicates that it is very likely a cis-isomer of β-carotene. However, the % AB:AII is higher than that found by De Rosso [29], but it should be noted that they used a different solvent. This % should also be lower than that of 13-cis-β-carotene, which is not the case. Thus, position 9 of the cis-isomerization can only be presumed.
  • All-trans-lycopene (peak 11): all-trans lycopene was precisely identified, especially due to its characteristic set of λmax values [28].

3.3. Details on UV-Visible and Fluorescence Spectra of Purified Carotenoids

Figure 3 and Figure 4, respectively, show the Ultraviolet-Visible (UV-Vis) absorption spectra and fluorescence spectra of carotenoids extracted and purified from persimmon fruits in petroleum ether. As explained after Table 1, fine UV-Vis spectral characteristics were derived from Figure 3. Regarding fluorescence spectra, different excitation wavelengths were chosen from 439 nm to 451 nm in order to excite in a maximum of absorption. It is often considered that the carotenoids are non-fluorescent compounds. In our conditions, we could nevertheless detect a very weak fluorescence with a band centered at around 500 nm.

3.4. Singlet Oxygen Quenching by Carotenoids

To evaluate the potential of 1O2 quenching, we used a porphyrin (P1COOH) to produce 1O2 and we added different amounts of carotenoids. We recorded the 1O2 decay of P1COOH without all-trans-β-carotene and with increasing amounts of all-trans-β-carotene (Figure 5).
Due to this decay, the Stern–Volmer plot can be built and the quenching constant evaluated. Data are already available in the literature and are reported in Table 2, but no evaluation of the quenching constants in hexane has been assessed to date. Our results are reported in Table 3.
Some photosensitizers have been reported in the literature to evaluate the fluorescence potential or 1O2 quenching of various carotenoids (Table 2). Dreuw et al. [31] studied chlorophyll fluorescence quenching by xanthophylls using time-dependent density-functional theory and configuration interaction singles. Naqvi et al. also used chlorophyll a (1O2 quantum yield is 0.77 in methanol) [32]. Cantrell et al. excited Bengal Rose ((1O2 quantum yield is 0.68 in ethanol) and (4-(1-pyrene)butyric acid) [33], like Ramel [34], to produce 1O2. Fiedor et al. used bacteriopheophytin-a (1O2 quantum yield is 0.2–0.3 in deuterated methanol) [35] and Scholz et al. 5,10,15,20-tetraphenyl-porphin (1O2 quantum yield is 0.60–0.70 in benzene) to evaluate β-carotene quenching [36]. Chanterell et al. also used another porphyrin, the protoporphyrin IX dimethyl ester ((1O2 quantum yield is 0.67 in dimethylsulfoxide (DMF)) [37]. In these published data, we observed a myriad of recorded Kq values for investigated carotenoids. The differences between the values might be attributed to the different experimental conditions used for 1O2 production and detection.
In our study, we decided to also choose a photosensitizer able to produce large amounts of 1O2, the porphyrin (P1COOH). Indeed, this molecule was synthesized by our group [25] and we found that the 1O2 quantum yield is 0.70 in ethanol and 0.87 in toluene. These quantum yields are similar to quantum yields of the photosensitizers used in the literature.
To detect 1O2 quenching, we recorded the decay of 1O2 in the presence of increasing amounts of carotenoids. In our study, all of the compounds were in hexane. We found a value similar to that of the literature (1.1 × 109 M−1·s−1 in our case, 14 × 109 M−1·s−1 in [38]). For lycopene, we found 1.1 × 109 M−1·s−1 versus 19 × 109 M−1·s−1 in [38]. The other carotenoids have not been studied in hexane so no comparison can be performed. The difference observed can be due to the different techniques used to evaluate 1O2 quenching. In [42], singlet oxygen was also detected via its luminescence at 1270 nm, but with a germanium photodiode linked to a Judson amplifier (Teledyne, Thousand Oaks, CA, USA) with a variable load resistor. In our case, we evaluated the lifetime of 1O2 with a single photon counting FluoroHub-B controller.
According to our results, the rate constants of quenching of 1O2 by β-cryptoxanthin, β-carotene, and lycopene recorded in hexane showed that these pigments efficiently quench 1O2 production, with associated (Kq) values of 1.6 × 109, 1.3 × 109 and 1.1 × 109 M−1·s−1. β-cryptoxanthin showed the higher quenching efficiency (1.6 × 109 M−1·s−1). In organic solvents, carotenoids with 10 or 11 conjugated double bond quench singlet oxygens near the diffusion limit. β-cryptoxanthin, β-carotene, lycopene contained between 11 C=C and 13 C=C conjugated double bonds (Table 3), which explain the high Kq values obtained. The little differences in the rate constants between the three carotenoids in hexane can be explained by the similarity of the number of C=C conjugated double bonds for these carotenoids. The hydroxylated forms of carotenoids lutein, zeaxanthin, violaxanthin, 5,6-epoxy-α-carotene and the not identified xanthophyll (peak 4) revealed smaller quenching rate constants. This could be explained by a possible chemical reaction between 1O2 and the hydroxyl group of xanthophylls. Moreover, xanthophylls are well dissolved in polar solvents but not in non-polar solvents such as hexane. This might lead to aggregation and decrease of the quenching rate constants.

4. Conclusions

Persimmon fruit cultivated in Tunisia is known for its richness in carotenoid pigments. In this study, it was possible to show that at least 11 carotenoids are present in the fruit, suggesting that persimmon fruit can be considered as a potential source of carotenoids. The photophysical properties (absorption, fluorescence, quenching of singlet oxygen) of carotenoids extracted from persimmon fruits and purified were investigated in hexane, using porphyrin as the photosensitizer for 1O2 production. Of the eight carotenoids investigated, β-cryptoxanthin, β-carotene and lycopene showed high quenching efficiencies compared to the xanthophylls (lutein, violaxanthin, zeaxanthin, 5,6-epoxy-α-carotene and the unidentified xanthophyll). The kinetic analysis of results obtained from the Stern–Volmer plot showed that β-cryptoxanthin quenches singlet oxygen as efficiently as β-carotene and lycopene, with Kq values of 1.6 × 109, 1.3 × 109 and 1.1 × 109 M−1·s−1. Xanthophylls in hexane gave lower values than 109 M−1·s−1. The bimolecular rate constants of 1O2 quenching (Kq) depend on the carotenoid structures such as the number of conjugated double bonds and the presence or absence of hydroxyl and epoxy groups. The efficiency of singlet oxygen quenching of β-cryptoxanthin can thus change the consideration that β-carotene and lycopene are the most singlet oxygen quenchers acting as catalysts for deactivation of the harmful 1O2. In the near future, photodynamic molecular beacons (PMB) using β-cryptoxanthin as 1O2 quencher will be synthesized and tested, hoping to show that natural molecules extracted from food matrices may be of interest for photodynamic therapy.

Author Contributions

Y.G. designed the extraction experiments; K.Z., O.N. and B.M. performed the experiments; P.A. and C.F. analyzed the photophysical data; and V.R. and C.F. wrote the paper and supervised the research and K.Z.’s thesis.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

APCIatmospheric pressure chemical ionization
ASEaccelerated solvent extraction
BHTbutylhydroxytoluene
Carcarotenoid
DHIR3,3-dihydroxyisorenieratene
DMA9,10-dimethylanthrancene
DMNO21,4-dimethyl-1,4-naphthalene endoperoxide
DPBF1,3-Diphenylisobenzofuran
EP-13-(1,4-epidioxy-4-methyl-1,4-dihydro-1-naphtyl propionic acid)
EtOHethanol
Hexhexane
1-HP1H-Phenalen-1-one
LC-MSliquid chromatography- mass spectrometry
LDLlow-density lipoprotein
MBmethylene blue
MTBEmethyl-tert-butyl-ether
NDPO23,3′-(1,4-naphthalylene dipropionate)
1-NN(1-nitronaphtalene)
PBA4-(1-pyrene)butyric acid
P1COOH5-(4-carboxyphenyl)-10,15,20-triphenylporphyrin
PDTphotodynamic therapy
PMBphotomolecular beacons
RBRose Bengal
RMsystem of sodium (bis-2-ethylhexanyl)sulfosuccinate in (hexane/H2O)
TEAtriethylamine
TPPtetraphenylporphyrin

References

  1. Rodriguez-Amaya, D.B.; Rodriguez, E.B.; Amaya-Farfan, J. Advances in food carotenoids research: Chemical and technological aspects, implications in human health. Malays. J. Nutr. 2006, 12, 101–121. [Google Scholar]
  2. Rodriguez-Concepcion, M. Supply of precursors for carotenoid biosynthesis in plants. Arch. Biochem. Biophys. 2010, 504, 118–122. [Google Scholar] [CrossRef] [PubMed]
  3. Britton, G.; Liaaen-Jensen, S.; Pfander, H. Carotenoids. Nutrition and Health; Birkhäuser Verlag: Basel, Switzerland, 2009; Volume 5, p. 431. [Google Scholar]
  4. LaFountain, A.M.; Pacheco, C.; Prum, R.O.; Frank, H.A. Nuclear magnetic resonance analysis of carotenoids from the burgundy plumage of the Pompadour Cotinga (Xipholena punicea). Arch. Biochem. Biophys. 2013, 539, 133–141. [Google Scholar] [CrossRef] [PubMed]
  5. Stahl, W.; Sies, H. Antioxidant activity of carotenoids. Mol. Asp. Med. 2003, 24, 345–351. [Google Scholar] [CrossRef]
  6. Ibañez, E.; Herrero, M.; Mendiola, J.; Castro-Puyana, M. Extraction and Characterization of Bioactive Compounds with Health Benefits from Marine Resources: Macro and Micro Algae, Cyanobacteria, and Invertebrates. In Marine Bioactive Compounds; Springer: New York, NY, USA, 2011; pp. 55–98. [Google Scholar]
  7. Rivera, S.; Canela, R. Influence of sample processing on the analysis of carotenoids in maize. Molecules 2012, 17, 11255–11268. [Google Scholar] [CrossRef] [PubMed]
  8. Fiedor, J.; Burda, K. Potential role of carotenoids as antioxidants in human health and disease. Nutrients 2014, 6, 466–488. [Google Scholar] [CrossRef] [PubMed]
  9. Namitha, K.K.; Negi, P.S. Chemistry and biotechnology of carotenoids. Crit. Rev. Food Sci. Nutr. 2010, 50, 728–760. [Google Scholar] [CrossRef] [PubMed]
  10. Perera, C.O.; Yen, G.M. Functional properties of carotenoids in human health. Int. J. Food Prop. 2007, 10, 201–230. [Google Scholar] [CrossRef]
  11. Koh, W.P.; Yuan, J.M.; Wang, R.; Lee, Y.P.; Lee, B.L.; Yu, M.C.; Ong, C.N. Plasma carotenoids and risk of acute myocardial infarction in the singapore chinese health study. Nutr. Metab. Cardiovasc. Dis. 2011, 21, 685–690. [Google Scholar] [CrossRef] [PubMed]
  12. Fuhrman, B.; Elis, A.; Aviram, M. Hypocholesterolemic effect of lycopene and beta-carotene is related to suppression of cholesterol synthesis and augmentation of LDL receptor activity in macrophages. Biochem. Biophys. Res. Commun. 1997, 233, 658–662. [Google Scholar] [CrossRef] [PubMed]
  13. Tanaka, T.; Shnimizu, M.; Moriwaki, H. Cancer chemoprevention by carotenoids. Molecules 2012, 17, 3202–3242. [Google Scholar] [CrossRef] [PubMed]
  14. Wertz, K.; Siler, U.; Goralczyk, R. Lycopene: Modes of action to promote prostate health. Arch. Biochem. Biophys. 2004, 430, 127–134. [Google Scholar] [CrossRef] [PubMed]
  15. Grubbs, C.J.; Eto, I.; Juliana, M.M.; Whitaker, L.M. Effect of canthaxanthin on chemically-induced mammary carcinogenesis. Oncology 1991, 48, 239–245. [Google Scholar] [CrossRef] [PubMed]
  16. Onogi, N.; Okuno, M.; Matsushima-Nishiwaki, R.; Fukutomi, Y.; Moriwaki, H.; Muto, Y.; Kojima, S. Antiproliferative effect of carotenoids on human colon cancer cells without conversion to retinoic acid. Nutr. Cancer 1998, 32, 20–24. [Google Scholar] [CrossRef] [PubMed]
  17. Kuno, T.; Hirose, Y.; Yamada, Y.; Hata, K.; Qiang, S.H.; Asano, N.; Oyama, T.; Zhi, H.L.; Iwasaki, T.; Kobayashi, H.; et al. Chemoprevention of mouse urinary bladder carcinogenesis by fermented brown rice and rice bran. Oncol. Rep. 2006, 15, 533–538. [Google Scholar] [CrossRef] [PubMed]
  18. Tanaka, T.; Kawamori, T.; Ohnishi, M.; Makita, H.; Mori, H.; Satoh, K.; Hara, A. Suppression of azoxymethane-induced rat colon carcinogenesis by dietary administration of naturally occurring xanthophylls astaxanthin and canthaxanthin during the postinitiation phase. Carcinogenesis 1995, 16, 2957–2963. [Google Scholar] [CrossRef] [PubMed]
  19. Britton, G.; Liaaen-Jensen, S.; Pfander, H. Carotenoids Natural Function; Birkhäuser Verlag: Basel, Switzerland, 2008; Volume 4, p. 370. [Google Scholar]
  20. Foote, C.S.; Denny, R.W.; Weaver, L.; Chang, Y.; Peters, J. Quenching of singlet oxygen. Ann. N. Y. Acad. Sci. 1970, 171, 139–148. [Google Scholar] [CrossRef]
  21. Zheng, G.; Chen, J.; Stefflova, K.; Jarvi, M.; Li, H.; Wilson, B.C. Photodynamic molecular beacon as an activatable photosensitizer based on protease-controlled singlet oxygen quenching and activation. Proc. Natl. Acad. Sci. USA 2007, 104, 8989–8994. [Google Scholar] [CrossRef] [PubMed]
  22. Chen, J.; Stefflova, K.; Niedre, M.J.; Wilson, B.C.; Chance, B.; Glickson, J.D.; Zheng, G. Protease-triggered photosensitizing beacon based on singlet oxygen quenching and activation. J. Am. Chem. Soc. 2004, 126, 11450–11451. [Google Scholar] [CrossRef] [PubMed]
  23. Verhille, M.; Benachour, H.; Ibrahim, A.; Achard, M.; Arnoux, P.; Barberi-Heyob, M.; Andre, J.C.; Allonas, X.; Baros, F.; Vanderesse, R.; et al. Photodynamic molecular beacons triggered by MMP-2 and MMP-9: Influence of the distance between photosensitizer and quencher onto photophysical properties and enzymatic activation. Curr. Med. Chem. 2012, 19, 5580–5594. [Google Scholar] [CrossRef] [PubMed]
  24. Zaghdoudi, K.; Pontvianne, S.; Framboisier, X.; Achard, M.; Kudaibergenova, R.; Ayadi-Trabelsi, M.; Kalthoum-Cherif, J.; Vanderesse, R.; Frochot, C.; Guiavarc’h, Y. Accelerated solvent extraction of carotenoids from: Tunisian kaki (Diospyros kaki L.), peach (Prunus persica L.) and apricot (Prunus armeniaca L.). Food Chem. 2015, 184, 131–139. [Google Scholar] [CrossRef] [PubMed]
  25. Chouikrat, R.; Champion, A.; Vanderesse, R.; Frochot, C.; Moussaron, A. Microwave-assisted synthesis of zinc 5-(4-carboxyphenyl)-10,15,20-triphenylporphyrin and zinc 5-(4-carboxyphenyl)-10,15,20-triphenylchlorin. J. Porphyr. Phthalocyanines 2015, 19, 595–600. [Google Scholar] [CrossRef]
  26. Thermo Scientific, Technical note 208. 2013. Available online: http://www.dionex.com/en-us/webdocs/4736-TN208_FINAL.pdf (accessed on 9 July 2016).
  27. Mustafa, A.; Turner, C. Pressurized liquid extraction as a green approach in food and herbal plants extraction: A review. Anal. Chim. Acta 2011, 703, 8–18. [Google Scholar] [CrossRef] [PubMed]
  28. Rodriguez-Amaya, D.B. International Life Sciences. In A Guide to Carotenoid Analysis in Foods; Omni, I., Ed.; ILSI Press: Washington, DC, USA, 2001. [Google Scholar]
  29. De Rosso, V.V.; Mercadante, A.Z. Identification and quantification of carotenoids, by HPLC-PDA-MS/MS, from amazonian fruits. J. Agric. Food Chem. 2007, 55, 5062–5072. [Google Scholar] [CrossRef] [PubMed]
  30. Zhou, C.H.; Zhao, D.Q.; Sheng, Y.L.; Tao, J.; Yang, Y. Carotenoids in fruits of different persimmon cultivars. Molecules 2011, 16, 624–636. [Google Scholar] [CrossRef] [PubMed]
  31. Dreuw, A.; Fleming, G.R.; Head-Gordon, M. Chlorophyll fluorescence quenching by xanthophylls. PCCP 2003, 5, 3247–3256. [Google Scholar] [CrossRef]
  32. Naqvi, K.R.; Melo, T.B.; Raju, B.B.; Javorfi, T.; Simidjiev, I.; Garab, G. Quenching of chlorophyll a singlets and triplets by carotenoids in light-harvesting complex of photosystem II: Comparison of aggregates with trimers. Spectrochim. Acta A Mol. Biomol. Spectrosc. 1997, 53, 2659–2667. [Google Scholar] [CrossRef]
  33. Cantrell, A.; McGarvey, D.J.; Truscott, T.G.; Rancan, F.; Böhm, F. Singlet oxygen quenching by dietary carotenoids in a model membrane environment. Arch. Biochem. Biophys. 2003, 412, 47–54. [Google Scholar] [CrossRef]
  34. Ramel, F.; Birtic, S.; Cuine, S.; Triantaphylides, C.; Ravanat, J.-L.; Havaux, M. Chemical quenching of singlet oxygen by carotenoids in plants. Plant Physiol. 2012, 158, 1267–1278. [Google Scholar] [CrossRef] [PubMed]
  35. Fiedor, J.; Fiedor, L.; Haessner, R.; Scheer, H. Cyclic endoperoxides of beta-carotene, potential pro-oxidants, as products of chemical quenching of singlet oxygen. Biochim. Biophys. Acta 2005, 1709, 1–4. [Google Scholar] [CrossRef] [PubMed]
  36. Scholz, M.; Dedic, R.; Svoboda, A.; Hala, J. TPP and singlet oxygen quenching by carotene in solution. J. Mol. Struct. 2011, 993, 474–476. [Google Scholar] [CrossRef]
  37. Chantrell, S.J.; McAuliffe, C.A.; Munn, R.W.; Pratt, A.C.; Land, E.J. Excited-states of protoporphyrin-IX dimethyl ester—Reaction of triplet with carotenoids. J. Chem. Soc. Faraday Trans. 1 1977, 73, 858–865. [Google Scholar] [CrossRef]
  38. Conn, P.F.; Schalch, W.; Truscott, T.G. The singlet oxygen and carotenoid interaction. J. Photochem. Photobiol. B Biol. 1991, 11, 41–47. [Google Scholar] [CrossRef]
  39. Devasagayam, T.P.A.; Werner, T.; Ippendorf, H.; Martin, H.D.; Sies, H. Synthetic carotenoids, novel polyene polyketones and new capsorubin isomers as efficient quenchers of singlet molecular oxygen. Photochem. Photobiol. 1992, 55, 511–514. [Google Scholar] [CrossRef] [PubMed]
  40. Ouchi, A.; Aizawa, K.; Iwasaki, Y.; Inakuma, T.; Terao, J.; Nagaoka, S.-I.; Mukai, K. Kinetic study of the quenching reaction of singlet oxygen by carotenoids and food extracts in solution. Development of a singlet oxygen absorption capacity (SOAC) assay method. J. Agric. Food Chem. 2010, 58, 9967–9978. [Google Scholar] [CrossRef] [PubMed]
  41. Di Mascio, P.; Kaiser, S.; Sies, H. Lycopene as the most efficient biological carotenoid singlet oxygen quencher. Arch. Biochem. Biophys. 1989, 274, 532–538. [Google Scholar] [CrossRef]
  42. Di Mascio, P.; Sundquist, A.R.; Devasagayam, T.P.A.; Sies, H. Assay of Lycopene and Other Carotenoids as Singlet Oxygen Quenchers. In Methods in Enzymology; Carotenoids, Part A: Chemistry, Separation, Quantitation, and Antioxidation; Academic Press: New York, NY, USA, 1992; Volume 213, pp. 429–438. [Google Scholar]
  43. Beutner, S.; Bloedorn, B.; Frixel, S.; Blanco, I.H.; Hoffmann, T.; Martin, H.D.; Mayer, B.; Noack, P.; Ruck, C.; Schmidt, M.; et al. Quantitative assessment of antioxidant properties of natural colorants and phytochemicals: Carotenoids, flavonoids, phenols and indigoids. The role of ss-carotene in antioxidant functions. J. Sci. Food Agric. 2001, 81, 559–568. [Google Scholar] [CrossRef]
  44. Böhm, F.; Edge, R.; Truscott, T.G. Interactions of dietary carotenoids with singlet oxygen (1O2) and free radicals: Potential effects for human health. Acta Biochim. Pol. 2011, 59, 27–30. [Google Scholar]
  45. Montenegro, M.A.; Nazareno, M.A.; Durantini, E.N.; Borsarelli, C.D. Singlet molecular oxygen quenching ability of carotenoids in a reverse-micelle membrane mimetic system. Photochem. Photobiol. 2002, 75, 353–361. [Google Scholar] [CrossRef]
  46. Stahl, W.; Sies, H. Physical quenching of singlet oxygen and cis-trans isomerization of carotenoids. Ann. N. Y. Acad. Sci. 1993, 691, 10–19. [Google Scholar] [CrossRef]
  47. Lee, S.H.; Min, D.B. Effects, quenching mechanisms, and kinetics of carotenoids in chlorophyll-sensitized photooxidation of soybean oil. J. Agric. Food Chem. 1990, 38, 1630–1634. [Google Scholar] [CrossRef]
  48. Rodgers, M.A.J.; Bates, A.L. Kinetic and spectroscopic features of some carotenoid triplet-states—Sensitization by singlet oxygen. Photochem. Photobiol. 1980, 31, 533–537. [Google Scholar] [CrossRef]
  49. Wilkinson, F.; Ho, W.T. Electronic-energy transfer from singlet molecular oxygen to carotenoids. Spectrosc. Lett. 1978, 11, 455–463. [Google Scholar] [CrossRef]
  50. Farmilo, A.; Wilkinson, F. On the mechanism of quenching of singlet oxygen in solution. Photochem. Photobiol. 1973, 18, 447–450. [Google Scholar] [CrossRef] [PubMed]
  51. Foote, C.S.; Chang, Y.C.; Denny, R.W. Chemistry of singlet oxygen 10. Carotenoid quenching parallels biological protection. J. Am. Chem. Soc. 1970, 92, 5216–5218. [Google Scholar] [CrossRef] [PubMed]
  52. Foote, C.S.; Denny, R.W. Chemistry of singlet oxygen .7. Quenching by beta-carotene. J. Am. Chem. Soc. 1968, 90, 6233–6235. [Google Scholar] [CrossRef]
  53. Foote, C.S. Quenching of Singlet Oxygen; Academic Press: New York, NY, USA, 1979; Volume 5, pp. 139–171. [Google Scholar]
  54. Murasecco-Suardi, P.; Oliveros, E.; Braun, A.M.; Hansen, H.J. Singlet-oxygen quenching by carotenoids—Steady-state luminescence experiments. Helv. Chim. Acta 1988, 71, 1005–1010. [Google Scholar] [CrossRef]
  55. Krasnovsky, A.A. Photo-luminescence of singlet oxygen in pigment solutions. Photochem. Photobiol. 1979, 29, 29–36. [Google Scholar] [CrossRef]
  56. Mathews-Roth, M.M.; Wilson, T.; Fujimori, E.; Krinsky, N.I. Carotenoid chromophore length and protection against photosensitization. Photochem. Photobiol. 1974, 19, 217–222. [Google Scholar] [CrossRef] [PubMed]
  57. Li, T.L.; King, J.M.; Min, Y.B. Quenching mechanisms and kinetics of carotenoids in riboflavin photosensitized singlet oxygen oxidation of vitamin D-2. J. Food Biochem. 2000, 24, 477–492. [Google Scholar] [CrossRef]
  58. Oliveros, E.; Braun, A.M.; Aminiansaghafi, T.; Sliwka, H.R. Quenching of singlet oxygen (1Dg) by carotenoid derivatives—Kinetic analysis by near-infrared luminescence. New J. Chem. 1994, 18, 535–539. [Google Scholar]
Figure 1. Particle size distribution of persimmon freeze dried powder.
Figure 1. Particle size distribution of persimmon freeze dried powder.
Foods 06 00004 g001
Figure 2. C30 semi preparative chromatogram of carotenoids extracted from persimmon.
Figure 2. C30 semi preparative chromatogram of carotenoids extracted from persimmon.
Foods 06 00004 g002
Figure 3. Absorption spectra of all the purified carotenoids in petroleum ether.
Figure 3. Absorption spectra of all the purified carotenoids in petroleum ether.
Foods 06 00004 g003
Figure 4. Fluorescence spectra of all compounds in petroleum ether (λex varying from 439 to 451 nm).
Figure 4. Fluorescence spectra of all compounds in petroleum ether (λex varying from 439 to 451 nm).
Foods 06 00004 g004
Figure 5. 1O2 decay of P1COOH without all-trans-β-carotene and with increasing amounts of all-trans-β-carotene (Molar concentration), λexc = 408 nm. c = 2.25 × 10−6 mol/L in hexane. The ratio of (Car)/(P1COOH) was increasing from 0 to 670 (blue: 0 mol/L, red: 4 × 10-7 mol/L, green: 4 × 10-5 mol/L, pink: 2 × 10-4 mol/L, yellow: 4 × 10-4 mol/L).
Figure 5. 1O2 decay of P1COOH without all-trans-β-carotene and with increasing amounts of all-trans-β-carotene (Molar concentration), λexc = 408 nm. c = 2.25 × 10−6 mol/L in hexane. The ratio of (Car)/(P1COOH) was increasing from 0 to 670 (blue: 0 mol/L, red: 4 × 10-7 mol/L, green: 4 × 10-5 mol/L, pink: 2 × 10-4 mol/L, yellow: 4 × 10-4 mol/L).
Foods 06 00004 g005
Table 1. Characteristics of carotenoids from persimmon obtained from preparative chromatography. λmax and fine spectral data were observed on the UV-Vis spectrum of each collected carotenoid, in a powder form, and in petroleum ether.
Table 1. Characteristics of carotenoids from persimmon obtained from preparative chromatography. λmax and fine spectral data were observed on the UV-Vis spectrum of each collected carotenoid, in a powder form, and in petroleum ether.
PeakCarotenoidOH/EpoxytR (min)λmax (nm) a% III/II a,b% AB/AII a,c[M + H]+ (m/z)
1all-trans-violaxanthin2/218416/440/465 (416/440/465)50 (98)0601
2all-trans-lutein2/023420/443/469 (421/445/474)25 (60)0569
3all-trans-zeaxanthin2/028424/448/475 (424/449/476)25 (25)0569
4not identified (cis-xanthophyll with 4 oxygens)432330/422/444/4694314601
55,6-epoxy-α-carotene suspected0/143421/445/470 (418/441/469) *0 (10) *0553
6all-trans-β-cryptoxanthin1/052426/449/476 (425/449/476)36 (25)0553
7not identified (mixture of monoxygenated xanthophyll and carotene)?62330/423/444/4701018mixture 537/553
8cis-isomere of β-carotene, position 13 (supposed)0/067332/421/443/470 (338/420/444/470) *16 (12) *20 (47) *537
9all-trans-β-carotene0/074426/449/476 (425/450/477)23 (25)0537
10cis-isomere of β-carotene, position 9 (supposed)0/075331/424/445/469 (330/420/444/472) *11 (20) *26 (18) *537
11all-trans-lycopene0/086444/470/502 (444/470/501)66 (65)0537
a values between brackets without * correspond to spectral characteristics also found in petroleum ether and compiled by Rodriguez-Amaya [28]. Values between brackets with * correspond to spectral characteristics along a methanol/ methyl-tert-butyl-ether (MTBE) gradient observed by De Rosso and Mercadante [29], who also used a C30 column; b ratio of the height of the longest-wavelength absorption peak, designated III, and that of the middle absorption peak, designated II, taking the minimum between the two peaks as baseline, multiplied by 100; c the relative intensity of the cis-peak is expressed as % AB/AII (Figure 3), which is the ratio of the height of the cis-peak, designated AB, and that of the middle main absorption peak, designated AII, multiplied by 100.
Table 2. Kq of carotenoids in different solvents, methods of 1O2 production and 1O2 quenching.
Table 2. Kq of carotenoids in different solvents, methods of 1O2 production and 1O2 quenching.
CarotenoidskQ 109 M−1·s−1 (Solvent)1O2 ProductionMethod of Detection of 1O2 QuenchingReferences
lycopene17 (C6H6)Phenazine1O2 emission[38]
18 (C6H5CH3)Phenazine1O2 emission[38]
19(C6H14)Phenazine1O2 emission[38]
9.0 (CHCl3)NDPO21O2 emission[39]
19 (CHCl3)Phenazine1O2 emission[38]
14 (CCl4)Phenazine1O2 emission[38]
13.8 (EtOH/CHCl3/D2O 50/50/1)EP-1DPBF[40]
31(EtOH/CHCl3/D2O 50/50/1)NDPO21O2 emission[41,42]
8.8 (EtOH/CHCl3/D2O 50/50/1)DMNO21O2 emission[43]
17.5(EtOH/CHCl3/H2O 50/50/1)Phenazine1O2 emission[38]
31 (EtOH/CHCl3/H2O 50/50/1)NDPO21O2 emission[41]
0.13 (ascorbic acid in methanol)1-NN1O2 emission[44]
23–25 (reverse micelle (RM))RBDMA[45]
8.8 (CHCl3)DMNO21O2 emission[43]
9 (EtOH/CHCl3/H2O 50/50/1)NDPO21O2 emission[46]
6.93 (soybean oil)ChlorophyllHeadspace oxygen depletion by gas chromatography[47]
β-carotene13.0 (C6H6)Phenazine1O2 emission[38]
13.8 (C6H6)Anthracene/NaphthaleneRadiolysis/1O2 emission[48]
12.5–14 (C6H6)Anthracene/NaphthaleneRadiolysis/1O2 emission[49]
13 (C6H6)AnthraceneDPBF[50]
14 (toluene)Phenazine1O2 emission[38]
14 (C6H14)Phenazine1O2 emission[38]
5.0 (CHCl3)NDPO21O2 emission[39]
14.0 (EtOH/CHCl3/H2O 50/50/1)NDPO21O2 emission[41]
12 (EtOH/CHCl3/H2O: 50/50/1)Phenazine1O2 emission[38]
10.8 (EtOH/CHCl3/D2O 50/50/1)EP-1DPBF[40]
4.2 (EtOH/CHCl3/D2O 50/50/1)NDPO21O2 emission[41,42]
8.4 (EtOH/CHCl3/D2O 50/50/1)DMNO21O2 emission[43]
30 (MeOH/C6H6 1/4)MB2-methyl-2-penten[51]
5 (MeOH/C6H6 1/4)Anthracene/NaphthaleneRadiolysis/1O2 emission[52]
13 (MeOH/C6H6 1/4)Anthracene/NaphthaleneRadiolysis/1O2 emission[53]
10.9 ± 0.5 (THF)TPP1O2 emission[36]
2.3 (DPPC)RB and PBA1O2 emission[33]
9.9 (CCl4)Phenazine1O2 emission[38]
5.9 (CCl4)1H-P + RB1O2 emission[54]
0.7 (CCl4)Porphyrin1O2 emission[55]
11 (CHCl3)Phenazine1O2 emission[38]
8.1 (CHCl3)DMNO21O2 emission[43]
23 (C6H6/MeOH: 3/2)MB/RB1O2 emission[56]
1.5 (CD3OD)1H-P/RB1O2 emission[54]
0.35 (Ascorbic acid in MeOH)1-NN1O2 emission[44]
5 (H2O/ (CH3)2CO 12/88)RiboflavinGC with thermal conductivity[57]
12.67 (reverse micelle (RM)) RBDMA[45]
5 (EtOH/CHCl3/D2O 50/50/1)NDPO21O2 emission[46]
Lutein11.0 (C6H6)TPP1O2 emission[58]
16 (C6H6)Phenazine1O2 emission[38]
8.0 (EtOH/CH2Cl2/H2O; 50:50:1)NDPO21O2 emission[41]
0.11 (DPPC)RB and PBA1O2 emission[33]
9.24 (EtOH/CHCl3/D2O 50/50/1)EP-1DPBF[40]
2.4 (EtOH/CHCl3/D2O 50/50/1)NDPO21O2 emission[42]
21 (MeOH/C6H6 1/4)Anthracene/NaphthaleneRadiolysis/1O2 emission[53]
1.3 (ascorbic acid in methanol)1-NN1O2 emission[44]
10-33 reverse micelle (RM)RBDMA[45]
5.72 (soybean oil)ChlorophyllHeadspace oxygen depletion by gas chromatography[47]
β-cryptoxanthin7.31 (EtOH/CHCl3/D2O 50/50/1)EP-1DPBF[40]
1.8–6 (EtOH/CHCl3/D2O 50/50/1)NDPO21O2emission[41,42]
Zeaxanthin12.6 (C6H6)Phenazine1O2emission[38]
12 (C6H6)Phenazine1O2emission[38]
2.8 (C6H6)Anthracene/NaphthaleneRadiolysis/1O2 emission[48]
10 (EtOH/CHCl3/D2O 50/50/1)NDPO21O2emission[41]
0.23 (DPPC)RB and PBA1O2emission[33]
10.5 (EtOH/CHCl3/D2O 50/50/1)EP-1DPBF[40]
3.0 (EtOH/CHCl3/D2O 50/50/1)NDPO21O2emission[41]
10 (EtOH/CHCl3/H2O 50/50/1)NDPO21O2emission[41]
0.77 (ascorbic acid in methanol)1-NNRadiolysis/1O2 emission[44]
6.79 (soybean oil)ChlorophyllHeadspace oxygen depletion by gas chromatography[47]
Violaxanthin16 (C6H6)Phenazine1O2 emission[38]
9 reverse micelle (RM)RBDMA[45]
Car: carotenoid; DHIR: -3,3-dihydroxyisorenieratene; DMA: 9,10-dimethylanthrancene; DMNO2: 1,4-dimethyl-1,4-naphthalene endoperoxide; DPBF: 1,3-Diphenylisobenzofuran; EP-1: 3-(1,4-epidioxy-4-methyl-1,4-dihydro-1-naphtyl propionic acid); EtOH: ethanol; Hex: hexane; 1-HP: 1H-Phenalen-1-one; MB: methylene blue; NDPO2: 3,3′-(1,4-naphthalylene dipropionate); 1-NN: (1-nitronaphtalene); PBA: 4-(1-pyrene)butyric acid; RB: Rose Bengal; RM: system of sodium (bis-2-ethylhexanyl)sulfosuccinate in (hexane/H2O); TPP: tetraphenylporphyrin.
Table 3. Bimolecular rate constants Kq of the quenching of 1O2 by carotenoids from persimmon fruits in hexane.
Table 3. Bimolecular rate constants Kq of the quenching of 1O2 by carotenoids from persimmon fruits in hexane.
CarotenoidsNb of C=C and OH GroupKq (M−1·s−1)
β-carotene in hexane11 C=C1.1 × 109
β-cryptoxanthin11 C=C and 1 OH1.6 × 109
lycopene13 C=C1.1 × 109
Lutein11 C=C and 2 OH8.0 × 108
Zeaxanthin11 C=C and 2 OH6.0 × 107
(Peak 4)-7.2 × 107
5,6-epoxy-α-carotene10 C=C and 1 OH, 2 epoxy3.8 × 107
Violaxanthin9 C=C, 2 OH and 2 epoxy group5.8 × 107

Share and Cite

MDPI and ACS Style

Zaghdoudi, K.; Ngomo, O.; Vanderesse, R.; Arnoux, P.; Myrzakhmetov, B.; Frochot, C.; Guiavarc’h, Y. Extraction, Identification and Photo-Physical Characterization of Persimmon (Diospyros kaki L.) Carotenoids. Foods 2017, 6, 4. https://doi.org/10.3390/foods6010004

AMA Style

Zaghdoudi K, Ngomo O, Vanderesse R, Arnoux P, Myrzakhmetov B, Frochot C, Guiavarc’h Y. Extraction, Identification and Photo-Physical Characterization of Persimmon (Diospyros kaki L.) Carotenoids. Foods. 2017; 6(1):4. https://doi.org/10.3390/foods6010004

Chicago/Turabian Style

Zaghdoudi, Khalil, Orleans Ngomo, Régis Vanderesse, Philippe Arnoux, Bauyrzhan Myrzakhmetov, Céline Frochot, and Yann Guiavarc’h. 2017. "Extraction, Identification and Photo-Physical Characterization of Persimmon (Diospyros kaki L.) Carotenoids" Foods 6, no. 1: 4. https://doi.org/10.3390/foods6010004

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop