Rice Bran Protein as a Potent Source of ... - ACS Publications

Sep 20, 2016 - Masayuki Taniguchi , Mitsuhiro Kameda , Toshiki Namae , Akihito Ochiai , Eiichi Saitoh , Takaaki Tanaka. Journal of Functional Foods 20...
1 downloads 0 Views 4MB Size
Article pubs.acs.org/jnp

Rice Bran Protein as a Potent Source of Antimelanogenic Peptides with Tyrosinase Inhibitory Activity Akihito Ochiai,*,†,‡ Seiya Tanaka,‡ Takaaki Tanaka,†,‡ and Masayuki Taniguchi†,‡ †

Department of Materials Science and Technology, Faculty of Engineering, and ‡Graduate School of Science and Technology, Niigata University, Niigata 950-2181, Japan S Supporting Information *

ABSTRACT: Rice (Oryza sativa) is consumed as a staple food globally, and rice bran, the byproduct, is an unused biomass that is ultimately discarded as waste. Thus, in the present study, a technique for producing tyrosinase inhibitory peptides from rice bran protein (RBP) was developed. Simultaneous treatment of RBP with chymotrypsin and trypsin produced numerous peptides. Subsequently, six tyrosinase inhibitory peptides were isolated from the hydrolysate fractions in a multistep purification protocol, and their amino acid sequences were determined. Three of these peptides had a C-terminal tyrosine residue and exhibited significant inhibitory effects against tyrosinase-mediated monophenolase reactions. Furthermore, peptide CT-2 (Leu−Gln− Pro−Ser−His−Tyr) potently inhibited melanogenesis in mouse B16 melanoma cells without causing cytotoxicity, suggesting the potential of CT-2 as an agent for melanin-related skin disorder treatment. The present data indicate that RBP is a potent source of tyrosinase inhibitory peptides and that simultaneous treatment of RBP with chymotrypsin and trypsin efficiently produces these peptides.

R

ice (Oryza sativa) is consumed as a staple food globally, with an estimated production of 645 million tons per year. The demand for rice in Asia is predicted to increase by 70% over the next 30 years, driven primarily by population growth.1 Whole rice grains comprise 35% outside husk (hull), 60% inner starchy endosperm, and 5% bran layer over the endosperm.2 Most of the rice bran is produced by postproduction milling and is ultimately discarded as waste. However, rice bran contains protein, fiber, minerals, and physiologically active phytochemicals such as γ-oryzanol, ferulic acid, and tocopherols.1 Therefore, the utilization of rice bran biomass warrants investigation. Human skin is repeatedly exposed to environmental DNAdamaging agents, such as ultraviolet radiation (UVR),3,4 and therefore requires numerous endogenous protective systems.5,6 Among these, skin pigmentation follows melanin deposition in the expression epidermis, and among melanin subtypes, eumelanin and pheomelanin are induced by UVR to provide a physical barrier that scatters UVR and an absorbent filter that limits UVR penetration. Whereas pheomelanin is a reddish pigment with weak UVR absorption capacity, eumelanin is a black or brown pigment and is the predominant component of UVR protective systems.7 Although skin pigmentation is indispensable for human health,6,8 excess deposition of melanin can lead to skin disorders and diseases, such as freckles and melanoderma, respectively.9 Tyrosinase (EC 1.14.18.1) is the rate-limiting enzyme in melanin biosynthesis and catalyzes the hydroxylation of L-tyrosine to 3,4-dihydroxy-L-phenylalanine (L© 2016 American Chemical Society and American Society of Pharmacognosy

dopa) and subsequent oxidation of L-dopa to L-dopaquinone.10,11 These steps are defined as monophenolase and diphenolase reactions, respectively, and the L-dopaquinone generated by the diphenolase reaction is subsequently converted to eumelanin through a series of nonenzymatic processes.7 Accordingly, because tyrosinase activity is critical during melanin biosynthesis, several tyrosinase inhibitors have been proposed as therapeutic agents for melanin-related skin disorders.12,13 Among these, kojic acid, arbutin, and hydroquinone are currently used in cosmetics and produce strong therapeutic effects, but they also have serious side effects. In particular, kojic acid and hydroquinone are well-known genotoxic and carcinogenic compounds.14−16 Moreover, the glucosylated hydroquinone arbutin, which is extracted from wheat, pear skins, and blueberry, is readily converted on the skin surface to harmful hydroquinone.17 Therefore, safe and stable cosmetic agents that control melanin production are strongly desired. Several tyrosinase inhibitory peptides have been identified over the past 10 years, and their safety has been emphasized in terms of the absence of inherent toxicity.18−21 However, only the peptide P4 has been used as a therapeutic agent for skin disorders.22 Previously, we showed that the decapeptide TH10 exhibits strong inhibitory activity against the monophenolase Received: May 17, 2016 Published: September 20, 2016 2545

DOI: 10.1021/acs.jnatprod.6b00449 J. Nat. Prod. 2016, 79, 2545−2551

Journal of Natural Products

Article

reaction of mushroom tyrosinase, with a half-maximal inhibitory concentration (IC50) of 102 μM, similar to that of P4.23 TH10 is a peptide comprising 10 amino acids with a Cterminal tyrosine residue, and docking simulation analysis of various TH10 variants revealed that the C-terminal tyrosine residue is essential for inhibitory activities and functions as a substrate analogue in the tyrosinase active site.23 Accordingly, approximately half of the known tyrosinase inhibitory peptides carry C-terminal tyrosine residues and appear to have similar mechanisms of action. In the present study, a technique for producing tyrosinase inhibitory peptides from rice bran protein (RBP) hydrolysates is demonstrated. Enzymatic hydrolysis is a predominant approach for producing bioactive peptides from food sources such as milk, soybean, wheat, rice, and barley proteins.24 Accordingly, the present simultaneous treatment with chymotrypsin and trypsin efficiently produced the above-mentioned “tyrosine-type” tyrosinase inhibitory peptides. This is the first report showing the production of tyrosinase inhibitory peptides from a food protein source.



RESULTS AND DISCUSSION Tyrosinase Inhibition by RBP Hydrolysates. Previously, it was demonstrated that the C-terminal tyrosine residue plays an important role in the tyrosinase inhibitory effects of peptides.23 Accordingly, the focus was on the substrate specificity of the proteinase chymotrypsin to produce inhibitory peptides. Chymotrypsin preferentially cleaves peptide bonds at the carboxyl sides of large hydrophobic and aromatic amino acids, such as phenylalanine, tyrosine, and tryptophan, leading to the production of peptides with a C-terminal tyrosine residue. However, the frequency of these aromatic residues in general protein sequences leads to a tendency for longer average lengths of chymotrypsin-generated peptides. Thus, because most previously characterized inhibitory peptides have short sequences of 2−10 amino acid residues,18−21 the proteinase trypsin was used in addition. These two proteinases are optimally catalytic under similar physical conditions of 37− 50 °C and pH 6−9. The high absorbance of RBP suspensions at 470 nm precluded the determinations of tyrosinase inhibitory activity. Thus, after hydrolysis by double digestion with chymotrypsin and trypsin, P-6 gel column chromatography was performed and inhibitory activities of each eluted fraction were confirmed. Fractions from nonhydrolyzed RBP suspensions showed no significant inhibition against monophenolase and diphenolase reactions of tyrosinase (Figure S1, Supporting Information). Slight inhibition of the diphenolase reaction likely reflected the presence of nonprotein compounds that lack significant absorbance at 210 nm. Chromatograms for RBP hydrolysates that were produced by double digestion with chymotrypsin and trypsin indicated the moderate digestion of RBP (Figure 1A), with peptide molecular weights from the main peak of 1000− 3000 Da, based on elution times of the standard protein (data not shown). However, a significant inhibition of both tyrosinase reactions was observed in fractions containing low-molecularweight peptides of 50% at 500 μM and was as effective as the arbutin positive control (Figure 5A). Furthermore, none of the present isolated peptides were cytotoxic in mouse B16 melanoma cells, and significant increases in cell growth were observed following peptide treatments, particularly in the presence of peptide CT-2 (Figure 5B). However, effective concentrations of CT-2 (micromolar) were much higher than those of epidermal growth factor (picomolar), which is a wellknown growth factor that stimulates cell growth, proliferation, and differentiation.25 Hence, although CT-2 is not applicable as a growth factor, the present cytotoxicity data indicate that CT-2 can be safely applied as a skin-whitening agent. Overall, these data demonstrate that CT-2 inhibits mushroom tyrosinase activity and potently inhibits melanogenesis in mouse B16 melanoma cells without causing cytotoxicity. In contrast, peptides CT-1 and -3 did not inhibit melanin production in mouse B16 melanoma cells, suggesting that these peptides do not inhibit mammalian tyrosinase. These observations may reflect differences in active site structures between mammalian and mushroom tyrosinases, which are well conserved but differ in some active site amino acid residues.26,27 Furthermore, PeptideCutter analysis (http://web.expasy.org/

peptide_cutter/) indicated that CT-2 has almost no cleavage sites for major proteinases, except for pepsin and proteinase K, which readily lose specificity under certain conditions. In contrast, CT-1 and -3 carry cleavage sites for other major proteinases such as Arg-C and glutamyl endopeptidases. This property may affect stability and the ensuing inhibition of melanogenesis after absorption into cells. Alternatively, appropriate hydrophobicity and molecular weight are considered requirements of skin permeability,28,29 and whereas identified peptides have similar hydrophobicity, the molecular weight of CT-2 (743.8) is lower than those of CT-1 (928.9) and CT-3 (831.8). Previously, Bos and co-workers proposed a “500 Da rule” for skin penetration of chemical compounds and drugs.28 Hence, if this rule is applicable to peptide drugs, low molecular weight CT-2 may be a more efficient drug candidate. A total of 4.0 g of RBP powder was used, and fraction 9, which was separated by Superdex Peptide 10/300 GL column chromatography, contained 19.2 mg of dry weight of peptide (data not shown). As calculated from the eluted peak area in the subsequent RP-HPLC separation, peak G (3.84% of total peak area) contained 0.73 mg of CT-2, indicating that 4.0 g of RBP powder contained at least 0.73 mg of CT-2. 2548

DOI: 10.1021/acs.jnatprod.6b00449 J. Nat. Prod. 2016, 79, 2545−2551

Journal of Natural Products

Article

In conclusion, in this study it was shown that RBP is a potent source of tyrosinase inhibitory peptides, and a technique was developed for isolating these peptides from RBP. Specifically, simultaneous treatments of RBP with chymotrypsin and trypsin led to efficient production of tyrosinase inhibitory peptides. Among identified peptides, CT-1, -2, and -3 significantly inhibited the monophenolase reaction of tyrosinase, and CT-2 potently inhibited melanogenesis in mouse B16 melanoma cells without causing cytotoxicity. This is the first report to show production of tyrosinase inhibitory peptides from a food protein source. Furthermore, the present data (Figures 1, 2, and 3) show that double digestion of RBP hydrolysates with chymotrypsin and trypsin increased the concentrations of tyrosinase inhibitory peptides, as indicated by numerous peptide peaks with inhibitory activity in chromatograms. Hence, the present data warrant further investigations of this technique for wide application to other natural protein sources.



EXPERIMENTAL SECTION

Materials. RBP 55 (protein contents, 55%) was kindly provided by Tsuno Food Industrial Co., Ltd. Trypsin (trypsin from porcine pancreas, lyophilized powder, 36 USP trypsin units/mg protein, catalog number: 207-09891) and chymotrypsin (α-chymotrypsin from bovine pancreas, lyophilized powder, 60 BTEE units/mg protein, catalog number: C4129) were purchased from Wako Pure Chemicals Ltd. (Osaka, Japan) and Sigma−Aldrich (St. Louis, MO, USA), respectively. Mushroom tyrosinase (EC 1.14.18.1) (lyophilized powder, catalogue number: T3824), L-tyrosine (catalogue number: T3754), and L-dopa (catalogue number: D9628) were also purchased from Sigma−Aldrich. The specific activity of the tyrosinase was 1320 U/mg (one unit = ΔA280 of 0.001 per min at pH 6.5 at 25 °C in a 3 mL reaction mixture containing L-tyrosine). All other regents were of analytical grade and were purchased from Wako Pure Chemicals Ltd. Preparation of RBP Hydrolysate. RBP powder (1.0 g) was suspended in ultrapure water (30 mL), and the pH was adjusted to 6.5 using NaOH prior to homogenization in a Polytron homogenizer (Kinematica; Bohemia, NY, USA). After dialysis against ultrapure water using Spectra/Por 1 dialysis tubing (molecular weight cutoff, 6− 8 kDa; Spectrum Laboratories, Inc., Rancho Dominguez, CA, USA), RBP solution was hydrolyzed by incubating with 20 mg of trypsin and 20 mg of chymotrypsin at 50 °C for 6 h. These enzymes were inactivated by incubating at 90 °C for 10 min, and after centrifugation at 12000g for 30 min at 4 °C, the clear supernatant was freeze-dried and used in subsequent isolation steps. Because trypsin from Wako Pure Chemicals Ltd. and chymotrypsin from Sigma−Aldrich have similar optimum pH (6−9) and temperature (37−50 °C), these conditions were used in all experiments with these proteases. Isolation of Tyrosinase Inhibitory Peptides. Dried RBP hydrolysate (656 mg) was dissolved in ultrapure water (15 mL) and was applied to a Bio-Gel P-6 gel column (2.5 × 110 cm; Bio-Rad; Hercules, CA, USA) that had been equilibrated using ultrapure water. Hydrolyzed peptides were eluted using ultrapure water (150 mL), and 6 mL fractions were collected every 6 min. Peptides in each fraction were freeze-dried and dissolved into small volumes of ultrapure water (300 mg/mL peptide concentration), and inhibitory activities against tyrosinase were determined. RBP hydrolysis and P-6 gel isolation steps were repeated, and a total of 4.0 g of RBP powder was treated as described above. Subsequently, all fractions containing tyrosinase inhibitory peptides were combined and applied to a Superdex Peptide 10/300 GL column (GE Healthcare; Uppsala, Sweden) that had been equilibrated with ultrapure water. Peptides were eluted using ultrapure water (25 mL) with an Ä KTA purifier (GE Healthcare), and 1 mL fractions were collected every 2 min. Peptides in each fraction were freeze-dried and dissolved in ultrapure water (30 mg/mL peptide concentration). Active fractions were combined and freeze-dried again, and the peptides were dissolved in 0.1% TFA in H2O and were further separated using RP-HPLC with a Cadenza CD-C18 column (150 × 10 mm, 3 μm; Imtakt Co., Kyoto, Japan). The HPLC system was

Figure 4. Dose−response curves of isolated peptides CT-1, -2, and -3 on inhibition of tyrosinase activity. (A) Inhibitory activity of each peptide against the monophenolase reaction of tyrosinase. (B) Inhibitory activity of each peptide against the diphenolase reaction of tyrosinase; remaining activity rates in the presence of indicated concentrations of CT-1 (open squares), -2 (open diamonds), -3 (open triangles), and arbutin (solid circles).

The other isolated peptides CT-4, -5, and -6 carry no tyrosine residues and do not inhibit tyrosinase activity, indicating that the strategy for producing peptides with a Cterminal tyrosine residue is efficient for producing tyrosinase inhibitory peptides. Most of the bran byproduct of rice milling is discarded as waste. However, enzymatic hydrolysates of proteins from food sources have various reported bioactive functions, including as inhibitors of angiotensin-I-converting enzyme (ACE), immunomodulating agents, antioxidants, antibiotics, and opioids.30 Hence, considerable efforts have been made to utilize rice bran as a source of bioactive compounds. Accordingly, various RBP hydrolysates have been produced during the past decade using specific proteinases such as pepsin, trypsin, and papain, and various antioxidant peptides have been isolated.31−34 Moreover, RBP hydrolysates produced using alcalase and protamax demonstrated strong inhibitory activities against α-amylase, βglucosidase, and ACE.35 Accordingly, the previous report and the present study showed that simultaneous treatment with trypsin and chymotrypsin efficiently produces tyrosinase inhibitory peptides from RBP. 2549

DOI: 10.1021/acs.jnatprod.6b00449 J. Nat. Prod. 2016, 79, 2545−2551

Journal of Natural Products

Article

Figure 5. Effects of isolated peptides CT-1, -2, and -3 on melanin contents and cell viability in mouse B16 melanoma cells. (A) Relative melanin contents 3 days after treatment with peptides. (B) Cell viability was determined using MTT-based assays 3 days after treatment with peptides; cells were treated with each peptide at indicated concentrations. Arbutin was used as a positive control. Data are presented as means of three independent experiments ± standard deviations (SD). initial monophenolase reaction, tyrosinase hydroxylates L-tyrosine into L-dopa in the presence of oxygen, and in the subsequent diphenolase reaction, L-dopa is oxidized into L-dopaquinone, which spontaneously decays into dopachrome. Dopachrome is a brown compound with a peak absorbance at 475 nm (molar extinction coefficient, ε = 3700 M−1 cm−1). Monophenolase activity was measured using a method similar to that reported by Chen and co-workers.39 In brief, 100 μL of 2.0 mM Ltyrosine, 45 μL of 0.1 M phosphate buffer (pH 6.5), and 5 μL of a peptide inhibitor solution or ultrapure water were placed in a 96-well plate and were preincubated at 30 °C for 10 min. After addition of 50 μL of 1320 units/mL mushroom tyrosinase to the mixture, the plates were incubated at 30 °C and were monitored at 475 nm using a microplate reader. Lag times were estimated by extrapolation of the linear portion of the L-dopa chrome accumulation curve to the abscissa axis. The remaining monophenolase activity (%) after treatment with the peptide inhibitor was calculated as follows: (A/B) × 100, where A represents the reciprocal of lag time (s) in the presence of peptide inhibitor, and B represents the reciprocal of the lag time (s) without peptide inhibitor (ultrapure water). IC50 values were manually estimated from inhibition curves. Diphenolase activity was determined using a similar method to that reported by Chen and co-workers.39 In brief, 100 μL of 2.0 mM Ldopa, 45 μL of 0.1 M phosphate buffer (pH 6.5), and 5 μL of peptide inhibitor solution or ultrapure water were placed in a 96-well plate and preincubated at 30 °C for 10 min. After addition of 50 μL of 1320 units/mL mushroom tyrosinase to the mixture, the reaction proceeded at 30 °C and was monitored at 475 nm using a microplate reader. The remaining diphenolase activity (%) after peptide inhibitor treatment was calculated as follows: (C/D) × 100, where C represents the absorbance change per min at 475 nm (ΔABS475/min) in the presence of peptide inhibitor, and D represents the reaction rate (ΔABS475/ min) without peptide inhibitor (ultrapure water). The tyrosinase inhibitor arbutin was used as a positive control. All measurements were performed three times. Cell Viability Assay. Mouse B16 melanoma cells were maintained in Eagle’s minimum essential medium with Earle’s salts supplemented

equipped with an SCL-10AVP system controller, LC-10AD pump, DGU-14A degasser, SPD-10AVP detector, and FRC-10A fraction collector (Shimadzu; Kyoto, Japan). Separation was performed at room temperature using a stepwise gradient elution with H2O−MeCN (supplemented with 0.1% TFA) under the following elution conditions: step 1, 0% MeCN for 5 min; step 2, linear gradient of 5.2−33.6% MeCN for 45 min; and step 3, 80% MeCN for 5 min; 3 mL fractions were collected every 3 min and freeze-dried. After determination of tyrosinase inhibitory activities of eluates, peptides from active fractions were further purified by rechromatography using a Cadenza CD-C18 column. Purification was performed using the following conditions: step 1, 0% MeCN for 3 min; step 2, linear gradient of 0−30.8% MeCN for 52 min; and step 3, 80% MeCN for 5 min; 3 mL fractions were collected every 3 min and freeze-dried. Absorbance of amide bonds (peptide bonds) was determined at 210 nm during all chromatographic elution. Identification of Amino Acid Sequences of Tyrosinase Inhibitory Peptides. Peptides were identified using MALDI-TOF/ MS as previously described.36 In brief, molecular weights of isolated peptides were investigated, and subsequent MS/MS analyses were performed using an Autoflex III TOF/TOF instrument (Bruker; Billerica, MA, USA) according to the manufacturer’s instructions. Amino acid sequences were identified from comparisons of experimentally obtained fragment masses and theoretical peptide masses of proteins in the Rice Annotation Project database.37 MS/MS ion searches were performed using the Mascot system (Matrix Science Ltd., London, UK);38 all identified peptides exceeded threshold values for 95% confidence interval levels (P < 0.05). Peptide Synthesis. Identified peptides were synthesized using the Fmoc solid-phase method with a PSSM-8 automated peptide synthesizer (Shimadzu) and were purified to >98% purity using RPHPLC with a Cadenza CD-C18 column. Molecular masses of purified peptides were confirmed using MALDI-TOF/MS data with an Autoflex III TOF/TOF (Bruker). Stock solutions of peptides were prepared in 10% (v/v) DMSO. Enzyme Assays. Tyrosinase catalyzes the two-step oxidative reaction from L-tyrosine to L-dopaquinone. Specifically, during the 2550

DOI: 10.1021/acs.jnatprod.6b00449 J. Nat. Prod. 2016, 79, 2545−2551

Journal of Natural Products

Article

(8) Park, H. Y.; Kosmadaki, M.; Yaar, M.; Gilchrest, B. A. Cell. Mol. Life Sci. 2009, 66, 1493−1506. (9) Lynde, C. B.; Kraft, J. N.; Lynde, C. W. Skin Therapy Lett. 2006, 11, 1−6. (10) Sanchez-Ferrer, A.; Rodriguez-Lopez, J. N.; Garcia-Canovas, F.; Garcia-Carmona, F. Biochim. Biophys. Acta, Protein Struct. Mol. Enzymol. 1995, 1247, 1−11. (11) Hearing, V. J., Jr.; Ekel, T. M.; Montague, P. M.; Nicholson, J. M. Biochim. Biophys. Acta 1980, 611, 251−268. (12) Solano, F.; Briganti, S.; Picardo, M.; Ghanem, G. Pigm. Cell Res. 2006, 19, 550−571. (13) Casanola-Martin, G. M.; Le-Thi-Thu, H.; Marrero-Ponce, Y.; Castillo-Garit, J. A.; Torrens, F.; Rescigno, A.; Abad, C.; Khan, M. T. Curr. Top. Med. Chem. 2014, 14, 1494−1501. (14) Westerhof, W.; Kooyers, T. J. J. Cosmet. Dermatol. 2005, 4, 55− 59. (15) Nohynek, G. J.; Kirkland, D.; Marzin, D.; Toutain, H.; LeclercRibaud, C.; Jinnai, H. Food Chem. Toxicol. 2004, 42, 93−105. (16) Takizawa, T.; Imai, T.; Onose, J.; Ueda, M.; Tamura, T.; Mitsumori, K.; Izumi, K.; Hirose, M. Toxicol. Sci. 2004, 81, 43−49. (17) Bang, S. H.; Han, S. J.; Kim, D. H. J. Cosmet. Dermatol. 2008, 7, 189−193. (18) Hsiao, N. W.; Tseng, T. S.; Lee, Y. C.; Chen, W. C.; Lin, H. H.; Chen, Y. R.; Wang, Y. T.; Hsu, H. J.; Tsai, K. C. J. Chem. Inf. Model. 2014, 54, 3099−3111. (19) Schurink, M.; van Berkel, W. J.; Wichers, H. J.; Boeriu, C. G. Peptides 2007, 28, 485−495. (20) Ubeid, A. A.; Do, S.; Nye, C.; Hantash, B. M. Biochim. Biophys. Acta, Gen. Subj. 2012, 1820, 1481−1489. (21) Ubeid, A. A.; Zhao, L.; Wang, Y.; Hantash, B. M. J. Invest. Dermatol. 2009, 129, 2242−2249. (22) Hantash, B. M.; Jimenez, F. J. Drugs Dermatol. 2012, 11, 660− 662. (23) Ochiai, A.; Tanaka, S.; Imai, Y.; Yoshida, H.; Kanaoka, T.; Tanaka, T.; Taniguchi, M. J. Biosci. Bioeng. 2016, 121, 607−613. (24) Wang, W.; De Mejia, E. G. Compr. Rev. Food Sci. Food Saf. 2005, 4, 63−78. (25) Carpenter, G.; Cohen, S. Annu. Rev. Biochem. 1979, 48, 193− 216. (26) Ismaya, W. T.; Rozeboom, H. J.; Weijn, A.; Mes, J. J.; Fusetti, F.; Wichers, H. J.; Dijkstra, B. W. Biochemistry 2011, 50, 5477−5486. (27) Matoba, Y.; Kumagai, T.; Yamamoto, A.; Yoshitsu, H.; Sugiyama, M. J. Biol. Chem. 2006, 281, 8981−8990. (28) Bos, J. D.; Meinardi, M. M. H. M. Exp. Dermatol. 2000, 9, 165− 169. (29) Yano, T.; Nakagawa, A.; Tsuji, M.; Noda, K. Life Sci. 1986, 39, 1043−1050. (30) Kitts, D. D.; Weiler, K. Curr. Pharm. Des. 2003, 9, 1309−1323. (31) Wattanasiritham, L.; Theerakulkait, C.; Wickramasekara, S.; Maier, C. S.; Stevens, J. F. Food Chem. 2016, 192, 156−162. (32) Adebiyi, A. P.; Adebiyi, A. O.; Ogawa, T.; Muramoto, K. Int. J. Food Sci. Technol. 2008, 43, 35−43. (33) Chanput, W.; Theerakulkait, C.; Nakai, S. J. Cereal Sci. 2009, 49, 422−428. (34) Adebiyi, A.; Adebiyi, A.; Yamashita, J.; Ogawa, T.; Muramoto, K. Eur. Food Res. Technol. 2009, 228, 553−563. (35) Uraipong, C.; Zhao, J. J. Sci. Food Agric. 2016, 96, 1101−1110. (36) Ahmed, F. E. Expert Rev. Proteomics 2008, 5, 841−864. (37) Tanaka, T.; Antonio, B. A.; Kikuchi, S.; Matsumoto, T.; Nagamura, Y.; Numa, H.; Sakai, H.; Wu, J.; Itoh, T.; Sasaki, T.; et al. Nucleic Acids Res. 2008, 36, D1028−D1033. (38) Perkins, D. N.; Pappin, D. J.; Creasy, D. M.; Cottrell, J. S. Electrophoresis 1999, 20, 3551−3567. (39) Chen, Q. X.; Song, K. K.; Wang, Q.; Huang, H. J. Enzyme Inhib. Med. Chem. 2003, 18, 491−496. (40) Arung, E. T.; Shimizu, K.; Kondo, R. Chem. Biodiversity 2007, 4, 2166−2171.

with 5% or 10% (v/v) fetal bovine serum (FBS; Moregate Biotech; Bulimba, Australia) and 1% penicillin−streptomycin (10 000 U/mL; Thermo Fisher Scientific; Waltham, MA, USA) in 5% CO2 at 37 °C. Cells were seeded at 1.0 × 103 cells/well in 200 μL of medium in 96well culture plates and were incubated for 24 h. Medium was replaced with 200 μL of fresh medium containing 5% FBS and various concentrations (125, 250, or 500 μM) of each identified tyrosinase inhibitory peptide. Cells were maintained under these conditions for 3 days, and viability was then assayed using a Premix WST-1 cell proliferation assay system (Takara Bio Inc.; Shiga, Japan) according to the manufacturer’s instructions. As described above, each peptide solution contains DMSO. Therefore, cells under the same DMSO conditions (and without peptides) were used as a background control in each assay. All assays were performed three times. Measurements of Melanin Contents. Melanin contents were measured using a previously described method 40 with some modifications. In brief, mouse B16 melanoma cells were seeded at 1.5 × 104 cells/well in 3 mL of medium in six-well culture plates and were incubated for 24 h. Subsequently, media was replaced with 3 mL of fresh medium containing 5% FBS and various concentrations (125, 250, or 500 μM) of identified tyrosinase inhibitory peptides. Cells were maintained under these conditions for 3 days and were recovered by treating with 750 μL of 0.25% trypsin solution containing 1 mM EDTA. Recovered cells were washed with PBS and were lysed with 225 μL of 1 M NaOH for 1 h at 60 °C, and absorbance was measured at 420 nm. The tyrosinase inhibitor arbutin was used as a positive control, and the cells under the same DMSO conditions without peptides were used as a background control in each assay. All measurements were performed three times.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jnatprod.6b00449. P-6 gel column chromatography for tyrosinase inhibitory peptides following the treatment of RBP with or without thermolysin; docking simulation analyses (PDF)



AUTHOR INFORMATION

Corresponding Author

*Tel (A. Ochiai): +81-252627722. Fax: +81-252626716. Email: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported in part by Cosmetology Research Foundation and Yamaguchi Scholarship Foundation funding to A.O. We would like to thank H. Yoshida for assistance in docking simulation analyses and T. Tomiya, Y. Imai, and T. Kanaoka for insightful advice regarding the manuscript.



REFERENCES

(1) Sharif, M. K.; Butt, M. S.; Anjum, F. M.; Khan, S. H. Crit. Rev. Food Sci. Nutr. 2014, 54, 807−816. (2) Lu, R.; Siebenmorgen, T. J. Trans. ASABE 1992, 35, 1955−1961. (3) Mancebo, S. E.; Wang, S. Q. Rev. Environ. Health 2014, 29, 265− 273. (4) Gallagher, R. P.; Lee, T. K.; Bajdik, C. D.; Borugian, M. Chronic Dis. Can. 2010, 29 (Suppl 1), 51−68. (5) Kraemer, K. H.; DiGiovanna, J. J. J. Invest. Dermatol. 2014, 134, E8−17. (6) Natarajan, V. T.; Ganju, P.; Ramkumar, A.; Grover, R.; Gokhale, R. S. Nat. Chem. Biol. 2014, 10, 542−551. (7) Hearing, V. J. J. Invest. Dermatol. 2011, 131, E8−E11. 2551

DOI: 10.1021/acs.jnatprod.6b00449 J. Nat. Prod. 2016, 79, 2545−2551