Sunteți pe pagina 1din 6

International Journal of Food Microbiology 192 (2015) 7–12

Contents lists available at ScienceDirect

International Journal of Food Microbiology


journal homepage: www.elsevier.com/locate/ijfoodmicro

Glutaminase-producing Meyerozyma (Pichia) guilliermondii isolated from


Thai soy sauce fermentation
Phichayaphorn Aryuman a, Sittiwat Lertsiri a, Wonnop Visessanguan b, Nuttawee Niamsiri a,
Amaret Bhumiratana a, Apinya Assavanig a,⁎
a
Department of Biotechnology, Faculty of Science, Mahidol University, Rama VI Rd., Bangkok 10400, Thailand
b
National Center for Genetic Engineering and Biotechnology (BIOTEC), Thailand Science Park, Phaholyothin Rd., Khlong Nueng, Khlong Luang, Pathumthani 12120, Thailand

a r t i c l e i n f o a b s t r a c t

Article history: In this study, 34 yeast isolates were obtained from koji and moromi samples of Thai soy sauce fermentation. How-
Received 25 March 2014 ever, the most interesting yeast strain was isolated from the enriched 2 month-old (M2) moromi sample and
Received in revised form 17 August 2014 identified as Meyerozyma (Pichia) guilliermondii EM2Y61. This strain is a salt-tolerant yeast that could tolerate
Accepted 15 September 2014
up to 20% (w/v) NaCl and produce extracellular and cell-bound glutaminases. Interestingly, its glutaminases
Available online 22 September 2014
were more active in 18% (w/v) NaCl which is a salt concentration in moromi. The extracellular glutaminase's
Keywords:
activity was found to be much higher than that of cell-bound glutaminase. The highest specific activity and
Soy sauce stability of the extracellular glutaminase were found in 18% (w/v) NaCl at pH 4.5 and 37 °C. A challenge test by
Salt-tolerant yeast adding partially-purified extracellular glutaminase from M. guilliermondii EM2Y61 into 1 month-old (M1)
Meyerozyma (Pichia) guilliermondii moromi sample showed an increased conversion of L-glutamine to L-glutamic acid. This is the first report of
Glutaminase glutaminase producing M. guilliermondii isolated from the moromi of Thai soy sauce fermentation. The results
Moromi suggested the potential application of M. guilliermondii EM2Y61 as starter yeast culture to increase L-glutamic
acid during soy sauce fermentation.
© 2014 Published by Elsevier B.V.

1. Introduction that converts L-glutamine originated from soy protein to L-glutamic


acid (Fukushima, 2004). Therefore, increasing production of glutaminase
Meyerozyma (Pichia) guilliermondii belongs to the group of so-called during soy sauce fermentation could result in an increase of L-glutamic
“flavogenic yeasts” (Tanner et al., 1945). Although, it is regarded as a acid which imparts so-called “umami” taste and improves the flavor in
contaminant producing off-flavor compound, i.e. 4-ethylphenol, during soy sauce (Ohshita et al., 2000; Yano et al., 1988). L-Glutaminase in soy
red wine production (Chatonnet et al., 1995; Dias et al., 2003), it is one sauce fermentation has been known to be produced by A. oryzae, howev-
of the dominant yeasts found in koji and moromi of Japanese and er this fungal enzyme is markedly inhibited by high NaCl concentration
Chinese fermented soybeans (Kim et al., 2010). Wah et al. (2013) of 20–22% (w/v) which is the common condition in moromi fermenta-
reported that M. guilliermondii EM1Y52 produced volatile phenol tion (Yano et al., 1988). Meanwhile, salt-tolerant yeast involving in
compounds and could also enhance the volatile flavor compound moromi fermentation, Z. rouxii has been reported for its glutaminase
production of Zygosaccharomyces rouxii in Thai soy sauce fermentation. that can tolerate high salt concentration (Iyer and Singhal, 2008). Many
Soy sauce is made by fermentation of soybean combined with wheat studies on the physiological roles, enzymatic properties, and application
flour, rice flour, and brine (Fukushima, 2004; Mongkolwai et al., 1997). of glutaminase from various microorganisms have been conducted,
Its production involves koji culture using Aspergillus oryzae and moromi including Micrococcus luteus (Moriguchi et al., 1994), Bacillus pasteurii
fermentation by adding brine solution into the koji. During koji culture, (Klein et al., 2002), Lactobacillus rhamnosus (Weingand-Ziade et al.,
A. oryzae mainly secretes protease and amylase to break down protein 2003), Lactobacillus reuteri KCTC3594 (Jeon et al., 2009), Saccharomyces
and carbohydrate in soybean and flour into peptides, amino acids and cerevisiae (Penninckx and Jaspers, 1985), Cryptococcus albidus (Iwasa
sugars, respectively (Mongkolwai et al., 1997). In addition to those en- et al., 1987) and Debaryomyces sp. (Durá et al., 2002). In this study,
zymes, glutaminase is an important enzyme in soy sauce fermentation the production of glutaminase in salt-tolerant yeast M. guilliermondii
isolated from Thai soy sauce moromi is first reported. Since
M. guilliermondii is GRAS (Generally Recognized as Safe) organism
⁎ Corresponding author at: Department of Biotechnology, Faculty of Science, Mahidol
University, 272 Rama VI Rd., Rachathevi, Bangkok 10400, Thailand. Tel.: + 66 2
(Sibirny and Boretsky, 2009), understanding of its glutaminase produc-
2015300; fax: +66 2 3547160. tion would be beneficial in the development of the starter culture for
E-mail address: apinya.ass@mahidol.ac.th (A. Assavanig). soy sauce fermentation.

http://dx.doi.org/10.1016/j.ijfoodmicro.2014.09.019
0168-1605/© 2014 Published by Elsevier B.V.
8 P. Aryuman et al. / International Journal of Food Microbiology 192 (2015) 7–12

2. Materials and methods stability to NaCl concentration was monitored during 48 h of dialysis
against 100 volumes of Britton–Robinson's buffer with various NaCl
2.1. Isolation of salt-tolerant yeasts from koji and moromi samples concentrations at 37 °C. The remaining glutaminase activity was
assayed at pH 4.5, 37 °C in the presence of 18% (w/v) NaCl.
Salt-tolerant yeasts were isolated from koji and moromi fermented
at 1 month (M1), 2 months (M2), and 3 months (M3) samples were 2.4. Production and partial purification of extracellular glutaminase from
collected from a soy sauce factory in Samutprakan province, Thailand. M. guilliermondii EM2Y61
Isolation process was performed at 30 °C using G-medium (per L: 10 g
L-glutamine, 1 g K2HPO4, 1 g KH2PO4, 1 g NaCl, 0.5 g yeast extract, Crude extracellular glutaminase was harvested from culture of the
0.1 g MgSO4·7H2O, 15 g agar, pH 7.0) (Wakayama et al., 2005) and selected yeast isolate in YPD broth containing 1% L-glutamine (w/v),
enrichment in YPD containing 10% (w/v) NaCl (10YPD) (Shumin et al., pH 4.5 by centrifugation. The enzyme was partially purified according
2012; Renouf and Lonvaud-Funel, 2007; Wah et al., 2013). All yeast col- to the method of Kingcha et al. (2012). Briefly, 1000 mL of the superna-
onies found were microscopically investigated. The isolated yeasts were tant was mixed with 20 g of Amberlite XAD-16 (Sigma-Aldrich, St. Louis,
further determined for salt tolerance by measuring the growth USA) and stirred at 4 °C for 12 h. Amberlite was activated with 50% (v/v)
(OD600nm ≥ 0.5) in YPD broth supplemented with 0, 10, 15, 18, 20% isopropanol and rinsed with deionized water prior to use. The slurry
(w/v) NaCl, at 30 °C under shaking condition for 28 days. The yeast was then loaded onto an Econo fast flow column (2.5 i.d. × 50 cm)
isolates grown in NaCl concentration (≥15%) were determined for glu- and washed with deionized water. Glutaminase fraction was eluted
taminase production. Molecular identification of selected salt-tolerant with 5 mL 50% (v/v) ethanol and concentrated by a rotary evaporation.
yeast producing the highest glutaminase was performed based on
sequence analysis of the D1/D2 domain on 26S rDNA (Ferreira et al., 2.5. Glutaminase activity assay
2010).
Glutaminase activity assay was based on the measurement of
2.2. Investigation of extracellular, intracellular, and cell-bound glutaminase L -glutamic acid produced under the defined conditions (Durá et al.,
activity in salt-tolerant yeast isolates 2002). The reaction mixture containing 1.8 mL of 100 mM L-glutamine
in 50 mM sodium phosphate buffer (pH 7.0) with or without 20% (w/v)
The selected salt-tolerant yeast isolates were cultured in 125 mL NaCl was pre-incubated at 37 °C for 3 min before adding 0.2 mL of
Erlenmeyer flask containing 25 mL of YPD broth, incubated at 30 °C crude enzyme sample. The reaction mixture was incubated at 37 °C for
with shaking (180 rpm) for 12 to 16 h. Cells were harvested by centrifu- 30 min and stopped by adding 2 mL of 10% (w/v) trichloroacetic acid
gation (3000 ×g, 4 °C, 15 min) and washed twice with 0.05 M phos- (TCA) and incubating at 25 °C for 2 h (Sivaraman et al., 1997). After
phate buffer, pH 7.0. The cell concentration was adjusted to OD600 nm centrifugation at 1000 ×g, 4 °C for 20 min, the supernatant was adjusted
1.5 before inoculation at 1% (v/v) inoculum into 100 mL of G-broth to pH 8.6 with 10 N KOH prior to determination of L-glutamic acid using
(Wakayama et al., 2005) in 250 mL Erlenmeyer flask. After incubation L-glutamic acid assay kit (R-Biopharm AG, Germany). For the control
at 30 °C with shaking for 3 days, the supernatant was separated from experiment, a sample was firstly added with 10% TCA to denature the
the yeast cells by centrifugation and determined for extracellular gluta- enzyme prior to adding L-glutamine substrate and measurement of
minase activity. The yeast cells were washed once in sterile distilled L -glutamic acid by the test kit. One unit (U) of glutaminase activity
water. One gram of cell pellet was suspended in 1 mL of 50 mM Tris– was defined as the amount of enzyme that catalyzed the formation of
HCl buffer containing 5 mM EDTA, pH 7.5 and then disrupted by homog- 1 μmolL-glutamic acid per min at 37 °C, pH 7.0. Protein concentration
enization at 4 °C with equal weight of glass beads at 1-min burst with an was determined by dyeing method using bovine serum albumin
intermediary pause of 1 min. The disrupted cells were observed under (BSA) as a standard (Bradford, 1976).
microscope (Durá et al., 2002). After centrifugation at 8000 × g, 4 °C
for 10 min, the liquid portion of cell lysate was subjected to determine 2.6. Demonstration of glutaminase activity of M. guilliermondii EM2Y61 in
the intracellular glutaminase activity. Cell-bound glutaminase fraction moromi model system
was prepared according to Soberón and González (1987). Briefly, cell
debris was separated from the cell lysate and then solubilized in the Fifty-mL portion of 1-month moromi (M1) was added with 100 mM
Tris–HCl buffer containing 0.1% (v/v) Triton X-100 followed by centrifu- L-glutamine and partially-purified glutaminase to obtain the final con-
gation at 8000 ×g, 4 °C for 10 min. The supernatant obtained was deter- centration of 0.5 U, and incubated at 37 °C for 60 min. Glutaminase
mined for cell-bound glutaminase activity. was partially purified as mentioned above. Samples were collected at
0, 15, 30, 45 and 60 min of incubation and centrifuged at 3000 × g,
2.3. Investigation of optimal pH, temperature, NaCl concentration, and 4 °C for 20 min (Kijima and Suzuki, 2007) prior to the determination
stability of crude extracellular glutaminase activity from M. guilliermondii of L-glutamic acid by L-glutamic acid test kit.
EM2Y61
2.7. Statistical analysis
The optimum pH for crude extracellular glutaminase activity was
assayed at 37 °C in Britton–Robinson's buffer (equi-volume of 0.04 M All experiments were run in triplicate. A completely randomized de-
boric acid, acetic acid, and phosphoric acid solution) containing sign was used throughout this study. Data were subjected to one-way
100 mM L-glutamine with and without 18% (w/v) NaCl in the pH analysis of variance and mean comparison with Duncan's test using
range of 3 to 9. For pH stability study, 10 mL of filter-sterilized extracel- Statistic Package for Social Sciences (SPSS for Windows: SPSS Inc.,
lular glutaminase fraction (Minisart Filters, 0.45 μm) were mixed with Chicago, IL, USA).
40 mL of Britton–Robinson's buffer containing 18% (w/v) NaCl and incu-
bated at 37 °C for 72 h (Durá et al., 2004). The remaining glutaminase 3. Results and discussion
activity was assayed under 18% (w/v) NaCl, at pH 4.5, 37 °C. The optimum
temperature and temperature stability were determined by incubating 3.1. Isolation of M. guilliermondii from soy sauce fermentation process and
the reaction mixture at 25, 30, 37, 40, 45 and 50 °C, pH 4.5, with and its production of glutaminase
without 18% (w/v) NaCl. The optimum NaCl concentration and stability
to NaCl concentration (i.e. 0, 5, 10, 15, 18 and 20% w/v) for extracellular Among 34 salt-tolerant yeasts (11 isolates from G-medium and 23
glutaminase activity were assayed at 37 °C, pH 4.5. Glutaminase isolates from enrichment 10YPD medium), only 26 isolates (76.5%)
P. Aryuman et al. / International Journal of Food Microbiology 192 (2015) 7–12 9

Fig. 1. Distribution of glutaminase activity in various cellular fractions of salt-tolerant yeast isolates. The activity was analyzed at pH 7.0, 37 °C for 30 min in 18% (w/v) NaCl. 1 U is defined as
the amount of glutaminase that catalyses the formation of 1 μmolL-glutamic acid per minute at 37 °C, pH 7.0. Total activity of extracellular glutaminase was defined as total units in
supernatant harvested. Total activity of intracellular glutaminase was defined as total units in total intracellular fraction obtained from cell lysis. Total activity of cell-bound glutaminase
was defined as total units in supernatant obtained after solubilization of cell debris.

Fig. 2. (a) pH profile; (b) pH stability of crude extracellular glutaminase activity of M. guilliermondii EM2Y61. Bars represent the standard deviation from triplicate experiments.
10 P. Aryuman et al. / International Journal of Food Microbiology 192 (2015) 7–12

produced glutaminase in G-medium. They were identified as from enriched culture of M2 moromi exhibited the highest extracellular
Zygosaccharomyces sp. (4 isolates, 15.4%), Debaryomyces sp. (15 isolates, glutaminase activity (Fig. 1). It was noticed that these yeast isolates
57.7%) and Pichia sp. (7 isolates, 26.9%) based on their cell morphology were moderately halotolerant similar to Z. rouxii (Iyer and Singhal,
and arrangement under light microscope and colonial appearance, ac- 2008), S. cerevisiae (Silva-Graca et al., 2003), D. hansenii (Durá et al.,
cording to Kurtzman et al. (2011). All 26 yeast isolates grew best in 2002) and P. guilliermondii (Butinar et al., 2005) which able to grow in
YPD broth, however they could slowly grow in YPD containing NaCl up to 18% (w/v) NaCl. Based on the partially amplified D1/D2 domain
up to 15% (w/v) and could survive in NaCl higher than 15%. Among of 26S rDNA sequence, these isolates of EM2Y61, EM1Y52 and EKY62
these isolates, only 3 isolates (EM2Y61, EM1Y52 and EKY62) could were positioned within the genus Meyerozyma (Pichia) and were
grow in NaCl up to 20% and produced remarkably high glutaminase identical to U45709 of the type strain (NRRL Y-2075) of M.(Pichia)
activity. guilliermondii. Hence, this study is the first report of salt-tolerant
From the results, glutaminase activity was variously distributed in M. guilliermondii producing the extracellular glutaminase.
different cellular fractions as the extracellular, intracellular, and cell- To explore the possibility of using M.(Pichia) guilliermondii EM2Y61
bound enzymes depending on the isolates (Fig. 1) when they were (GenBank accession number: JX568888) for starter culture develop-
cultivated in G-medium (Wakayama et al., 2005). Interestingly, ment, optimal conditions and stability of crude extracellular glutamin-
among 26 yeast isolates, 9 isolates produced 2 types and 17 isolates pro- ase of EM2Y61 were investigated. The optimum pH which was 4.5 of
duced only one type of glutaminase. All of these glutaminase enzymes this glutaminase was lower than those reported in other yeasts, i.e.
were active in the presence of 18% (w/v) NaCl which is the NaCl concen- pH 6.0 of glutaminase from Candida albicans (Iwasa et al., 1987);
tration employed in soy sauce moromi fermentation (Mongkolwai et al., pH 7.5 of membrane-bound glutaminase A and pH 8.1 of intracellular
1997). Moreover, the extracellular glutaminase activity was shown to glutaminase B from S. cerevisiae (Soberón and González, 1987); and
be the highest compared to intracellular and cell-bound glutaminase pH 8.5 of glutaminase from Debaryomyces spp. (Durá et al., 2002). In
activity. In yeasts, glutaminase was mostly found in intracellular frac- addition, the higher glutaminase activity of EM2Y61 was observed in
tions, i.e. Debaryomyces spp. (Durá et al., 2002); C. albidus; S. cerevisiae the reaction containing18% (w/v) NaCl at pH 4.5 (Fig. 2a). During
(Iwasa et al., 1987) and membrane-bound in S. cerevisiae (Iwasa et al., moromi fermentation, the pH of moromi liquid drops from 6.5–7.0 to
1987). Only halophilic yeast, i.e. Z. rouxii, which involves in moromi fer- 4.0–5.0 due to lactic acid bacteria such as Tetragenococcus halophilus,
mentation, has been reported to produce extracellular glutaminase converting sugars to acids. This acidic pH range is the suitable condition
(Iyer and Singhal, 2008). A. oryzae which takes part in koji fermentation for yeast growth during moromi fermentation (Kanbe and Uchida,
also produces extracellular glutaminase (Iyer and Singhal, 2008; Yano 1982). The crude extracellular glutaminase of M. guilliermondii
et al., 1988). But A. oryzae glutaminase is markedly inhibited by EM2Y61was stable with the half-life longer than 12 h when incubated
18% (w/v) NaCl whereas Z. rouxii glutaminase is partially inhibited by at pH 3 to 5, at 37 °C. The activity lasted long for 24 h at pH 3 to 4.5
18–20% (w/v) NaCl (Nandakumar et al., 2003). Among 3 isolates (Fig. 2b). The activity rapidly decreased at pH higher than 4.5. Since
(EM2Y61, EM1Y52 and EKY62), the isolate EM2Y61 that was obtained the activity was relatively stable at the pH close to the optimum pH

Fig. 3. (a) Temperature profile; (b) temperature stability of crude extracellular glutaminase activity of M. guilliermondii (EM2Y61). Bars represent the standard deviation from triplicate
experiments.
P. Aryuman et al. / International Journal of Food Microbiology 192 (2015) 7–12 11

(i.e. pH 4.5), this finding suggested that the native conformation and bi- The EM2Y61 glutaminase in our study was relatively stable when
ological activity of the enzyme might be maintained in the acidic pH incubated in the presence of 15–20% (w/v) NaCl concentration for
(Frankenberger and Johnson, 1982). The glutaminase stability has 12 h, with the residual activities of 85–90%, compared to 65–75% in 0–
been reported: glutaminase from Debaryomyces spp. was isolated 10% (w/v) NaCl (Fig. 4b). The glutaminase activity remained more
from cure meat products showing the stability at pH 7.5–8.5 but it than 50% after 24 h in the presence of NaCl concentration higher than
was not stable at the acidic pH (Durá et al., 2004); glutaminase from 10% (w/v). The results showed that the specific glutaminase activity of
Stenotrophomonas maltophilia NYW-81 was isolated from soil samples EM2Y61 was cultured in YPD medium, and could be increased from
being active at pH 9 and having relative activities from 60–80% between 0.36 to 0.43 U/mg protein when increasing the concentration of NaCl
pH 5 and 10 (Wakayama et al., 2005). Such acid-stable property proba- from 0 to 18% (w/v) (Fig. 4a). However, the activity slightly declined
bly depended upon where the enzyme-producing microorganisms under 20% (w/v) NaCl (0.36 U/mg protein). The optimal NaCl concen-
were isolated. The pH of our moromi liquid at 4.5–5 was probably suit- tration for glutaminase activity during cultivation of M. guilliermondii
able for optimal activity and stability of EM2Y61 glutaminase. EM2Y61 was in agreement with those from C. albicans (Iwasa et al.,
The optimum temperature of crude extracellular glutaminase was 1987), Cryptococcus nodaensis and Cryptococcus famata(Sato et al.,
37 °C which was lower than those reported in other yeasts, such as 1999). On the other hand, no activity of glutaminase A (membrane-
70 °C in C. albicans (Iwasa et al., 1987) and 40 °C in Debaryomyces spp. bound glutaminase) and B (intracellular glutaminase) from S. cerevisiae
(Durá et al., 2002). The optimal activity at 37 °C was observed in both (Soberón and González, 1987) and intracellular glutaminase from
0 and 18% (w/v) NaCl but the latter was higher (Fig. 3a). The glutamin- Debaryomyces spp. were observed in 3% (w/v) NaCl (Durá et al., 2002).
ase activity was gradually declined during 12 h of incubation at
25–37 °C with the residual activities of 90% at below 30 °C, and 85% at
37 °C. The half-life of the activity at 40–50 °C was longer than 12 h, 3.2. Implication of M. guilliermondii glutaminase in moromi model system
whereas it was more stable at below 37 °C (Fig. 3b). In contrast, gluta-
minase from Debaryomyces spp. was unstable at 37 °C with a half-life Contents of L-glutamine in moromi samples determined by HPLC
as short as 7.4 h comparedto the highest stability at 25 °C with a half- were 2.10, 1.95, and 1.43 mM for M1, M2, and M3, respectively. A
life of 135.8 h (Durá et al., 2004). C. albidus glutaminase has the activity large amount of L-glutamine in moromi was liberated from the soy pro-
that remained higher than 90% after incubation at 60 °C (Sato et al., tein, (Nakadai and Nasuno, 1989) into the mash through the actions of
1999). On the other hand, S. cerevisiae glutaminase A (membrane- proteinases and peptidases. Hence, we assumed that L-glutamine in the
bound glutaminase) was reported to be thermostable, but the B isozyme moromi was a possible substrate for glutaminase that produced during
(intracellular glutaminase) is thermolabile (Iwasa et al., 1987; Soberón moromi fermentation. Since the highest concentration of L-glutamine
and González, 1987). was observed in 1 month-moromi, it is possible to increase L-glutamic
acid by addition of extracellular glutaminase at this stage. Therefore,
the addition of partially-purified extracellular glutaminase from
EM2Y61 was challenged in the 1-month moromi model with a supple-
mentation of 100 mM L-glutamine. In the presence of 0.5 U of
M. guilliermondii EM2Y61 glutaminase, the concentration of L-glutamic
acid was significantly increased with the increasing time and reached
the maximum concentration at 60 min, which was 3 times higher than
the control with no addition of glutaminase (Fig. 5). The results demon-
strate the action of glutaminase to convert L-glutamine to L-glutamic acid
during moromi fermentation. However, the amount of L-glutamine
would be a limiting factor for the enhancement of L-glutamic acid
production. The results suggested that glutaminase activity could be
intensified by direct addition of glutaminase or inoculation of
glutaminase-producing starter culture. The application could be done
by using M. guilliermondii EM2Y61 as one of the yeasts in the starter
culture to increase L-glutamic acid during soy sauce fermentation.

Fig. 4. (a) Salt concentration profile; (b) salt stability of extracellular glutaminase activity
of M. guilliermondii (EM2Y61). Bars represent the standard deviation from triplicate Fig. 5. Effect of the addition of partially-purified extracellular glutaminase from
experiments. M. guilliermondii (EM2Y61) in 1-month moromi sample.
12 P. Aryuman et al. / International Journal of Food Microbiology 192 (2015) 7–12

4. Conclusions Kijima, K., Suzuki, H., 2007. Improving the umami taste of soy sauce by the addition of
bacterial γ-glutamyltranspeptidase as a glutaminase to the fermentation mixture.
Enzym. Microb. Technol. 41, 80–84.
Halotolerant yeast, M.(Pichia) guilliermondii EM2Y61, isolated from Kim, T.W., Lee, J.H., Park, M.H., Kim, H.Y., 2010. Analysis of bacterial and fungal communi-
Thai soy sauce fermentation could grow and produce extracellular glu- ties in Japanese- and Chinese-fermented soybean pastes using nested PCR-DGGE.
Curr. Microbiol. 60, 315–320.
taminase with high activity under high salt concentration. The highest Kingcha, Y., Tosukhowong, A., Zendo, T., Roytrakul, S., Luxananil, P., Chareonpornsook, K.,
enzyme activity and stability were observed at pH 4.5, in 18% (w/v) Valyasevi, R., Sonomoto, R., Visessanguan, W., 2012. Anti-listeria activity of
NaCl, at 37 °C. The results suggested that this enzyme could be active Pediococcus pentosaceus BCC 3772 and application as starter culture for Nham, a
traditional fermented pork sausage. Food Control 25, 190–196.
during the moromi fermentation of Thai soy sauce production. There- Klein, M., Kaltwasser, H., Jahns, T., 2002. Isolation of a novel, phosphate-activated gluta-
fore, M. guilliermondii EM2Y61 has potential to be used as a starter minase from Bacillus pasteurii. FEMS Microbiol. Lett. 206, 63–67.
culture at the early stage of moromi fermentation. Kurtzman, C.P., Fell, J.W., Boekhout, T., 2011. The yeasts, a taxonomic study, fifth ed.
Elsevier, London.
Mongkolwai, T., Assavanig, A., Annajsongsiri, C., Flegel, T.W., Bhumiratana, A., 1997. Tech-
Acknowledgments nology transfer for small and medium soy sauce fermentation factories in Thailand: a
consortium approach. Food Res. Int. 30, 555–563.
This work was supported by the Commission of Higher Education Moriguchi, M., Sakai, K., Tateyama, R., Furuta, Y., Wakayama, M., 1994. Isolation and char-
acterization of salt-tolerant glutaminase from Micrococcus luteus. J. Ferment. Bioeng.
in Thailand (CHE) and The Thailand Research Fund (grant no. 77, 621–625.
DBG5380022). Special thanks go to Dr. Thunyarat Pongtharangkul, Nakadai, T., Nasuno, S., 1989. Use of glutaminase for soy sauce made by Koji or a prepara-
Department of Biotechnology, Faculty of Science, Mahidol University tion of proteases from Aspergillus oryzae. J. Ferment. Bioeng. 67, 158–162.
Nandakumar, R., Yoshimune, K., Wakayama, M., Moriguchi, M., 2003. Microbial glutamin-
for helpful suggestions on identification techniques, Ms. Supawan ase: biochemistry, molecular approaches and applications in the food industry. J. Mol.
Walaisri for co-operating in various microbiological techniques and Catal. B Enzym. 23, 87–100.
soy sauce factory in Thailand for kindly providing the koji and moromi Ohshita, K., Nakajima, Y., Yamakoshi, J., Kataoka, S., Kikuchi, M., Pariza, M.W., 2000. Safety
evaluation of yeast glutaminase. Food Chem. Toxicol. 38, 661–670.
samples. Penninckx, M.J., Jaspers, C.J., 1985. Characterization γ-glutamylarylamidase glutaminase
by Saccharomyces cerevisiae. Biochimie 67, 999–1006.
References Renouf, V., Lonvaud-Funel, A., 2007. Development of an enrichment medium to detect
Dekkera/Brettanomyces bruxellensis, a spoilage wine yeast, on the surface of grape
Bradford, M.M., 1976. A rapid and sensitive method for the quantitation of microgram berries. Microbiol. Res. 162, 154–167.
quantities of protein utilizing the principle of protein-dye binding. Anal. Biochem. Sato, I., Kobayashi, H., Hanya, Y., Abe, K., Murakami, S., Scorzetti, G., 1999. Cryptococcus
72, 248–254. nodaensis sp. nov., a yeast isolated from soil in Japan that produces a salt-tolerant
Butinar, L., Santos, S., Spencer-Martins, I., Oren, A., Gunde-Cimerman, N., 2005. Yeast and thermostable glutaminase. J. Ind. Microbiol. Biotechnol. 22, 127–132.
diversity in hypersaline habitats. FEMS Microbiol. Lett. 244, 229–234. Shumin, Y., Na, N., Zhaoyanga, X., Jinglia, T., 2012. Screening and identification of
Chatonnet, P., Dubourdieu, D., Boidron, J.N., 1995. The influence of Brettanomyces/Dekkera halotolerant yeast for hydrocarbon degrading and its properties studies. Afr. J.
sp. yeasts and lactic acid bacteria on the ethylphenol content of red wines. Am. J. Enol. Microbiol. Res. 6, 1819–1828.
Vitic. 46, 463–468. Sibirny, A.A., Boretsky, Y.R., 2009. Yeast biotechnology: diversity and application. In:
Dias, L., Dias, S., Sancho, T., Stender, H., Querol, A., Malfeito-Ferreira, M., Loureiro, V., 2003. Satyanarayana, T.G., Kunze, G. (Eds.), Pichia guilliermondii. Springer Science and
Identification of yeasts originated from wine-related environments and capable of Business Media B.V., New York, pp. 113–134.
producing 4-ethylphenol. Food Microbiol. 20, 567–574. Silva-Graca, M., Neves, L., Lucas, C., 2003. Outlines for the definition of halotolerance/
Durá, M.A., Flores, M., Toldra, F., 2002. Purification and characterisation of a glutaminase halophile in yeasts: Candida versatilis (halophila) CBS4019 as the archetype. FEMS
from Debaryomyces spp. Int. J. Food Microbiol. 76, 117–126. Yeast Res. 3, 347–362.
Durá, M.A., Flores, M., Toldra, F., 2004. Effects of curing agents and the stability of a Sivaraman, T., Kumar, T.K.S., Jayaraman, G., Yu, C., 1997. The mechanism of 2,2,2 trichlo-
glutaminase from Debaryomyces spp. Food Chem. 86, 385–389. roacetic acid-induced protein precipitation. J. Protein Chem. 16, 291–297.
Ferreira, N., Belloch, C., Querol, A., Manzanares, P., Vallez, S., Santos, A., 2010. Yeast Soberón, M., González, A., 1987. Physiological role of glutaminase activity in Saccharomyces
microflora isolated from Brazilian cassava roots: taxonomical classification based on cerevisiae. J. Gen. Microbiol. 133, 1–8.
molecular identification. Curr. Microbiol. 60, 287–293. Tanner, F., Vojnovich, C., Lane, J.M., 1945. Riboflavin production by Candida species.
Frankenberger, W.T., Johnson, J.B., 1982. Influence of crude oil and refined petroleum Science 101, 180–185.
products on soil dehydrogenate activity. J. Environ. Qual. 11, 602–607. Wah, T.T., Walaisri, S., Assavanig, A., Niamsiri, N., Lertsiri, S., 2013. Co-culturing of Pichia
Fukushima, D., 2004. Industrialization of fermented soy sauce production centering guilliermondii enhanced volatile flavor compound formation by Zygosaccharomyces
around Japanese shoyu. In: Steinkraus, K.H. (Ed.), Industrialization of Indigenous rouxii in the model system of Thai soy sauce fermentation. Int. J. Food Microbiol.
Fermented Foods. Marcel Dekker, New York, pp. 1–88. 160, 282–289.
Iwasa, T., Fuji, M., Yokotsuka, T., 1987. Glutaminase produced by Cryptococcus albidus Wakayama, M., Yamagata, T., Kamemura, A., Bootim, N., Yano, S., Tachiki, T., Yoshimune,
ATCC20293. Purification and some properties of the enzyme. Nippon Shoyu K., Moriguchi, M., 2005. Characterization of salt-tolerant glutaminase from
Kenkyusho Zasshi 13, 205–210. Stenotrophomonas maltophilia NYW-81 and its application in Japanese soy sauce
Iyer, P., Singhal, R.S., 2008. Production of glutaminase (E.C.3.2.1.5) from Zygosaccharomyces fermentation. J. Ind. Microbiol. Biotechnol. 32, 383–390.
rouxii: statistical optimization using response surface methodology. Bioresour. Weingand-ziade, A., Gerber-decombaz, C., Affolter, M., 2003. Functional characterization
Technol. 99, 4300–4307. of a salt- and thermotolerant glutaminase from Lactobacillus rhamnosus. Enzym.
Jeon, J., Lee, H., So, J., 2009. Glutaminase activity of Lactobacillus reuteriKCTC3594 and Microb. Technol. 32, 862–867.
expression of the activity in other Lactobacillus spp.by introduction of the glutamin- Yano, T., Ito, M., Tomita, M., Kumagai, H., 1988. Purification and properties of glutaminase
ase gene. Afr. J. Microbiol. Res. 3, 605–609. from Aspergillus oryzae. J. Ferment. Technol. 66, 137–143.
Kanbe, C., Uchida, K., 1982. Diversity in the metabolism of organic acids by Pediococcus
halophilus. Agric. Biol. Chem. 46, 2357–2359.

S-ar putea să vă placă și