Next Article in Journal
An Invertebrate Host to Study Fungal Infections, Mycotoxins and Antifungal Drugs: Tenebrio molitor
Next Article in Special Issue
Chemical Diversity and Biological Activities of Phaeosphaeria Fungi Genus: A Systematic Review
Previous Article in Journal
Caenorhabditis elegans as a Model Host to Monitor the Candida Infection Processes
Previous Article in Special Issue
Oryzines A & B, Maleidride Congeners from Aspergillus oryzae and Their Putative Biosynthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Development of an Improved Carotenoid Extraction Method to Characterize the Carotenoid Composition under Oxidative Stress and Cold Temperature in the Rock Inhabiting Fungus Knufia petricola A95

1
Department of Plant Physiology, Institute of Biology, Leipzig University, Johannisallee 21-23, 04103 Leipzig, Germany
2
Department of Materials & Environment, Bundesanstalt für Material-forschung und-prüfung, BAM, Unter den Eichen 87, 12205 Berlin, Germany
3
Department of Solar Materials, Applied Biocatalytics, Helmholtz Centre for Environmental Research, Permoser Strasse 15, 04318 Leipzig, Germany
*
Author to whom correspondence should be addressed.
J. Fungi 2018, 4(4), 124; https://doi.org/10.3390/jof4040124
Submission received: 2 October 2018 / Revised: 7 November 2018 / Accepted: 8 November 2018 / Published: 9 November 2018
(This article belongs to the Special Issue Fungal Polyketides and Other Secondary Metabolites)

Abstract

:
Black yeasts are a highly specified group of fungi, which are characterized by a high resistance against stress factors. There are several factors enabling the cells to survive harsh environmental conditions. One aspect is the pigmentation, the melanin black yeasts often display a highly diverse carotenoid spectrum. Determination and characterization of carotenoids depend on an efficient extraction and separation, especially for black yeast, which is characterized by thick cell walls. Therefore, specific protocols are needed to ensure reliable analyses regarding stress responses in these fungi. Here we present both. First, we present a method to extract and analyze carotenoids and secondly we present the unusual carotenoid composition of the black yeast Knufia petricola A95. Mechanical treatment combined with an acetonitrile extraction gave us very good extraction rates with a high reproducibility. The presented extraction and elution protocol separates the main carotenoids (7) in K. petricola A95 and can be extended for the detection of additional carotenoids in other species. K. petricola A95 displays an unusual carotenoid composition, with mainly didehydrolycopene, torulene, and lycopene. The pigment composition varied in dependency to oxidative stress but remained relatively constant if the cells were cultivated under low temperature. Future experiments have to be carried out to determine if didehydrolycopene functions as a protective agent itself or if it serves as a precursor for antioxidative pigments like torulene and torularhodin, which could be produced after induction under stress conditions. Black yeasts are a promising source for carotenoid production and other substances. To unravel the potential of these fungi, new methods and studies are needed. The established protocol allows the determination of carotenoid composition in black yeasts.

1. Introduction

A wide variety of natural pigments are produced by a wide spectrum of organisms, including bacteria, plants, and fungi. Carotenoids are the main pigments besides the essential photosynthetic chlorophyll pigments. The carotenoid synthesis pathways of most organisms share two steps, which start with the synthesis of geranylgeranyl pyrophosphate (GGPP) from the head-to-tail condensation of four C5 isoprene units and the tail-to-tail condensation of two GGPP units to produce the colorless precursor phytoene. These reactions are catalyzed by the enzymes GGPP synthase and phytoene synthase respectively. The introduction of conjugated double bonds into the phytoene backbone yields molecules able to absorb visible light and provides the characteristic yellow, orange to red colors of the carotenoids. Four desaturation steps are necessary for the conversion of phytoene to the maximally de-saturated red lycopene. These steps are carried out by a single enzyme [1,2], a carotene cyclase encoded by Al-2 gene in Neurospora crassa. Similar genes were found also in Exophiala strains [3]. Later reactions including cyclisation, isomerization, hydroxylation, and oxidation result in the more than 600 different natural carotenoids known so far [1,4,5]. Those can be divided into a class of hydrocarbons (carotenes) and their oxygenated derivatives (xanthophylls). In general, carotenoids have several functions, ranging from supporting light harvesting during photosynthesis, to protecting the cell from high light exposure, and the structural support of the photosynthetic apparatus [6]. They are important for the cells to survive UV radiation, oxidative stress, and water or salt stress. These properties make carotenoids interesting for industrial applications, for example as food colorants, nutritional supplements, cosmetics, or health purposes. Especially in fungi, oxidative stress protection seems to be the major function and is intensively investigated [6]. Furthermore, fungi and yeasts were recently screened for commercial carotenoid production. Several Basidiomycetous yeast genera (e.g., Rhodotorula or Phaffia) were found to be candidates for a commercial utilization [7,8,9]. Interestingly it seems that extremophilic fungi produce higher amounts and a higher diversity of carotenoids [10,11,12,13]. One important group of extremophilic fungi are the so called black fungi or black yeasts, which can cause human diseases. Such yeasts can be found in almost any environment and often form subaerial biofilms (SABs) [14,15,16,17]. Among black fungi, several salt and cold tolerant fungi were found, such species could be candidates for industrial carotenoid production [10,12]. The most obvious feature of black fungi is their dark pigmentation, which results from inclusion of melanin into the cell wall. Formation of melanin together with the production of secondary metabolites including carotenoids and mycosporines, are passive physiological adaptations that help these fungi resisting environmental stresses [18,19,20,21,22]. Secondly black fungi are characterized by a thick cell wall providing a high mechanical protection especially under water stress [22].
The rock inhabiting fungus K. petricola A95(syn. Sarcinomyces petricola), which was isolated from a marble rock near the Philopappos monument on Musaios Hill, Athens (Greece) [23] belongs to an ancestral lineage of the order Chaetothyriales [24] and possesses all characteristic features of black fungi, including meristematic growth, melaninized cell walls and extensive secondary metabolite production [22]. The desiccation resistant black yeast K. petricola A95 is recently under investigation regarding stress response mechanisms and adaptation processes [25]. Since it produces melanin, carotenoids and mycosporins [21], among other stress related substances, it is an ideal candidate to investigate the role of the single components regarding stress response in black fungi. The high variety of carotenoids in fungi and the growing fungal related biotechnology demand techniques and protocols to investigate pigment composition and production under various environmental conditions. Recent reviews show the growing interest of biotechnology industry to produce such pigments in yeast cultures or other fungi and listed important factors (carbon source, aeration, light etc.) and limiting steps (production rates, enzymatic activity etc.) in carotenoid production [7,8,26]. To analyze carotenoid concentration, accumulation, and conversion a highly efficient, cost effective reliable system with a low detection limit is needed. HPLC is one of the standard methods used for such analyses. However, sample preparation, especially the difficulties of extracting specific carotenoids (low concentration or derivatives), requires a specific sample treatment (cell disruption), extraction medium, and solvent mixture. Several studies discussed separation of xanthophylls with HPLC, but they reported difficulties for some carotenoids. Most studies propose a C18 column in combination with a wide variety of solvents; e.g., studies used water-acetone mixtures [27,28], acetonitrile or hexane as extraction medium or solvent. All of the extraction solutions or solvents face the same problem: that they are not equally suitable for all pigments. Additionally, a comparison of carotenoids by spectroscopic measurements is difficult, since the absorption spectra depend on the solvent used [29] and the solvents can chemically react with pigments and such reactions can result in pigment derivatives [30].
Several methods were tested to increase extraction efficiency, mechanical or chemical treatment, and the use of specific solvents. The enhancement of pigment extraction was demonstrated by using mechanical or chemical treatment [31,32], ultra-sonication for cell disruption, and by using different extraction media [33]. Cell consumption is critical for quantitative pigment analysis. As [34,35] demonstrated the extraction efficiency depends on two parameters: an efficient cell disruption and a proper extraction medium. Mechanical treatment seems to be the preferred method for cells with thick cell walls [31,32]. Other parameters that influence pigment stability during extraction and therefore should be considered are pH-values, oxygen concentration, light and temperature.
The main fungal pigments are β-carotene, astaxanthin, lycopene, and neurosporene. Recently the production of torulene and torularhodin in fungi has gained some interest, especially as stress response to UV radiation [10,36,37]. The exact function of carotenoids in fungi is still under discussion, since mutants lacking carotenoids display the same growth rates as wild type strains. In the cases investigated, a lack of carotenoids has no apparent phenotypic consequences on growth or morphology in laboratory cultures. However, secondary effects of carotenoids, especially in regard to membrane integrity, are under discussion [6]. It is known in fungi that carotenoid formation is activated by blue light and H2O2, which generates oxidative stress [38]. The rock inhabiting fungus K. petricola A95 produces: β-carotene, γ-carotene, phytoene, torulene, and torularhodin and desiccation/rehydration stresses affect the formation of the colorless phytone as well as the other carotenoids [22]. A new pigment extraction protocol was used to investigate the carotenoid composition of the rock inhabiting yeast K. petricola A95 under control conditions but also under temperature and oxidative stress. This new, robust, and reproducible method allows quantitative analysis of carotenoids in black yeasts to be used to determine the adaptation responses to extreme environmental conditions in high stress resistant fungi.

2. Material and Methods

Cell cultures of K. petricola A95 were grown on MEA (malt extract agar) plates at 25 °C and 4 °C. Cell colonies were harvested after 7 days and freeze dried. Oxidative stress related pigment accumulation was tested with cell cultures grown (equal cell number) in liquid BG11 (cyanobacteria media,) media with 2% malt extract and 0.2% glucose and with subsequent addition of 20 mM H2O2 for 2 h. K. petricola A95 was cultivated in BG11 media to investigate biofilm formation and interaction with cyanobacteria as described in [39], especially since such biofilms are exposed to harsh environmental conditions, leading to adaptations in pigment compositions in both organisms.
All cells were freeze dried after harvesting (Labconco FreeZone 2.5), dry weights of the cells were determined and the cells were homogenized. A mill for glass beads with integrated CO2 cooling was used for the extraction of the pigments. The freeze dried material was transferred into a glass vessel with ground glass joint (Sartorius). The glass beads used had a mixing ratio of 1:3 (w/w), including 1 part of cells with a diameter of 1 mm and 3 parts of beads with an average diameter of 0.25 mm (Carl ROTH GmbH). Simultaneously the samples were disrupted with 100% HPLC grade acetonitrile (Sigma-Aldrich) for 3 min. To reduce the amount of fine particles in the supernatant, the samples were centrifuged at 20,800 g (Centrifuge 5417 C, Eppendorf) twice for 1 min after the disrupting procedure. The remaining pellets were tested for extraction efficiency by applying extraction solvents afterwards, to make sure, the pigment extraction was complete.
The determination of the different carotenoids was realized via HPLC (DIONEX, Ultimate 3000) A reversed phase C18 column (300-5, C18-250 × 4 mm, from CS Chromatography Service) was used. A two eluent protocol was used with following gradient (Table 1)
Eluent A was composed of 80% acetonitrile (HPLC grade; Sigma-Aldrich) and 20% aqua dest. (v/v) and Eluent B was 100% ethyl acetate (HPLC grade; Sigma-Aldrich). The total protocol is 30 min long and was carried out with a flow rate of 0.8 mL/min. The used solvents were degassed for 15 min by ultrasonication prior use.
Pigment identification and quantification was established for lycopene, didehydrolycopene, dehydrolycopene, β-carotene, γ-carotene, torulene, and as well for torularhodin. Authentic standards were obtained for β-carotene and lycopene from Sigma-Aldrich. The quantification was based on the β-carotene calibration curve, only, because lycopene is not soluble in the extraction solvent. Lycopene was dissolved in 100% hexane (HPLC grade; Sigma-Aldrich). Calculation of pigment concentrations were based on β-carotene calibration and dry weight as reference values. Therefore, peak areas were related to β-carotene concentration and then divided by the specific extinction coefficients of the pigments in acetonitrile [40,41]. The pigments were detected at 440 nm for β-carotene, at 460 nm for of γ-carotene, lycopene and dehydrolycopene at 470 nm, torulene at 480 nm and torularhodin was detected at 497 nm. The dry weight was subsequently included. During the HPLC measurements, absorption spectra were ascertained by the HPLC-detector itself, supported by a tungsten lamp. The absorption spectra were determined from 402 to 762 nm. Furthermore, concentrations of purchased standards (β-carotene and lycopene) were verified by photometric measurements (Carl Zeiss Specord M500), to prevent overloading of the column. In all determined conditions three different biological replicates with a dry weight of about 0.1 g were analyzed. Pigment concentrations were subjected to the one-way analysis of variance (ANOVA) test in order to establish comparisons between different stress treatments. All datas generated or analyzed during this study are included in this published article.

3. Results

Pigment extraction was carried out with K. petricola A95 WT in comparison to a spontaneous melanin deficient mutant of K. petricola A95 (hereafter mdK) (Figure 1).
The mutant allowed us to determine the quality of the carotenoid extraction since the containing carotenoids were visible in this mutant and not masked by melanin as in the WT of K. petricola A95. Pigment extraction was carried out for freeze dried cell pellets with a minimum weight of 15 mg. Torularhodin was the pigment with the lowest concentration and minimum weight of the used material was adjusted to determine the specific amount of torularhodin accordingly. The thick cell wall of K. petricola A95 caused the use of relative long cell disruption times, and the shorter extraction lead to a higher variation in extraction efficiency. Therefore to ensure quantitative pigment extraction, cell pellets were tested for the remaining pigments in a second extraction step with 100% acetonitrile and 100% hexane. No significant amounts of pigments were detectable via HPLC analysis (<2 µg/g dry weight), once the method was established. The only pigment left in the pellet was melanin in the WT. The efficient pigment extraction after cell disruption was confirmed by using mdK of K. petricola A95, leaving a pale pellet after extraction. Ideally, pigment extraction and cell disruption must be carried out under low temperatures and in the dark without oxygen to prevent pigment degradation.
To verify the usability of our system the described standards of all the pigments were tested in our system. We determined the retention times of β-carotene and lycopene. Although lycopene was not completely soluble in 100% acetonitrile, a mixture of 105:30:25 (v/v/v) hexane, acetone and acetonitrile was used for this standard. A second, isocratic 100% hexane, gradient was established and in some cases the isocratic separation with hexane was used for the re-extraction of the pellets.
Several extraction solvents as well as HPLC separation set-ups were tested. Chemical extraction, for 2 h at 60 °C with either dimethyl sulfoxide or n-propanol, was proven to be not successful. Although a high extraction efficiency for both solvents was expected according to [34] and [42] and DMSO was frequently used for studies with fungi [43]. In this study, acetone was used with the described mill with integrated CO2 cooling combined with the Precellys 24 of Peqlab (at 6000 rpm for 5 up to 10 min) to ensure total cell disruption. In the latter cases the heating of the samples due to the rotation was the biggest problem and since sample cooling is crucial for stable pigment analysis, the mill was used exclusively. Furthermore, a solvent mixture of 40:20:40 (v/v/v) of acetonitrile, ethyl acetate, and n-propanol was tested, this mixture yielded good extraction results, but interacted negatively with the HPLC separation.
HPLC separation was tested with three eluents, acetonitrile/water, ethyl acetate, and ethyl acetate/acetone. Due to problems with a drifting baseline, eluents were reduced to eluent A (acetonitrile/distilled water) and eluent B (100% ethyl acetate). Eluent A was adapted, consisting of 80% acetonitrile (HPLC grade; Sigma-Aldrich)/20% distilled water. (v/v) was used. In combination with 100% acetonitrile as the extraction solvent the developed protocol ensures the detection of all pigments. All pigments show separate and distinct peaks. This was the most effective in extracting of the carotenoids from K. petricola A95 and was used as the established method. The total carotenoids in this approach were the highest, changing from, for example 153.7 µg g cdw−1 to 461.1 µg g cdw−1.
In the final analysis, seven different pigments could be observed and quantified (Figure 2) in K. petricola A95. The pigments are determined in the following order: dehyrolycopene (1), didehydrolycopene (2), which was always the largest peak, torulene (3), lycopene (4), γ-carotene (5) and β-carotene (6). The pigment torularhodin (7) was only detectable after prolonged incubation under low temperatures in trace amounts. Standards and absorption spectra were used to confirm the identity of the different pigments (Figure 2 and Figure 3 and Table 2).
Short-term incubation (2 h) of WT K. petricola A95 under oxidative stress (treatment with 20 mM H2O2) showed an increase in concentration for several carotenoids compared to growth under normal conditions for the same time period (Figure 4). The concentrations of lycopene (38.7 to 113.2 µg g cdw−1), γ-carotene (6.6 to 18.2 µg g cdw−1) and dihydrolycopene (8.5 to 36.2 µg g cdw−1) increased strongly under the applied conditions (differences significant for p = 0.05). The concentrations of β-carotene and torulene did not increase during the incubation and for didehydrolycopene an increase was detectable, however, the high variation between the samples did not allow us to draw conclusions. Didehydrolycopene showed in general the highest variation in pigment concentration in all samples.
Incubation under low temperature did not influence the pigment composition (Figure 5), compared to cultures grown under normal temperatures (the ANOVA Test showed no significant differences between the control samples and temperature stressed cells for all pigments). However, a strong increase in the didehydrolycopene concentration (up to 300 µg g cdw−1) under both conditions was visible after long term incubation. Therefore, the total measured pigment concentration increased to ~500 µg g cdw−1. Identical results for all experiments are exhibited for the spontaneous mdK of K. petricola A95. The spontaneous mutant displays a similar amount of pigments per dry weight and the carotenoid concentrations did not change as a response to temperature stress (data not shown).

4. Discussion

For an efficient pigment analysis, two steps are important: First, a solvent for extraction should be selected that allows the quantitative determination of all relevant pigments. In specific cases and with unknown samples, a selection and/or a mixture of media should be used, like acetonitrile, ethyl acetate, isopropanol, and hexane. The most suitable solvents seemed to be isopropanol and hexane, which were used to verify the extraction efficiency (re-extraction of the pellet), but caused a baseline shift in the HPLC. It is noteworthy to state that DMSO is to our opinion not a suitable extraction solvent; the difficulties to remove DMSO from the system limit the application in an adequate eluent system. After centrifugation and removal, the supernatant residues of DMSO still remain in the sample, which is problematic for HPLC analyses. Additionally, the proposed DMSO based method of pigment extraction [37,44] should be tested accordingly, depending on the systems used, to ensure optimal pigment analysis results. Thus, DMSO extraction is always a time consuming method and the results are hardly comparable with other methods. Secondly an efficient cell rupture method under conditions that ensure the pigment stability should be selected. Black yeasts are characterized by a thick cell wall with high melanin concentration. In general pigment extraction should be performed under low oxygen, and if possible under oxygen free conditions. The major challenge in pigment analysis is to recover pigments with minimum risk of damage but with high efficiency. A mechanical treatment is needed in the most cases but the method applied should be tested carefully due to the high sensitivity of the molecules. Especially for black yeasts and other fungi with a thick cell wall, a mechanical cell disruption seems necessary and provides better results as a chemical treatment [31,32].
K. petricola A95 displayed an uncommon pigment composition; especially in the high concentrations of dihydrolycopene, didehydrolycopene and torulene, which are not so often described in fungi. A normally high abundance of carotene derivatives and lycopene can be observed in pigmented fungi [45,46]. Several yeasts showed a high torulene concentration but no dihydrolycopene was detected in yeast so far [46,47]. The abundance of didehydrolycopene and torulene indicates a similar production pathway as described for N. crassa, leading in this fungus to neurosporaxanthin [2,48], a carotenoid that was not detectable in K. petricola A95 under the applied conditions. It is noteworthy to mention that the determination of the abundance of astaxanthin and neurosporaxanthin is technically possible with the described protocol. The only pigment left in the pellet was melanin in the WT. Hence comparing the results with methods developed for pigments in algae or higher plants [26,27,31,49,50,51] is not advisable due to the thick cultured cell wall of the supervised fungi. Although [49] compared different methods using maize seeds, they are not characterized by a thick cell wall like black yeast. Similar to our study, the decision regarding efficacy was made with respect to the color of the pellet. In general, these studies provide a wide method spectra and were useful regarding extraction, elution, and calibration of the authentic standards, see also [31,51]. Fungi pigments were mainly analyzed, focusing on certain valuable pigments only. Knowledge regarding pigment composition in fungi is needed for species description, however, such studies used standard methods and did not compare different methods [5,6,7,10,33,34,52]. Methodological overviews for fungi pigment extraction are still missing [38]. Therefore, the established method was compared to [19]. The combination of extraction media and the eluent system of [19] resulted in a strong injection peak in the presented HPLC system and was adjusted afterwards, concerning the different tested method combinations shown.
The detected carotenoids are consistent with [22], they found also carotene derivatives, torulene, and torularhodin. Our extraction method, especially the extraction medium and the mechanical treatment, gave us additionally the possibility to quantify lycopene and didehydrolycopene, unusual pigments in fungi and not often found in yeast strains.
To our opinion didehydrolycopene functions as a synthesis interface, comparable to β-carotene in photosynthetic organisms. Such a pool of didehydrolycopene can be converted fast to stress response pigments and helps cells to respond quickly to environmental stresses. The two main pigments in K. petricola A95, lycopene and didehydrolycopene, display much higher concentrations, as in other fungi detected [53] and β-carotene concentrations are lower compared to other fungi and yeasts. The detected lycopene derivatives are the major carotenoids in K. petricola A95. Future experiments must be carried out to find out if these pigments have a photoprotective role or if they function as a precursor for other pigments. In N. crassa the conversion of didehydrolycopene to torulene by cyclisation was described [54]. The already described torularhodin and torulene could be involved in photoprotection [6,13,37,55]. The antioxidant properties of torulene are attributed to its conjugated double bond system; in fact, torulene has more antioxidant efficiency than β-carotene, which presents less of a double bond on its chemical structure than torulene [56]. However, several other carotenoids were also described to have an antioxidant activity.
Another explanation for the stable carotenoid concentrations in K. petricola A95 would be that the mechanism of carotenoid action is more likely to consist of shielding sensitive molecules or organelles than of the neutralization of harmful oxidants [52]. Therefore, carotenoids do not play a major physiological role in fungal cells, but they may have beneficial effects under specific conditions. Several studies showed an increase in carotenoid concentration under stress [45,57]. The potential photoprotective pigment torulene [9,58] showed low concentrations in K. petricola A95, which would imply a less important role in K. petricola A95, but an inducible production under a stress situation out of the more stable dehydrolycopene seems possible. However, we could not detect an increase in torulene concentrations if the cells were exposed to oxidative or temperature stress. Such an induction was described for a black yeast under oxidative stress [59], in red yeast under salt stress [58], and a light dependent increase of the overall carotenoid concentration was described for N. crassa [38] incubated with H2O2. Such an inducible protective system would enable K. petricola A95 to react fast to stress situations. A similar system was described for the red yeast Dioszegia with the xanthophyll plectaniaxanthin [60]. The photoprotective carotenoid torularhodin was induced in K. petricola A95 under cold temperatures, but very low amounts and other stresses or longer incubation times have to be investigated to unravel the function of this pigment in K. petricola A95.
The major changes under the applied conditions were detected for lycopene and derivatives, which should therefore be considered as the most important pigments in K. petricola A95. The high concentration of the unusual carotenoid didehydrolycopene in K. petricola A95 should be investigated in detail, especially since the photoprotective role and the biotechnological potential of lycopene derivatives was described before [49]. Screening other black yeasts regarding their pigmentation could result in a wide diversity of pigments with multiple promising applications. Black fungi are remarkable in their stress resistance and we showed that their carotenoid pigmentation is complex and future studies have to be performed to determine the specific function of the specific pigments. The protocol presented will allow the quantitative analysis of black fungi, characterized by thick cell walls and melanin pigmentation, regarding stress response and adaptation to extreme environmental conditions. Future experiments have to unravel the function of melanin, carotenoids, and the cell wall structure regarding the stress responses in detail.

5. Conclusions

Extremotolerant and extremophile black yeasts are a promising source of pigments and other chemicals. New protocols and studies are needed to determine the capacity for production of these high stress resistance fungi. The established protocol allows the determination of carotenoid composition in black yeasts. Oxidative stress results in an adaptation in pigment composition. Future experiments have to be carried out to determine if didehydrolycopene functions as a protective agent itself or if it serves as a precursor for antioxidative pigments like torulene and torularhodin, which could be produced after induction under stress conditions.

Author Contributions

J.T. and N.K. performed the growth experiments and drafted this manuscript. K.F. developed the HPLC methods and participated in the conception of the study, analysis of data and revision of the manuscript. J.T. and N.K. participated in the conception of the study, supervised the experiments, and revised the manuscript. All authors read and approved the final manuscript.

Funding

This work was supported financially by internal funds of the BAM.

Acknowledgments

We thank Anna Gorbushina (BAM Berlin) and Christian Wilhelm (University of Leipzig) for their support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Diaz-Sanchez, V.; Estrada, A.F.; Trautmann, D.; Limon, M.C.; Al-Babili, S.; Avalos, J. Analysis of al-2 mutations in Neurospora. PLoS ONE 2011, 6. [Google Scholar] [CrossRef] [PubMed]
  2. Hausmann, A.; Sandmann, G. A single five-step desaturase is involved in the carotenoid biosynthesis pathway to β-carotene and torulene in Neurospora crassa. Fungal Genet. Biol. 2000, 30, 147–153. [Google Scholar] [CrossRef] [PubMed]
  3. Tafer, H.; Lopandic, K.; Blasi, B.; Poyntner, C.; Sterflinger, K. Draft genome sequence of Exophiala mesophila, a black yeast with high bioremediation potential. Genome Announc. 2015, 3. [Google Scholar] [CrossRef] [PubMed]
  4. Sandmann, G. Carotenoid biosynthesis in microorganisms and plants. Eur. J. Biochem. 1994, 223, 7–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Armstrong, G.A.; Hearst, J.E. Carotenoids 2: Genetics and molecular biology of carotenoid pigment biosynthesis. FASEB J. 1996, 10, 228–237. [Google Scholar] [CrossRef] [PubMed]
  6. Avalos, J.; Carmen Limon, M. Biological roles of fungal carotenoids. Curr. Genet. 2014, 61, 309–324. [Google Scholar] [CrossRef] [PubMed]
  7. Frengova, G.I.; Beshkova, D.M. Carotenoids from Rhodotorula and Phaffia: Yeasts of biotechnological importance. J. Ind. Microbiol. Biotechnol. 2009, 36, 163–180. [Google Scholar] [CrossRef] [PubMed]
  8. Mata-Gomez, L.C.; Montanez, J.C.; Mendez-Zavala, A.; Aguilar, C.N. Biotechnological production of carotenoids by yeasts: An overview. Microb. Cell Fact. 2014, 13, 12. [Google Scholar] [CrossRef] [PubMed]
  9. Moline, M.; Libkind, D.; van Broock, M. Production of torularhodin, torulene, and β-carotene by Rhodotorula yeasts. Methods Mol. Biol. 2012, 898, 275–283. [Google Scholar] [PubMed]
  10. Amaretti, A.; Simone, M.; Quartieri, A.; Masino, F.; Raimondi, S.; Leonardi, A.; Rossi, M. Isolation of carotenoid-producing yeasts from an alpine glacier. Chem. Eng. Trans. 2014, 38, 217–222. [Google Scholar]
  11. Arcangeli, C.; Cannistraro, S. In situ raman microspectroscopic identification and localization of carotenoids: Approach to monitoring of UV-B irradiation stress on antarctic fungus. Biopolymers 2000, 57, 179–186. [Google Scholar] [CrossRef]
  12. Libkind, D.; Moline, M.; Sampaio, J.P.; van Broock, M. Yeasts from high-altitude lakes: Influence of UV radiation. FEMS Microbiol. Ecol. 2009, 69, 353–362. [Google Scholar] [CrossRef] [PubMed]
  13. Moline, M.; Libkind, D.; Dieguez Mdel, C.; van Broock, M. Photoprotective role of carotenoids in yeasts: Response to UV-B of pigmented and naturally-occurring albino strains. J. Photochem. Photobiol. B 2009, 95, 156–161. [Google Scholar] [CrossRef] [PubMed]
  14. Teixeira, M.M.; Moreno, L.F.; Stielow, B.J.; Muszewska, A.; Hainaut, M.; Gonzaga, L.; Abouelleil, A.; Patane, J.S.; Priest, M.; Souza, R.; et al. Exploring the genomic diversity of black yeasts and relatives (Chaetothyriales, Ascomycota). Stud. Mycol. 2017, 86, 1–28. [Google Scholar] [CrossRef] [PubMed]
  15. De Hoog, G.S.; Vicente, V.A.; Gorbushina, A.A. The bright future of darkness—the rising power of black fungi: Black yeasts, microcolonial fungi, and their relatives. Mycopathologia 2013, 175, 365–368. [Google Scholar] [CrossRef] [PubMed]
  16. Gunde-Cimerman, N.; Grube, M.; de Hoog, G.S. The emerging potential of melanized fungi: Black yeast between beauty and the beast. Fungal Biol. 2011, 115, 935–936. [Google Scholar] [CrossRef] [PubMed]
  17. Gorbushina, A.A.; Broughton, W.J. Microbiology of the atmosphere-rock interface: How biological interactions and physical stresses modulate a sophisticated microbial ecosystem. Annu. Rev. Microbiol. 2009, 63, 431–450. [Google Scholar] [CrossRef] [PubMed]
  18. Sterflinger, K. Temperature and NaCl-tolerance of rock-inhabiting meristematic fungi. Antonie Van Leeuwenhoek 1998, 74, 271–281. [Google Scholar] [CrossRef] [PubMed]
  19. Kogej, T.; Stein, M.; Volkmann, M.; Gorbushina, A.A.; Galinski, E.A.; Gunde-Cimerman, N. Osmotic adaptation of the halophilic fungus hortaea werneckii: Role of osmolytes and melanization. Microbiol-Sgm 2007, 153, 4261–4273. [Google Scholar] [CrossRef] [PubMed]
  20. Kogej, T.; Wheeler, M.H.; Rizner, T.L.; Gunde-Cimerman, N. Evidence for 1,8-dihydroxynaphthalene melanin in three halophilic black yeasts grown under saline and non-saline conditions. FEMS Microbiol. Lett. 2004, 232, 203–209. [Google Scholar] [CrossRef] [Green Version]
  21. Volkmann, M.; Gorbushina, A.A. A broadly applicable method for extraction and characterization of mycosporines and mycosporine-like amino acids of terrestrial, marine and freshwater origin. FEMS Microbiol. Lett. 2006, 255, 286–295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Gorbushina, A.A.; Kotlova, E.R.; Sherstneva, O.A. Cellular responses of microcolonial rock fungi to long-term desiccation and subsequent rehydration. Stud. Mycol. 2008, 61, 91–97. [Google Scholar] [CrossRef] [PubMed]
  23. Wollenzien, U.; de Hoog, G.S.; Krumbein, W.; Uijthof, J.M. Sarcinomyces petricola, a new microcolonial fungus from marble in the mediterranean basin. Antonie Van Leeuwenhoek 1997, 71, 281–288. [Google Scholar] [CrossRef] [PubMed]
  24. Gueidan, C.; Villasenor, C.R.; de Hoog, G.S.; Gorbushina, A.A.; Untereiner, W.A.; Lutzoni, F. A rock-inhabiting ancestor for mutualistic and pathogen-rich fungal lineages. Stud. Mycol. 2008, 61, 111–119. [Google Scholar] [CrossRef] [PubMed]
  25. Nai, C.; Wong, H.Y.; Pannenbecker, A.; Broughton, W.J.; Benoit, I.; de Vries, R.P.; Gueidan, C.; Gorbushina, A.A. Nutritional physiology of a rock-inhabiting, model microcolonial fungus from an ancestral lineage of the Chaetothyriales (Ascomycetes). Fungal Genet. Biol. 2013, 56, 54–66. [Google Scholar] [CrossRef] [PubMed]
  26. Sandmann, G. Carotenoids of biotechnological importance. Adv. Biochem. Eng. Biotechnol. 2015, 148, 449–467. [Google Scholar] [PubMed]
  27. Weber, R.W.; Anke, H.; Davoli, P. Simple method for the extraction and reversed-phase high-performance liquid chromatographic analysis of carotenoid pigments from red yeasts (Basidiomycota, fungi). J. Chromatogr. A 2007, 1145, 118–122. [Google Scholar] [CrossRef] [PubMed]
  28. Bona-Lovasz, J.; Bona, A.; Ederer, M.; Sawodny, O.; Ghosh, R. A rapid method for the extraction and analysis of carotenoids and other hydrophobic substances suitable for systems biology studies with photosynthetic bacteria. Metabolites 2013, 3, 912–930. [Google Scholar] [CrossRef] [PubMed]
  29. Zang, L.Y.; Sommerburg, O.; VanKuijk, F.J.G.M. Absorbance changes of carotenoids in different solvents. Free Radic. Biol. Med. 1997, 23, 1086–1089. [Google Scholar] [CrossRef]
  30. Takaichi, S. Characterization of carotenes in a combination of a C (18) HPLC column with isocratic elution and absorption spectra with a photodiode-array detector. Photosynthesis Res. 2000, 65, 93–99. [Google Scholar] [CrossRef] [PubMed]
  31. Geciova, J.; Bury, D.; Jelen, P. Methods for disruption of microbial cells for potential use in the dairy industry—A review. Int. Dairy J. 2002, 12, 541–553. [Google Scholar] [CrossRef]
  32. Gu, Z.; Deming, C.; Yongbin, H.; Zhigang, C.; Feirong, G. Optimization of carotenoids extraction from Rhodobacter sphaeroides. LWT-Food Sci. Technol. 2008, 41, 1082–1088. [Google Scholar] [CrossRef]
  33. Papaioannou, E.; Roukas, T.; Liakopoulou-Kyriakides, M. Effect of biomass pre-treatment and solvent extraction on β-carotene and lycopene recovery from Blakeslea trispora cells. Prep. Biochem. Biotechnol. 2008, 38, 246–256. [Google Scholar] [CrossRef] [PubMed]
  34. Craft, N.E.; Soares, J.H. Relative solubility, stability, and absorptivity of lutein and β-carotene in organic-solvents. J. Agric. Food Chem. 1992, 40, 431–434. [Google Scholar] [CrossRef]
  35. Su, Q.; Rowley, K.G.; Balazs, N.D.H. Carotenoids: Separation methods applicable to biological samples. J. Chromatogr. B 2002, 781, 393–418. [Google Scholar] [CrossRef]
  36. Sperstad, S.; Lutnaes, B.F.; Stormo, S.K.; Liaaen-Jensen, S.; Landfald, B. Torularhodin and torulene are the major contributors to the carotenoid pool of marine Rhodosporidium babjevae (golubev). J. Ind. Microbiol. Biotechnol. 2006, 33, 269–273. [Google Scholar] [CrossRef] [PubMed]
  37. Moline, M.; Flores, M.R.; Libkind, D.; Dieguez Mdel, C.; Farias, M.E.; van Broock, M. Photoprotection by carotenoid pigments in the yeast rhodotorula mucilaginosa: The role of torularhodin. Photochem. Photobiol. Sci. 2010, 9, 1145–1151. [Google Scholar] [CrossRef] [PubMed]
  38. Iigusa, H.; Yoshida, Y.; Hasunuma, K. Oxygen and hydrogen peroxide enhance light-induced carotenoid synthesis in Neurospora crassa. FEBS Lett. 2005, 579, 4012–4016. [Google Scholar] [CrossRef] [PubMed]
  39. Seiffert, F.; Bandow, N.; Bouchez, J.; von Blanckenburg, F.; Gorbushina, A.A. Microbial colonization of bare rocks: Laboratory biofilm enhances mineral weathering. Procedia Earth Planet. Sci. 2014, 10, 123–129. [Google Scholar] [CrossRef]
  40. Britton, G.; Liaaen-Jensen, S.; Pfander, H. Carotenoids Handbook; Birkhäuser Verlag: Basel, Switzerland, 2004; ISBN 3764361808. [Google Scholar]
  41. Foppen, F.H. Tables for the identification of carotenoid pigments. Chromatogr. Rev. 1971, 14, 133–298. [Google Scholar] [CrossRef]
  42. Choi, S.K.; Kim, J.H.; Park, Y.S.; Kim, Y.J.; Chang, H.I. An efficient method for the extraction of astaxanthin from the red yeast Xanthophyllomyces dendrorhous. J. Microbiol. Biotechnol. 2007, 17, 847–852. [Google Scholar] [PubMed]
  43. Sedmak, J.J.; Weerasinghe, D.; Jolly, S. Extraction and quantitation of astaxanthin from Phaffia rhodozyma. Biotechnol. Technol. 1990, 4, 107–112. [Google Scholar] [CrossRef]
  44. Libkind, D.; Perez, P.; Sommaruga, R.; Dieguez, M.D.; Ferraro, M.; Brizzio, S.; Zagarese, H.; van Broock, M. Constitutive and UV-inducible synthesis of photoprotective compounds (carotenoids and mycosporines) by freshwater yeasts. Photochem. Photobiol. Sci. 2004, 3, 281–286. [Google Scholar] [CrossRef] [PubMed]
  45. Gmoser, R.; Ferreira, J.A.; Lennartsson, P.R.; Taherzadeh, M.J. Filamentous ascomycetes fungi as a source of natural pigments. Fungal Biol. Biotechnol. 2017, 4, 4. [Google Scholar] [CrossRef] [PubMed]
  46. Zoz, L.; Carvalho, J.C.; Soccol, V.T.; Casagrande, T.C.; Cardoso, L. Torularhodin and torulene: Bioproduction, properties and prospective applications in food and cosmetics—A review. Braz. Arch. Biol. Technol. 2015, 58, 278–288. [Google Scholar] [CrossRef]
  47. Buzzini, P.; Innocenti, M.; Turchetti, B.; Libkind, D.; van Broock, M.; Mulinacci, N. Carotenoid profiles of yeasts belonging to the genera Rhodotorula, Rhodosporidium, Sporobolomyces, and Sporidiobolus. Can. J. Microbiol. 2007, 53, 1024–1031. [Google Scholar] [CrossRef] [PubMed]
  48. Avalos, J.; Prado-Cabrero, A.; Estrada, A.F. Neurosporaxanthin production by Neurospora and Fusarium. Methods Mol. Biol. 2012, 898, 263–274. [Google Scholar] [PubMed]
  49. Albrecht, M.; Takaichi, S.; Steiger, S.; Wang, Z.-Y.; Sandmann, G. Novel hydroxycarotenoids with improved antioxidative properties produced by gene combination in Escherichia coli. Nat. Biotechnol. 2000, 18, 843. [Google Scholar] [CrossRef] [PubMed]
  50. Olives Barba, A.I.; Camara Hurtado, M.; Sanchez Mata, M.C.; Fernadez Ruiz, V.; Lopez Saenz de Tejada, M. Application of a UV–vis detection-HPLC method for a rapid determination of lycopene and β-carotene in vegetables. Food Chem. 2006, 95, 328–336. [Google Scholar] [CrossRef]
  51. Busch, M.; Seuter, A.; Hain, R. Functional analysis of the early steps of carotenoid biosynthesis in tobacco. Plant Physiol. 2012, 128, 439–453. [Google Scholar] [CrossRef] [PubMed]
  52. Schnitzler, N.; Peltroche-Llacsahuanga, H.; Bestier, N.; Zundorf, J.; Lutticken, R.; Haase, G. Effect of melanin and carotenoids of Exophiala (Wangiella) dermatitidis on phagocytosis, oxidative burst, and killing by human neutrophils. Infect. Immun. 1999, 67, 94–101. [Google Scholar] [PubMed]
  53. Zhang, Z.; Zhang, X.; Tan, T. Lipid and carotenoid production by Rhodotorula glutinis under irradiation/high-temperature and dark/low-temperature cultivation. Bioresour. Technol. 2014, 157, 149–153. [Google Scholar] [CrossRef] [PubMed]
  54. Arrach, N.; Schmidhauser, T.J.; Avalos, J. Mutants of the carotene cyclase domain of al-2 from Neurospora crassa. Mol. Genet. Genom. 2002, 266, 914–921. [Google Scholar]
  55. Du, C.; Li, Y.; Guo, Y.; Han, M.; Zhang, W.; Qian, H. The suppression of torulene and torularhodin treatment on the growth of PC-3 xenograft prostate tumors. Biochem. Biophys. Res. Commun. 2016, 469, 1146–1152. [Google Scholar] [CrossRef] [PubMed]
  56. Sakaki, H.; Nochide, H.; Komemushi, S.; Miki, W. Effect of active oxygen species on the productivity of torularhodin by Rhodotorula glutinis No. 21. J. Biosci. Bioeng. 2002, 93, 338–340. [Google Scholar] [CrossRef]
  57. Nanou, K.; Roukas, T.; Papadakis, E.; Kotzekidou, P. Carotene production from waste cooking oil by Blakeslea trispora in a bubble column reactor: The role of oxidative stress. Eng. Life Sci. 2017, 17, 775–780. [Google Scholar] [CrossRef]
  58. Li, C.; Zhang, N.; Li, B.; Xu, Q.; Song, J.; Wei, N.; Wang, W.; Zou, H. Increased torulene accumulation in red yeast Sporidiobolus pararoseus NGR as stress response to high salt conditions. Food Chem. 2017, 237, 1041–1047. [Google Scholar] [CrossRef] [PubMed]
  59. Geis, P.A.; Szaniszlo, P.J. Carotenoid-pigments of the dematiaceous fungus Wangiella-dermatitidis. Mycologia 1984, 76, 268–273. [Google Scholar] [CrossRef]
  60. Madhour, A.; Anke, H.; Mucci, A.; Davoli, P.; Weber, R.W.S. Biosynthesis of the xanthophyll plectaniaxanthin as a stress response in the red yeast Dioszegia (Tremellales, Heterobasidiomycetes, fungi). Phytochemistry 2005, 66, 2617–2626. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Colonies of K. petricola A95 (a) WT and (b) the spontaneous melanin deficient mutant (mdK) grown on MEA plates for 14 days at 25 °C. Colony size approximately 0.5 cm.
Figure 1. Colonies of K. petricola A95 (a) WT and (b) the spontaneous melanin deficient mutant (mdK) grown on MEA plates for 14 days at 25 °C. Colony size approximately 0.5 cm.
Jof 04 00124 g001
Figure 2. Representative HPLC chromatogram of carotenoids in K. petricola A95 detected in our experiments. The peaks were determined via reference substances and identified as dehyrolycopene (1), didehydrolycopene (2), torulene (3), lycopene (4), γ-carotene (5) and β-carotene (6) and torularhodin (7).
Figure 2. Representative HPLC chromatogram of carotenoids in K. petricola A95 detected in our experiments. The peaks were determined via reference substances and identified as dehyrolycopene (1), didehydrolycopene (2), torulene (3), lycopene (4), γ-carotene (5) and β-carotene (6) and torularhodin (7).
Jof 04 00124 g002
Figure 3. Absorption spectra of main carotenoids in K. petricola A95 (determined via HPLC detector in according eluent mixture). The maxima of each pigment was determined: didehydrolycopene (463 nm, 492 nm, 523 nm); torulene (454 nm, 484 nm, 515 nm); dehydrolycopene (448 nm, 476 nm, 507 nm); lycopene (440 nm, 465 nm, 496 nm).
Figure 3. Absorption spectra of main carotenoids in K. petricola A95 (determined via HPLC detector in according eluent mixture). The maxima of each pigment was determined: didehydrolycopene (463 nm, 492 nm, 523 nm); torulene (454 nm, 484 nm, 515 nm); dehydrolycopene (448 nm, 476 nm, 507 nm); lycopene (440 nm, 465 nm, 496 nm).
Jof 04 00124 g003
Figure 4. Carotenoid composition of WT K. petricola A95 incubated with and without 20 mM H2O2 for 2 h, determined via HPLC analysis (n = 3). Significant differences at p < 0.05 with respect to the control are marked by asterisks (one-way ANOVA).
Figure 4. Carotenoid composition of WT K. petricola A95 incubated with and without 20 mM H2O2 for 2 h, determined via HPLC analysis (n = 3). Significant differences at p < 0.05 with respect to the control are marked by asterisks (one-way ANOVA).
Jof 04 00124 g004
Figure 5. Pigment composition of K. petricola A95 (WT) determined via HPLC after growth under normal conditions (25 °C) and under low temperature (4 °C) for 7 days (n = 3). Significant differences at p < 0.05 with respect to the wild type are marked by asterisks (one-way ANOVA).
Figure 5. Pigment composition of K. petricola A95 (WT) determined via HPLC after growth under normal conditions (25 °C) and under low temperature (4 °C) for 7 days (n = 3). Significant differences at p < 0.05 with respect to the wild type are marked by asterisks (one-way ANOVA).
Jof 04 00124 g005
Table 1. Eluent protocol used for carotenoid separation.
Table 1. Eluent protocol used for carotenoid separation.
Time [min]Eluent AEluent B
060%40%
750%50%
1740%60%
2130%70%
28.520%80%
29.510%90%
30.5 to 42 min60%40%
Table 2. Pigments absorption maxima and concentrations of the black yeast Knufia petricola A95 grown on MEA plates under standard growth conditions.
Table 2. Pigments absorption maxima and concentrations of the black yeast Knufia petricola A95 grown on MEA plates under standard growth conditions.
PigmentAbsorption Maxima [nm]Concentration [µg/g dry weight]
dihydrolycopene448, 476, 50730.74
didehydrolycopene463, 492, 523383.12
torulene454, 484, 51585.39
lycopene440, 465, 496100.30
γ-carotene460, 48923.88
β-carotene457, 4837.53
torularhodin469, 497, 5255.84

Share and Cite

MDPI and ACS Style

Flieger, K.; Knabe, N.; Toepel, J. Development of an Improved Carotenoid Extraction Method to Characterize the Carotenoid Composition under Oxidative Stress and Cold Temperature in the Rock Inhabiting Fungus Knufia petricola A95. J. Fungi 2018, 4, 124. https://doi.org/10.3390/jof4040124

AMA Style

Flieger K, Knabe N, Toepel J. Development of an Improved Carotenoid Extraction Method to Characterize the Carotenoid Composition under Oxidative Stress and Cold Temperature in the Rock Inhabiting Fungus Knufia petricola A95. Journal of Fungi. 2018; 4(4):124. https://doi.org/10.3390/jof4040124

Chicago/Turabian Style

Flieger, Kerstin, Nicole Knabe, and Jörg Toepel. 2018. "Development of an Improved Carotenoid Extraction Method to Characterize the Carotenoid Composition under Oxidative Stress and Cold Temperature in the Rock Inhabiting Fungus Knufia petricola A95" Journal of Fungi 4, no. 4: 124. https://doi.org/10.3390/jof4040124

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