ArticlePDF Available

Screening For Yeast Phytase Leads to the Identification of a New Cell-Bound and Secreted Activity in Cyberlindnera jadinii CJ2

Frontiers
Frontiers in Bioengineering and Biotechnology
Authors:

Abstract and Figures

Phytic acid is an anti-nutritional compound able to chelate proteins and ions. For this reason, the food industry is looking for a convenient method which allows its degradation. Phytases are a class of enzymes that catalyze the degradation of phytic acid and are used as additives in feed-related industrial processes. Due to their industrial importance, our goal was to identify new activities that exhibit best performances in terms of tolerance to high temperature and acidic pH. As a result of an initial screening on 21 yeast species, we focused our attention on phytases found in Cyberlindnera jadinii, Kluyveromyces marxianus, and Torulaspora delbrueckeii. In particular, C. jadinii showed the highest secreted and cell-bound activity, with optimum of temperature and pH at 50°C and 4.5, respectively. These characteristics suggest that this enzyme could be successfully used for feed as well as for food-related industrial applications.
Content may be subject to copyright.
fbioe-09-662598 May 18, 2021 Time: 17:20 # 1
ORIGINAL RESEARCH
published: 24 May 2021
doi: 10.3389/fbioe.2021.662598
Edited by:
Lucia Gardossi,
University of Trieste, Italy
Reviewed by:
Hasan Bugra Coban,
Dokuz Eylül University, Turkey
Farshad Darvishi,
Alzahra University, Iran
*Correspondence:
Concetta Compagno
concetta.compagno@unimi.it
Specialty section:
This article was submitted to
Industrial Biotechnology,
a section of the journal
Frontiers in Bioengineering and
Biotechnology
Received: 01 February 2021
Accepted: 29 March 2021
Published: 24 May 2021
Citation:
Capusoni C, Serra I, Donzella S
and Compagno C (2021) Screening
For Yeast Phytase Leads to the
Identification of a New Cell-Bound
and Secreted Activity in Cyberlindnera
jadinii CJ2.
Front. Bioeng. Biotechnol. 9:662598.
doi: 10.3389/fbioe.2021.662598
Screening For Yeast Phytase Leads
to the Identification of a New
Cell-Bound and Secreted Activity in
Cyberlindnera jadinii CJ2
Claudia Capusoni, Immacolata Serra, Silvia Donzella and Concetta Compagno*
Department of Food, Environmental and Nutritional Sciences, University of Milan, Milan, Italy
Phytic acid is an anti-nutritional compound able to chelate proteins and ions. For
this reason, the food industry is looking for a convenient method which allows its
degradation. Phytases are a class of enzymes that catalyze the degradation of phytic
acid and are used as additives in feed-related industrial processes. Due to their industrial
importance, our goal was to identify new activities that exhibit best performances in
terms of tolerance to high temperature and acidic pH. As a result of an initial screening
on 21 yeast species, we focused our attention on phytases found in Cyberlindnera
jadinii,Kluyveromyces marxianus, and Torulaspora delbrueckeii. In particular, C. jadinii
showed the highest secreted and cell-bound activity, with optimum of temperature and
pH at 50C and 4.5, respectively. These characteristics suggest that this enzyme could
be successfully used for feed as well as for food-related industrial applications.
Keywords: phytic acid, yeast, Cyberlindnera jadinii, feed additive, food production, phytase
INTRODUCTION
Phytic acid (myo-inositol 1,2,3,4,5,6-hexakis dihydrogen phosphate) is the main source of stored
phosphorus in grains, oil seeds, and nuts (Mullaney and Ullah, 2003), typically representing up to
60–80% of total phosphorus in seed, and playing an important role during seed germination and
growth (Shi et al., 2005). The presence of phytic acid creates problems in breeding, being feeds
mainly composed by vegetal materials rich in this acid. Polygastric animals are able to degrade
phytate, thanks to their particular gut microbiota (Nakashima et al., 2007), but this process does
not occur in the monogastric ones, like poultry, pigs, fishes, and also humans. Since phytate
cannot be metabolized, feed for monogastric animals are often fortified with inorganic phosphorus,
increasing their final cost. In addition, accumulation of phytic acid has a negative effect on animal
health, because it represents an anti-nutritional and chelating agent, that reduces bioavailability of
proteins and ions like Fe3+, Ca2+, Zn2+, and Mg2+forming insoluble complexes (Reddy et al.,
1982;Coban and Demirci, 2017). The undigested phytate then accumulates in manure and liquid
effluents, leading to phosphorus pollution and water eutrophication. Also for human nutrition,
there is now an increasing attention about these aspects. Phytate degradation in food is mediated
mainly by fermentation processes led by phytate-degrading microorganisms (De Angelis et al.,
2003;Rizzello et al., 2010) or during the food processing by endogenous phytases present in food
Abbreviations: aa, aminoacid; EDTA,ethylene-diamine-tetraacetic acid; OD, optical density; TCA, tricloracetic acid; gDNA,
genomic DNA; U, unit.
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org 1May 2021 | Volume 9 | Article 662598
fbioe-09-662598 May 18, 2021 Time: 17:20 # 2
Capusoni et al. Phytase From Cyberlindnera jadinii
matrix (Leenhardt et al., 2005). For the reasons described above,
science community has focused attentions on phytate-degrading
enzymes (Lei et al., 2013;Mrudula Vasudevan et al., 2019;
Pires et al., 2019).
Phytases are a class of enzymes that catalyze the hydrolytic
degradation of phytic acid to free inorganic phosphorus, to yield
lower myo-inositol phosphate esters and, in some case, free
myo-inositol (Vats and Banerjee, 2004). Enzymes described as
phytases show different structures: histidine acid phosphatase
(HAP), βpropeller phytase (BPP), and purple acid phosphatase
(PAP) (Mullaney and Ullah, 2003). The most known and wide
class is HAPs (EC 3.1.3.8). This class is ubiquitous, indeed
HAPs can be found not only in bacteria, yeasts, and filamentous
fungi but also in upper eukaryotes (Lei et al., 2013). All the
proteins belonging to this class maintain two common domains:
a conserved N-terminal heptapeptide active site RHGXRXP (aa
38–44) and a C-terminal catalytically active dipeptide HD (aa
325–326) (Mullaney and Ullah, 2003). Among HAPs exist a
variety of specific activities. Wyss and coworkers analyzed several
fungal phytases dividing them in two different subclasses: one
with broad substrate specificity but low specific activity on phytic
acid (PhyBp), and the second with narrow substrate specificity
but high activity on phytic acid (PhyAp) (Wyss et al., 1999).
Curiously, some organisms as Aspergillus niger possess both
forms (Mullaney and Ullah, 2003).
Although phytases have been reported in a wide range of
bacteria, not many of them have been used so far as feed
supplement, since their neutral/alkaline pH optimum and their
optimal temperature could preclude their activity in these
processes. For industrial applications, the ideal phytase should
display in fact three characteristics: ability to hydrolyze phytic
acid in the upper digestive tract of the animals, resilience up
to 65–90C and cheap production cost. In particular, to work
properly in the digestive tract, phytase needs to have a pH
optimum between 3.5 and 5.5 and optimum of temperature in the
range 37–40C. Furthermore, phytases should actively work also
at higher temperatures, required in feed production processes,
as during pelletting and heat treatment to control Salmonella
spoilage (Mrudula Vasudevan et al., 2019). In addition, it would
need to be resistant to protease activity and to show low
sensitivity to ions (Wyss et al., 1999).
Addition of phytases is currently not yet applied in human
food production for increasing the mineral bioavailability of the
food. This is mainly due to the fact that, so far, all the commercial
phytase-producing organisms are genetically modified organisms
(GMOs), which are commonly not well accepted for human
food production.
Yeasts are good candidates for phytase production and
some of them have been already characterized (Ragon et al.,
2009;Nuobariene et al., 2011;Greppi et al., 2015;Pires et al.,
2019). These enzymes show different localizations: in some
species, phytases are extracellular enzymes, in others are cell
bound or released in the periplasmic space (In et al., 2009;
Olstorpe et al., 2009;Kaur et al., 2010;Hellström et al., 2015;
Kłosowski et al., 2018).
With the aim to find new phytases with characteristics suitable
for feed and food industrial applications, in this study, we
screened 28 yeast strains isolated from different environments,
terrestrial and marine. We investigated cellular localization
of phytase activity and, for some of them, the effects of
different phosphorus sources on their expression. In addition, we
characterized optimal temperature and pH.
MATERIALS AND METHODS
Yeast Strains
The yeast strains studied in this work (Table 1 and
Supplementary Table 1) belong to CBS collection, UBO
Culture Collection (UBO-CC1), DBVPG collection, and private
collections at the University of Milan (CML and UMY). Yeasts
were stored in YPD-20% (vol/vol) glycerol stocks at 80C.
Media and Cultivation
Yeasts were cultivated on different media.
YPD: 10 g/L yeast extract, 20 g/L peptone, 20 g/L glucose.
MMPhy: 20 g/L glucose, 5 g/L (NH4)2SO4, 0.5 g/L
MgSO47H20, 0.11 g/L phytic acid sodium salt hydrate (Sigma
Aldrich, Milano, Italy), trace metals (di-sodic EDTA 15 mg/L,
ZnSO44H2O 4.5 mg/L, MnCl4H2O 0.1 mg/L, CoCl26 H2O
0.3 mg/L, CuSO45H2O 0.3 mg/L, Na2MoO42H20 0.4 mg/L,
CaCl22H2O 4.5 mg/L, FeSO47H2O 3 mg/L, H3BO31 mg/L,
KI 0.1 g/L) and vitamins, d-biotin 0.05 mg/L, calcium D-
pantothenate 1 mg/L, nicotinic acid 1 mg/L, myoinositol 25 mg/L,
thiamine hydrochloride 1 mg/L, pyridoxine hydrochloride,
p-aminobenzoic acid 0.2 mg/L) as reported in Merico et al. (2007)
with some modification.
MM-: 20 g/L glucose, 5 g/L (NH4)2SO4, 0.5 g/L MgSO47H2O,
trace metals, and vitamins as in MMPhy.
MMP: 20 g/L glucose, 5 g/L (NH4)2SO4, 0.5 g/L
MgSO47H2O, KH2PO41 g/L, trace metals, and
vitamins as in MMPhy.
MMP/Phy: 20 g/L glucose, 5 g/L (NH4)2SO4, 0.5 g/L
MgSO47H2O, 0.11 g/L phytic acid sodium salt hydrate, KH2PO4
1 g/L, trace metals, and vitamins as in MMPhy.
In all media, pH was adjusted to 4.5 using H2SO4.
Yeast cells were cultivated at 28C in a rotary shaker at
150 rpm in bluffed flasks (100 ml) containing 20 ml of medium.
Optical density was monitored at 600 nm (OD600). Cells were
precultured for 24 h in YPD, harvested by centrifugation
at 5,000 rpm, and washed three times with sterile NaCl
solution (9 g/L). Then they were used to inoculate alternatively
MMPhy, MMP, MMP/Phy, and YPD at an initial OD600 = 1
(corresponding to 107cells/ml approximately).
Dry Weight Determination
For dry weight measurements (DW), samples from different
culture conditions were collected (in triplicate at each point).
Cells were filtered through a glass microfiber GF/A filter
(Whatman), washed with three volumes of deionized water and
dried at 100C for 24 h.
1http://www.univ-brest.fr/ubocc
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org 2May 2021 | Volume 9 | Article 662598
fbioe-09-662598 May 18, 2021 Time: 17:20 # 3
Capusoni et al. Phytase From Cyberlindnera jadinii
TABLE 1 | Results of phytase screening performed on 28 yeast strains grown on medium containing phytic acid as sole phosphorus source.
Yeast Strain OD600 nm Cell-bound Extracellular
Debaryomyces hansenii MI 1 5.5 BDL BDL
Ciberlindnera jadinii CJ2 28 YES YES
Schizosaccharomyces pombe Y709 4 BDL BDL
Kluyveromyces lactis CBS 2359 3 YES BDL
Kluyveromyces lactis Y1356 1 BDL BDL
Kluyveromyces marxianus Y1058 14 YES BDL
Hanseniaspora uvarum UMY 514 8 YES BDL
Hanseniaspora uvarum UMY 571 14 YES BDL
Saccharomyces cerevisiae CENPK 113 7D 7.78 BDL BDL
Saccharomyces cerevisiae LALVIN T73 8.36 BDL BDL
Brettanomyces bruxellensis CBS 2499 10 YES BDL
Lachancea thermotolerans CBS 6340 8 YES BDL
Torulaspora delbruekii CBS1466 8.56 YES BDL
Candida humilis CBS 5658 8.6 BDL BDL
Candida milleri CBS 6897 11 BDL BDL
Rhodosporidium azoricum DBVPG 4620 12 BDL BDL
Zygosaccharomyces kombutchaensis CBS 8849 7 YES BDL
Kazakistania unispora CML133 3 BDL BDL
Meyerozyma guilliermondii UBOCC-A-214008 26 YES BDL
Meyerozyma guilliermondii UBOCC-A-214143 20 YES BDL
Pichia guilliermondii EX15 UBOCC-A-208004 21 YES BDL
Rhodotolura mucilaginosa UBOCC-A-214025 10 BDL BDL
Rhodotolura mucilaginosa UBOCC-A-214036 7.9 YES BDL
Candida atlantica Mo31 UBOCC-A-208026 25 YES BDL
Candida oceani Mo39 UBOCC-A-208034 4.2 BDL BDL
Debaryomyces hansenii BIO2 UBOCC-A-208002 11 YES BDL
Debaryomyces hansenii Mo40 UBOCC-A-208035 8.2 BDL BDL
Rhodotorula diobovata Mo38 UBOCC-A-208033 12.5 YES BDL
References for strains origin are reported in Supplementary Table 1. BDL, below detection limit.
Phytase Activity Determination
Enzymatic activity was determined on supernatants (for
extracellular activity) and whole cells (for cell-bound activity) of
early exponential phase cultures. The activity was measured by
ortho-phosphate production, following ammonium molibdate
blue method as reported by Schimizu (Shimizu, 1992), with
some modifications.
For extracellular activity determination, cell cultures were
centrifuged at 13,000 rpm and 1 ml of supernatant was added
to 4 ml of buffer composed by 0.2 M Na acetate/acetic acid,
8 mM phytic acid at pH 4.5. To determine cell-bound activity,
a standard amount of cells (corresponding to 10 mg of dry
weight) was collected, washed twice with 0.2 M Na acetate/acetic
acid at pH 4.5, and resuspended in a final volume of 1 ml of
water. Cell suspension was added to 4 ml of buffer in 0.2 M
Na acetate/acetic acid, 8 mM phytic acid at pH 4.5. All buffers
employed to test enzymatic activity were prewarmed at the
reaction temperature. Blank was assembled using 1 ml of water
and 4 ml of 0.2 M Na acetate/acetic acid at pH 4.5, 8 mM phytic
acid, and treated as sample.
For enzymatic activity determination, 5 ml of reaction mixture
were incubated in 15 ml tube at 37C and stirred at 300 rpm.
The reaction was immediately stopped (for time 0) and then
stopped after 15, 30, 60, and 120 min by mixing 0.5 ml of
reaction mixture with 0.5 ml of TCA 5% solution. Samples
were centrifuged for 3 min at 13,000 rpm and the supernatants
collected. To determine orthophosphate concentration, 0.4 ml
of supernatant was added to 0.4 ml of molibdate solution.
This solution was prepared daily, by mixing solution A and B
in a ratio of 4:1 (solution A: 2.6% N6H24Mo7O24 4H2O and
5.5% H2SO4; solution B: 4.6% FeSO47H2O). The sample was
incubated 10 min at 25C and read against blank at OD700.
Phosphate concentration was determined using a standard curve
for KH2PO4. One unit of phytase is defined as the amount
of protein that hydrolyses 1 µmol of phosphorus/min. Specific
activity is expressed as milliunits per milligram of cell dry weight.
To determine the effect of temperature, samples prepared with
prewarmed buffer (pH 4.5) were incubated at 50 and 60C. To
determine the effect of pH on enzymatic activity, pH buffers
were adjusted at pH 4 and pH 5.5, and the reactions incubated
at 37C.
Genomic Extraction
To isolate genomic DNA, pellets corresponding to 30 OD of cells
were resuspended in 0.5 ml of 50 mM Tris–HCl, 20 mM M EDTA
at pH 7.5. This suspension was transferred to a precooled tube
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org 3May 2021 | Volume 9 | Article 662598
fbioe-09-662598 May 18, 2021 Time: 17:20 # 4
Capusoni et al. Phytase From Cyberlindnera jadinii
with an equal volume of glass beads (425–600 µm). Mechanical
lysis was performed using a TissueLyser LT (Qiagen) alternating
2 min of agitation at 50 Hz with 1 min in ice for four cycles.
The supernatant was added with 25 µl of SDS 20% (w/v) and
incubated at 65C for 30 min. Immediately, 0.2 ml of 5 M
potassium acetate was added and the tubes were placed on ice
for 30 min. Samples were centrifuged at 13,000 rpm for 5 min
and supernatants transferred to a fresh microcentrifuge tube.
The DNA was precipitated by adding 1 vol of isopropanol.
After incubation at room temperature for 5 min, the tubes were
centrifuged for 10 min. The DNA was washed with 70% ethanol
and dissolved in 50 µl of TE RNAse (10 mM Tris–HCl, 1 mM
EDTA, pH 7.5 RNAse 100 µg/ml). Samples were incubated at
37C for 30 min.
Strain Identification
gDNA was amplified with Phusion taq polymerase employing
universal primers for amplification on D1/D2 domain of the
26S rDNA (NL1: GCATATCAATAAGCGGAGGAAAAG, NL4:
GGTCCGTGTTTCAAGACGG) 0.2 µM each, 200 µM dNTP,
and MgCl22.5 mM (Fliegerová et al., 2006). PCR amplification
was carried out by denaturing at 98C for 7 min, followed
by 30 cycles of denaturing at 98C for 30 s, annealing at
52C for 30 s, extension at 72C for 1 min, and a final
extension at 72C for 5 min. The produced PCR amplicon
was sequenced using the Sanger method at Microsynth Seqlab
(Germany), and the strain was identified by the sequence
similarity using basic local alignment search tool against the
NCBI databases2.
Phylogenetic and Bioinformatics
Analysis
Phytase sequence of D. hansenii Mo40 and Bio2 were obtained in
this work. Phytase gene was amplified from gDNA using primers:
Forward: Phy1 CCGACCATGGATGGTATCGATTTCC,
Reverse: Phy2 CATCGGATCCTAATT GTCACCGGA. Primers
were designed based on D. hansenii CBS 767 (GeneID: 2900382;
XP_460696.1). PCR amplification was carried out by denaturing
at 98C for 7 min, followed by 30 cycles of denaturing at 98C
for 10 s, annealing at 59C for 30 s, extension at 72C for
45 s, and a final extension at 72C for 10 min. PCR amplicons
were sequenced using the Sanger method at Microsynth
Seqlab (Germany).
The aminoacidic sequences (accession numbers on
Supplementary Table 2) were identified through a BLASTp
by XP_460696.1 of D. hansenii CBS 767 against the NCBI
databases. For phylogenetic analysis, multiple alignments
of aminoacidic sequences were performed using MUSCLE
(EMBL-EBI tool on3) and a maximum likelihood tree was
built using Mega X 10.1.74. Analysis of signal secretion
sequence was performed employing SignalIP-5.0 available on
http://www.cbs.dtu.dk/services/SignalP/.
2http://www.ncbi.nlm.nih.gov/BLAST/
3https://www.ebi.ac.uk/Tools/msa/muscle/
4www.megasoftware.net
RESULTS AND DISCUSSION
Growth in Presence of Phytic Acid as
Sole Phosphorus Source
Twenty-one yeast species (28 strains) belonging to Debaryomyces,
Cyberlindnera,Schizosaccharomyces,Kluyveromyces,
Saccharomyces,Brettanomyces,Candida,Torulaspora,
Rhodosporidium,Meyerozyma,Hanseniaspora,Pichia,
Lachancea,Kazakistania, and Rhodotorula genera were
characterized for their ability to grow using phytic acid as
sole phosphorus source (Table 1 and Supplementary Figure 1).
Strain CJ2 was identified in this work as Cyberlindnera jadinii.
All strains were cultivated on medium MMPhy, and their growth
was monitored for 72 h. In parallel, as negative control, growth
on MM- (without any phosphorus source) was carried out,
and no appreciable growth was detected (data not shown). We
avoided to perform screening on solid medium due to ambiguous
results reported sometimes in literature.
All strains except Kluyveromyces lactis Y1356 were able to
grow using phytic acid as sole phosphorus source, but with
variable extent (Table 1). Some strains, like C. jadinii and
Meyerozyma guilliermondii were able to exceed 20 OD after 24 h
of incubation, reaching 26 OD and 25 OD, respectively, after 48 h.
Other strains reached lower OD values after 72 h of incubation,
and K. lactis CBS 2359 duplicated only two times reaching 4 OD
(Table 1 and Supplementary Figure 1). These differences reflect
species-dependent efficiency of phytase activity, as well as species-
specific mechanism of phytic acid hydrolysis. Literature reports
that in some yeasts, like Debaryomyces castellii, phytase is able to
completely hydrolyze phosphate from phytic acid (Ragon et al.,
2008), but others, like in Kodamaea ohmeri are, on the contrary,
not able to perform a complete hydrolysis of this acid, leaving
some phosphate group not bioavailable (Li et al., 2009).
Screening for Phytase Activity
In order to identify enzymes that show characteristics suitable
for feed/food industrial processes, meaning a phytase able to
work in the upper digestive tract of monogastric animals, with
optimal temperature range between 37 and 40C and pH range
between 3.5 and 5.5, we decided to screen phytase activity at 37C
and pH 4.5. Extracellular and cell-bound activity was assayed on
cells grown using phytate as sole phosphorus source, on MMPhy
medium. To correctly compare phytase activity in various strains,
we expressed the activity as milliunits per milligram d.w., instead
of milliunits per milliliter as often reported in literature. In this
way, we avoided the bias due to different ability of strains to grow
in presence of phytate (Table 1), reaching different amount of
biomass per milliliter.
Under these conditions, we detected extracellular phytase
activity only in C. jadinii (26.25 mU/mgd.w.). On the other hand,
16 out of the 28 tested strains showed a detectable cell-bound
activity (Figure 1), and C. jadinii was identified as the one
with the highest (58.36 mU/mgd.w.). Lower levels of cell-bound
activity were found in fact in the other species (Figure 1), like
Kluyveromyces marxianus (4.17 mU/mgd.w.), Meyerozyma
guilliermondii (10.49 mU/mgd.w.), Pichia guillermondii
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org 4May 2021 | Volume 9 | Article 662598
fbioe-09-662598 May 18, 2021 Time: 17:20 # 5
Capusoni et al. Phytase From Cyberlindnera jadinii
FIGURE 1 | Cell-bound phytase activity (mU/mgd.w.) detected using whole cells grown on medium containing phytate.
(7.1 mU/mgd.w.), Rhodotorula diobovata (7.99 mU/mgd.w.),
and Torulaspora delbrueckii (6.1 mU/mgd.w.).
Analysis of Conserved Domain
Aminoacidic sequences of phytase were recovered from
RefSeq/GenBank database on https://www.ncbi.nlm.nih.gov/,
except for D. hansenii Mo40 and Bio2, which were sequenced
in this work. All strains analyzed possess a protein that shows
a good homology with phytase of D. hansenii Mo40 and
Bio2, and some of them, like M. guilliermondii,C. jadinii,
and K. marxianus, exhibit even two sequences encoding for
this enzyme. In the case of K. unispora, we could not identify
any sequence even if this strain can make a duplication on
MMPhy medium. Probably, this ability is due to the presence of
a phosphatase that does not specifically cleave phosphate from
phytic acid (Table 1).
Generally, HAP phytase brings two common domains: a
conserved N-terminal heptapeptide active site RHGXRXP (aa
38–44) and a C-terminal catalytically active dipeptide HD
(aa 325–326) (Mullaney and Ullah, 2003). In the analyzed
phytase sequences, the C-terminal domain is always conserved.
N-terminal domain is maintained in all except in R. toruloides,
in which the active site shows mismatch XHGHRXP, leading
us to conclude that all belong to HAP family. These sequences
were used to build a phylogenetic tree (Figure 2). In the red
box, we included sequences that contain a signal peptide for
protein secretion. Literature data (Li et al., 2009) report that
K. ohmeri enzyme contains this sequence, but the tool we
used was not able to recognize it. As reported in Figure 2, all
sequences showing a signal peptide can be clustered, suggesting
that extracellular localization can be phylogenetically related. The
fact that we detected extracellular activity only in C. jadinii,
could probably depend on the use of culture supernatants
without any step to concentrate it. It is possible to speculate
that extracellular phytase activity in other strains was too low
to be detected. This hypothesis is corroborated also by studies
that report in other species extracellular activity detected only in
concentrated samples (Olstorpe et al., 2009;Ragon et al., 2009;
Hellström et al., 2015).
Regulation of Phytase Expression
To carry out this investigation, we selected three species,
C. jadinii,K. marxianus, and T. delbrueckii that showed
high phytase activity and could be interesting for food-related
applications. C. jadinii and K. marxianus are in fact included in
QPS EFSA list (Koutsoumanis et al., 2019), and T. delbrueckii is a
wine starter with commercial name (BIODIVATM—Lallemand).
Phytase activity was analyzed by cultivating yeast cells in the
presence of different phosphorus sources: phytic acid (in MMPhy
medium), phosphate salt (KH2PO4in MMP medium), mixture
of phytic acid and phosphate salt (in MMP/Phy medium), and
also in rich medium (YPD). This allowed to understand the
regulation of phytase expression based on the type of phosphate
available (Table 2).
When C. jadinii cells were cultivated in medium containing
phytic acid as the sole phosphorus source, we detected
58.36 mU/mgd.w.of cell-bound activity (Table 2) and
26.25 mU/mgd.w.of extracellular activity. These activities
were not appreciable when inorganic phosphate was the only
source (MMP), suggesting that under this condition, the enzyme
was not expressed. In addition, the concomitant presence of
phytic acid and KH2PO4(MMP/Phy medium) was not able
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org 5May 2021 | Volume 9 | Article 662598
fbioe-09-662598 May 18, 2021 Time: 17:20 # 6
Capusoni et al. Phytase From Cyberlindnera jadinii
FIGURE 2 | Phylogenetic tree of phytase aminoacidic sequences. The red rectangle includes proteins that contain a signal sequence for secretion (detected using
SignalIP-5.0).
to induce phytase activity (Table 2). On the contrary, when
C. jadinii was cultivated in the presence of organic phosphate
(YPD medium), a reduced level of cell-bound activity was
detected: it decreased in fact from 58.36 mU/mgd.w., measured
TABLE 2 | Effects of phosphorus source on phytase activity.
YPD MMPhy MMP MMP/Phy
C. jadinii 2.02 ±0.4 58.36 ±6.24 BDL BDL
T. delbrueckii 2.39 ±0.29 6.1 ±1.25 BDL BDL
K. marxianus 1.71 ±0.35 4.17 ±1.17 BDL BDL
Cell-bound activity (mU/mgd.w.) was assayed at 37C, pH 4.5. BDL, below
detection limit.
in MMPhy, to 2.02 mU/mgd.w.in YPD (Table 2). Under this
condition, extracellular activity was under the detection limit.
The same behavior was observed in T. delbrueckii and
K. marxianus. The highest activities (6.1 and 4.17 mU/mgd.w.,
respectively) were detected in the presence of sole phytic acid
and decreased in YPD (Table 2). As observed in C. jadinii, also
in T. delbrueckii and in K. marxianus, the presence of phosphate
inhibited expression of phytase, being no activity detectable in
cells growing in MMP as well as in MMP/Phy.
In conclusion, a high level of phytase activity can be
expressed only when cells grow using phytate as sole phosphorus
source. On the contrary, the presence of inorganic phosphate
completely inhibits expression of phytase. This happens in
fact also in medium with concomitant presence of phosphate
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org 6May 2021 | Volume 9 | Article 662598
fbioe-09-662598 May 18, 2021 Time: 17:20 # 7
Capusoni et al. Phytase From Cyberlindnera jadinii
FIGURE 3 | Effects of temperature on phytase activity, measured at pH 4.5. (A) Cell-bound (full line) and extracellular (dotted line) activity detected on C. jadinii.
(B) Cell-bound activity detected on T. delbruekii and K. marxianus.
and phytic acid. This indicates that sole presence of phytic
acid is not enough to induce phytase activity, and lead us to
conclude that lack of inorganic phosphate is requested for its
expression. The presence of low activity in cells grown in medium
not containing phosphate salt, namely YPD, corroborates
this hypothesis.
An analogous phenomenon has been observed previously
by Olstorpe and colleagues in 2009 (Olstorpe et al., 2009).
They reported that in some Candida species, in Pichia anomala
and in S. cerevisiae, phytase activity was repressed in presence
of inorganic phosphate, but this did not occur in other
species like Arxula adeninivorans, as well as in Cryptococcus
laurentii (Pavlova et al., 2008). By employing mutagenesis, an
improved strain with reduced phosphate repression was obtained
in Pichia kudriavzevii (Qvirist et al., 2017). Understanding
which role plays phosphorus source on expression of phytase
activity is pivotal for set-up of industrial processes, due
to the fact that feed/food matrices as well as cultivation
media based on agrifood wastes can contain different types
of this nutrient.
Characterization of Phytase Activity
Phytases suitable for feed/food-related processes need to work
under conditions present in the upper digestive tract of
monogastric animals, and/or need to be resilient during
production processes. To work in the digestive tract, a
good phytase should exhibit a pH optimum between 3.5
and 5.5 and high activity at 37C. Resilience at higher
temperature can be requested because heat treatments are
commonly adopted to contain spoilage and during pelleting
processes. This treatment permits also the incorporation of
ingredients in the feed to “lock” the feed mixture. Unfortunately,
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org 7May 2021 | Volume 9 | Article 662598
fbioe-09-662598 May 18, 2021 Time: 17:20 # 8
Capusoni et al. Phytase From Cyberlindnera jadinii
FIGURE 4 | Effect of pH on phytase activity detected in C. jadinii. Activity was
measured at 37C.
heat treatment could reduce phytase activity, and for this
reason, it is important to select a thermostable enzyme
(Mrudula Vasudevan et al., 2019).
In order to select enzymes suitable for these applications, the
effects of temperature and pH were investigated (Figures 3,4).
The optimal temperature for phytase activity in C. jadinii
was found to be 50C. When the assay was performed at
this temperature, cell-bound activity reached 146 mU/mgd.w.
and extracellular activity was 51.95 mU/mgd.w.At 60C, the
values for cell-bound and extracellular activity were lower:
105.2 and 37.2 mU/mgd.w., respectively (Figure 3A). Similar
results were obtained for K. marxianus. Also in this case, the
optimum of temperature was observed at 50C, with activity
of 7.11 mU/mgd.w.A similar temperature activity profile was
found for phytase from Rhodotorula mucilaginosa JMUY14 (Yu
et al., 2015). On the other hand, in T. delbrueckii, the highest
activity of 6.1 mU/mgd.w.was detected at 37C (Figure 3B).
Data reported in Figure 3 lead us to hypothesize that phytase
activity in C. jadinii could be suitable for industrial purpose. High
phytase activity detected even at 60C suggests that this enzyme
could be resilient also at high temperatures used during feeds
production. In addition, at 37CC. jadinii phytase activity is
higher in comparison with K. marxianus and T. delbrueckii ones.
With the aim to investigate the effect of pH on C. jadinii
activity, we selected temperature of 37C (Figure 4). Even if
this temperature is not the optimum for C. jadinii, it is the
one requested to work at gastric level. As reported in Figure 4,
pH optimum for this phytase was 4.5, similarly to phytases
found in other yeasts (In et al., 2009;Caputo et al., 2015;
Ogunremi et al., 2020). As reported by Lei et al. (2013), this is
an important characteristic for the development of enzymes as
feed/food additive.
Comparing our results with some data reported in literature,
it is possible to observe that phytase activity found in C. jadinii
could be promising for future applications (Supplementary
Table 3). To the best of our knowledge, C. jadinii cell-bound
activity is one of the highest observed on cells grown in mineral
media with phytate as sole phosphorus sources.
In conclusion, we think that these results found for C. jadinii
phytase activity could represent a good starting point to
set-up optimization of cultural conditions, in order to improve
phytase production. As reported, with a statistical approach
for medium optimization, phytase production can be easily
increased (Puppala et al., 2018). The productivity of phytase
using Aspergillus niger under submerged fermentation conditions
was improved by 3.97 times employing a statistical media
optimization strategy and Box-Behnken experimental designs
(Shah et al., 2017). Optimization of temperature, pH, and
aeration allowed the successful production of phytase with
A. ficuum in submerged fermentation as opposed to the
traditional solid-state fermentation (Coban and Demirci, 2014).
In S. cerevisiae modulation of media components, like addition
of magnesium sulfate, manganese sulfate, and ferrous sulfate, and
scaling up in 10 L fermenter could increase phytase activity from
45 to 164 mU/mgd.w..InK. marxianus, phytase activity could be
easily increased adding to fermentation media cheap substrates
rich in phytic acid like rice bran (Pires et al., 2019). Similar
behavior was observed in W. anomalus, where the presence of
cane molasses in media can increase enzymatic activity from 6
up to 176 mU/mgd.w.reducing enzyme production cost from
0.25 to 0.006 £/1,000 U (Vohra and Satyanarayana, 2004). In
our case, performing media optimization could be the right
approach in order to increase phytase productivity, reducing its
production cost.
Furthermore, Cruz and colleagues demonstrated that biomass
of C. jadinii can partially replace feed protein content (generally
consistent in soybean meal, fish meal, rapeseed meal) in swine
and poultry formulation (Cruz et al., 2020a,b). The possibility
to have active enzymes in feed could be very appealing in order
to decrease their phytate content. This phenomenon has been
observed in W. anomalus, whose biomass-containing phytase
was added to aquaculture feed (1,000 U/kg feed), with results
comparable with commercial phytase (Vohra et al., 2011). The
safety statement of C. jadinii could open the possibility for
applications of this no-GMO microorganism also in food-related
processes to produce functional foods (Handa et al., 2020).
CONCLUSION
The combined effect of phytate as antinutritional factor and as
cause of environmental pollution makes phytase an industrially
interesting target. We identified new phytase activities in “safe”
yeasts, like C. jadinii,K. marxianus, and T. delbrueckii. In
particular, C. jadinii shows the highest enzymatic activity
localized both extracellularly and cell-bound. Our results suggest
that this phytase is suitable as additive in feed/food-related
processes. Indeed, its activity showed characteristics in terms of
temperature and pH suitable to work efficiently under conditions
compatible with the upper digestive tract of monogastric animals
as well as to be used in feed industrial production processes. The
safety statement of C. jadinii could open the possibility for its
application to reduce phytate content in food matrices.
Furthermore, a released phytase significantly increases the
interaction between phytate present in food matrix and the
enzyme. In addition, an extracellular enzyme improves greatly
the downstream during industrial production.
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org 8May 2021 | Volume 9 | Article 662598
fbioe-09-662598 May 18, 2021 Time: 17:20 # 9
Capusoni et al. Phytase From Cyberlindnera jadinii
DATA AVAILABILITY STATEMENT
The original contributions presented in the study are included
in the article/Supplementary Material, further inquiries can be
directed to the corresponding author/s.
AUTHOR CONTRIBUTIONS
ClC: investigation and writing the original draft. IS: investigation
and review and editing the draft. SD: investigation. CoC:
conceptualization, supervision, and review and editing the
draft. All authors contributed to the article and approved the
submitted version.
SUPPLEMENTARY MATERIAL
The Supplementary Material for this article can be found
online at: https://www.frontiersin.org/articles/10.3389/fbioe.
2021.662598/full#supplementary-material
REFERENCES
Caputo, L., Visconti, A., and De Angelis, M. (2015). Selection and use of a
Saccharomyces cerevisae strain to reduce phytate content of wholemeal flour
during bread-making or under simulated gastrointestinal conditions. LWT-
Food Sci. Technol. 63, 400–407. doi: 10.1016/j.lwt.2015.03.058
Coban, H. B., and Demirci, A. (2014). Screening of phytase producers and
optimization of culture conditions for submerged fermentation. Bioprocess
Biosyst. Eng. 37, 609–616. doi: 10.1007/s00449-013-1028-x
Coban, H. B., and Demirci, A. (2017). “Phytase as a diet ingredient: from microbial
production to its applications in food and feed industry,” in Handbook of Food
Bioengineering, Microbial Production of Food Ingredients and Additives, eds
A. M. Holban and A. M. Grumezescu (Cambridge: Academic Press), 33–55.
doi: 10.1016/B978-0- 12-811520-6.00002-7
Cruz, A., Sterten, H., Steinhoff, F. S., Mydland, L. T., and Øverland, M. (2020a).
Cyberlindnera jadinii yeast as a protein source for broiler chickens: effects on
growth performance and digestive function from hatching to 30 days of age.
Poultry Sci. 99, 3168–3178. doi: 10.1016/j.psj.2020.01.023
Cruz, A., Tauson, A. H., Matthiesen, C. F., Mydland, L. T., and Øverland, M.
(2020b). Cyberlindnera jadinii yeast as a protein source for growing pigs: Effects
on protein and energy metabolism. Livest. Sci. 231:103855. doi: 10.1016/j.livsci.
2019.103855
De Angelis, M., Gallo, G., Corbo, M. R., McSweeney, P. L. H., Faccia, M., Giovine,
M., et al. (2003). Phytase activity in sourdough lactic acid bacteria: Purification
and characterization of a phytase from Lactobacillus sanfranciscensis CB1. Int.
J. Food Microbiol. 87, 259–270. doi: 10.1016/S0168-1605(03)00072-2
Fliegerová, K., Mrázek, J., and Voigt, K. (2006). Differentiation of anaerobic
polycentric fungi by rDNA PCR-RFLP. Folia Microbiol. 51, 273–277. doi: 10.
1007/BF02931811
Greppi, A., Krych, Ł, Costantini, A., Rantsiou, K., Hounhouigan, D. J., Arneborg,
N., et al. (2015). Phytase-producing capacity of yeasts isolated from traditional
African fermented food products and PHYPK gene expression of Pichia
kudriavzevii strains. Int. J. Food Microbiol. 205, 81–89. doi: 10.1016/j.
ijfoodmicro.2015.04.011
Handa, V., Sharma, D., Kaur, A., and Arya, S. K. (2020). Biotechnological
applications of microbial phytase and phytic acid in food and feed industries.
Biocatal. Agric. Biotechnol. 25:101600. doi: 10.1016/j.bcab.2020.101600
Hellström, A., Qvirist, L., Svanberg, U., Veide Vilg, J., and Andlid, T. (2015).
Secretion of non-cell-bound phytase by the yeast Pichia kudriavzevii TY13.
J. Appl. Microbiol. 118, 1126–1136. doi: 10.1111/jam.12767
In, M. J., Seo, S. W., Kim, D. C., and Oh, N. S. (2009). Purification and biochemical
properties of an extracellular acid phytase produced by the Saccharomyces
cerevisiae CY strain. Process Biochem. 44, 122–126. doi: 10.1016/j.procbio.2008.
10.006
Kaur, P., Singh, B., Böer, E., Straube, N., Piontek, M., Satyanarayana, T., et al.
(2010). Pphy-A cell-bound phytase from the yeast Pichia anomala: Molecular
cloning of the gene PPHY and characterization of the recombinant enzyme.
J. Biotechnol. 149, 8–15. doi: 10.1016/j.jbiotec.2010.06.017
Kłosowski, G., Mikulski, D., and Jankowiak, O. (2018). Extracellular phytase
production by the wine yeast S. cerevisiae (Finarome strain) during submerged
fermentation. Molecules 23:848. doi: 10.3390/molecules23040848
Koutsoumanis, K., Allende, A., Alvarez-Ordóñez, A., Bolton, D., Bover-Cid, S.,
Chemaly, M., et al. (2019). Update of the list of QPS-recommended biological
agents intentionally added to food or feed as notified to EFSA 10: Suitability
of taxonomic units notified to EFSA until March 2019. EFSA J. 17:e05753.
doi: 10.2903/j.efsa.2019.5753
Leenhardt, F., Levrat-Verny, M. A., Chanliaud, E., and Rémésy, C. (2005).
Moderate decrease of pH by sourdough fermentation is sufficient to reduce
phytate content of whole wheat flour through endogenous phytase activity.
J. Agric. Food Chem. 53, 98–102. doi: 10.1021/jf049193q
Lei, X. G., Weaver, J. D., Mullaney, E., Ullah, A. H., and Azain, M. J. (2013).
Phytase, a new life for an “old” enzyme. Annu. Rev. Anim. Biosci. 1, 283–309.
doi: 10.1146/annurev-animal- 031412-103717
Li, X., Liu, Z., Chi, Z., Li, J., and Wang, X. (2009). Molecular cloning,
characterization, and expression of the phytase gene from marine yeast
Kodamaea ohmeri BG3. Mycol. Res. 113, 24–32. doi: 10.1016/j.mycres.2008.07.
003
Merico, A., Sulo, P., Piškur, J., and Compagno, C. (2007). Fermentative lifestyle
in yeasts belonging to the Saccharomyces complex. FEBS J. 274, 976–989. doi:
10.1111/j.1742-4658.2007.05645.x
Mrudula Vasudevan, U., Jaiswal, A. K., Krishna, S., and Pandey, A. (2019).
Thermostable phytase in feed and fuel industries. Bioresour. Technol. 278,
400–407. doi: 10.1016/j.biortech.2019.01.065
Mullaney, E. J., and Ullah, A. H. J. (2003). The term phytase comprises several
different classes of enzymes. Biochem. Biophys. Res. Commun. 312, 179–184.
doi: 10.1016/j.bbrc.2003.09.176
Nakashima, B. A., McAllister, T. A., Sharma, R., and Selinger, L. B. (2007). Diversity
of phytases in the rumen. Microb. Ecol. 53, 82–88. doi: 10.1007/s00248-006-
9147-4
Nuobariene, L., Hansen, A. S., Jespersen, L., and Arneborg, N. (2011). Phytase-
active yeasts from grain-based food and beer. J. Appl. Microbiol. 110, 1370–1380.
doi: 10.1111/j.1365-2672.2011.04988.x
Ogunremi, O. R., Agrawal, R., and Sanni, A. (2020). Production and
characterization of volatile compounds and phytase from potentially probiotic
yeasts isolated from traditional fermented cereal foods in Nigeria. J. Genet. Eng.
Biotechnol. 18:16. doi: 10.1186/s43141-020-00031-z
Olstorpe, M., Schnürer, J., and Passoth, V. (2009). Screening of yeast strains for
phytase activity. FEMS Yeast Res. 9, 478–488. doi: 10.1111/j.1567-1364.2009.
00493.x
Pavlova, K., Gargova, S., Hristozova, T., and Tankova, Z. (2008). Phytase from
antarctic yeast strain Cryptococcus laurentii AL27. Folia Microbiol. 53, 29–34.
doi: 10.1007/s12223-008- 0004-3
Pires, E. B. E., de Freitas, A. J., Souza, F. F. E., Salgado, R. L., Guimarães, V. M.,
Pereira, F. A., et al. (2019). Production of fungal phytases from agroindustrial
byproducts for pig diets. Sci. Rep. 9, 1–9. doi: 10.1038/s41598-019- 45720-z
Puppala, K. R., Naik, T., Shaik, A., Dastager, S., Kumar, V. R., Khire, J., et al.
(2018). Evaluation of Candida tropicalis (NCIM 3321) extracellular phytase
having plant growth promoting potential and process development. Biocatal.
Agric. Biotechnol. 13, 225–235. doi: 10.1016/j.bcab.2017.12.013
Qvirist, L., Vorontsov, E., Veide Vilg, J., and Andlid, T. (2017). Strain improvement
of Pichia kudriavzevii TY13 for raised phytase production and reduced
phosphate repression. Microb. Biotechnol. 10, 341–353. doi: 10.1111/1751-7915.
12427
Ragon, M., Hoh, F., Aumelas, A., Chiche, L., Moulin, G., and Boze, H. (2009).
Structure of Debaryomyces castellii CBS 2923 phytase. Acta Crystallogr F. Struct.
Biol. Cryst. Commun. 65, 321–326. doi: 10.1107/S1744309109008653
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org 9May 2021 | Volume 9 | Article 662598
fbioe-09-662598 May 18, 2021 Time: 17:20 # 10
Capusoni et al. Phytase From Cyberlindnera jadinii
Ragon, M., Neugnot-Roux, V., Chemardin, P., Moulin, G., and Boze, H. (2008).
Molecular gene cloning and overexpression of the phytase from Debaryomyces
castellii CBS 2923. Protein Expr. Purif. 58, 275–283. doi: 10.1016/j.pep.2007.12.
003
Reddy, N. R., Sathe, S. K., and Salunkhe, D. K. (1982). Phytates in legumes and
cereals. Adv. Food Res. 28, 1–92. doi: 10.1016/S0065-2628(08)60110-X
Rizzello, C. G., Nionelli, L., Coda, R., De Angelis, M., and Gobbetti, M. (2010).
Effect of sourdough fermentation on stabilisation, and chemical and nutritional
characteristics of wheat germ. Food Chem. 119, 1079–1089. doi: 10.1016/j.
foodchem.2009.08.016
Shah, P. C., Kumar, V. R., Dastager, S. G., and Khire, J. M. (2017). Phytase
production by Aspergillus niger NCIM 563 for a novel application to degrade
organophosphorus pesticides. AMB Express 7:66. doi: 10.1186/s13568-017-
0370-9
Shi, J., Wang, H., Hazebroek, J., Ertl, D. S., and Harp, T. (2005). The maize low-
phytic acid 3 encodes a myo-inositol kinase that plays a role in phytic acid
biosynthesis in developing seeds. Plant J. 42, 708–719. doi: 10.1111/j.1365-
313X.2005.02412.x
Shimizu, M. (1992). Purification and characterization of phytase from Bacillus
subtilis (natto) N-77. Biosci. Biotechnol. Biochem. 56, 1266–1269. doi: 10.1271/
bbb.56.1266
Vats, P., and Banerjee, U. C. (2004). Production studies and catalytic properties
of phytases (myo-inositolhexakisphosphate phosphohydrolases): An overview.
Enzyme Microb. Technol. 35, 3–14. doi: 10.1016/j.enzmictec.2004.03.010
Vohra, A., Kaur, P., and Satyanarayana, T. (2011). Production, characteristics
and applications of the cell-bound phytase of Pichia anomala.Antonie Van
Leeuwenhoek 99, 51–55. doi: 10.1007/s10482-010- 9498-1
Vohra, A., and Satyanarayana, T. (2004). A cost-effective cane molasses medium
for enhanced cell-bound phytase production by Pichia anomala.J. Appl.
Microbiol. 97, 471–476. doi: 10.1111/j.1365-2672.2004.02327.x
Wyss, M., Brugger, R., Kronenberger, A., Rémy, R., Fimbel, R., Oesterhelt, G.,
et al. (1999). Biochemical characterization of fungal phytases (myo-inositol
hexakisphosphate phosphohydrolases): Catalytic properties. Appl. Environ.
Microbiol. 65, 367–373. doi: 10.1128/aem.65.2.367-373.1999
Yu, P., Wang, X., and Liu, J. (2015). Purification and characterization of a novel
cold-adapted phytase from Rhodotorula mucilaginosa strain JMUY14 isolated
from Antarctic. J. Basic Microbiol. 55, 1029–1039. doi: 10.1002/jobm.201400865
Conflict of Interest: The authors declare that the research was conducted in the
absence of any commercial or financial relationships that could be construed as a
potential conflict of interest.
Copyright © 2021 Capusoni, Serra, Donzella and Compagno. This is an open-access
article distributed under the terms of the Creative Commons Attribution License
(CC BY). The use, distribution or reproduction in other forums is permitted, provided
the original author(s) and the copyright owner(s) are credited and that the original
publication in this journal is cited, in accordance with accepted academicpractice. No
use, distribution or reproduction is permitted which does not comply withthese terms.
Frontiers in Bioengineering and Biotechnology | www.frontiersin.org 10 May 2021 | Volume 9 | Article 662598
... In another study, 50°C was shown to be an appropriate temperature for a bacterial strain named DR 6 to produce 387 U/ml of phytase (9). The optimum temperature for phytase activity in Cyberlindnera jadinii CJ2 was found to be 50°C with a highest activity of 51.95 mU/mgdw, while the activity decreased with increase in temperature (24). Capusoni et al., (24) have shown pH 4.5 as optimum for highest phytase activity by Cyberlindnera jadinii CJ2. ...
... The optimum temperature for phytase activity in Cyberlindnera jadinii CJ2 was found to be 50°C with a highest activity of 51.95 mU/mgdw, while the activity decreased with increase in temperature (24). Capusoni et al., (24) have shown pH 4.5 as optimum for highest phytase activity by Cyberlindnera jadinii CJ2. According to Alves et al., (25) Muscodor sp. was shown to synthesize phytase at pH 5.0. ...
Article
Full-text available
Introduction and Aim: Phytic acid (myo-inositol 1, 2, 3, 4, 5, 6- hexakis dihydro­gen phosphate), a storage form of phosphorus is present in legumes, cereals and oilseeds. By forming insoluble complexes with proteins and divalent cations, it serves as an anti-nutrient in animal feed. Phytase (EC 3.1.3.8), a special class of phosphomonoesterases, catalyse the conversion of inorganic phosphate into mono-, di-, tri-, tetra-, and penta-phosphates of derivatives of myo-inositol. It is an important enzyme in food and feed industries. The present study aimed at isolation, molecular identification and optimization of physico-chemical parameters for phytase production by a red pigmented yeast Cystobasidium minutum. Materials and Methods: The phytase activity was estimated by using sodium phytate as substrate and the production of phytase under varied temperature, pH, agitation, incubation time, carbon and nitrogen sources. Results: The results showed maximum activity of 91.86 and 27.13 U/ml at 35°C and pH 5.5. At an agitation speed of 150 rpm and 120h of incubation time the enzyme activity was 18.9 and 21.4 U/ml respectively. Among the carbon sources tested sucrose served for highest enzyme activity of 26.33 U/ml and ammonium sulphate served the sole source of nitrogen and showed an activity of 50.12 U/ml. Conclusion: Cystobasidium minutum (Rhodotorula minutum) produced maximum phytase enzyme in an optimized physical and chemical condition. Hence, from the present investigation, it is found that optimization of physic-chemical conditions may be a promising tool for the growth of organisms and also maximum yield of any metabolite.
... The ANOVA results from the Duncan's test showed that the F-value of FeSO 4 Á7H 2 O was 1917.335, which was highly significant. Merico et al. [43] and Capusoni et al. [44] reported the use of MgSO 4 at 0.5 to 5 g/L for yeast phytase production. ...
... Capusoni et al. [43] reported that 20 g/L of yeast extract was optimum for yeast phytase. Another report by Kammoun et al. [44] tested several organic and inorganic nitrogenous sources and the highest phytase titer was attained in the presence of 0.75 % (w/v) yeast extract using Bacillus subtilis US417 [16]. In addition, K 2 HPO 4 was varied in five levels ranging from 0 to 45 g/L. ...
Article
Medium components and their solubility for improvement of recombinant phytase production in Escherichia coli DH5α. was optimized and reported in this study. Solubility study was supported by the computational works conducted using COSMO-RS, a quick tool to predict the thermophysical and chemical properties of fluid mixture. Solubility concept through thermodynamic calculation was introduced to study the interaction between medium components with water and selected medium components with each other. Based on the σ-profile and σ-potential, together with Gibbs free energy, a less soluble medium component showed a negative effect on phytase production due to the formation of insoluble metal-ion phytate complexes. Highest phytase production was obtained under optimum conditions of as 30 g/L yeast extracts, 1 g/L (NH4)2HPO4, 3 g/L MgSO4.7H2O, 0.3 g/L FeSO4.7H2O and 0.1 % (v/v) glycerol. Both experimental and computational works were conducted to investigate the interaction between medium components. The melting point and phase appearance of the mixture were studied. From experimental results, the mixtures were seemed to be unstable due to its higher melting point compared to individual medium components. The analysis from Gibbs free energy reflected the total immiscibility between medium components. It was concluded that medium components worked individually to promote the phytase production. Yeast extract, MgSO4.7H2O and FeSO4.7H2O were selected for optimization by central composite design, whereas other factors were maintained at their optimal level. Statistical analysis showed that the optimum media containing 34.06 g/L yeast extract, 3.6 g/L MgSO4.7H2O and 0.32 g/L FeSO4.7H2O gave the maximum phytase production of 118.91 U/mL. The optimization of process condition such as temperature, agitation speed and seed age were further investigated by employing Full Factorial Experimental (FUFE) Design after fixing the media composition. The maximum phytase production of 200 U/mL was obtained at optimum condition (Temperature 37 °C, agitation speed 500 rpm and seed age 1.55 at OD600nm). The phytase production in a larger scale was conducted by submerged fermentation in 2 L and 30 L bioreactor. In large scale production, constant tip speed suited the best for scale-up strategy comparing to the constant power number. The optimum process condition improved the phytase production and plasmid stability.
... Even though some of the already described microbial phytases have found industrial applications, they are still unable to meet all the requirements of the feed industry. The search for new phytases with high activity and stability at temperatures above 37°C and acidic pH, accompanied by low production cost, is the subject of increased scientific interest [9]. Yeasts are good candidates for enzyme production due to their ease of cultivation, rapid growth, and genetic stability. ...
Article
Full-text available
Phytases, which perform the stepwise hydrolysis of phytic acid to myo-inositol and inorganic phosphate, are used worldwide to reduce phosphorus pollution and improve nutrition in monogastric animals and humans. Yeasts isolated from their natural environments represent rich and still underexplored sources of industrially valuable enzymes, including phytases; therefore, they are widely studied for the production of these enzymes. In this regard, thirteen yeast pure cultures were isolated from the microbial consortium of four types of sourdough obtained during the natural fermentation of different grain-based flours. Ten of the newly isolated yeast strains were selected as potential phytase producers based on their growth in liquid culture media with sodium phytate as the sole source of phosphorus. Using 18S rDNA and D1/D2 26S rDNA analyses, the species affiliation of the selected isolates was established. They referred to seven yeast species from 3 families, with the most significant representation of the family Saccharomycetaceae.Intracellular phytate-degrading activity was found in 8 isolates, the highest being in Nakaseomyces glabratus strain 7-4. The highest level of extracellular phytase was measured in Pichia membranifaciens strain 5-2. Both isolates showed significant antioxidant capacity higher than those of ascorbic acid.
... Reliable screening methods have been reported to elucidate the ability of different yeast strains to utilize phytic acid as the sole source of phosphorus (Olstorpe et al., 2009;Capusoni et al., 2021). Furthermore, a link between exoelectrogenicity and phytase production has also been established (Hubenova et al., 2014;Hubenova et al., 2014). ...
Article
Research into phytase production is useful for improving the efficiency of animal production, reducing envi- ronmental impact, and contributing to the development of sustainable and efficient animal production systems. This study aims to investigate the potential of yeast strains for phytase biosynthesis in nutrient media. Phytase is a phosphomonoesterase (E.C 3.1.3.8) catalyzing in a ladder-like manner the dephosphorylation of phytic acid and its salts, with various resulting myo-inositol phosphates and phosphoric acid. Yeasts of the genera Saccha- romyces, Zygosaccharomyces, Candida, and Pichia were evaluated in a two-step screening procedure for phytase production. One hundred and eighteen strains were screened in the first stage, which was conducted on four types of solid culture media containing calcium phytate as the selected background. On PSM medium, many strains were found to form halos as early as the 24th hour of development. Several strains with significant po- tential for enzyme production were evaluated in the second step of the screening. It was conducted in a liquid culture medium. In conclusion, the strain C. melibiosica 2491 was selected for further studies when cultured in a YPglu culture medium. Further research will focus on finding suitable conditions that increase the biosynthesis of the enzyme, which is of significant technological and practical interest for animal nutrition.
... These enzymes show different localizations: in some species, phytases are extracellular enzymes, and in others are cell-bound or released in the periplasmic space. The extracellular phytase and cell-bound activity in Cyberlindnera jadinii was detected together with cell-bound activity in Kluyveromyces marxianus, and Torulaspora delbrueckeii (Capusoni et al., 2021). Extracellular phytase in S. cerevisiae was reported by some authors (Roopashri and Varadaraj, 2015;Kłosowski et al., 2018). ...
Article
Full-text available
Phytases are a large group of enzymes that hydrolyze phytate and its complexes. This most abundantorganic phosphate in the world is commonly found in plant-based foods. It can bind to essential minerals,making them less available for absorption. Enzymatic hydrolysis of phytates is the most beneficial methodfor reducing their content in foods and feeds. Phytase supplementation enables more efficient utilization ofphytate phosphorus. The enzyme is produced by prokaryotic and eukaryotic microorganisms, plants, andanimals. Several types of phytases, depending on certain structural and kinetic properties are described.Phytase activity is influenced by metal ions, surfactants, and various plant extracts.
... The advanced developments in molecular engineering especially in gene mining aid to identify an extensively active phytase (LpPHY233) from the Lactobacillus plantarum. The catalytic and temperature efficiency of the same has been achieved through the 'C-terminal truncation' and 'disulfide bond engineering' respectively (Capusoni et al., 2021). ...
Article
Phytases are a group of digestive enzymes which are commonly used as feed enzymes. These enzymes are used exogenously in the feeds of monogastric animals thereby it improves the digestibility of phosphorous and thus reduces the negative impact of inorganic P excretion on the environment. Even though these enzymes are widely distributed in many life forms, microorganisms are the most preferred and potential source of phytase. Despite the extensive availability of the phytase-producing microbial consortia, only a few microorganisms have been known to be exploited at industrial level. The high costs of the enzyme along with the incapability to survive high temperatures followed by the poor storage stability are noted to be the bottleneck in the commercialization of enzymes. For this reason, besides the conventional fermentation approaches, the applicability of cloning, expression studies and genetic engineering has been implemented for the past few years to accomplish the abovesaid benefits. The site-directed mutagenesis as well as knocking out have also validated their prominent role in microbe-based phytase production with enhanced levels. The present review provides detailed information on recent insights on the modification of phytases through heterologous expression and protein engineering to make thermostable and protease-resistant phytases.
... The advanced developments in molecular engineering especially in gene mining aid to identify an extensively active phytase (LpPHY233) from the Lactobacillus plantarum. The catalytic and temperature efficiency of the same has been achieved through the 'C-terminal truncation' and 'disulfide bond engineering' respectively (Capusoni et al., 2021). ...
... D. hansenii is able to grow using phytic acid as a sole phosphorus source [45]. To characterize phytase activity in the Mo40 strain, we started sequencing the gene encoding in order to compare a marine phytase sequence with its terrestrial counterpart (reference strain CBS 767, isolated from beer [46]). ...
Article
Full-text available
The possibility to perform bioprocesses with reduced ecological footprint to produce natural compounds and catalyzers of industrial interest is pushing the research for salt tolerant microorganisms able to grow on seawater-based media and able to use a wide range of nutrients coming from waste. In this study we focused our attention on a Debaryomyces hansenii marine strain (Mo40). We optimized cultivation in a bioreactor at low pH on seawater-based media containing a mixture of sugars (glucose and xylose) and urea. Under these conditions the strain exhibited high growth rate and biomass yield. In addition, we characterized potential applications of this yeast biomass in food/feed industry. We show that Mo40 can produce a biomass containing 45% proteins and 20% lipids. This strain is also able to degrade phytic acid by a cell-bound phytase activity. These features represent an appealing starting point for obtaining D. hansenii biomass in a cheap and environmentally friendly way, and for potential use as an additive or to replace unsustainable ingredients in the feed or food industries, as this species is included in the QPS EFSA list (Quality Presumption as Safe—European Food Safety Authority).
Article
Full-text available
Phosphorous actively participates in numerous metabolic and regulatory activities of almost all living organisms including animals and humans. Therefore, it is considered as an essential macronutrient required supporting their proper growth. On contrary, phytic acid (PA), an anti-nutritional substance is widely known for its strong affinity to chelate essential mineral ions including PO43-, Ca2+, Fe2+, Mg2+ and Zn2+. Being one the major reservoir of PO43- ions, PA has great potential to bind PO43- ions in diverse range of foods. Once combined with P, PA transform into an undigested and insoluble complex namely phytate. Produced phytate leads to a notable reduction in the bio-availability of P due to negligible activity of phytases in monogastric animals and humans. This highlights the importance and consequent need of enhancement of phytase level in these life-forms. Interestingly, phytases, catalyzing the breakdown of phytate complex and recycling the phosphate into ecosystem to its available form, have naturally been reported in a variety of plants and microorganisms over past few decades.In pursuit of a reliable solution, the focus of this review is to explore the keynote potential of bacterial phytases for sustainable management of phosphorous via efficient utilization of soil phytate. The core of the review covers detailed discussion on bacterial phytases along with their widely reported applications viz. biofertlizers, phosphorus acquisition and plant growth promotion. Moreover, meticulous description on fermentation based strategies and future trends on bacterial phytases have also been included. This article is protected by copyright. All rights reserved.
Article
Full-text available
Background: Probiotic strains are incorporated into food substrates to contribute to fermentation process. The technological suitability of such strains to improve the flavor and nutritional value of fermented food is strain-specific. Potentially probiotic yeasts isolated from Nigerian traditional fermented foods were assessed for production of volatile compounds by gas chromatography-mass spectrophotometry. Phytases were characterized for activity and stability at different pH (3-8) and temperatures (25-50 °C). Results: A total of 45 volatiles compounds were identified from intracellular cell-free extracts of Pichia kluyveri LKC17, Issatchenkia orientalis OSL11, P. kudriavzevii OG32, P. kudriavzevii ROM11, and Candida tropicalis BOM21. They include alcohols (14), carbonyls (13), esters (10), and organic acids (8). Phenylethyl alcohol was the highest higher-alcohol in Issatchenkia orientalis OSL11 (27.51 %). The largest proportion of esters was detected in P. kudriavzevii OG32 (17.38 %). Pichia kudriavzevii OG32 and C. tropicalis BOM21 showed vigorous gowth in minimal medium supplemented with sodium phytate (2 g L-1). Extracellular phytases from P. kudriavzevii OG32 and Candida tropicalis BOM2 showed optimal activiy at pH 4.6 (104.28 U) and pH 3.6 (81.43 U) respectively. Conclusions: Results obtained revealed species- and strain-specific potentials of the yeast strains to improve flavor and mineral bioavailability of fermented food products. Therefore, the application of these yeasts as starter cultures during food fermentation process is a very promising method to enhance the flavor profile and enhance mineral bioavailability in indigenous cereal-based fermented food products.
Article
Full-text available
Europe is heavily dependent on imported feed protein sources such as soybean meal (SBM); thus, investigating local sustainable alternatives is crucial to increase self-sufficiency. This study evaluated the effects of the inactivated yeast Cyberlindnera jadinii grown on local lignocellulosic sugars on the growth performance and digestive function of Ross 308 broiler chickens. A total of 1,000 male chicks were allocated to 20 pens. There were 5 replicate pens with 50 birds each, from 1 to 30 D after hatch. The birds were offered one conventional wheat–oat–SBM–based control diet and 3 diets with increasing levels of C. jadinii replacing 10, 20, and 30% of dietary crude protein (CP), whereas SBM levels were gradually decreased. The feed intake and weight gain of the birds decreased linearly, and feed conversion ratio increased linearly (P < 0.01) with increasing dietary levels of C. jadinii. Nevertheless, growth performance and feed intake were similar between the birds fed with control diets and diets containing 10% CP from C. jadinii in the starter and grower periods. The apparent ileal digestibility (AID) of dry matter, crude fat, organic matter, and carbohydrates was higher in control diets than in diets with 30% C. jadinii CP (P < 0.05) and decreased (P < 0.01) with incremental levels of dietary C. jadinii. Regardless, the AID of CP, starch, ash, and phosphorus was unaffected. Ileal villus height on day 10 was maintained in birds fed with diets containing 30% C. jadinii CP compared with the birds fed with control diets but was lower for birds fed with diets containing 10 and 20% C. jadinii protein (P < 0.05). To conclude, up to 10% C. jadinii CP can replace SBM CP in broiler chicken diets, maintaining growth performance and digestive function, whereas higher levels of C. jadinii may decrease bird performance. Altogether, this suggests the potential of C. jadinii as a local-based protein source in broiler chicken diets, contributing to a more sustainable feed.
Article
Full-text available
Inactivated Cyberlindnera jadinii yeast (previously classified as Candida utilis) produced from local lignocellulosic biomass-based sugars is an alternative protein source in diets for young pigs. The objective of this study was to evaluate the effects of diets containing C. jadinii on the nitrogen and energy metabolism and apparent total tract digestibility (ATTD) of major nutrients and energy in young pigs. Twenty-four intact boars with a mean initial body weight of 16.7 ± 4.5 kg were assigned to four diets: a conventional control diet for young pigs with soybean meal, fish meal, rapeseed meal and potato protein concentrate as major protein sources or one of three experimental diets containing 10, 20, and 40%% of crude protein (CP) from C. jadinii. The pigs were equally distributed to the dietary treatments according to initial body weight and litter, comprising a total of six replicates per diet. The experiment was carried out during three periods of 14 days. Each period comprised an energy and nitrogen balance experiment of four consecutive days including a 22 h respiration experiment by means of indirect calorimetry. Adding C. jadinii to diets did not affect the ATTD of nutrients and energy in the diets. The energy and nitrogen metabolism was not affected by partially replacing the main protein sources with C. jadinii. Collectively, the results indicate that CP from C. jadinii can replace up to 40% of dietary CP from conventional protein sources while maintaining the efficiency of nitrogen and energy metabolism in young pigs.
Article
Full-text available
Abstract The qualified presumption of safety (QPS) procedure was developed to provide a harmonised generic pre‐evaluation to support safety risk assessments of biological agents performed by EFSA's Scientific Panels. The taxonomic identity, body of knowledge, safety concerns and antimicrobial resistance were assessed. Safety concerns identified for a taxonomic unit (TU) are, where possible and reasonable in number, reflected by ‘qualifications’ which should be assessed at the strain level by the EFSA's Scientific Panels. During the current assessment, no new information was found that would change the previously recommended QPS TUs and their qualifications. The list of microorganisms notified to EFSA from applications for market authorisation was updated with 47 biological agents, received between October 2018 and March 2019. Of these, 19 already had QPS status, 20 were excluded from the QPS exercise by the previous QPS mandate (11 filamentous fungi) or from further evaluations within the current mandate (9 notifications of Escherichia coli). Sphingomonas elodea, Gluconobacter frateurii, Corynebacterium ammoniagenes, Corynebacterium casei, Burkholderia ubonensis, Phaeodactylum tricornutum, Microbacterium foliorum and Euglena gracilis were evaluated for the first time. Sphingomonas elodea cannot be assessed for a possible QPS recommendation because it is not a valid species. Corynebacterium ammoniagenes and Euglena gracilis can be recommended for the QPS list with the qualification ‘for production purposes only’. The following TUs cannot be recommended for the QPS list: Burkholderia ubonensis, due to its potential and confirmed ability to generate biologically active compounds and limited of body of knowledge; Corynebacterium casei, Gluconobacter frateurii and Microbacterium foliorum, due to lack of body of knowledge; Phaeodactylum tricornutum, based on the lack of a safe history of use in the food chain and limited knowledge on its potential production of bioactive compounds with possible toxic effects.
Article
Full-text available
The application of phytases for animal feed in developing countries is limited due to the high cost of these enzymes, determined by the importation fees and the expensive substrates used for their production. In this work, we have used agroindustrial byproducts for the production of extracts containing phytases, which were accessed for their stability focusing on the conditions found in the gastrointestinal tract of pigs. The fungus Acremonim zeae presented higher phytase production in medium containing cornmeal, while the yeast Kluyveromyces marxianus produced 10-fold more phytase when cultivated on rice bran. Process optimization increased the difference in productivity to more than 300 fold. The phytase from A. zeae was thermostable, with higher activity at neutral pH and 50 °C, but was inhibited at pH 2.5 and by various ions. The phytase activity in the K. marxianus extract was stable at a wide range of conditions, which indicates the presence of at least two enzymes. As far as we know, this manuscript describes for the first time the phytase production and the characteristics of the extracts produced by both these microbial species. These enzymes could be produced at low cost and have potential to replace enzymes currently imported for this purpose.
Article
Full-text available
One of the key steps in the production of phytases of microbial origin is selection of culture parameters, followed by isolation of the enzyme and evaluation of its catalytic activity. It was found that conditions for S. cerevisiae yeast culture, strain Finarome, giving the reduction in phytic acid concentration of more than 98% within 24 h of incubation were as follows: pH 5.5, 32 °C, continuous stirring at 80 rpm, the use of mannose as a carbon source and aspartic acid as a source of nitrogen. The highest catalytic activity of the isolated phytase was observed at 37 °C, pH 4.0 and using phytate as substrate at concentration of 5.0 mM. The presence of ethanol in the medium at a concentration of 12% v/v reduces the catalytic activity to above 60%. Properties of phytase derived from S. cerevisiae yeast culture, strain Finarome, indicate the possibility of its application in the form of a cell's free crude protein isolate for the hydrolysis of phytic acid to improve the efficiency of alcoholic fermentation processes. Our results also suggest a possibility to use the strain under study to obtain a fusant derived with specialized distillery strains, capable of carrying out a highly efficient fermentation process combined with the utilization of phytates.
Article
The enzyme phytases facilitates in degradation of phytate. Phytate as a natural compound serving as primary source for storing phosphate among plants. From the biotechnological prospects, there has been a considerable leap in the Enzyme technology, which has massively broadened the commercial aspects of phytase. Their impact in the food and feed industry has become much more quintessential in the recent times. For nearly two decades, there has been a wide array of commercially available microbial phytases in market with commercial significance as it facilitates the farmers with essential. Phytases in particular cannot be neglected from being a threat for human diet due to its anti-nutrient activity as they served as strong chelating agent against many divalent minerals. Similar to phytases activity, PA also was found to showcase a potential towards binding positively charged proteins, amino acids, and/or multivalent cations or minerals in foods. Besides the food industry has overlooked on the very fact of phytase significance as its supplementation results in improving the net availability of the essential trace elements and minerals to humans. Similarly they serve as an essential feed source for mongastric animals.
Article
Phytase with wide ranging biochemical properties has long been utilized in a multitude of industries, even so, thermostability plays a crucial factor in choosing the right phytase in a few of the sectors. Mesophilic phytases are not considered to be a viable option in the feed industry owing to its limited stability in the required feed processing temperature. In the recent past, inclusion of thermostable phytase in fuel ethanol production from starch based raw material has been demonstrated with economic benefits. Therefore, considerable emphasis has been placed on using complementary approaches such as mining of extremophilic microbial wealth, encapsulation and using enzyme engineering for obtaining stable phytase variants. This article means to give an insight on role of thermostable phytases in feed and fuel industries and methods for its development, highlighting molecular determinants of thermostability.
Article
Phytase is known to provide a solution for depletion of phosphorus (P). It helps it by hydrolyzing the insoluble P source in soil which is phytate. In this study, provides insight on yeast Candida tropicalis (NCIM 3321) which produces cell bound and extracellular thermostable phytase. The media components were optimized to enhance the enzyme production and checked for plant growth promoting activity. On optimization the isolate exhibited enhanced cell bound and extracellular phytase activity by four folds (from 236 to 1024 IU DCG⁻¹) and by five folds (from 0.46 to 1.95 IU ml⁻¹) respectively in 36 h. The production time decreased to 24 h compare to shake flask on Up-scaling the production process upto 10 L scale, thus increasing the productivity of cell bound (1810 IU DCG⁻¹day⁻¹) and extracellular phytase (6.08 IU ml⁻¹ day⁻¹). The crude phytase (12 IU) from NCIM 3321 strain was studied for plant growth promotion activity in lab scale and field level experiments with maize crop. Findings of the study revealed that the extracellular phytase derived from non pathogenic C. tropicalis (NCIM 3321) was found to be plant growth stimulating by increasing the available P in soil. Our findings of phytase isolated from non-pathogenic yeast C. tropicalis NCIM 3321 exhibited dephytinization potential. Therefore, current study may have profound application in sustainable agriculture.