Next Article in Journal
An Overview of the Safety Profile and Clinical Impact of CDK4/6 Inhibitors in Breast Cancer—A Systematic Review of Randomized Phase II and III Clinical Trials
Previous Article in Journal
N-Glycome Profile of the Spike Protein S1: Systemic and Comparative Analysis from Eleven Variants of SARS-CoV-2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

RP-HPLC Separation and 1H NMR Identification of a Yellow Fluorescent Compound—Riboflavin (Vitamin B2)—Produced by the Yeast Hyphopichia wangnamkhiaoensis

by
Raziel Arturo Jiménez-Nava
1,2,
Luis Gerardo Zepeda-Vallejo
3,
Fortunata Santoyo-Tepole
2,4,
Griselda Ma. Chávez-Camarillo
2,* and
Eliseo Cristiani-Urbina
1,*
1
Departamento de Ingeniería Bioquímica, Instituto Politécnico Nacional, Escuela Nacional de Ciencias Biológicas, Avenida Wilfrido Massieu s/n, Unidad Profesional Adolfo López Mateos, Ciudad de Mexico 07738, Mexico
2
Departamento de Microbiología, Instituto Politécnico Nacional, Escuela Nacional de Ciencias Biológicas, Prolongación de Carpio y Plan de Ayala s/n, Colonia Santo Tomás, Ciudad de Mexico 11340, Mexico
3
Departamento de Química Orgánica, Instituto Politécnico Nacional, Escuela Nacional de Ciencias Biológicas, Prolongación de Carpio y Plan de Ayala s/n, Colonia Santo Tomás, Ciudad de Mexico 11340, Mexico
4
Departamento de Investigación, Instituto Politécnico Nacional, Escuela Nacional de Ciencias Biológicas, Prolongación de Carpio y Plan de Ayala s/n, Colonia Santo Tomás, Ciudad de Mexico 11340, Mexico
*
Authors to whom correspondence should be addressed.
Biomolecules 2023, 13(9), 1423; https://doi.org/10.3390/biom13091423
Submission received: 29 August 2023 / Revised: 15 September 2023 / Accepted: 18 September 2023 / Published: 20 September 2023

Abstract

:
The yeast Hyphopichia wangnamkhiaoensis excretes a brilliant yellow fluorescent compound into its growth culture. In this study, we isolated and identified this compound using reverse-phase high-performance liquid chromatography-diode array detector (RP-HPLC-DAD) as well as 1H NMR and UV–Vis spectroscopy. Two of the three RP-HPLC-DAD methods used successfully separated the fluorescent compound and involved (1) a double separation step with isocratic flow elution, first on a C18 column and later on a cyano column, and (2) a separation with a linear gradient elution on a phenyl column. The wavelengths of maximum absorption of the fluorescent compound-containing HPLC fractions (~224, 268, 372, and 446 nm) are in good agreement with those exhibited by flavins. The 1H NMR spectra revealed methyl (δ 2.30 and 2.40) and aromatic proton (δ 7.79 and 7.77) signals of riboflavin. The 1H NMR spectra of the samples spiked with riboflavin confirmed that the brilliant yellow fluorescent compound is riboflavin. The maximum excitation and emission wavelengths of the fluorescent compound were 448 and 528 nm, respectively, which are identical to those of riboflavin.

1. Introduction

Fluorescent compounds, also known as fluorochromes or fluorophores, are widely used as markers in microscopy of biological samples, histology, molecular biology, immunology, material sciences, and chemistry. Several fluorochromes have also been used as drugs [1,2,3,4]. Fluorochromes are classified as intrinsic or extrinsic based on their luminescence principle and synthetic or organic (natural) based on their chemical origin [5]. Synthetic fluorochromes are the most commonly used fluorochromes owing to their higher stability in reaction to physicochemical changes. Notably, the vitamins riboflavin (B2), ergocalciferol (D2), phytomenadione (K1), and pyridoxine (B6); the amino acids phenylalanine, tyrosine, and tryptophan; and the vitamin precursors ergosterol and β-carotene are some important organic fluorochromes for human health [3,5,6].
Several organic fluorochromes are produced industrially by microorganisms [7,8,9,10] in high yields and with high specificity. Yeasts have attracted significant attention in recent years for their ability to biosynthesize fluorochromes. Yeasts are metabolically versatile, low maintenance, easy to cultivate, safe, and have a high specific growth rate and short duplication time [11,12]. Some of the fluorochromes produced by yeasts are as follows: (1) ergosterol, a precursor of vitamin D2, is produced by Cystofilobasidium capitatum, Rhodotorula glutinis, and Sporobolomyces roseus; (2) β-carotene, a precursor of vitamin A, is produced by the same yeast species mentioned previously (C. capitatum, R. glutinis, and S. roseus) [9,10,13,14,15], and (3) riboflavin, an essential vitamin for human nutrition, is produced by some yeast species such as Candida famata and Meyerozyma guilliermoindii (formerly known as Candida or Pichia guilliermondii) [16,17,18,19,20].
Nowadays, high-performance liquid chromatography (HPLC) is one of the most reliable analytical techniques for separating, quantifying, and identifying fluorochromes [21,22,23]. Likewise, nuclear magnetic resonance spectroscopy and mass spectrometry are powerful analytical tools that have been used to quantify known fluorochromes, as well as to elucidate the chemical identity, structure, and properties of fluorochromes [24,25,26].
Hyphopichia wangnamkhiaoensis, formerly known as Candida wangnamkhiaoensis and Wickerhamia sp. X-Fep, is a dimorphic yeast species capable of producing high levels of extracellular α-amylase [12,27,28,29] and oleic acid [30]. During the course of our investigation into the above-mentioned biosynthesis, we observed that the cultures of H. wangnamkhiaoensis turned yellow during the production of α-amylase and oleic acid, which was attributed to a yellow fluorescent compound. However, this compound has not yet been characterized in detail. Therefore, in this study, we isolated the bright yellow fluorescent compound produced by H. wangnamkhiaoensis using reverse-phase high-performance liquid chromatography (HPLC) and characterized it using nuclear magnetic resonance (NMR) and UV–visible and fluorescence spectroscopy. To the best of our knowledge, this is the first report of the isolation and structural characterization of a fluorescent compound produced by H. wangnamkhiaoensis.

2. Materials and Methods

2.1. Reagents

HPLC-grade methanol, water, acetonitrile, and other chemical reagents used in the culture medium were purchased from JT Baker (Avantor Performance Materials, Inc., Xalostoc, Estado de México, Mexico). Analytical standards of deuterated water (deuterium oxide) (D2O), biotin, and riboflavin were purchased from Sigma-Aldrich (Sigma-Aldrich, Co., Santa Clara, CA, USA).

2.2. Batch Cultivation of H. wangnamkhiaoensis

The H. wangnamkhiaoensis yeast strain was obtained from the Industrial Microbiology Laboratory Culture Collection of the National School of Biological Sciences, National Polytechnic Institute (ENCB-IPN), Mexico City, Mexico.
Modified Castañeda-Agulló’s [31] culture medium (10 g/L glucose, 4.85 g/L (NH4)2SO4, 0.625 g/L dibasic ammonium citrate, 1.0114 g/L KH2PO4, 0.275 g/L MgSO4·7H2O, 0.375 g/L Na2CO3, 0.250 g/L NaCl, and 0.02 mg/L biotin) was used for yeast cultivation.
The yeast was batch-cultivated in a bubble column pneumatic bioreactor for 30 h at 28 ± 2 °C, with a sterile air supply of 1.11 vvm. After incubation, the yeast cells were separated through centrifugation (5000 rpm, 10 min), and the supernatant was sterilized with microfiltration using mixed cellulose esters Millipore® (Merck KGaA, Darmstadt, Germany) membranes of 0.22 μm pore size.

2.3. Concentration, Extraction, and Separation of the Fluorescent Compound Produced by H. wangnamkhiaoensis

The membrane-sterilized supernatant (400 mL) was lyophilized to concentrate the fluorescent compound and remove all the water. Subsequently, the lyophilized powder was leached out with 100 mL of HPLC-grade methanol and vortexed for 5 min at ambient temperature (25 ± 1 °C). The resulting mixture was separated through centrifugation (5000 rpm, 10 min). The methanolic extract was collected, and the methanol-insoluble phase was discarded. The methanolic extract was then subjected to HPLC.
Reverse-phase HPLC (RP-HPLC) analysis of the methanolic extract was performed using two isocratic elution methods (Methods 1 and 2) and one linear gradient elution method (Method 3), which are described below:
Method 1: RP-HPLC analysis was performed in an Agilent 1260 Infinity series system (Agilent Technologies, Inc., Santa Clara, CA, USA) equipped with a Zorbax SB-C18 column (250 mm × 4.6 mm, 5 μm particle size) and a diode array detector (DAD). The isocratic mobile phase was 20:80 (v/v) acetonitrile and water at a flow rate of 1 mL/min. Chromatograms were obtained at 280 nm because of the aromatic nature of several known fluorochromes and at 440 nm because a preliminary assay showed that the fluorescent compound-containing supernatant has the highest absorption in the visible region at this wavelength.
The fraction with the longest retention time on the HPLC column, a well-defined absorbance peak with a small width, and/or a high absorption at both 280 and 440 nm was then selected and dried in an oven at 40 °C. One portion of this fraction was used as a feedstock for Method 2, and another portion of the fraction was analyzed spectroscopically.
Method 2: The previously selected fraction from Method 1 was subjected to another analytical RP-HPLC separation using an Agilent 1260 Infinity series system equipped with a cyano (CN) column (150 mm × 4.6 mm, 5 μm particle size Zorbax SB-CN) connected to a DAD. UV–Vis signals were collected at 280 and 440 nm. The isocratic mobile phase was 10:90 (v/v) acetonitrile and formic acid 0.1% (v/v) in water at a flow rate of 1 mL/min. The fraction with the longest retention time on the HPLC column, a narrow absorbance peak, and/or a high absorption at both 280 and 440 nm was selected and dried in an oven at 40 °C for further spectroscopic analysis and characterization.
Method 3: A modified version of the method reported by Odanaka et al. [23] was used. Briefly, 50 mL of the methanolic extract was evaporated to dryness at 35 ± 1 °C, and the resulting powder was dissolved in 10 mL of HPLC-grade water. Next, 3 mL of the resulting solution was loaded in triplicate onto C18 Alltech® Maxi-CleanTM solid phase extraction cartridges (Thermo Fisher Scientific, Inc., Waltham, MA, USA). The cartridges were eluted under vacuum conditions with 1.5 mL of methanol containing 0.002% (v/v) HCl. The eluate was collected and processed by analytical RP-HPLC using an Agilent 1260 Infinity series system equipped with a DAD (UV–Vis signals were traced at 280 and 440 nm) and a phenyl column (150 mm × 4.6 mm, 5 μm particle size, Zorbax SB-Phenyl (Agilent Technologies, Inc., Santa Clara, CA, USA)). The mobile phase was a mixture of acetonitrile and water containing 0.05% (v/v) phosphoric acid at a flow rate of 0.5 mL/min. The column was eluted using a linear gradient from 10 to 100% acetonitrile applied for 15 min. The fraction that exhibited the longest retention time on the HPLC column, a narrow absorbance peak, and/or a high absorption at both 280 and 440 nm was selected and dried in an oven at 40 °C for further spectroscopic analysis and characterization.

2.4. UV–Vis Spectroscopy Analysis of the Selected Fractions from RP-HPLC-DAD Analyses

The UV–Vis absorption spectra of the fractions selected from the three different RP-HPLC methods were obtained with the information provided by DAD and plotted using Agilent OpenLab CDS EZChrom A.0404 software (Agilent Technologies, Inc., Santa Clara, CA, USA).

2.5. The 1D 1H NMR Analysis

The one-dimensional (1D) 1H NMR spectra of the fractions selected from the three different RP-HPLC methods and a riboflavin standard solution were measured at 499.85 MHz using a Varian NMR 500 system (Agilent Technologies, Inc., Santa Clara, CA, USA) operating at 11.7 T equipped with a 5 mm OneNMR probe (Agilent Technologies, Inc., Santa Clara, CA, USA) at 25 °C, using deuterium oxide (D2O) as solvent. The 1D 1H NMR spectra were recorded without spinning using the PRESAT pulse sequence to suppress the residual H2O signal. The acquisition parameters for the 1H NMR observations were: 32k data points (np), 8012.8 Hz spectral width (sw), 2.0047 s acquisition time (at), 3.0 s delay time (d1), and 256 scans (ns). The data were zero-filled to 64k data points before Fourier transformation (FT).

2.6. The 1H NMR Spectra Data Processing

The 1H NMR spectra were processed using MestReNova 14.2.0 software from Mestrelab Research S.L. (Santiago de Compostela, Coruña, Spain).

2.7. Spectrofluorometric Characterization of the Fractions Selected from RP-HPLC Analyses

The excitation and emission spectra of the fractions selected from RP-HPLC-DAD analyses were measured at 24 ± 1 °C using a SpectraMax M3 fluorometer (Molecular Devices LCC, San Jose, CA, USA). A sweep from 450 to 550 nm at an emission wavelength (λem) of 525 nm was used for excitation measurements, and a sweep from 300 to 500 nm at an excitation wavelength (λex) of 450 nm was used for emission measurements. These maximum fluorescence excitation and emission wavelengths were previously optimized to enhance the selectivity and sensitivity of the method [32].

3. Results and Discussion

3.1. RP-HPLC-DAD Analysis of the Supernatant of the H. wangnamkhiaoensis Liquid Culture

The yellow fluorescent compound excreted into the culture supernatant of H. wangnamkhiaoensis was analyzed by three RP-HPLC-DAD methods. Several fractions were obtained from RP-HPLC-DAD Method 1 (Figure 1A). The fraction with the longest retention time (3.8 min), a high absorption at both 280 and 440 nm, and a narrow absorbance peak width (marked with an arrow in Figure 1A) was selected and named FCHw-M1. Similarly, major narrow peaks with retention times of 5.3 min (marked with an arrow in Figure 1B) and 8.2 min (marked with an arrow in Figure 1C) and exhibiting the highest absorption at both 280 and 440 nm were selected from RP-HPLC-DAD Methods 2 and 3 and named FCHw-M2 and FCHw-M3, respectively.

3.2. UV–Vis Characterization of the Fractions Selected from RP-HPLC-DAD Analyses

The UV–Vis absorption spectra of FCHw-M1, FCHw-M2, and FCHw-M3 are shown in Figure 1E–G. All UV–Vis spectra showed comparable profiles and four peaks with similar wavelengths of maximum absorption (λmax). FCHw-M1 and FCHw-M2 exhibited peaks with λmax at 224, 268, 372, and 446 nm (Figure 1E,F), whereas FCHw-M3 showed peaks with similar λmax at 225, 270, 370, and 446 nm (Figure 1G).
Flavins exhibit four characteristic absorption peaks at λmax of approximately 220, 265, 375, and 445 nm [33,34], which are consistent with the spectra of the fluorescent compound produced by H. wangnamkhiaoensis. Furthermore, flavins are pale yellow, water-soluble, fluorescent organic compounds [34,35,36]; such characteristics are also exhibited by the yellow fluorescent compound produced by the yeast strain. Flavins such as riboflavin (RF) and its derivatives, flavin mononucleotide (FMN) and flavin adenine dinucleotide (FAD), are commonly found in all living organisms. These are the most biologically important flavins owing to their essential role in oxidation–reduction (redox) reactions involved in energy production, cellular antioxidant functions, and numerous metabolic pathways, all of which impact human health [34,35,37]. Flavins such as formylmethylflavin (FMF), carboxylmethylflavin (CMF), lumiflavin (LF), and lumichrome (LC) derived from the photolysis and oxidation of RF are found in milk, dairy products, and the culture supernatant of some flavinogenic microorganisms [33,34,38,39,40,41,42,43,44,45,46,47]. LC is biosynthesized by organisms of diverse origins such as the bacterium Nocardia alba [48], the ascidian Halocynthia roretzi [49], and the herbaceous plant Galactites tomentosa Moench [50].
The fluorescent compound excreted by H. wangnamkhiaoensis was successfully separated by the three RP-HPLC-DAD methods used. Similarly, flavins from foods, milk, dairy products, beverages [42,43,50,51,52,53,54,55,56,57,58], plant extracts [59], pharmaceutical products [46,60], culture supernatants of microorganisms [23,41,45], Keroplatus larvae [25] and the ascidian H. roretzi [49] have been previously separated using HPLC methods.
The UV–Vis spectra of most of the above-mentioned flavins were comparable to each other [33,38,40,42,43,44,46,61]. The λmax of RF produced by Bordetella pertussis (225, 270, 370, and 450 nm) [23], dissolved in methanol (270–271, 344–358, and 440–450 nm) [61], dissolved in phosphate and borate buffers (362 and 440 nm in the visible region) [62], and separated from multivitamin tablets (222, 267, 369, and 445 nm) [60] and from pig liver (220, 267, 361, and 462 nm) [43] are nearly identical to those reported for CMF (223, 266, 376, and 445 nm) [63] but slightly different from those reported for LC (225, 258, 353, and 388 nm [41] and 218, 261, 355, and 382 nm [48]). The slight differences in the λmax of RF could be attributed to the solvent effect and hydrogen bonding interactions [33,61].
The fluorescent compound was further characterized by 1H NMR to confirm whether it was a flavin.

3.3. The 1H NMR Characterization of the Fractions Selected from RP-HPLC-DAD Analysis

The 1H NMR spectrum of FCHw-M1 showed a large number of signals (Figure 2A), making it difficult to identify the signals corresponding to the fluorescent compound. These results suggest that RP-HPLC-DAD Method 1 is not suitable for separating the fluorescent compound from the culture supernatant of H. wangnamkhiaoensis. Hence, FCHw-M1 was no longer analyzed further.
In contrast, the 1H NMR spectrum of FCHw-M2 revealed two single signals at δ 2.30 and 2.40 ppm (Figure 2B), which correspond to methyl protons. The other group of signals corresponds to aromatic protons at δ 7.79 and 7.77 (Figure 2B). Similarly, the 1H NMR analysis of FCHw-M3 shows methyl proton signals at δ 2.27 and 2.37 (Figure 2C) and aromatic proton signals at δ 7.76 and 7.75. Moreover, the spectrum of FCHw-M3 exhibited signals of citrate (δ 2.58, 2.61, 2.80, and 2.83) present in the culture medium (Figure 2C), which were not observed for FCHw-M2 (Figure 2B).
The 1H NMR spectroscopy has been previously used to elucidate the structure of flavins, such as FMN [64,65,66,67,68,69,70], FAD [66,67,68,71], LF [72], LC [48,49,50,72], and RF [25,73,74]. The 1H NMR spectral characteristics in this study are consistent with those of RF in deuterated dimethyl sulfoxide (DMSOd6) reported by Malele et al. [73], where methyl proton signals at δ 2.36 and 2.45 and aromatic proton signals at δ 7.84 and 7.78 were observed. The 1H NMR spectrum of an RF reference standard dissolved in a deuterated trifluoroacetic acid/trifluoroacetic anhydride mixture (TFAd:TFAA) showed methyl group protons at δ 2.11 and 2.24 and aromatic proton signals at δ 7.63 and 7.44 [74]. Similarly, the 1H NMR spectrum of RF isolated from K. testaceus larvae and dissolved in deuterated water showed methyl protons at δ 2.5 and 2.6 and aromatic protons at δ 7.93 and 7.96 [25]. These chemical shift variations can be attributed to the solvent used to dissolve RF for 1H NMR spectral analysis. Furthermore, the chelating effect of the citrate ion present in FCHw-M3 can affect the NMR chemical shifts of RF and other chemical compounds [75,76].
The 1H NMR spectra of FMN [64,66,69,70,77], LC [48,49,50,72], and LF [72] are easily distinguishable from the RF spectrum due to their intrinsic structural differences. Compared to RF, FMN has a phosphate group, LC has an additional imino group, LF has an additional methyl group, and both LC and LF do not contain the d-ribityl side chain. The 1H NMR spectrum of the fluorescent compound produced by H. wangnamkhiaoensis is consistent with that of RF and substantially differs from the spectra of the other flavins [25,73,74]. Similarly, the UV–Vis spectra of the fluorescent compound in FCHw-M2 and FCHw-M3 are comparable to those of other flavins, particularly RF [23,33,53,60,61,62]. Therefore, the brilliant yellow fluorescent compound produced by H. wangnamkhiaoensis is confirmed to be RF.
Most yeasts can synthesize RF only to the extent of their own requirements; only a few yeasts overproduce and excrete RF into the fermentation broth [7,8,78,79,80,81,82,83]. Candida, Schizosaccharomyces, Schwanniomyces, Meyerozyma, Pichia, Hyphopichia, and Debaryomyces genera produce RF [7,80,84]. However, to the best of our knowledge, the production of RF by H. wangnamkhiaoensis has not been described thus far.

3.4. RP-HPLC-DAD, UV–Vis, and 1H NMR Analyses of Riboflavin Standard

To support our structural conclusion, we analyzed an RF reference standard (Sigma-Aldrich) via RP-HPLC-DAD Method 3 (Figure 1D), UV–Vis spectroscopy (Figure 1H), and 1H NMR (Figure 2D). A single peak at the retention time of 8.4 min (Figure 1D) was observed, which is similar to that obtained for FCHW-M3 (8.2 min). Furthermore, the UV–Vis spectrum of the RF standard revealed four peaks of maximum absorption at 225, 270, 370, and 446 nm (Figure 1H), which were also observed for FCHw-M2 and FCHw-M3.
The 1H NMR spectrum of the RF standard (Figure 2D) exhibited methyl proton signals at δ 2.30 and 2.40, d-ribityl proton signals (δ 3.50–4.50), and an aromatic proton signal (δ 7.77), which were also observed for FCHW-M2 and FCHW-M3. The chemical shift variations may be ascribed to changes in sample concentration and measurement temperature, as previously reported for the aromatic protons in flavins [66,67,68,71,85].

3.5. Spike-In 1H NMR Experiments

To confirm the identity of the fluorescent compound, spike-in 1H NMR experiments with the RF standard were performed. The RF standard solution was added to the FCHw-M2 and FCHw-M3 1H NMR samples, and the corresponding spectra were recorded. In the 1H NMR spectrum of the unspiked FCHw-M2 sample, a group of signals was observed between δ 1.09 and 1.14, corresponding to an impurity. These signals were used as the internal reference, and their integral was normalized to a value of 100. Consequently, the integrals of the methyl group and aromatic protons were estimated to be 41.36 and 24.67 (Figure 3A). The 1H NMR spectrum of the RF-spiked FCHw-M2 sample showed no new signal; however, the integrals of the methyl (114.06) and aromatic protons (68.10) increased substantially (Figure 3B).
For FCHw-M3, the citrate signal appearing at δ 2.80 was considered the internal reference (Figure 4A). The signals of the methyl group (218.34) and aromatic protons (59.87) increased in the spectrum enriched with the RF standard (239.28 and 65.68, respectively) (Figure 4B). These results confirm that the fluorescent compound produced by the H. wangnamkhiaoensis yeast strain is RF.

3.6. Spectrofluorometric Characterization of the Riboflavin Standard, FCHw-M2, and FCHw-M3

The fluorescence excitation and emission spectra of the RF-containing FCHw-M2 and FCHw-M3 and the RF standard were recorded and compared. All fluorescence excitation and emission spectra showed considerable similarities (Figure 5). The maximum emission wavelength (λem_max) of FCHw-M2, FCHw-M3, and the RF standard was 528 nm. The maximum excitation wavelength (λex_max) of FCHw-M2 was 448 nm, while that for FCHw-M3 and the RF standard was 449 nm. These maximum wavelengths are also in strong agreement with those previously reported for RF of different origins [36,38,42,43,53,86]. These results also support that the yellow fluorescent compound produced by H. wangnamkhiaoensis is RF.
From all the above, the HPLC and 1H NMR analytical techniques used in our work are powerful and robust analytical tools that provide truly reliable, precise, reproducible, and fast results for the separation and identification of riboflavin. However, their main drawbacks are as follows: (1) Expensive equipment is required, (2) high-quality components are needed, (3) the solvents and/or columns are expensive, (4) regular maintenance and calibration are needed, which add extra cost, and (5) sophisticated software is required for data analysis [87,88].

4. Conclusions

A brilliant yellow fluorescent compound excreted into the culture supernatant of H. wangnamkhiaoensis was isolated and identified for the first time in the literature. The culture supernatants of H. wangnamkhiaoensis were analyzed using three different RP-HPLC-DAD methods, and the desired compound was successfully separated using two out of the three RP-HPLC-DAD methods. The maximum UV–Vis absorption wavelengths (~224, 268, 372, and 448 nm) and the 1H NMR signals of methyl groups (δ 2.30 and 2.40) and aromatic protons (δ 7.79 and 7.77) revealed that the fluorescent compound is riboflavin. The identity of this compound was further confirmed by spiking the 1H NMR spectra with riboflavin and spectrofluorometric measurements. This work proposes simple, modern, fast, precise, reliable, sensitive, and reproducible methods for separating and identifying riboflavin, an essential vitamin for overall good health. Furthermore, it also broadens the spectrum of riboflavin-overproducing yeasts, opens new possibilities and perspectives for practical applications of riboflavin production, and triggers new innovative actions. Studies on riboflavin production by batch and single-stage steady-state continuous cultures of the novel H. wangnamkhiaoensis yeast strain in a bubble column pneumatic bioreactor are in progress.

Author Contributions

Conceptualization, E.C.-U., G.M.C.-C. and L.G.Z.-V.; methodology, R.A.J.-N., L.G.Z.-V. and F.S.-T.; software, R.A.J.-N.; validation, R.A.J.-N.; formal analysis, R.A.J.-N., E.C.-U., G.M.C.-C., L.G.Z.-V. and F.S.-T.; investigation, R.A.J.-N.; resources, E.C.-U., G.M.C.-C. and L.G.Z.-V.; data curation, R.A.J.-N.; writing—original draft preparation, R.A.J.-N., E.C.-U. and L.G.Z.-V.; visualization, R.A.J.-N.; supervision, E.C.-U., G.M.C.-C. and L.G.Z.-V.; project administration, E.C.-U. and G.M.C.-C.; funding acquisition, E.C.-U. and G.M.C.-C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Secretaría de Investigación y Posgrado del Instituto Politécnico Nacional, grant numbers: 20220923, 20220889, 20231441, and 20232093.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All relevant data are within the paper.

Acknowledgments

CONAHCYT (Consejo Nacional de Humanidades, Ciencias y Tecnologías) awarded a graduate scholarship to the coauthor R.A.J.-N. G.M.C.-C., E.C.-U., L.G.Z.-V. and F.S.-T. received research grants from the EDI-IPN, COFAA-IPN, and SNI-CONAHCYT.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kasten, F.H. Introduction to fluorescent probes: Properties, history and applications. In Fluorescent and Luminescent Probes for Biological Activity, 2nd ed.; Mason, W., Ed.; Academic Press: Cambridge, MA, USA, 1999; pp. 17–39. [Google Scholar] [CrossRef]
  2. Kim, J.; Oh, J.H.; Kim, D. Recent advances in single-benzene-based fluorophores: Physicochemical properties and applications. Org. Biomol. Chem. 2021, 19, 933–946. [Google Scholar] [CrossRef]
  3. Marcu, L.; French, P.M.W.; Elson, D.S. Fluorescence Lifetime Spectroscopy and Imaging, 1st ed.; Marcu, L., French, P.M.W., Elson, D.S., Eds.; CRC Press: Boca Raton, FL, USA, 2014. [Google Scholar]
  4. Sugden, J.K. Photochemistry of dyes and fluorochromes used in biology and medicine: Some physicochemical background and current applications. Biotech. Histochem. 2004, 79, 71–90. [Google Scholar] [CrossRef] [PubMed]
  5. Lakowicz, J.R. Principles of Fluorescence Spectroscopy, 3rd ed.; Lakowicz, J.R., Ed.; Springer: New York, NY, USA, 2006. [Google Scholar] [CrossRef]
  6. Croce, A.C.; Bottiroli, G. Autofluorescence spectroscopy and imaging: A tool for biomedical research and diagnosis. Eur. J. Histochem. 2014, 58, 2461. [Google Scholar] [CrossRef]
  7. Averianova, L.A.; Balabanova, L.A.; Son, O.M.; Podvolotskaya, A.B.; Tekutyeva, L.A. Production of vitamin B2 (riboflavin) by microorganisms: An overview. Front. Bioeng. Biotechnol. 2020, 8, 570828. [Google Scholar] [CrossRef]
  8. Liu, S.; Hu, W.; Wang, Z.; Chen, T. Production of riboflavin and related cofactors by biotechnological processes. Microb. Cell Fact. 2020, 19, 31. [Google Scholar] [CrossRef]
  9. Singh, R.V.; Sambyal, K. An overview of β-carotene production: Current status and future prospects. Food Biosci. 2022, 47, 101717. [Google Scholar] [CrossRef]
  10. Wang, L.; Liu, Z.; Jiang, H.; Mao, X. Biotechnology advances in β-carotene production by microorganisms. Trends Food Sci. Technol. 2021, 111, 322–332. [Google Scholar] [CrossRef]
  11. Andreieva, Y.; Lyzak, O.; Liu, W.; Kang, Y.; Dmytruk, K.; Sibirny, A. SEF1 and VMA1 genes regulate riboflavin biosynthesis in the flavinogenic yeast Candida famata. Cytol. Genet. 2020, 54, 379–385. [Google Scholar] [CrossRef]
  12. Chávez-Camarillo, G.M.; Lopez-Nuñez, P.V.; Jiménez-Nava, R.A.; Aranda-García, E.; Cristiani-Urbina, E. Production of extracellular α-amylase by single-stage steady-state continuous cultures of Candida wangnamkhiaoensis in an airlift bioreactor. PLoS ONE 2022, 17, e0264734. [Google Scholar] [CrossRef] [PubMed]
  13. Moliné, 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] [CrossRef]
  14. Petrik, S.; Marova, I.; Haronikova, A.; Kostovova, I.; Breierova, E. Production of biomass, carotenoid and other lipid metabolites by several red yeast strains cultivated on waste glycerol from biofuel production—A comparative screening study. Ann. Microbiol. 2013, 63, 1537–1551. [Google Scholar] [CrossRef]
  15. Sun, Z.J.; Lian, J.Z.; Zhu, L.; Jiang, Y.Q.; Li, G.S.; Xue, H.L.; Wu, M.B.; Yang, L.R.; Lin, J.P. Combined biosynthetic pathway engineering and storage pool expansion for high-level production of ergosterol in industrial Saccharomyces cerevisiae. Front. Bioeng. Biotechnol. 2021, 9, 681666. [Google Scholar] [CrossRef] [PubMed]
  16. Andreieva, Y.; Petrovska, Y.; Lyzak, O.; Liu, W.; Kang, Y.; Dmytruk, K.; Sibirny, A. Role of the regulatory genes SEF1, VMA1 and SFU1 in riboflavin synthesis in the flavinogenic yeast Candida famata (Candida flareri). Yeast 2020, 37, 497–504. [Google Scholar] [CrossRef]
  17. Lyzak, O.O.; Ledesma-Amaro, R.; Dmytruk, K.V.; Sibirny, A.A.; Revuelta, J.L. Molecular studies of the flavinogenic fungus Ashbya gossypii and the flavinogenic yeast Candida famata. In Biotechnology of Yeasts and Filamentous Fungi; Springer International Publishing: Berlin/Heidelberg, Germany, 2017; pp. 281–296. [Google Scholar] [CrossRef]
  18. Petrovska, Y.; Lyzak, O.; Dmytruk, K.; Sibirny, A. Effect of gene SFU1 on riboflavin synthesis in flavinogenic yeast Candida famata. Cytol. Genet. 2020, 54, 408–412. [Google Scholar] [CrossRef]
  19. Prokopiv, T.M.; Fedorovych, D.V.; Boretsky, Y.R.; Sibirny, A.A. Oversynthesis of riboflavin in the yeast Pichia guilliermondii is accompanied by reduced catalase and superoxide dismutases activities. Curr. Microbiol. 2013, 66, 79–87. [Google Scholar] [CrossRef] [PubMed]
  20. Tsyrulnyk, A.O.; Fedorovych, D.V.; Dmytruk, K.V.; Sibirny, A.A. Overexpression of riboflavin excretase enhances riboflavin production in the yeast Candida famata. Methods Mol. Biol. 2021, 2280, 31–42. [Google Scholar] [CrossRef]
  21. Sommer, K.; Hillinger, M.; Eigenmann, A.; Vetter, W. Characterization of various isomeric photoproducts of ergosterol and vitamin D2 generated by UV irradiation. Eur. Food. Res. Technol. 2023, 249, 713–726. [Google Scholar] [CrossRef]
  22. Somai, B.M.; Belewa, V.; Frost, C. Tulbaghia violacea (Harv) exerts its antifungal activity by reducing ergosterol production in Aspergillus flavus. Curr. Microbiol. 2021, 78, 2989–2997. [Google Scholar] [CrossRef]
  23. Odanaka, K.; Iwatsuki, M.; Satho, T.; Watanabe, M. Identification and characterization of a brilliant yellow pigment produced by Bordetella pertussis. Microbiol. Immunol. 2017, 61, 490–496. [Google Scholar] [CrossRef]
  24. Mitrofanov, D.A.; Nazarov, G.V.; Babkin, I.Y.; Galan, S.E.; Goncharov, V.M.; Vostrukhov, S.V. Using HPLC with mass-spectrometric detection for riboflavin determination in complex medicinal forms. Pharm. Chem. J. 2009, 43, 176–179. [Google Scholar] [CrossRef]
  25. Kotlobay, A.A.; Dubinnyi, M.A.; Polevoi, A.V.; Kovalchuk, S.I.; Kaskova, Z.M. Riboflavin as one of possible components of Keroplatus (Insecta: Diptera: Keroplatidae) fungus gnat bioluminescence. Russ. J. Bioorg. Chem. 2022, 48, 1215–1220. [Google Scholar] [CrossRef]
  26. Maswanna, T.; Maneeruttanarungroj, C. Identification of major carotenoids from green alga Tetraspora sp. CU2551: Partial purification and characterization of lutein, canthaxanthin, neochrome, and β-carotene. World J. Microbiol. Biotechnol. 2022, 38, 129. [Google Scholar] [CrossRef] [PubMed]
  27. Chávez-Camarillo, G.M.; Santiago-Flores, U.M.; Mena-Vivanco, A.; Morales-Barrera, L.; Cortés-Acosta, E.; Cristiani-Urbina, E. Transient responses of Wickerhamia sp. yeast continuous cultures to qualitative changes in carbon source supply: Induction and catabolite repression of α-amylase synthesis. Ann. Microbiol. 2018, 68, 625–635. [Google Scholar] [CrossRef]
  28. Hernández-Montañez, Z.F.; Juárez-Montiel, M.; Velázquez-Ávila, M.; Cristiani-Urbina, E.; Hernández-Rodríguez, C.; Villa-Tanaca, L.; Chávez-Camarillo, G. Production and characterization of extracellular α-amylase produced by Wickerhamia sp. X-Fep. Appl. Biochem. Biotechnol. 2012, 167, 2117–2129. [Google Scholar] [CrossRef] [PubMed]
  29. Hossain, T.; Miah, A.B.; Mahmud, S.A.; Al Mahin, A.A. Enhanced bioethanol production from potato peel waste via consolidated bioprocessing with statistically optimized medium. Appl. Biochem. Biotechnol. 2018, 186, 425–442. [Google Scholar] [CrossRef]
  30. Pérez-Rodríguez, A.; Flores-Ortiz, C.M.; Chávez-Camarillo, G.M.; Cristiani-Urbina, E.; Morales-Barrera, L. Potential capacity of Candida wangnamkhiaoensis to produce oleic acid. Fermentation 2023, 9, 443. [Google Scholar] [CrossRef]
  31. Castañeda-Agulló, M. Studies on the biosynthesis of extracellular proteases by bacteria. I. Serratia marcescens, synthetic and gelatin media. J. Gen. Physiol. 1956, 39, 369–375. [Google Scholar] [CrossRef]
  32. Linares-Martínez, L. Producción y Caracterización Parcial de un Compuesto Fluorescente de Candida wangnamkhiaoensis. Bachelor’s Thesis, Escuela Nacional de Ciencias Biológicas, Mexico City, Mexico, 2019. [Google Scholar]
  33. Ahmad, I.; Vaid, F.H.M. Photochemistry of flavins in aqueous and organic solvents. In Flavins; Silva, E., Edwards, A.M., Eds.; RSC Publishing: Cambridge, UK, 2006; Volume 6. [Google Scholar] [CrossRef]
  34. Edwards, A.M. Structure and general properties of flavins. In Flavins and Flavoproteins: Methods in Molecular Biology; Weber, S., Schleicher, E., Eds.; Humana Press: Totowa, MJ, USA, 2014; Volume 1146. [Google Scholar] [CrossRef]
  35. Gadda, G. Flavins. In Encyclopedia of Biophysics; Roberts, G.C.K., Ed.; Springer: Berlin/Heidelberg, Germany, 2013; pp. 771–775. [Google Scholar] [CrossRef]
  36. Weimar, W.R.; Neims, A.H. Physical and chemical properties of flavins; binding of flavins to protein and conformational effects; biosynthesis of riboflavin. In Riboflavin; Rivlin, R.S., Ed.; Springer: New York, NY, USA, 1975; pp. 1–47. [Google Scholar] [CrossRef]
  37. Pedrolli, D.B.; Jankowitsch, F.; Schwarz, J.; Langer, S.; Nakanishi, S.; Mack, M. Natural riboflavin analogs. In Flavins and Flavoproteins: Methods in Molecular Biology; Weber, S., Schleicher, E., Eds.; Humana Press: Totowa, MJ, USA, 2014; Volume 1146. [Google Scholar] [CrossRef]
  38. Ahmad, I.; Anwar, Z.; Sheraz, M.A.; Ahmed, S.; Khattak, S.U. Stability-indicating spectrofluorimetric method for the assay of riboflavin and photoproducts: Kinetic applications. Luminescence 2018, 33, 1070–1080. [Google Scholar] [CrossRef]
  39. Ahmad, I.; Mirza, T.; Anwar, Z.; Ejaz, M.A.; Sheraz, M.A.; Ahmed, S. Multicomponent spectrofluorimetric method for the assay of formylmethylflavin and its hydrolytic products: Kinetic applications. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2018, 205, 540–550. [Google Scholar] [CrossRef]
  40. Ball, G.F.M. Water-Soluble Vitamin Assays in Human Nutrition; Springer: New York, NY, USA, 1994. [Google Scholar] [CrossRef]
  41. Bretzel, W.; Schurter, W.; Ludwig, B.; Kupfer, E.; Doswald, S.; Pfister, M.; Van Loon, A.P.G.M. Commercial riboflavin production by recombinant Bacillus subtilis: Down-stream processing and comparison of the composition of riboflavin produced by fermentation or chemical synthesis. J. Ind. Microbiol. Biotechnol. 1999, 22, 19–26. [Google Scholar] [CrossRef]
  42. Gliszczyńska, A.; Koziołowa, A. Chromatographic identification of a new flavin derivative in plain yogurt. J. Agric. Food Chem. 1999, 47, 3197–3201. [Google Scholar] [CrossRef] [PubMed]
  43. Gliszczyńska-Świgło, A.; Koziołowa, A. Chromatographic determination of riboflavin and its derivatives in food. J. Chromatogr. A 2000, 881, 285–297. [Google Scholar] [CrossRef] [PubMed]
  44. Mirza, T.; Anwar, Z.; Ejaz, M.A.; Ahmed, S.; Sheraz, M.A.; Ahmad, I. Multicomponent spectrofluorimetric method for the assay of carboxymethylflavin and its hydrolytic products: Kinetic applications. Luminescence 2018, 33, 1314–1325. [Google Scholar] [CrossRef] [PubMed]
  45. Russo, P.; Capozzi, V.; Arena, M.P.; Spadaccino, G.; Dueñas, M.T.; López, P.; Fiocco, D.; Spano, G. Riboflavin-overproducing strains of Lactobacillus fermentum for riboflavin-enriched bread. Appl. Microbiol. Biotechnol. 2014, 98, 3691–3700. [Google Scholar] [CrossRef] [PubMed]
  46. Trang, H.K. Development of HPLC Methods for the Determination of Water-Soluble Vitamins in Pharmaceuticals and Fortified Food Products. Master’s Thesis, Clemson University, Clemson, SC, USA, 2013. Available online: https://tigerprints.clemson.edu/all_theses/1745 (accessed on 22 May 2023).
  47. Wold, J.P.; Jørgensen, K.; Lundby, F. Nondestructive measurement of light-induced oxidation in dairy products by fluorescence spectroscopy and imaging. J. Dairy Sci. 2002, 85, 1693–1704. [Google Scholar] [CrossRef] [PubMed]
  48. Ding, Z.G.; Zhao, J.Y.; Yang, P.W.; Li, M.G.; Huang, R.; Cui, X.L.; Wen, M.L. 1H and 13C NMR assignments of eight nitrogen containing compounds from Nocardia alba sp. nov. (YIM 30243T). Magn. Reson. Chem. 2009, 47, 366–370. [Google Scholar] [CrossRef] [PubMed]
  49. Tsukamoto, S.; Kato, H.; Hirota, H.; Fusetani, N. Lumichrome. A larval metamorphosis-inducing substance in the ascidian Halocynthia roretzi. Eur. J. Biochem. 1999, 264, 785–789. [Google Scholar] [CrossRef]
  50. Tuberoso, C.I.G.; Bifulco, E.; Caboni, P.; Sarais, G.; Cottiglia, F.; Floris, I. Lumichrome and phenyllactic acid as chemical markers of thistle (Galactites tomentosa Moench) honey. J. Agric. Food Chem. 2011, 59, 364–369. [Google Scholar] [CrossRef]
  51. Ashoor, S.H.; Seperich, G.J.; Monte, W.C.; Welty, J. HPLC determination of riboflavin in eggs and dairy products. J. Food 1983, 48, 92–94. [Google Scholar] [CrossRef]
  52. Fracassetti, D.; Limbo, S.; D’Incecco, P.; Tirelli, A.; Pellegrino, L. Development of a HPLC method for the simultaneous analysis of riboflavin and other flavin compounds in liquid milk and milk products. Eur. Food Res. Technol. 2018, 244, 1545–1554. [Google Scholar] [CrossRef]
  53. Gliszczyńska-Świgło, A.; Rybicka, I. Simultaneous determination of caffeine and water-soluble vitamins in energy drinks by HPLC with photodiode array and fluorescence detection. Food Anal. Methods 2015, 8, 139–146. [Google Scholar] [CrossRef]
  54. Jakobsen, J. Optimisation of the determination of thiamin, 2-(1-hydroxyethyl)thiamin, and riboflavin in food samples by use of HPLC. Food Chem. 2008, 106, 1209–1217. [Google Scholar] [CrossRef]
  55. Johnsson, H.; Branzell, C. High performance liquid chromatographic determination of riboflavin in food—A comparison with a microbiological method. Int. J. Vitam. Nutr. Res. 1987, 57, 53–58. [Google Scholar]
  56. Russell, L.F.; Vanderslice, J.T. Non-degradative extraction and simultaneous quantitation of riboflavin, flavin mononucleotide, and flavin adenine dinucleotide in foods by HPLC. Food Chem. 1992, 43, 151–162. [Google Scholar] [CrossRef]
  57. Stancher, B.; Zonta, F. High performance liquid chromatographic analysis of riboflavin (vitamin B2) with visible absorbance detection in Italian cheeses. J. Food Sci. 1986, 51, 857–858. [Google Scholar] [CrossRef]
  58. Torres-Sequeiros, R.A.; Garda-Falcón, M.S.; Sirnai-Gandara, J. Analysis of fluorescent vitamins riboflavin and pyridoxine in beverages with added vitamins. Chromatographia 2001, 53, 236–239. [Google Scholar] [CrossRef]
  59. Sunarić, S.; Pavlović, D.; Stanković, M.; Živković, J.; Arsić, I. Riboflavin and thiamine content in extracts of wild-grown plants for medicinal and cosmetic use. Chem. Pap. 2020, 74, 1729–1738. [Google Scholar] [CrossRef]
  60. Garmonov, S.Y.; Salakhov, I.A.; Nurislamova, G.R.; Ismailova, R.N.; Irtuganova, É.A.; Sopin, V.F. Assay of ascorbic acid, thiamine, riboflavin, nicotinamide, and pyridoxine in “hexavit” by HPLC. Pharm. Chem. J. 2011, 45, 440–443. [Google Scholar] [CrossRef]
  61. Ahmad, I.; Anwar, Z.; Ahmed, S.; Sheraz, M.A.; Bano, R.; Hafeez, A. Solvent effect on the photolysis of riboflavin. AAPS PharmSciTech 2015, 16, 1122–1128. [Google Scholar] [CrossRef] [PubMed]
  62. Bartzatt, R.; Wol, T. Detection and assay of vitamin B-2 (riboflavin) in alkaline borate buffer with UV/Visible spectrophotometry. Int. Sch. Res. Not. 2014, 2014, 453085. [Google Scholar] [CrossRef]
  63. Ahmad, I.; Mirza, T.; Musharraf, S.G.; Anwar, Z.; Sheraz, M.A.; Ahmed, S.; Ejaz, M.A.; Khurshid, A. Photolysis of carboxymethylflavin in aqueous and organic solvent: A kinetic study. RSC Adv. 2019, 9, 26559–26571. [Google Scholar] [CrossRef] [PubMed]
  64. Bastian, M.; Sigel, H. The self-association of flavin mononucleotide (FMN2−) as determined by 1H NMR shift measurements. Biophys. Chem. 1997, 67, 27–34. [Google Scholar] [CrossRef] [PubMed]
  65. Bullock, F.J.; Jardetzky, O. An experimental demonstration of the nuclear magnetic resonance assignments in the 6,7-dimethylisoalloxazine nucleus*. J. Org. Chem. 1965, 30, 2056–2057. [Google Scholar] [CrossRef]
  66. Kainosho, M.; Kyogoku, Y. High-resolution proton and phosphorus nuclear magnetic resonance spectra of flavin-adenine dinucleotide and its conformation in aqueous solution. Biochemistry 1972, 11, 741–752. [Google Scholar] [CrossRef]
  67. Kotowycz, G.; Teng, N.; Klein, M.P.; Calvin, M. The 220 MHz nuclear magnetic resonance study of a solvent-induced conformational change in flavin adenine dinucleotide. J. Biol. Chem. 1969, 244, 5656–5662. [Google Scholar] [CrossRef]
  68. Sarma, R.H.; Dannies, P.; Kaplan, N.O. Investigations of inter- and intramolecular interactions in flavine-adenine dinucleotide by proton magnetic resonance. Biochemistry 1968, 7, 4359–4367. [Google Scholar] [CrossRef]
  69. Tachibana, S.; Murakami, T.; Ninomiya, T. Identification of the chemical structures of schizoflavins as 7,8-dimethyl-10-(2,3,4-trihydroxy-4-formylbutyl)isoalloxazine and 7,8-dimethyl-10-(2,3,4-trihydroxy-4-carboxybutyl)isoalloxazine. J. Nutr. Sci. Vitaminol. 1975, 21, 347–353. [Google Scholar] [CrossRef]
  70. Tachibana, S.; Murakami, T. Isolation and identification of schizoflavins. Methods Enzymol. 1980, 66, 333–338. [Google Scholar] [CrossRef] [PubMed]
  71. Raszka, M.; Kaplan, N.O. Intramolecular hydrogen bonding in flavin adenine dinucleotide. Proc. Natl. Acad. Sci. USA 1974, 71, 4546–4550. [Google Scholar] [CrossRef]
  72. Grande, H.J.; Van Schagen, C.G.; Jarbandhan, T.; Müller, F. An 1H-NMR. Spectroscopic study of alloxazines and isoalloxazines. Helv. Chim. Acta 1977, 60, 348–366. [Google Scholar] [CrossRef]
  73. Malele, C.N.; Ray, J.; Jones, W.E. Synthesis, characterization and spectroscopic study of riboflavin–molybdenum complex. Polyhedron 2010, 29, 749–756. [Google Scholar] [CrossRef]
  74. Pluta, P.L.; Crespi, H.L.; Klein, M.; Blake, M.I.; Studier, M.H.; Katz, J.J. Biosynthesis of deuterated riboflavin: Structure determination by NMR and mass spectrometry. J. Pharm. Sci. 1976, 65, 362–366. [Google Scholar] [CrossRef] [PubMed]
  75. Lindon, J.C.; Nicholson, J.K.; Everett, J.R. NMR spectroscopy of biofluids. Annu. Rep. NMR Spectrosc. 1999, 38, 1–88. [Google Scholar] [CrossRef]
  76. Tredwell, G.D.; Bundy, J.G.; De Iorio, M.; Ebbels, T.M.D. Modelling the acid/base 1H NMR chemical shift limits of metabolites in human urine. Metabolomics 2016, 12, 152. [Google Scholar] [CrossRef]
  77. Fan, P.; Suri, A.K.; Fiala, R.; Live, D.; Patel, D.J. Molecular recognition in the FMN–RNA aptamer complex. J. Mol. Biol. 1996, 258, 480–500. [Google Scholar] [CrossRef]
  78. Abbas, C.A.; Sibirny, A.A. Genetic control of biosynthesis and transport of riboflavin and flavin nucleotides and construction of robust biotechnological producers. Microbiol. Mol. Biol. Rev. 2011, 75, 321–360. [Google Scholar] [CrossRef]
  79. Demain, A.L.; Phaff, H.J.; Kurtzman, C.P. The industrial and agricultural significance of yeasts. In The Yeasts, 4th ed.; Kurtzman, C.P., Fell, J.W., Eds.; Elsevier: Amsterdam, The Netherlands, 1998; pp. 13–19. [Google Scholar] [CrossRef]
  80. Fedorovych, D.; Kszeminska, H.; Babjak, L.; Kaszycki, P.; Kołoczek, H. Hexavalent chromium stimulation of riboflavin synthesis in flavinogenic yeast. BioMetals 2001, 14, 23–31. [Google Scholar] [CrossRef]
  81. Hohmann, H.-P.; Stahmann, K.-P. Biotechnology of riboflavin production. In Comprehensive Natural Products II; Elsevier: Amsterdam, The Netherlands, 2010; pp. 115–139. [Google Scholar] [CrossRef]
  82. Lim, S.H.; Choi, J.S.; Park, E.Y. Microbial production of riboflavin using riboflavin overproducers, Ashbya gossypii, Bacillus subtilis, and Candida famate: An overview. Biotechnol. Bioprocess Eng. 2001, 6, 75–88. [Google Scholar] [CrossRef]
  83. Schwechheimer, S.K.; Becker, J.; Peyriga, L.; Portais, J.C.; Wittmann, C. Metabolic flux analysis in Ashbya gossypii using 13C-labeled yeast extract: Industrial riboflavin production under complex nutrient conditions. Microb. Cell Fact. 2018, 17, 162. [Google Scholar] [CrossRef]
  84. Kurtzman, C.P. New species and a new combination in the Hyphopichia and Yarrowia yeast clades. Antonie Leeuwenhoek 2005, 88, 121–130. [Google Scholar] [CrossRef]
  85. Müller, F. NMR spectroscopy on flavins and flavoproteins. In Flavins and Flavoproteins: Methods in Molecular Biology; Weber, S., Schleicher, E., Eds.; Humana Press: Totowa, MJ, USA, 2014; Volume 1146. [Google Scholar] [CrossRef]
  86. Pinto, J.T.; Rivlin, R.S. Riboflavin (vitamin B2). In Handbook of Vitamins, 5th ed.; Zempleni, J., Suttie, J.W., Gregory, J.F., III, Stover, P.J., Eds.; CRC Press: Boca Raton, FL, USA, 2013; pp. 191–266. [Google Scholar]
  87. Timchenko, Y.V. Advantages and disadvantages of High-Performance Liquid Chromatography (HPLC). J. Environ. Anal. Chem. 2021, 8, 335. [Google Scholar]
  88. Chen, D.; Wang, Z.; Guo, D.; Orekhov, V.; Qu, X. Review and prospect: Deep learning in Nuclear Magnetic Resonance Spectroscopy. Chem. Eur. J. 2020, 26, 10391–10401. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Reverse-phase high-performance liquid chromatography-diode array detector (RP-HPLC-DAD) chromatograms of (A) FCHw-M1, (B) FCHw-M2, (C) FCHw-M3, and (D) riboflavin standard (green plot: 280 nm signal; blue plot: 440 nm signal); UV–Vis spectra of (E) FCHw-M1, (F) FCHw-M2, (G) FCHw-M3, and (H) riboflavin standard. The arrow marks in (AD) show the selected HPLC fractions whose UV–Vis spectra are displayed in (EH).
Figure 1. Reverse-phase high-performance liquid chromatography-diode array detector (RP-HPLC-DAD) chromatograms of (A) FCHw-M1, (B) FCHw-M2, (C) FCHw-M3, and (D) riboflavin standard (green plot: 280 nm signal; blue plot: 440 nm signal); UV–Vis spectra of (E) FCHw-M1, (F) FCHw-M2, (G) FCHw-M3, and (H) riboflavin standard. The arrow marks in (AD) show the selected HPLC fractions whose UV–Vis spectra are displayed in (EH).
Biomolecules 13 01423 g001
Figure 2. The 1H NMR (500 MHz, D2O) spectra of (A) FCHw-M1, (B) FCHw-M2, (C) FCHw-M3, and (D) riboflavin standard.
Figure 2. The 1H NMR (500 MHz, D2O) spectra of (A) FCHw-M1, (B) FCHw-M2, (C) FCHw-M3, and (D) riboflavin standard.
Biomolecules 13 01423 g002
Figure 3. The 1H NMR spectra of (A) FCHw-M2 and (B) FCHw-M2 spiked with riboflavin.
Figure 3. The 1H NMR spectra of (A) FCHw-M2 and (B) FCHw-M2 spiked with riboflavin.
Biomolecules 13 01423 g003
Figure 4. The 1H NMR spectra of (A) FCHw-M3 and (B) FCHw-M3 spiked with riboflavin.
Figure 4. The 1H NMR spectra of (A) FCHw-M3 and (B) FCHw-M3 spiked with riboflavin.
Biomolecules 13 01423 g004
Figure 5. Fluorescence excitation (blue plots) and emission (green plots) spectra of the riboflavin standard (–), FCHw-M2 (---), and FCHw-M3 (···).
Figure 5. Fluorescence excitation (blue plots) and emission (green plots) spectra of the riboflavin standard (–), FCHw-M2 (---), and FCHw-M3 (···).
Biomolecules 13 01423 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Jiménez-Nava, R.A.; Zepeda-Vallejo, L.G.; Santoyo-Tepole, F.; Chávez-Camarillo, G.M.; Cristiani-Urbina, E. RP-HPLC Separation and 1H NMR Identification of a Yellow Fluorescent Compound—Riboflavin (Vitamin B2)—Produced by the Yeast Hyphopichia wangnamkhiaoensis. Biomolecules 2023, 13, 1423. https://doi.org/10.3390/biom13091423

AMA Style

Jiménez-Nava RA, Zepeda-Vallejo LG, Santoyo-Tepole F, Chávez-Camarillo GM, Cristiani-Urbina E. RP-HPLC Separation and 1H NMR Identification of a Yellow Fluorescent Compound—Riboflavin (Vitamin B2)—Produced by the Yeast Hyphopichia wangnamkhiaoensis. Biomolecules. 2023; 13(9):1423. https://doi.org/10.3390/biom13091423

Chicago/Turabian Style

Jiménez-Nava, Raziel Arturo, Luis Gerardo Zepeda-Vallejo, Fortunata Santoyo-Tepole, Griselda Ma. Chávez-Camarillo, and Eliseo Cristiani-Urbina. 2023. "RP-HPLC Separation and 1H NMR Identification of a Yellow Fluorescent Compound—Riboflavin (Vitamin B2)—Produced by the Yeast Hyphopichia wangnamkhiaoensis" Biomolecules 13, no. 9: 1423. https://doi.org/10.3390/biom13091423

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