Anda di halaman 1dari 61

BIOTEKNOLOGI DALAM PEMBUATAN WINE RENDAH ALKOHOL

Eklesia Vida Kurnia (200802101)


Fakultas Teknobiologi, Universitas Atma Jaya Yogyakarta.

PENDAHULUAN
Wine yang dijual di pasaran telah mengalami peningkatan kadar alkohol rata-rata
sekitar 1% (v/v) setiap 10 tahun, dan hingga saat ini telah terjadi peningkatan kadar total
3-4%. Hal ini dapat terjadi akibat pemanasan global yang mempengaruhi proses
pematangan buah anggur, dimana akumulasi gula yang lebih besar. Kandungan gula yang
tinggi dalam buah anggur matang, menyebabkan kadar alkohol tinggi karena semua
kandungan gula akan difermentasi oleh yeast.
Kadar alkohol yang terlalu tinggi dapat menjadi masalah bagi industri wine. Dari
segi sensori, tingginya kadar alkhol akan meningkatkan rasa pahit yang berlebihan, serta
meningkatkan rasa manis, astringency, dan panas, sehingga menutupi rasa buah anggur itu
sendiri. Dari segi kesehatan, tingginya kadar alkohol pada wine meningkatkan risiko
kesehatan konsumen dan mengancam keselamatan dalam berkemudi. Selain itu, dari segi
ekonomi, kadar alkohol yang makin tinggi akan meningkatkan pajak di beberapa negara
yang berorientasi pada kadar alkohol produk. Oleh karena itu, perlu dilakukan tindakan
untuk mencegah kadar alkhol yang terlalu tinggi pada akhir proses fermentasi wine.

PENDEKATAN BIOTEKNOLOGI
Dalam beberapa tahun terakhir, telah dilakukan ada upaya pendekatan bioteknolgi untuk
membatasi kandungan akohol dalam wine. Pendekatan yang cukup popular yaitu dari sisi
mikrobiologi, seperti rekayasa genetika yeast dan pencarian spesies dan galur yeast yang
tepat dalam proses fermentasi.

1. Penghapusan PDC2 pada Saccharomyces cerevisiae


Salah satu proses dalam fermentasi yaitu piruvat menjadi asetaldehida, yang
dikatalisis oleh piruvat dekarboksilase (PDC). PDC2 merupakan pengkode faktor
transkripsi yang terlibat dalam gen PDC. Penghapusan PDC2, pengkodean faktor
transkripsi yang terlibat dalam ekspresi gen PDC, menghasilkan aktivitas PDC 19%,
dengan pengurangan yang jelas dalam kandungan etanol, dikombinasikan dengan
peningkatan pelepasan gliserol dan piruvat. Gliserol berkontribusi positif dalam
viskositas dan rasa manis wine. Efek ini ditingkatkan dengan peningkatan aktivitas
gliserol-3-fosfat dehidrogenase, melalui overekspresi GPD1, tanpa produksi berlebih
dari asam asetat dan asetaldehida.

2. Mutasi Faktor Transkripsi Spt15p


Global transcriptional machinery engineering (gTME) digunakan untuk
mengembangkan strain S. cerevisiae dengan kemampuan produksi etanol pada proses
fermentasi berkurang. Sasaran teknologi ini adalah memutasi faktor transkripsi umum
Spt15p (TATA binding protein) yang berperan dalam kerja RNA polimerase dan
sebagai salah satu protein pengikat DNA utama yang mengatur spesifisitas promotor
dalam ragi. Strain yang digunakan yaitu S. cerevisiae YS59 (MATα; ura3-52, leu2-3,
dan 5-519).
Gen SPT15 ditransfer ke plasmid pada situs restriksi antara BamHI dan EcoRI
menggunakan vektor pY16, yang diapit oleh promotor TEF1 dan terminator CYC1
(pY16-SPT15). Dilakukan mutagenesis acak (error-prone PCR) dari gen SPT15
menggunakan kit DiversifyTM PCR Random Mutagenesis (Clontech) dengan pY16-
SPT15 sebagai templat. Plasmid yang diperoleh ditransfer ke Escherichia coli JM109,
kemudian koloni dipilih secara acak untuk identifikasi mutasi.
Plasmid yang mengandung SPT15 termutasi, kembali ditransfer ke S. cerevisiae
YS59 dengan metode lithium asetat, kemudian diinkubasi pada 25◦C pada medium SD
padat untuk menghasilkan yeast library for SPT15 mutant. Setelah masa inkubasi,
diambil 20 mutan dengan biomassa tertinggi. Dipilih biomassa tertinggi karena strain
dengan biomassa tingi umumnya memiliki kapasitas produksi etanol yang lebih rendah
akibat persaingan untuk sumber karbon antara produksi biomassa dan etanol.
Dilakukan uji kapasitas produksi etanol dengan memfermentasi masing-masing dalam
10 mL media Triple M, kemudian strain dengan hasil etanol terendah diidentifikasi
dan diapatkan strain YS59-409. Strain ini dibandingkan dengan strain kontrol,
menunjukkan produksi etanol rendah yaitu berkurang secara signifikan sebesar 34,9%.

3. Mutasi Faktor Transkripsi Pdc2p


Sasaran teknologi ini merupakan protein transkripsi Pdc2p. Faktor transkripsi
Pdc2p yang diubah secara struktural dapat menunjukkan penurunan aktivitas regulasi
positif PDC1 dan PDC5. Penurunan aktivitas regulasi ini mengarah pada pengurangan
moderat aktivitas enzimatik PDC dan akibatnya pengurangan produksi etanol.
Protein Pdc2p mutasi, pdc2Δ519, diintegrasikan ke dalam genom yeast melalui
rekombinasi homolog. Yeast yang digunakan adalah Saccharomyces cerevisiae
BY4743. Strain mutan BY4743pdc2Δ519 menunjukkan penurunan sekitar 1-2% (v/v)
kandungan etanol wine, tanpa ada pengaruh signifikan terhadap konsentrasi sisa gula
dan asam asetat. Selanjutnya, dilakukan pemeriksaan terhadap fenotipe untuk dapat
direproduksi pada jenis yeast lain yang sering digunakan pada fermentasi wine pada
skala industri. Mutasi pdc2Δ519 diintegrasikan dengan transformasi dan rekombinasi
homolog ke dalam genom EC1118 komersial dan ragi anggur asli Mab2C.
Setelah penyisipan satu salinan mutasi pdc2Δ519, diamati kurva pertumbuhan
dalam kondisi aerobik dan vinifikasi skala lab untuk membandingkan EC1118Δ519
dan Mab2CΔ519 yang dimodifikasi secara genetik dengan galur kontrol berupa wild
type. Hasil menunjukkan bahwa mutan EC1118Δ519 dan Mab2CΔ519 tumbuh normal
di media YPD, serta tidak menunjukkan perbedaan dengan kontrol. Analisis parameter
fermentasi menunjukkan bahwa EC1118Δ519 dapat menghasilkan pengurangan total
etanol sebesar 15%, sedangkan Mab2CΔ519 dapat menghasilkan pengurangan total
etanol sebesar 10%, dibandingkan dengan kontrol. Konsentrasi asam asetat
EC1118Δ519 dan Mab2CΔ519 juga tidak terpengaruh oleh mutasi.

4. Delesi Gen ADH


Langkah enzimatik terakhir dalam pembentukan etanol yaitu kerja dari enzim alkohol
dehidrogenase (ADH). Penghapusan gen pengkode ADH menyebabkan pengalihan
seluruh karbon untuk digunakan dalam pembentukan gliserol dan asetal-dehida.
Namun, teknologi ini memiliki kekurangan karena memiliki sensori yang kurang baik
akibat tingginya kadar asetal dehida.

5. Penggunaan Spesies Saccharomyces Non-Konvensional


Salah satu spesies Saccharomyces yang dapat digunakan yaitu Saccharomyces
uvarum. S. uvarum telah terbukti menghasilkan wine dengan konsentrasi alkohol
(<0,7% v/v) dan asiditas volatil (asam asetat) yang lebih rendah dibandingkan wine
yang difermentasi dengan S. cerevisiae. Wine yang diproduksi dengan S. uvarum
memiliki atribut sensori “off-flavor” yang unik dan tidak dimiliki wine yang diproduksi
dengan S.cerevisiae. Namun, atribut ini dapat menjadi suatu hal negatif dalam industri
karena rasa yang berbeda dari wine konvensional.
6. Penggunaan Jenis Yeast Lain
Jenis yeast lain yang digunakan untuk menghasilkan wine rendah alkohol yaitu
Metschnikowia pulcherrima (FLAVIA MP346). Metschnikowia pulcherrima
diinokulasikan terlebih dahulu, kemudian setelah 48 jam, diinokulasikan
Saccharomyces bayanus, kemudian setelah 96 jam diinokulasikan Saccharomyces
cerevisiae. Hasil fermentasi dengan 3 yeast ini menghasilkan produksi anggur dengan
kandungan etanol yang rendah (penurunan 0,9% v/v dibandingkan wine kontrol).
Namun, teknologi ini memiliki kekurangan karena terdapat perbedaan kandungan
senyawa aromatik tertentu, serta rasa dan aroma yang sering dianggap negatif oleh
industri.

SIMPULAN
Berdasarkan studi pustaka yang telah dilakukan, diketahui bahwa penerapan bioteknologi,
khususnya pendekatan mikrobiologi yang paling efektif untuk produksi wine rendah
alkohol adalah dengan melakukan mutasi pada faktor transkripsi Pdc2p. Hal ini
dikarenakan teknologi mutasi Pdc2p berhasil menurunkan kadar alkohol pada produk
akhir wine pada berbagai strain, baik skala lab maupun skala industri. Teknologi ini
berhasil menurunkan kadar alkohol dalam wine sebesar 1-15% dibandingkan wine
komersil umumnya.

DAFTAR PUSTAKA
Cuello, R. A., Montero, K. J. F., Mercado, L. A., Combina, M. dan Ciklic, I. F. 2017.
Construction of low-ethanol–wine yeasts through partial deletion of the
Saccharomyces cerevisiae PDC2 gene. AMB Express 7 (67): 1-11.

Du, Q., Liu, Y., Song, Y. dan Qin, Y. 2020. Creation of a low-alcohol-production yeast by
a mutated SPT15 transcription regulator triggers transcriptional and metabolic
changes during wine fermentation. Frontiers in Microbiology 11 (597828): 1-10.

Gonzalez, R., Guindal, A. M., Tronchoni, J. dan Morales, P. 2021. Biotechnological


approaches to lowering the ethanol yield during wine fermentation. Biomolecules
11 (1569): 1-16.
Puškaš, V. S., Miljić, U. D., Djuran, J. J. dan Vučurović, V. M. 2019. The aptitude of
commercial yeast strains for lowering the ethanol content of wine. Food Science
& Nutrition 2020 (8):1489–1498.

Varela, J. dan Varela, C. 2019. Microbiological strategies to produce beer and wine with
reduced ethanol concentration. Current Opinion in Biotechnology 2019 (56):88–
96.
Cuello et al. AMB Expr (2017) 7:67
DOI 10.1186/s13568-017-0369-2

ORIGINAL ARTICLE Open Access

Construction of low‑ethanol–wine yeasts


through partial deletion of the Saccharomyces
cerevisiae PDC2 gene
Raúl Andrés Cuello1,2, Karina Johana Flores Montero2, Laura Analía Mercado2, Mariana Combina1,2
and Iván Francisco Ciklic2* 

Abstract 
We propose an alternative GMO based strategy to obtain Saccharomyces cerevisiae mutant strains with a slight reduc-
tion in their ability to produce ethanol, but with a moderate impact on the yeast metabolism. Through homologous
recombination, two truncated Pdc2p proteins Pdc2pΔ344 and Pdc2pΔ519 were obtained and transformed into
haploid and diploid lab yeast strains. In the pdc2Δ344 mutants the DNA-binding and transactivation site of the protein
remain intact, whereas in pdc2Δ519 only the DNA-binding site is conserved. Compared to the control, the diploid
BY4743pdc2Δ519 mutant strain reduced up to 7.4% the total ethanol content in lab scale-vinifications. The residual
sugar and volatile acidity was not significantly affected by this ethanol reduction. Remarkably, we got a much higher
ethanol reduction of 10 and 15% when the pdc2Δ519 mutation was tested in a native and a commercial wine yeast
strain against their respective controls. Our results demonstrate that the insertion of the pdc2Δ519 mutation in wine
yeast strains can reduce the ethanol concentration up to 1.89% (v/v) without affecting the fermentation performance.
In contrast to non-GMO based strategies, our approach permits the insertion of the pdc2Δ519 mutation in any locally
selected wine strain, making possible to produce quality wines with regional characteristics and lower alcohol con-
tent. Thus, we consider our work a valuable contribution to the problem of high ethanol concentration in wine.
Keywords:  Yeast, Ethanol, Metabolism, Genetic engineering, Wine

Introduction global warming, this effect is particularly exacerbated in


Nowadays there is a growing demand for softer wines regions with hot summers (Mira de Orduña 2010). A high
with reduced ethanol content. However, during the last concentration of alcohol in wine can have many negative
twenty years there has been an increment in the alcohol consequences. The quality of the product may be com-
concentration of wine of about 2% (v/v) (Kutyna et  al. promised, e.g. the perception of viscosity and hotness
2010; Tilloy et  al. 2014). Current viticultural practices could be enhanced, in detriment of acidity, aroma, inten-
favor the harvest of very mature grapes to obtain wines sity of flavors, sweetness and other organoleptic proper-
with sweet tannins, as demanded by the consumers. ties (Gawel et al. 2007a, b). Production costs may rise in
The biosynthesis of polyphenols is usually delayed with countries where taxes are applied according to the etha-
respect to sugar production, leading to a harvest of grapes nol content. Sluggish or stuck fermentations might also
with high sugar amounts. This sugar excess in turn, pro- happen as a result of yeast inhibition provoked by high
duces wines with high ethanol. As a consequence of ethanol levels (Buescher et  al. 2001). In addition, con-
sumption of wine with high ethanol content can also have
a negative impact on health. This combination of quality,
*Correspondence: ciklic.ivan@inta.gob.ar
2
economic and health problems associated with high alco-
Laboratorio de Biotecnología, Estación Experimental Agropecuaria
Mendoza, Instituto Nacional de Tecnología Agropecuaria (INTA), San hol in wine, has promoted a significant interest in devel-
Martín 3853, Luján de Cuyo, Mendoza, Argentina oping technologies to reduce ethanol, but conserving all
Full list of author information is available at the end of the article

© The Author(s) 2017. This article is distributed under the terms of the Creative Commons Attribution 4.0 International License
(http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium,
provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license,
and indicate if changes were made.
Cuello et al. AMB Expr (2017) 7:67 Page 2 of 11

desirable sensory characteristics of wine (Kutyna et  al. high levels of oxidized by-products like acetic acid, ace-
2010). Indeed, this problem represents a major chal- toin and acetaldehyde, and this effect could not be com-
lenge for the wine industry in its attempt to counteract pletely neutralized with additional genetic modifications.
some of the negative effects of global warming. Differ- In the present work, we propose an alternative GMO
ent approaches have been tried, such as the application based strategy to obtain S. cerevisiae mutant strains with
of alternative viticultural practices (Kontoudakis et  al. a slight reduction in their ability to produce ethanol, but
2011), or the implementation of physical methods like with a moderate impact on the yeast metabolism. For this
dealcoholization (Bogianchini et  al. 2011). Neverthe- purpose, we designed two functional strategic mutations
less, the microbiological approach is probably the most at the C-terminus of the PDC2 gene in order to alter the
attractive because of its easy implementation and low structure of the encoded protein. PDC2 encodes a tran-
costs (Kutyna et al. 2010). The microbiological approach scription factor (Hohmann 1993) that regulates the avail-
includes the use of non-GMO (genetically modified ability of the pyruvate decarboxylase (PDC) isozymes
organisms) strategies like sequential inoculations and Pdc1p and Pdc5p, which catalyze the reaction of pyruvate
co-inoculations of S. cerevisiae with non-Saccharomyces to acetaldehyde in the ethanol biosynthetic pathway [the
yeasts, as well as GMO-based strategies with the use of main active form during glucose catabolism is PDC1 and
genetically modified yeasts. Considering the first case, PDC5 is a secondary form which is only expressed under
sequential inoculations and co-inoculations have become thiamine starvation (Mojzita and Hohmann 2006)]. The
quite popular in recent years (Comitini et al. 2011; Mag- full Δpdc2 deletion has already been tested to study the
yar and Tóth 2011; Di Maio et  al. 2012; Sadoudi et  al. genetic factors affecting the glycerol formation and its
2012 and reviewed in Ciani et  al. 2016). Unfortunately, overproduction (Nevoigt and Stahl 1996). Although the
low yields of ethanol are usually the result of wines with glycerol formation is enhanced without affecting the ace-
high residual sugar concentrations (Ciani and Ferraro tic acid concentration, the Δpdc2 deletion exhibits a phe-
1996; Ciani et al. 2006). In some cases, reduction in the notype incompatible for yeasts with oenological purposes
ethanol concentration varied only between 0.2 and 0.7% (drastic reduction of PDC specific activity and ethanol
(v/v) (Benito et al. 2011; Gobbi et al. 2013). However, in concentration, as well as strong inhibition of growth in
recent works, ethanol reduction between 1.5 and 2.2% aerobic conditions). A structurally altered version of the
(v/v) has been achieved through different strategies like Pdc2p transcription factor may display a reduced posi-
sequential fermentation with immobilized non-Saccharo- tive regulatory activity of PDC1 and PDC5, leading to a
myces yeasts (Canonico et  al. 2016) and the use of non- moderate reduction of the PDC enzymatic activity and
Saccharomyces yeasts combined with S. cerevisiae under consequently a reduction in ethanol production. Through
controlled aeration conditions (Contreras et  al. 2015; homologous recombination, two truncated Pdc2p pro-
Morales et al. 2015). With regards to GMO-based strate- teins lacking 344 (pdc2Δ344) or 519 (pdc2Δ519) amino
gies, most of the studies have concentrated in redirecting acids at the C-terminus were obtained and transformed
part of the carbon flux from the ethanol metabolic path- into lab yeast strains. In the case of pdc2Δ344 the DNA-
way to other secondary products like glycerol. The over- binding and transactivation site are both intact, whereas
expression of GPD1 and/or GDP2 genes, which encode in pdc2Δ519 only the DNA-binding site is conserved.
the glycerol-3-phosphate dehydrogenase isozymes, Subsequently, these mutants were tested in lab-scale vini-
reduce the ethanol and enhance the glycerol production fications to select low-ethanol yeasts. The selected muta-
(Remize et al. 1999; Lopes et al. 2000). However, the con- tion was then tested in both native and commercial wine
centrations of some undesirable by-products, particularly yeast strains.
acetic acid, are also increased. There were some efforts to
inhibit the formation of acetic acid by deleting the ALD6 Materials and methods
gene, but this increased the synthesis of other oxidized Strains, media and growth conditions
compounds like acetoin, which has negative organoleptic Saccharomyces cerevisiae laboratory strains BY4741
properties (Cambon, et al. 2006; Eglinton, et al. 2002). In (haploid, MATa; his3Δ 1; leu2Δ 0; met15Δ 0; ura3Δ 0)
a previous study, a large number of genetic modifications and BY4743 (diploid, MATa/MATα his3Δ 1/his3Δ 1;
were generated and evaluated in order to reduce ethanol leu2Δ0/leu2Δ 0; met15Δ 0/MET15; LYS2/lys2Δ 0; ura3Δ
concentration in wine (Varela et al. 2012). Using the same 0/ura3Δ 0) were used to construct the mutants and as
genetic background, 41 different alterations in different control strains during the vinification experiments. The
combinations were tested. Of all the strategies tried, the commercial wine yeast EC1118 (Lallemand, Denmark)
most successful to reduce ethanol were those designed to and the native Mab2C strain previously selected in our
increase glycerol formation. Still, like in previous works, laboratory, were used to test the pdc2Δ519 mutation in
the increase of glycerol formation was accompanied with native genetic backgrounds. All yeast strains were grown
Cuello et al. AMB Expr (2017) 7:67 Page 3 of 11

at 30  °C and 150  rpm in YPD medium (2% glucose, 2% flasks plugged with glass fermentation traps so that only
peptone and 1% yeast extract). YPD supplemented with CO2 could evolve from the system, and they were kept
200  mg/L of G-418 was used for selection and mainte- at 28  °C without agitation (Vaughan-Martini and Mar-
nance of transformants. tini 1998). All vinification experiments were performed
under the described conditions but comparing different
Construction and genomic integration of the pdc2Δ344 strains (V1, V2, V3 and V4). Fermentation kinetics was
and pdc2Δ519 mutations monitored by measuring the daily CO2 weight loss. Alco-
The pdc2Δ344 and pdc2Δ519 mutations were integrated hol concentration, acetic acid and residual sugar were
into the yeast genome through homologous recombi- measured according to standard methods (OIV 2015),
nation following the method described by Güldener whereas glycerol was measured with spectrophotometry
et  al. (1996). The disrupting fragment was obtained by (WinescanTM Foss, Hillerød, Denmark). Several derived
PCR amplification of the KanMx resistance cassette fermentative parameters such as carbon balance, glucose
present in the pUG6 plasmid (Güldener et  al. 1996). and ethanol yield were calculated (Vazquez et  al. 2001).
The primers used were, forward 5′-ACAGAATACT- Carbon balance was calculated as the ratio between
GTTGATAATAGTACCAAAA CAGGTAACCCTT- carbon moles of fermentation by-products and carbon
GAAGCTGAAGCTCGTACGC-3 for pdc2Δ344 or moles of glucose. Meanwhile, glucose yield results from
5′-TTGGGAT GATATACCCGTTGATGCTATCAAA- the amount of glucose required (g) to produce 1% (v/v) of
GCAAATTGGTGAAGCTGAAGCTTCGTACGC-3 for ethanol, and ethanol yield, from the ratio between grams
pdc2Δ519 with the reverse primer 5′-CTAAAAAAA- of produced ethanol and grams of consumed glucose. All
GCCTGTGT TACCAGGTAAGTGTAAGTTATTAG- assays were performed at least in triplicates (three inde-
CATAGGCCACTAGTGGATCTG-3′. A specific PDC2 pendent cultures).
homologous sequence was inserted at the 5′-end of
each primer to generate the recombination event at the Statistical analysis of data
expected C-terminal region of the PDC2 gene. Besides, An ANOVA and a LSD Fisher test with a p value <0.05
two stop codons were placed downstream to the homolo- was performed for the analysis and media comparison
gous region of each forward primer to generate the spe- of the growth, fermentative and kinetics parameters.
cific truncated proteins at the C-terminus. After the PCR Growth parameters as well as kinetic parameters were
reaction, the disrupting fragments were transformed by estimated using the reparameterized Gompertz equation
a slightly modified version of the lithium acetate method as proposed by Zwietering et al. (1990):
(Gietz and Woods 2002) selecting the G-418 resistant
y = D∗ exp −exp[((µmax ∗ e)/D) ∗ ( − t) + 1]
 
transformants. The correct insertion of the fragment was
confirmed by PCR using the forward primer 5′-GCGTG- where y = ln (ODt/OD0) OD0 is the initial OD and ODt is
GTCGACTCAAAACCAATAGCTGCTTAAAAA-3′ the OD at time t; D = ln (ODmax/DO0) is the curve maxi-
which binds upstream of the PDC2 gene and the reverse mum asymptotic, μmax is the maximum specific growth
5′-GGATGTATGGGCTAAATG-3′ which binds inside rate (1/h), and λ is the lag phase period (h). Growth and
the KanMx resistance cassette. kinetics data were fitted by nonlinear regression proce-
dure, minimizing the sum of squares of the difference
Growth curves in YPD rich medium between the experimental data and the fitted model.
Yeast cultures were inoculated into 10  mL YPD and
grown overnight. These cultures were then used to Results
inoculate 50 mL of YPD in 100 mL Erlenmeyer flasks at Growth of haploid and diploid pdc2Δ519 and pdc2Δ344
an initial OD600 of 0.2 (Spectrophotometer UV–visible mutants in aerobic conditions
T60U PG Instruments, Leicestershire, UK). The cultures The Δpdc2 complete deletion causes a drastic reduction
were grown with 150 rpm at 30 °C and aliquots of 100 µL of the PDC specific activity, an accumulation of pyruvate
were taken at different intervals for the measurement of and a strong inhibition of growth in aerobic conditions
the OD600. (Hohmann 1993; Nevoigt and Stahl 1996). The ability to
grow in aerobic conditions is essential for a yeast strain
Laboratory scale vinifications to be selected to develop wine yeast starters. Although
Following the recommendation of Vazquez et al. (2001), the conditions of wine fermentation are predominantly
Lab-scale vinifications were carried out using concen- anaerobic, the biomass production is performed in
trated white must as substrate, diluted to a final sugar aerobic conditions. Hence, before quantifying the fer-
concentration of 20°Bx and supplemented with 1  g/L of mentative parameters of each strain we wanted to test
yeast extract. The vinifications were performed in 500 mL the growth capability of the pdc2Δ519 and pdc2Δ344
Cuello et al. AMB Expr (2017) 7:67 Page 4 of 11

mutants under aerobic conditions with glucose as carbon 0.1% (w/v) of yeast extract (Vazquez et al. 2001). The con-
source. Figure 1 shows the growth curves in YPD medium centration of ethanol, acetic acid, glycerol and residual
for the haploid mutant strains BY4741pdc2Δ344 and sugar were determined, and other derived fermentative
BY4741pdc2Δ519 as well as the diploid BY4743pdc2Δ344 parameters like carbon balance and glucose yield were
and BY4743pdc2Δ519 with their respective BY4741 and calculated. We searched for a mutant strain with a slight
BY4743 controls. Remarkably, all mutants displayed inefficiency to produce ethanol in order to reduce the
growth pattern similar to the controls, and consequently alcohol content of wine between 1 and 2% (v/v). During
no statistical difference was detected among the kinetic the first vinification experiment (V1) the haploid and dip-
parameters analyzed (λ, and µmax) (data not shown). This loid pdc2Δ519 mutants were assayed against their respec-
result demonstrates that all mutant strains grow as good tive BY4741 and BY4743 controls (Table 1). Considering
as the controls in aerobic conditions, and therefore they the ethanol production, BY4741pdc2Δ519 showed no
are suitable for further experimentation. statistical difference when compared to the control,
whereas BY4743pdc2Δ519 displayed a statistically sig-
Quantification of fermentative parameters from lab‑scale nificant ethanol reduction (p  <  0.05) of around 7%. The
vinifications and selection of low ethanol pdc2 mutants moderate reduction observed in BY4743pdc2Δ519 was
In order to select low ethanol pdc2 mutant strains, we within the expected, however, it is surprising we didn’t
performed a series of microvinifications with concen- observe any phenotype for the haploid BY4741pdc2Δ519,
trated must diluted to 20°Bx and supplemented with which carries only the mutant copy of the PDC2 gene.

Fig. 1  Growth of parental and mutant laboratory strains in YPD at 30 °C and 150 rpm shaking. Each point represents the average value of two
independent cultures
Table 1  Fermentative parameters of lab-scale vinification assays in 20°Bx white must by parental and mutant laboratory yeast strains
Parameter BY4741 BY4741 BY4743 BY4743 BY4741 BY4741 BY4743 BY4743 BY4743
pdc2Δ519 pdc2Δ519 pdc2Δ344 pdc2Δ344 pdc2Δ519
Cuello et al. AMB Expr (2017) 7:67

Main compounds V1 V1 V1 V1 V2 V2 V2 V2 V2
(g/Liter)
 Consumed 153.95 ± 0.87A 146.49 ± 0.63B 147.22 ± 0.00B 143.84 ± 1.66C 172.04 ± 4.64B,C 154.90 ± 4.22D 179.56 ± 3.08A 170.98 ± 1.98C 176.05 ± 4.71A,B
sugar
 CO2 71.11 ± 1.92A 73.33 ± 5.77A 66.67 ± 0.00A,B 62.22 ± 3.85B 81.88 ± 3.89A,B 72.48 ± 2.52C 83.45 ± 1.60A 78.68 ± 2.84B 83.75 ± 2.20A
A A A B A C A A,B
 Ethanol 75.61 ± 0.91 75.35 ± 2.23 74.56 ± 1.58 69.04 ± 0.79 87.58 ± 5.69 76.007 ± 2.28 87.74 ± 1.41 84.27 ± 2.58 81.27 ± 1.48B
A A,B A,B B A B,C C B
 Acetate 1.11 ± 0.16 0.91 ± 0.06 0.88 ± 0.24 0.78 ± 0.07 1.11 ± 0.11 0.63 ± 0.06 0.58 ± 0.12 0.83 ± 0.22 0.60 ± 0.09C
b B B,C B A
 Glycerol ND ND ND ND 2.67 ± 0.20 2.44 ± 0.074 2.58 ± 0.14 3.16 ± 0.11 2.24 ± 0.18C
a
Balance (%)
 Carbon 97.18 ± 1.27B 101.43 ± 3.18A 97.64 ± 0.72A,B 95.26 ± 1.13B 99.32 ± 2.50A 97.33 ± 1.27A,B 96.40 ± 0.96A,B 96.67 ± 2.66A,B 95.18 ± 2.37B
Yield
 Ethanol produc- 9.58 ± 0.12A 9.55 ± 0.20A 9.45 ± 0.20A 8.80 ± 0.10B 11.1 ± 0.72A 9.63 ± 0.29C 11.12 ± 0.18A 10.68 ± 0.33A,B 10.30 ± 0.19B
tion (%[v/v])
 EtOH (g/g 0.49 ± 0.008A 0.51 ± 0.011A 0.51 ± 0.009A 0.48 ± 0.009A 0.51 ± 0.02A 0.49 ± 0.02A,B 0.49 ± 0.008B 0.49 ± 0.019A,B 0.46 ± 0.012C
glucose
consumed)
 Glucose (g) 16.07 ± 0.28A,B 15.34 ± 0.27C 15.58 ± 0.33B,C 16.35 ± 0.32A 15.52 ± 0.62B 16.09 ± 0.59B 16.15 ± 0.25B 16.02 ± 0.62B 17.09 ± 0.42A
required
for 1% (v/v)
ethanol
production
 Glucose (g) ND ND ND ND 65.64 ± 4.21B 69.49 ± 2.30A,B 64.56 ± 3.83B 54.57 ± 1.99C 73.64 ± 3.38A
required for 62.65 ± 0.87C 70.11 ± 0.63B 69.38 ± 0.00B 72.76 ± 1.66A 44.56 ± 4.64B,C 61.7 ± 4.22A 37.04 ± 3.08D 45.62 ± 1.98B 40.55 ± 4.71C,D
1 g/L glycerol
Residual sugar
(g/Liter)
ND not determined
Distinct letters correspond to statistical significant difference for a Fischer test with p < 0.05
a
  Carbon balance represents the ratio between carbon moles of fermentation by-products and carbon moles of glucose
b
  Glycerol was measured in an independent experiment
Page 5 of 11
Cuello et al. AMB Expr (2017) 7:67 Page 6 of 11

In correspondence with the observed ethanol reduc- not increased but reduced, showing statistically sig-
tion, BY4743pdc2Δ519 also presented the lowest value nificant differences with the control in the final concen-
of carbon balance and it was the least efficient to pro- tration and the glucose yield for glycerol. In contrast,
duce ethanol. It is important to note that this ethanol BY4743pdc2Δ344 which previously showed no ethanol
reduction didn’t cause an increase of the volatile acid- reduction produced more glycerol than the control with
ity. Taken together, this was a promising result and as a statistically significant higher values. At least for the
consequence BY4743pdc2Δ519 was selected for further Δpdc2 diploid mutants, there seems to be no clear cor-
experimentation. Continuing with the mutant’s char- relation between ethanol and glycerol production and the
acterization we performed a second vinification experi- ethanol reduction observed in BY4743pdc2Δ519 is not a
ment (V2), with the selected BY4743pdc2Δ519, the still consequence of an increment of glycerol concentration.
uncharacterized BY4741pdc2Δ344, BY4743pdc2Δ344d
strains, and both haploid BY4741 and BY4743 diploid Kinetics analysis of the vinifications by CO2 weight loss
controls (Table  1). The same trend was observed for The progression of the vinifications was daily moni-
the mutant BY4743pdc2Δ519, which again is the least tored by measuring the CO2 weight loss. Figure 2 shows
efficient to produce ethanol, showing statistically sig- the curves of accumulated CO2 weight loss obtained for
nificant lower values of glucose yield compared with the each vinification experiment. As seen in the graphs (pan-
control and the rest of the strains. With regards to the els a, b), all fermentations were very slow, lasting around
ethanol production of the pdc2Δ344 mutants, the hap- three weeks (an industrial fermentation performed with
loid produced significantly less ethanol than the BY4741 wine yeasts usually last 7–10  days). Nevertheless, this
control, whereas the mutant diploid showed no statisti- result is not surprising considering we used lab strains,
cal difference comparing with the diploid BY4743 con- which are not specialized for wine fermentation. We cal-
trol. Nonetheless, the ethanol reduction observed for culated for each curve the mean value of the three main
BY4741pdc2Δ344 was not reflected in the efficiency of kinetic parameters (Zwietering et  al. 1990) lag phase
the mutant strains, since its glucose yield was not sig- (λ), maximum CO2 weight loss speed (μmax) and total
nificantly different from the control. This compensation accumulated CO2 weight loss (A, for asymptote). After
in the glucose yield production is explained by the fact performing an ANOVA, the mean values of the kinetic
that BY4741pdc2Δ344 displayed considerably higher val- parameters were compared (Table  2). With few excep-
ues of residual sugar compared with BY4741. Up to this tions, the kinetics of the fermentations was very similar
point, the BY4743pdc2Δ519 mutant showed a consistent between the mutants and their controls. In the first vini-
phenotype of ethanol reduction being the least efficient fication (V1) there was no statistical difference between
in both vinifications. According to the value obtained for the haploid strains BY4741pdc2Δ519h and BY4741, but
the glucose yield, BY4743pdc2Δ519 would reduce the they lost more CO2 than the diploid strains. The diploid
ethanol almost 1 degree (0.85) for a wine with a predic- mutant BY4743pdc2Δ519 lost less CO2 than its control
tion of 15.5% (v/v). The BY4743pdc2Δ344 mutant showed but the difference was not statistically significant. Con-
an intermediate phenotype producing 4% less ethanol sidering that this strain also produces less ethanol, we
than the control, but there was no statistically significant expected a higher CO2 reduction for BY4743pdc2Δ519,
difference in the glucose yield. In summary, the vinifica- which was not the case. In the second vinification,
tion experiments showed that BY4743pdc2Δ519 is the BY4743pdc2Δ519 showed again similar values of CO2
most promising mutant strain displaying a small but con- and μmax when compared to the control. It is interest-
sistent reduction of the ethanol concentration, as a result ing to note that despite the ethanol reduction observed
of a slight inefficiency for its production. Importantly, for BY4743pdc2Δ519, there was little (vinification 1) or
this ethanol reduction did not cause an increment of the no difference (vinification 2) in the total CO2 weight loss
concentration of acetic acid, which is perhaps the most comparing with the BY4743 control. Apparently, another
undesirable by-product of the wine fermentation. decarboxylation reaction is compensating the CO2
Most of the efforts to deviate the carbon flux away from formed during the ethanol biosynthesis, and this is a clue
ethanol have concentrated in the production of glycerol, which could help us to reveal how the carbon flux has
a desirable secondary metabolite of fermentation. There- been modified in the BY4743pdc2Δ519 mutant strain.
fore, the glycerol produced in lab-scale vinifications was
also measured (Table  1). The quantification of glycerol Insertion of the pdc2Δ519 mutation in the commercial
could give as a clue of what is happening to the carbon EC1118 and native Mab2C wine yeast strains
flux of the pdc2Δ mutants, particularly BY4743pdc2Δ519 The BY4743pdc2Δ519 mutant strain showed a pheno-
which showed a consistent ethanol reduction. Surpris- type according to our aim of reducing around 1–2%
ingly, the glycerol production of BY4743pdc2Δ519 was (v/v) the ethanol content of wine, without significantly
Cuello et al. AMB Expr (2017) 7:67 Page 7 of 11

Fig. 2  Accumulated CO2 weight loss curves along the experiments of vinifications 1 (a), 2 (b), 3 (c) and 4 (d) with parental and mutant strains. The
evolution of each vinification was daily monitored by measuring the CO2 weight loss. Each point represents the average value of three independent
cultures

Table 2  Quantification and statistical analyses of fermentation parameters


Parameter BY4741 BY4741 BY4743 BY4743
pdc2Δ519 pdc2Δ519

Vinification 1
 A 21.96 ± 0.59A,B 22.83 ± 1.57A 21.31 ± 0.52A,B 19.79 ± 1.74B
 µmax 1.8 ± 0.05A 1.74 ± 0.13A 1.44 ± 0.05B 1.39 ± 0.01B
 λ 2.66 ± 0.26A 2.38 ± 0.01A 2.47 ± 0.39A 2.86 ± 0.40A
BY4741 BY4741 BY4743 BY4743 BY4743
pdc2Δ344 pdc2Δ344 pdc2Δ519

Vinification 2
 A 25.24 ± 1.10A,B 22.05 ± 0.68C 26.11 ± 0.49A 24.38 ± 0.88B 26.07 ± 0.74A
B,C C A,B B
 µmax 1.78 ± 013 1.64 ± 0.09 1.91 ± 0.06 1.80 ± 0.09 1.98 ± 0.06A
A A,B B A,B
 λ 2.97 ± 1.11 2.30 ± 0.17 1.80 ± 0.05 1.97 ± 0.05 2.04 ± 0.23B
A = Asymptote, µmax = maximum specific growth rate and (λ) = lag time, for the three vinifications
Distinct letters correspond to statistical significant difference for a Fischer test with p < 0.05
Cuello et al. AMB Expr (2017) 7:67 Page 8 of 11

Table 3 Determination of  fermentative parameters for  mutant and  wild type wine yeast strains EC1118 and  Mab2C
in two independent lab-scale vinifications
Parameter EC1118 EC1118Δ519 Mab2C Mab2CΔ519
V3 V3 V4 V4

Main compounds (g/L)


 Consumed sugar 210.38 ± 0.64A 203.95 ± 2.51B 212.03 ± 0.44A 209.01 ± 0.63B
A A A
 CO2 94.84 ± 2.05 91.56 ± 3.01 95.60 ± 3.50 93.51 ± 1.40A
A B A
 Ethanol 102.31 ± 2.99 87.05 ± 2.77 95.73 ± 4.05 86.00 ± 1.58B
A A A
 Acetate 1.13 ± 0.16 1.12 ± 0.19 0.83 ± 0.17 0.86 ± 0.19A
Balance (%)
 Carbon 94.42 ± 1.81A 88.81 ± 1.48B 90.82 ± 3.19A 86.78 ± 0.47A
Yield
 Ethanol production (% [v/v]) 12.97 ± 0.38A 11.03 ± 0.35B 12.13 ± 0.51A 10.90 ± 0.20B
A B A
 EtOH (g/g glucose consumed) 0.49 ± 0.013 0.43 ± 0.08 0.47 ± 0.021 0.42 ± 0.008B
A B B
 Glucose (g) required for 1% (v/v) ethanol production 16.23 ± 0.42 18.49 ± 0.36 17.50 ± 0.71 19.18 ± 0.32A
A B B
 Residual sugar (g/L) 6.22 ± 0.64 12.65 ± 2.51 4.57 ± 0.44 7.59 ± 0.63A
Distinct letters correspond to statistical significant difference for a Fischer test with p < 0.05

affecting the concentration of residual sugar and ace- well adapted wine yeasts. Interestingly, the mutants were
tic acid. This is already a positive result, but we should not affected by the mutation and their kinetics was very
bear in mind that all the previous experiments were per- similar to the wild type strains. The analysis of the fer-
formed with laboratory yeast strains which are geneti- mentative parameters revealed a remarkable 15% total
cally quite different from wine yeasts. Therefore, our ethanol reduction for the mutant EC1118Δ519 and 10%
next challenge was to check whether this phenotype for Mab2CΔ519 comparing with the wild type controls.
could be reproduced or even improved in a wine yeast Both EC1118Δ519 and Mab2CΔ519 mutants were also
genetic background. This way, we would obtain wine less efficient to produce ethanol showing statistically
yeasts suitable for industrial fermentation and capable significant lower values of glucose yield. According to
of reducing the ethanol content of wine. To test this, the the values obtained for the glucose yield, EC1118Δ519
pdc2Δ519 mutation was first integrated by transforma- would reduce the ethanol almost 2 degrees (1.89) and
tion and homologous recombination into the genome Mab2CΔ519 almost 1.5 degrees (1.36) for a wine with a
of the commercial EC1118 and the native Mab2C wine prediction of 15.5% (v/v). As it happened with the labo-
yeasts. Considering the ploidy of these strains, EC1118 ratory BY4743pdc2Δ519 mutant strain, the acetic acid
has been reported to be a diploid (Novo et al. 2009) and concentrations of EC1118Δ519 and Mab2CΔ519 were
according to a PCR test performed in our lab, Mab2C also unaffected by the mutation. The ethanol reduction
would be at least a diploid since both mating types were obtained with EC1118Δ519 was about two-fold higher
present. After the insertion of one pdc2Δ519 mutation than that of the laboratory BY4743pdc2Δ519 mutant
copy, we performed a growth curve in aerobic condi- strain. The Mab2CΔ519 native strain also displayed a
tions and lab-scale vinifications to compare the geneti- higher ethanol reduction (about 1.5-fold) when com-
cally modified EC1118Δ519 and Mab2CΔ519 with their pared with the laboratory BY4743pdc2Δ519 mutant
respective wild type control strains. The experiments strain. Thus, we were not only able to reproduce the phe-
were performed with the same conditions used for the notype observed in the lab strain, but also we improved
lab strains. With respect to the growth in aerobic condi- it. In agreement with the ethanol reduction displayed,
tions, both EC1118Δ519 and Mab2CΔ519 mutants grew the carbon balance of the mutants are considerable lower
normally in YPD medium showing no difference with than the wild type controls indicating that part of the car-
their respective wild type controls (data not shown). As bon flux is being redirected away from the ethanol bio-
for the vinification experiments V3 and V4, Fig. 2 shows synthesis pathway.
the curve for the accumulated CO2 weight loss (panels
c, d) and Table 3 summarizes the value obtained for the Discussion
main fermentative parameters. In contrast with the lab Before quantifying the fermentative parameters of each
strains, the vinifications were in this case much faster and laboratory mutant strain we tested the growth capability
lasted around eight days, which is the expected time for of the pdc2Δ519 and pdc2Δ344 mutants under aerobic
Cuello et al. AMB Expr (2017) 7:67 Page 9 of 11

conditions with glucose as carbon source. All mutants strain, the acetic acid concentrations of EC1118Δ519
were able to grow normally showing no difference with and Mab2CΔ519 were also unaffected by the mutation.
their respective controls. This was already a positive The ethanol reduction obtained with EC1118Δ519 and
result considering the inability of the full Δpdc2 deletion Mab2CΔ519 was about two-fold higher than that of the
to grow under such conditions (Hohmann 1993; Nevoigt laboratory BY4743pdc2Δ519 mutant strain. Thus, we
and Stahl 1996). Nevertheless, it is quite surprising that were not only able to reproduce the phenotype observed
none of the mutants showed a growth defect, especially in the lab strain, but also we improved it. These results
BY4741pdc2Δ519 that carries only the mutant version of demonstrate that the wine yeasts mutants EC1118Δ519
Pdc2p with a deletion of 519 amino acids which accounts and Mab2CΔ519 are good candidates to develop a yeast
for 56% of the wild type protein. Apparently, the activity starter for the elaboration of wines with reduced etha-
of the binding site alone is sufficient to sustain a normal nol content. Still, it will be necessary to determine how
growth of the yeast. the carbon flux is being redirected. In an exploratory
With respect to the mutant’s fermentative param- experiment, we determined the glycerol concentra-
eters, they were determined by lab-scale vinifications. tion for the laboratory strain BY4743pdc2Δ519 and we
The phenotypic analysis showed that BY4743pdc2Δ519 found no significant increment compared to the con-
is the most interesting mutant strain, displaying a con- trol. Perhaps, a clue could come from the analysis of
sistent reduction of the ethanol concentration of up to the CO2 weight loss data. Despite the ethanol reduc-
7.4%, as a result of a slight inefficiency for its production. tion observed for BY4743pdc2Δ519, EC1118Δ519 and
It is important to remark that this ethanol reduction did Mab2CΔ519 mutants, there was little or no difference in
not provoke an increment of the concentration of acetic the total CO2 weight loss comparing with the wild type
acid, in contrast to previous GMO based strategies where controls. It seems that another decarboxylation reac-
the carbon flux was diverted to glycerol (Remize et  al. tion is compensating the CO2 formed during the etha-
1999; Lopes et  al. 2000; Varela et  al. 2012). The moder- nol biosynthesis. Following this reasoning, two good
ate reduction observed in BY4743pdc2Δ519 was within candidates could be acetoin and 2,3-butanediol. Both
the expectable, and it could be the result of a competence compounds are derived from the secondary metabolism
phenomenon between the wild type and the mutant of yeast, and are produced from pyruvate in a series
Pdc2p protein for the DNA binding site and some acti- of chemical reactions where at least one decarboxyla-
vating protein. Although it is just a speculation, there is tion reaction is involved (Romano and Suzzi 1996). In
some indirect evidence which supports this proposition. a recent work, an ethanol reduction of 1.3% (v/v) was
On one hand, it has been shown that the Pdc2p DNA- achieved by combining adaptive laboratory evolution
binding site alone retains some DNA binding activity, strategies with hybridization (Tilloy et  al. 2014). Inter-
and on the other hand, an activating protein has been estingly, the enhancement of glycerol production in
proposed at least for the regulation by PDC2 of the THI the selected yeast was accompanied by an increment in
genes (Nosaka et  al. 2012). In any case, the molecular 2,3-butanediol, which is consider a neutral organoleptic
mechanism underlying the regulation of PDC1 by PDC2 compound.
is still unknown (Brion et  al. 2014) and more investiga- In this study we present an alternative microbiological
tion would be required to clarify this matter. strategy to reduce the ethanol content in wine, through
In view of the good results obtained with the a genetic modification of the S. cerevisiae Pdc2p tran-
BY4743pdc2Δ519 strain our next goal was to test this scription factor. Our results demonstrate that the inser-
mutation in wine yeast strains. For this purpose the tion of the pdc2Δ519d mutation in a wine yeast strain can
pdc2Δ519 mutation was transformed into the com- reduce the ethanol concentration up to 1.89% v/v with-
mercial EC1118 and native Mab2C wine yeast strains. out affecting the fermentation performance. In contrast
EC1118 has been widely used in the wine industry and to non-GMO based strategies, our approach permits the
it is known for its reliability and excellent fermenta- insertion of the selected mutation in any locally selected
tion performance (Aceituno et  al. 2012). Meanwhile, wine strain, making possible to produce quality wines
Mab2C is a native strain previously selected in our with regional characteristics and lower alcohol content.
laboratory for its excellent oenological properties in This makes our work a valuable contribution to the prob-
the elaboration of Malbec wine. Both EC1118Δ519 and lem of high ethanol concentration in wine. Nevertheless,
Mab2CΔ519 mutants grew normally in aerobic condi- pilot-scale trials complemented with sensorial analysis
tions with glucose as a carbon source and displayed typ- of the produced wines are required for a full evaluation
ical fermentation kinetics for a wine yeast strain. As it of our strain’s potential for its application in the wine
happened with the laboratory BY4743pdc2Δ519 mutant industry.
Cuello et al. AMB Expr (2017) 7:67 Page 10 of 11

Abbreviations aeration as a strategy for the production of wine with reduced alcohol
ANOVA: analysis of variance; GMO: genetically modified organism; LSD: least content. Int J Food Microbiol 205:7–15
significant difference; PDC: pyruvate decarboxylase. Di Maio S, Genna G, Gandolfo V, Amore G, Ciaccio M, Oliva D (2012) Presence
of Candida zemplinina in sicilian musts and selection of a strain for wine
Authors’ contributions mixed fermentations. S Afr J Enol Vitic 33:80–87
RAC Participated in the design and coordination of the study, performed Eglinton JM, Heinrich AJ, Pollnitz AP, Langridge P, Henschke PA, de Barros
the experiments, interpreted the data, and drafted the manuscript. KJFM Lopes M (2002) Decreasing acetic acid accumulation by a glycerol
performed the experiments and the data analysis. LAM performed the experi- overproducing strain of Saccharomyces cerevisiae by deleting the ALD6
ments and the data analysis. MC participated in its design and coordination, aldehyde dehydrogenase gene. Yeast 19:295–301
interpreted the data, and helped to draft the manuscript. IFC conceived of the Gawel R, Francis L, Waters EJ (2007a) Statistical correlations between the
study, participated in its design and coordination, interpreted the data, and in-mouth textural characteristics and the chemical composition of Shiraz
drafted the manuscript. All authors read and approved the final manuscript. wines. J Agric Food Chem 55:2683–2687
Gawel R, Van Sluyter S, Waters EJ (2007b) The effects of ethanol and glycerol
Author details on the body and other sensory characteristics of Riesling wines. Aust J
1
 Consejo Nacional de Investigaciones Científicas y Tecnológicas (CONICET), Grape Wine Res 13:38–45
Buenos Aires, Argentina. 2 Laboratorio de Biotecnología, Estación Experimental Gietz DR, Woods RA (2002) Transformation of yeast by lithium acetate/single-
Agropecuaria Mendoza, Instituto Nacional de Tecnología Agropecuaria (INTA), stranded carrier DNA/polyethylene glycol method. Methods Enzymol
San Martín 3853, Luján de Cuyo, Mendoza, Argentina. 350:87–96
Gobbi M, Comitini F, Domizio P, Romani C, Lencioni L, Mannazzu I, Ciani M
Competing interests (2013) Lachancea thermotolerans and Saccharomyces cerevisiae in simulta-
The authors declare that they have no competing interests. neous and sequential co-fermentation: a strategy to enhance acidity and
improve the overall quality of wine. Food Microbiol 33:271–281
Funding Güldener U, Heck S, Fiedler T, Beinhauer J, Hegemann JH (1996) A new effi-
This worked was supported by a PICT 2008-206 project of the Fondo para la cient gene disruption cassette for repeated use in budding yeast. Nucleic
investigación Científica y Tecnológica (FonCyT) Argentina. R.A.C. is a fellow of Acids Res 24:2519–2524
CONICET. Hohmann S (1993) Characterization of PDC2, a gene necessary for high level
expression of pyruvate decarboxylase structural genes in Saccharomyces
Received: 7 March 2017 Accepted: 13 March 2017 cerevisiae. Mol Gen Genet 1:657–666
Kontoudakis N, Esteruelasa M, Forta F, Canalsa JM, De Freitas V, Zamora F
(2011) Influence of the heterogeneity of grape phenolic maturity on wine
composition and quality. Food Chem 124:767–774
Kutyna DR, Varela C, Henschke PA, Chambers PJ, Stanley GA (2010) Microbio-
logical approaches to lowering ethanol concentration in wine. Trends
References Food Sci Technol 21:293–302
Aceituno FF, Orellana M, Torres J, Mendoza S, Slater AW, Melo F, Agosin E (2012) Lopes M, Ur-Rehman A, Gockowiak H, Heinrich AJ, Langridge P, Henschke PA
Oxygen response of the wine yeast Saccharomyces cerevisiae EC1118 (2000) Fermentation properties of a wine yeast over-expressing the Sac-
grown under carbon-sufficient, nitrogen-limited enological conditions. charomyces cerevisiae glycerol 3-phosphate dehydrogenase gene (GPD2).
Appl Environ Microbiol 78:8340–8352 Aust J Grape Wine Res 6:208–215
Benito S, Morata A, Palomero F, González MC, Suárez-Lepe JA (2011) Formation Magyar I, Tóth T (2011) Comparative evaluation of some oenological proper-
of vinylphenolic pyranoanthocyanins by Saccharomyces cerevisiae and ties in wine strains of Candida stellata, Candida zemplinina, Saccharomyces
Pichia guillermondii in red wines produced following different fermenta- uvarum and Saccharomyces cerevisiae. Food Microbiol 28:94–100
tion strategies. Food Chem 124:15–23 Mojzita D, Hohmann S (2006) Pdc2 coordinates expression of the THI
Bogianchini M, Cerezo AB, Gomis A, López F, García-Parrilla MC (2011) Stabil- regulon in the yeast Saccharomyces cerevisiae. Mol Genet Genomics
ity, antioxidant activity and phenolic composition of commercial and 276(2):147–161
reverse osmosis obtained dealcoholised wines. LWT Food Sci Technol Morales P, Rojas V, Quirós M, Gonzalez R (2015) The impact of oxygen on the
44:1369–1375 final alcohol content of wine fermented by a mixed starter culture. Appl
Brion C, Ambroset C, Delobel P, Sanchez I, Blondin B (2014) Deciphering regu- Microbiol Biotechnol 9:3993–4003
latory variation of THI genes in alcoholic fermentation indicate an impact Nevoigt E, Stahl U (1996) Reduced pyruvate decarboxylase and increased
of Thi3p on PDC1 expression. BMC Genom 15:1085 glycerol-3-phosphate dehydrogenase [NAD+] levels enhance glycerol
Buescher WA, Siler CE, Morris JR, Threlfall RT, Main GL, Cone GC (2001) High production in Saccharomyces cerevisiae. Yeast 12:1331–1337
alcohol wine production from grape juice concentrates. Am J Enol Vitic Nosaka K, Esaki H, Onozuka M, Konno H, Hattori Y, Akaji K (2012) Facilitated
52:345–351 recruitment of Pdc2p, a yeast transcriptional activator, in response to
Cambon B, Monteil V, Remize F, Camarasa C, Dequin S (2006) Effects of GPD1 thiamin starvation. FEMS Microbiol Lett 330:140–147
overexpression in Saccharomyces cerevisiae commercial wine yeast strains Novo M, Bigey F, Beyne E, Galeote V, Gavory F, Mallet S, Cambon B, Legras JL,
lacking ALD6 genes. Appl Environ Microbiol 72:4688–4694 Wincker P, Casaregola S, Dequin S (2009) Eukaryote-to-eukaryote gene
Canonico L, Comitini F, Oro L, Ciani M (2016) Sequential fermentation with transfer events revealed by the genome sequence of the wine yeast Sac-
selected immobilized non-Saccharomyces yeast for reduction of ethanol charomyces cerevisiae EC1118. Proc Natl Acad Sci 38:16333–16338
content in wine. Front Microbiol 7:278 OIV International methods of analysis of wines and musts (2015) Compen-
Ciani M, Ferraro L (1996) Enhanced glycerol content in wines made with dium of international methods of analysis of wines and musts, vol 1,
immobilized Candida stellata cells. Appl Environ Microbiol 62:128–132 sections 3.1.1 Sugars, 3.1.2 Alcohols and 3.1.3 Acids
Ciani M, Beco L, Comitini F (2006) Fermentation behavior and metabolic Orduña RM (2010) Climate change associated effects on grape and wine qual-
interactions of multistarter wine yeast fermentations. Int J Food Microbiol ity and production. Food Res Intern 43:1844–1855
108:239–245 Remize F, Roustan JL, Sablayrolles JM, Barre P, Dequin S (1999) Glycerol over-
Ciani M, Morales P, Comitini F, Tronchoni J, Canonico L, Curiel JA, Oro L, Rodri- production by engineered Saccharomyces cerevisiae wine yeast strains
gues AJ, Gonzalez R (2016) Non-conventional yeast species for lowering leads to substantial changes in by-product formation and to a stimula-
ethanol content of wines. Front Microbiol 7:642 tion of fermentation rate in stationary phase. Appl Environ Microbiol
Comitini F, Gobbi M, Domizio P, Romani C, Lencioni L, Mannazzud I, Ciani M 65:143–149
(2011) Selected non-Saccharomyces wine yeasts in controlled multistarter Romano P, Suzzi G (1996) Origin and production of acetoin during wine yeast
fermentations with Saccharomyces cerevisiae. Food Microbiol 5:873–882 fermentation. Appl Environ Microbiol 62:309–315
Contreras A, Hidalgo C, Schmidt S, Henschke PA, Curtin C, Varela C (2015) The Sadoudi M, Tourdot-Maréchal R, Rousseaux S, Steyerd D, Gallardo-Chacóna
application of non-Saccharomyces yeast in fermentations with limited JJ, Ballesterc J, Vichie S, Guérin-Schneider R, Caixache J, Alexandre H
Cuello et al. AMB Expr (2017) 7:67 Page 11 of 11

(2012) Yeast–yeast interactions revealed by aromatic profile analysis of Vaughan-Martini A, Martini A (1998) Determination of ethanol production.
Sauvignon Blanc wine fermented by single or co-culture of non-Saccha- In: Kurtzman CP, Fell JW (eds) The yeasts: a taxonomic study. Elsevier,
romyces and Saccharomyces yeasts. Food Microbiol 32:243–253 Amsterdam, pp 358–371
Tilloy V, Ortiz-Julien A, Dequin S (2014) Reduction of ethanol yield and Vazquez F, Figueroa L, Toro ME (2001) Enological characteristics of yeasts. Food
improvement of glycerol formation by adaptive evolution of the wine Microbiol Protoc Methods Biotechnol 7:297–306
yeast Saccharomyces cerevisiae under hyperosmotic conditions. Appl Zwietering M, Jongenburger I, Rombouts F, van Riet K (1990) Modeling of the
Environ Microbiol 80:2623–2632 bacterial growth curve. Appl Environ Microbiol 56:1875–1881
Varela C, Kutyna DR, Solomon MR, Black CA, Borneman A, Henschke PA, Pre-
torius IS, Chambers PJ (2012) Evaluation of gene modification strategies
for the development of low-alcohol-wine yeasts. Appl Environ Microbiol
78:6068–6077
ORIGINAL RESEARCH
published: 14 December 2020
doi: 10.3389/fmicb.2020.597828

Creation of a
Low-Alcohol-Production Yeast by a
Mutated SPT15 Transcription
Regulator Triggers Transcriptional
and Metabolic Changes During Wine
Fermentation
Qing Du 1,2,3 , Yanlin Liu 1,2,3 , Yuyang Song 1,2,3 and Yi Qin 1,2,3*
1
College of Enology, Northwest A&F University, Yangling, China, 2 Shaanxi Engineering Research Center for Viti-Viniculture,
Edited by:
Yangling, China, 3 National Forestry and Grassland Administration Engineering Research Center for Viti-Viniculture, Yangling,
Hui Wu,
China
East China University of Science
and Technology, China
Reviewed by: There is significant interest in the wine industry to develop methods to reduce the ethanol
Xinqing Zhao, content of wine. Here the global transcription machinery engineering (gTME) technology
Shanghai Jiao Tong University, China
Estéfani García Ríos, was used to engineer a yeast strain with decreased ethanol yield, based on the mutation
Consejo Superior de Investigaciones of the SPT15 gene. We created a strain of Saccharomyces cerevisiae (YS59-409), which
Científicas (CSIC), Spain
Soo Rin Kim,
possessed ethanol yield reduced by 34.9%; this was accompanied by the increase
Kyungpook National University, in CO2 , biomass, and glycerol formation. Five mutation sites were identified in the
South Korea mutated SPT15 gene of YS59-409. RNA-Seq and metabolome analysis of YS59-409
*Correspondence: were conducted compared with control strain, suggesting that ribosome biogenesis,
Yi Qin
qinyi@nwsuaf.edu.cn nucleotide metabolism, glycolysis flux, Crabtree effect, NAD+ /NADH homeostasis and
energy metabolism might be regulated by the mutagenesis of SPT15 gene. Furthermore,
Specialty section:
two genes related to energy metabolism, RGI1 and RGI2, were found to be associated
This article was submitted to
Microbial Physiology and Metabolism, with the weakened ethanol production capacity, although the precise mechanisms
a section of the journal involved need to be further elucidated. This study highlighted the importance of applying
Frontiers in Microbiology
gTME technology when attempting to reduce ethanol production by yeast, possibly
Received: 22 August 2020
Accepted: 23 November 2020
reprogramming yeast’s metabolism at the global level.
Published: 14 December 2020
Keywords: low-alcohol, SPT15, RNA-seq, metabolome, wine, Saccharomyces cerevisiae
Citation:
Du Q, Liu Y, Song Y and Qin Y
(2020) Creation of a INTRODUCTION
Low-Alcohol-Production Yeast by
a Mutated SPT15 Transcription
Over recent decades, the alcohol concentration of wines produced by many warm regions around
Regulator Triggers Transcriptional
and Metabolic Changes During Wine
the world has increased by approximately 2% (v/v) (Goold et al., 2017). This is mainly due to
Fermentation. the increasing preference of consumers for well-structured, full-bodied, and ripe-fruit wines and
Front. Microbiol. 11:597828. those wines are generally made from more mature grapes. Also, this has been exacerbated by
doi: 10.3389/fmicb.2020.597828 global warming, which leads to higher sugar content in the grape varieties used to make wine

Frontiers in Microbiology | www.frontiersin.org 1 December 2020 | Volume 11 | Article 597828


Du et al. Low-Alcohol-Production Yeast

(Varela et al., 2015; Rolle et al., 2017). Prompted by various Seong et al., 2017). In the present study, we used gTME
reasons regarding wine quality, economic variables, and technology to weaken the capacity of yeast to produce ethanol
health concerns caused by high levels of alcohol, strategies and ultimately created a strain of S. cerevisiae (YS59-409) with
aimed at reducing ethanol concentrations without impairing a low yield of ethanol production. RNA-Seq and metabolomic
wine organoleptic quality have been performed in different analysis were also conducted in an attempt to understand the
ways (Varela et al., 2015; Dequin et al., 2017). Generally, metabolic mechanisms underlying the modified phenotype of
microbiological strategies relating to the isolation and/or YS59-409. This study highlighted the critical role of the SPT15
generation of the yeast strains used to make wine have proved regulator in reducing ethanol production in yeast and provided
to be the simplest and most economical methods, including comprehensive insights to understand the molecular mechanisms
Saccharomyces cerevisiae and non-conventional yeast species of a new low-ethanol yeast.
(Tilloy et al., 2015; Varela and Varela, 2018).
Engineering yeast strains with the capacity of redirecting
carbon away from ethanol production to other endpoints MATERIALS AND METHODS
is an effective approach and thus far, enhancing glycerol
production has testified to be the most effective method Plasmids, Strains, SPT15 Mutant Library
(Otterstedt et al., 2004; Varela et al., 2012; Tan et al., 2016).
Classical gene modification (GM) technologies have achieved Construction and Culture Conditions
increasing glycerol formation to reduce ethanol production by We used S. cerevisiae YS59 (MATα; ura3-52, leu2-3, and his 5-
the manipulation of single or several genes. However, high 519) (Liu et al., 2007) as the host strain and then amplify the
concentrations of by-products, such as acetate, acetaldehyde, and open reading frame of SPT15 gene from genomic DNA of YS59.
acetoin, were generated in these previous experiments (Cambon The SPT15 gene was inserted into the restriction sites between
et al., 2006; Varela et al., 2012). Since these by-products could BamHI and EcoRI using the pY16 vector, which was flanked with
have a significant negative effect on the flavor of wine, concerted TEF1 promoter and CYC1 terminator (pY16-SPT15); the plasmid
efforts have been made to reduce the formation of these by- has a URA3 selective marker, an ampicillin resistant marker and
products (Cambon et al., 2006; Ehsani et al., 2009). Additionally, a CEN/ARS element (low copy).
a combination of adaptive evolution and breeding strategies The yeast mutant library was created by random mutagenesis
is applied to develop a low-alcohol yeast for wine making (error-prone PCR) of the SPT15 gene. Firstly, the SPT15 mutant
with higher levels of glycerol production; this strain reduced library was generated using the DiversifyTM PCR Random
ethanol production by 1.3% (v/v) without the formation of Mutagenesis kit (Clontech) with pY16-SPT15 as template.
undesirable by-products (Tilloy et al., 2014). Other researchers Plasmids obtained were transformed into Escherichia coli JM109
have focused on non-Saccharomyces strains in an attempt to to produce a primary library for SPT15 mutants. From the
reduce the ethanol content of wine; this is because such strains sequencing of 20 randomly selected colonies, mutations were
are known to exhibit different respiro-fermentative regulatory found at 2–10 sites without preference. Then library plasmids
mechanisms when compared to S. cerevisiae (Quiros et al., 2014). were transformed into S. cerevisiae YS59 and incubated at 25◦ C
Nevertheless, those non-Saccharomyces strains generally possess on solid SD to generate a yeast library for SPT15 mutant.
a weak capacity to complete wine fermentation on their own The plasmid was transformed into yeast cells by the
and must be accompanied by S. cerevisiae (Hranilovic et al., lithium acetate method (Gietz and Woods, 2001). The strains
2020). Consequently, there is significant interest in developing and plasmids used in this procedure are summarized in
more effective strategies to balance low-ethanol wine production Supplementary Table 1. S. cerevisiae strains were pre-cultured
efficiency with good organoleptic qualities. overnight in YPD medium (1% yeast extract, 2% peptone, and
In this study, an alternative approach, global transcriptional 2% glucose) at 30◦ C for non-selective propagation. The selective
machinery engineering (gTME), was used to develop strains culture of engineered strains was conducted in SD medium
of S. cerevisiae with reduced ethanol-production ability. The (0.67% YNB, 0.077% Ura DO Supplement, and 2% glucose)
gTME technology was carried out by mutating the general at 30◦ C.
transcription factor Spt15p, the TATA-binding protein (Alper
et al., 2006); this protein plays a key role in the action of Alcoholic Fermentation
RNA polymerase and is one of the main DNA binding proteins Alcoholic fermentation was carried out using an inoculum of
that regulate promoter specificity in yeast (Eisenmann et al., 5 × 105 cells/mL in a shaking incubator at 25◦ C at 150 rpm;
1989). The gTME technology is first used to improve the all cultures were carried out in triplicate. The medium was
glucose/ethanol tolerance of S. cerevisiae, which shows the similar to Triple M medium as reported previously (Spiropoulos
ability to re-program global gene transcription and change the et al., 2000), and consisted of 75 g/L glucose, 75 g/L fructose,
complex phenotype of yeast strains (Alper et al., 2006; Alper 6 g/L tartaric acid, 3 g/L malic acid, 0.5 g/L citric acid, 1.7 g/L
and Stephanopoulos, 2007). Since then, several research studies yeast nitrogen base without amino acids, 1 g/L ammonium
have used the gTME approach to optimize the ethanol tolerance phosphate, 2 g/L casamino acids, 0.8 g/L L-arginine,1 g/L L-
and ethanol production capacity of S. cerevisiae, and successfully proline, 0.1 g/L Tryptophan, and 4 mL ergo stock (composed
demonstrated that gTME is advantageous when attempting to of 250 mL/L Tween80, 750 mL/L 95% ethanol and 2.5 g/L
regulate the ethanol metabolism of yeast strains (Yang et al., 2011; ergosterol), with pH 3.25.

Frontiers in Microbiology | www.frontiersin.org 2 December 2020 | Volume 11 | Article 597828


Du et al. Low-Alcohol-Production Yeast

Site-Directed Mutagenesis of the SPT15 mass spectrometer (Thermo, United States). The separation was
Gene and Fermentation Assays for performed on a Hypergod C18 column (100 mm × 4.6 mm
3 µm) at 40◦ C. A flow rate of 0.3 mL/min, and an injection
Recombinant Strains volume of 4 µL, were used in all analyses; this was followed by
Mutations in SPT15 in strain YS59-409 were identified by DNA auto-sampling at 4◦ C. Mass spectrometry detection used negative
sequencing. For site-directed mutagenesis, plasmid pY16-SPT15 polarity with the following parameters: heater temperature,
(containing the non-mutated SPT15 gene) was used as a template 300◦ C; sheath gas flow rate, 45 arb; auxiliary gas flow rate, 15
with primers designed to the target nucleotide substitutions arb; sweep gas flow rate, 1 arb; spray voltage, 3.2 kV; capillary
(Supplementary Table 2); these reactions were carried out with temperature, 350◦ C, and an S-Lens RF Level of 60%.
a Mut Express II Fast Mutagenesis Kit (Vazyme, China) in With regards to multivariate statistical analysis, principal
accordance with the manufacturer’s directions. The reconstructed component analysis (PCA) and orthogonal partial least squares
plasmids were then transformed into S. cerevisiae YS59 in order discriminant analysis (OPLS-DA) were performed using SIMCA-
to obtain recombinant strains (Supplementary Table 1). The P version 13.0 software (Umetrics AB, Sweden) to separate the
effects of mutation on the ethanol production capacity of strains two groups of data. We then searched for differential metabolites
were performed in Triple M media. using the variable importance in the projection (VIP) value of the
OPLS-DA model (VIP > 1) in combination with the p-value of
RNA-Seq Analysis the t-test (p < 0.05). Qualitative metabolites were characterized
RNA-Seq analysis was used to investigate the transcriptional by searching an online database1 . Metabolic pathway analysis was
differences between the low-ethanol-production strain YS59- carried out using MetaboAnalyst 3.02 .
409 and the control strain YS59-pY16. Three independent
samples were collected from the mid-log phase of fermentation Analytical Methods
for RNA extraction. Total RNA was extracted using the We measured the weight loss in CO2 from each sample
procedures described previously (Li et al., 2019). Agarose by weighing the fermenters on a daily basis. Cell growth
gel (1%) electrophoresis and a spectrophotometer (NanoDrop was recorded by a microplate reader at OD600 nm (BioTek,
ND1000, United States) were then used to detect the purity ELx800, United States). Ethanol yields were analyzed using
and concentration of samples. Bioanalyzer (Agilent 2100, an SBA-40C biological sensor analyzer (Biology Institute of
United States) was used to detect the RNA integrity numbers Shandong Academy of Sciences, China). The content of reducing
(>8.0) of these samples to satisfy the particular requirements of sugar in each sample was measured by the DNS method
RNA-seq. The cDNA library construction and RNA sequencing (Hu et al., 2008) with glucose as a standard using an
were carried out by Beijing Genomics Institute (Shenzhen, ultra-violet spectrophotometer. Concentrations of glycerol and
China) using standard protocols. The RNA-Seq data generated in acetic acid were determined with an Enology Analyzer Y15
this study were submitted to NCBI Sequence Read Archive (SRA) (BioSystems, Spain).
under the accession number PRJNA548495.
To compare the transcriptomes of the mutant and control Data Analysis
yeast strains, we used Bowtie2 to map clean reads to the Statistically significant differences between the wild-type (YS59-
reference gene and then used HISAT to reference the genome pY16) and mutant yeast strains (YS59-409) were determined
of S. cerevisiae S288c. Gene expression levels were quantified using the Student’s t-test. The effect of site-directed mutagenesis
using the FPKM method (Trapnell et al., 2012). Screening and gene knockout on the fermentation characteristics of the two
of differentially expressed genes (DEGs) between the two strains were further determined by one-way analysis of variance
strains was conducted using the NOISeq method (Tarazona (ANOVA) and Duncan’s test. The confidence level for both tests
et al., 2011) based on a foldchange (log2 Ratio) ≥1.5 and a was 95% and all analyses were carried out using SPSS version 19.0
divergent probability ≥0.8. Functional enrichment analysis of software (SPSS Inc., United States).
Gene Ontology (GO) and KEGG pathway enrichment analysis
of DEGs were conducted by comparison with the entire genome
background (Bonferroni-corrected P-value < 0.05). RESULTS AND DISCUSSION
Metabolomics Analysis Selection and the Characteristics of the
The method used to prepare samples (six independent replicates) Low-Ethanol-Yield Strain
from the low-ethanol-production strain for metabolomic analysis Previously, we produced a yeast mutant library (>1,000 clones)
was consistent with that used for RNA-seq analysis. Metabolites that was based on S. cerevisiae YS59 and created by random
were extracted by vortex blending approximately 1 g of cells mutagenesis (error-prone PCR) of the SPT15 gene with the
(fresh weight) in 1 mL of cold methyl alcohol for 30 s. Cells pY16 plasmid backbone using gTME technology. Preliminary
were then lysed by ultrasonication for 15 min and centrifuged screening using SD media in 24-well plates was performed to
for 15 min at 12,000 × g at 4◦ C; an aliquot of 200 µL of identify the 20 mutants with the highest biomass; these were
the supernatant was used for further analysis. Samples were
analyzed by Shanghai Biocluster Biotech Co., Ltd. (China) using 1
http://metlin.scripps.edu
the Ultimate 3000 LC system coupled with an Orbitrap Elite 2
http://www.metaboanalyst.ca

Frontiers in Microbiology | www.frontiersin.org 3 December 2020 | Volume 11 | Article 597828


Du et al. Low-Alcohol-Production Yeast

the strains that were supposed to have lower ethanol-production with the control under fermentation conditions (Table 2). Of all
capacity due to competition for carbon sources between these mutational sites, Y195H has been reported to work together
biomass and ethanol production (Supplementary Figure 1). To with F177S and K218R to improve glucose/ethanol tolerance and
determine the ethanol production capacity of these 20 strains, we the efficiency of glucose conversion to ethanol of yeast (Alper
fermented each in 10-mL of Triple M media (approximately one- et al., 2006); and Leu-205 site has shown to be important for
fifth of the tube’s total volume; this allowed the strains to grow DNA binding specificity of Spt15p and mutating this site to other
fully) in the tubes. Finally, the strain with the lowest ethanol yield different amino acids can cause various degrees of changes in
was identified (Supplementary Table 3). Compared with the yeast growth (Arndt et al., 1992). Although our results showed
control strain, the ethanol yield of the low-ethanol-production that single mutation of Y195H or L205S did not produce the low-
strain (YS59-409) was significantly reduced by 34.9% (Table 1, alcohol phenotype, we hypothesized that these two sites might
p < 0.05). Interestingly, the strain featuring SPT15 gene mutation play key roles in the mutant strain and that the desired phenotype
exhibited a greater CO2 weight loss than the control group might be obtained through double or multiple mutations based
(Figure 1). In the wine industry, CO2 weight loss is generally on previous reports (Arndt et al., 1992; Alper et al., 2006). In
used as an indicator of the fermentation capacity of yeast. It is addition, the same mutation on SPT15 may lead to different
worth noting that the YS59-409 strain produced a larger amount phenotypes in strains with different genetic backgrounds. In a
of CO2 than the control but had a lower ethanol production. This word, these results illustrate that the combined action of the
suggests that other decarboxylation pathways might contribute five separate mutations, or at least in part, conferred the desired
to the CO2 loss in the mutant strain. In addition, the glycerol phenotype to the mutant strain.
content of the YS59-409 strain was 43% higher than the control
strain, although there were no significant differences with regards Transcriptional and Metabolic Analysis of
to acetic acid production (Table 1). These findings were not the Mutated Strain
consistent with the classical theory that higher glycerol synthesis As described in previous research, gTME can trigger overall
is associated with increased acetic acid production (Nevoigt and disturbances at the transcriptional level and can be used to
Stahl, 1996). Commonly, the main by-product, glycerol, confers unravel complex phenotypes (Alper and Stephanopoulos, 2007).
positive sensory effects, body, and sweetness, to wines when In order to investigate the mechanisms underlying the observed
present in appropriate amounts and has been used as a target to phenotype, we used RNA-Seq and metabolomic analysis to
redirect carbon sources away from ethanol production by many analyze the differences in transcription and metabolism between
researchers in the wine industry (Tilloy et al., 2015; Goold et al., mutant strain (YS59-409) and the control (YS59-pY16). Under
2017). Our mutant strain (YS59-409) not only produced more fermentation conditions, we observed significant changes in
glycerol than the control strain, but also did not accumulate the transcription and metabolism of the mutant; these factors
overmuch acetic acid, an undesirable compound; these findings are likely to have made an important contribution to the
were also in agreement with those from previous studies (Ehsani target phenotype of the mutant. Subsequently, we would discuss
et al., 2009; Tilloy et al., 2014). During wine fermentation, most of the overall transcriptional and metabolic regulation and the
the sugars are used for the production of ethanol and CO2 , with specific pathways and function change associated with ethanol
a small part for biomass and glycerol forming, including minute metabolism answering to the mutation of SPT15 gene.
amounts of other byproducts. Therefore, we can speculate that
strain YS59-409 uses more sugar for CO2 , biomass, and glycerol The Overall Analysis of Transcription and Metabolism
synthesis, thus shunting the carbon source away from ethanol in Strain YS59-409
synthesis and resulting in low ethanol yield. A total of 964 genes showed significantly different expression
levels when compared between the mutant and control strains;
636 genes were up-regulated and 328 genes were down-regulated
Sequence Analysis of the SPT15 Gene in in the mutant YS59-409 strain (Supplementary Table 4). These
the YS59-409 Strain differentially expressed genes (DEGs) demonstrated that the
The SPT15 gene in YS59-409 was amplified from the pY16-409 mutation of the SPT15 gene, a transcription factor, led to the
plasmid and then was sequenced. There were 5 mutational sites global reprogramming of transcription at genes and may have
in the structural domain of the mutated SPT15 gene as shown contributed to the regulation of ethanol metabolism in this
in Figure 2, in which methionine is substituted for isoleucine mutant strain. GO analysis and KEGG pathway enrichment
(Ile46 Met), and similarly, Asp56 Gly, Ser118 Pro, Tyr195 His, and analysis were used to further analyze the functional enrichment of
Leu205 Ser (I46M, D56G, S118P, Y195H, and L205S, respectively). DEGs. KEGG pathway enrichment analysis revealed that most of
Three point mutations were located in the highly conserved the DEGs participated in those key metabolic pathways, including
domain (amino acids 61–240) and two mutations were located translation, transcription, carbohydrate metabolism, nucleotide
in the non-conserved domain (amino acids 1–60). To explore metabolism, and amino acid metabolism (Table 3).
which of these sites conferred the most desirable phenotype to the There were 41 differential metabolites identified (16
low-ethanol-yielding strain, we constructed five strains featuring upregulated and 25 downregulated), as shown in Supplementary
single point mutations (Supplementary Table 1). However, none Table 6. OPLS-DA and PCA analysis were conducted to
of these mutants led to a reduction in ethanol yield; instead, identify differences in the metabolome that were caused by
I46M, D56G, and L205S, led to enhanced ethanol yield compared random mutagenesis of the SPT15 gene. We found there

Frontiers in Microbiology | www.frontiersin.org 4 December 2020 | Volume 11 | Article 597828


Du et al. Low-Alcohol-Production Yeast

TABLE 1 | Fermentation characteristics of the mutant strain S. cerevisiae YS59-409.

Strains OD600 Residual sugar (g/L) Ethanol (g/L) Glycerol (g/L) Acetate (g/L) Ethanol yielda (g/g Sugar) Change in ethanol
yield

YS59-pY16 4.52 ± 0.34 1.13 ± 0.35 54.67 ± 0.94 2.13 ± 0.08 0.493 ± 0.109 0.367 ± 0.007 0
YS59-409 5.45 ± 0.40* 1.90 ± 0.30 35.33 ± 3.09* 3.05 ± 0.07* 0.487 ± 0.066 0.239 ± 0.021* −34.9%
a Ethanolyield represents the ratio of ethanol production (g) to the sugar consumption (g).
*Student’s t-test showed significant difference at the level of 0.05.

of SPT15 in the mutant YS59-409 strain (Supplementary


Table 4) for its regulatory effect on RNA polymerase and
gene transcription levels (Eisenmann et al., 1989). In addition,
RPL and RPS genes encoding ribosome biogenesis proteins
were significantly over-expressed in the low ethanol-production
strain (Supplementary Tables 4, 5), which reflected the
stimulation of relative translational activity at certain time
points (Backhus et al., 2001). Coincidently, the upregulation
of RNA polymerase metabolism and ribosome biogenesis were
reported in a low-ethanol-production yeast by a previous study
(Varela et al., 2018). Furthermore, most of the genes associated
with nucleotide metabolism were expressed at high levels
(Table 3 and Supplementary Tables 4, 5), of which purines
and pyrimidines are the basic components (Ljungdahl and
Daignan-Fornier, 2012), exhibiting the activation of purine and
pyrimidine metabolism. Collectively, these results above indicate
that transcription, translation, and nucleotide metabolism,
FIGURE 1 | CO2 weight loss (g). Black square represents S. cerevisiae
YS59-409 and black round represents S. cerevisiae YS59-pY16 (control were activated in the mutant YS59-409 strain. This probably
strain). demonstrates that the mutation of the SPT15 gene leads to the
reprogramming of global gene expression and metabolism.

were distinct discrepancies between the mutant and control The Fermentative Pathways Leading to Ethanol
strains in terms of metabolic profile (Supplementary Figure 2). Production Was Reduced in Strain YS59-409
Indeed, we found that several metabolites related to amino Glycolytic flux relies on glucose uptake rate, which is regulated
acid metabolism (L-glutamine, L-glutamate, L-tryptophan, by a family of hexose transporters encoded by HXT genes. In
L -histidine, L -lysine, kynurenine, glutathione, succinic acid the present study, we found that several hexose transporter genes
semialdehyde, and adenylosuccinate) (Supplementary Table 6) were significantly downregulated in the YS59-409 strain (HXT2,
showed significant differences when compared between the two HXT4, HXT5, HXT6, HXT7, HXT9, HXT11, and HXT13; Figure 3
strains, corresponding to transcriptional alterations in amino and Supplementary Table 4); of these, HXT4, HXT6, and HXT7,
acid metabolism. are known to be vital for the uptake of glucose (Özcan and
Johnston, 1999). Previous research has reported that a deficiency
Transcription, Translation, and Nucleotide of hexose transporters in S. cerevisiae could reduce the levels of
Metabolism, Were Activated in Strain YS59-409 sugar uptake and thus regulate the glycolysis flux, particularly
In total, 14 and 135 genes were functionally annotated with regards to HXT7, which eventually resulted in a reduction
into transcription and translation, respectively (Table 3 and in ethanol production (Otterstedt et al., 2004). Consequently, the
Supplementary Tables 4, 5). Genes related to RNA polymerase significant downregulation of several HXT genes in YS59-409
were upregulated, thus indicating the transcriptional activation is likely to result in a decrease of glycolysis flux. Besides, the
of the mutant strain. This could be explained by the upregulation expression of HXK1 was downregulated by 4.7-fold (Figure 3

FIGURE 2 | Mutation sites in the SPT15 gene of mutant strain (arrows). The schematic of structural domain is referred to the previous study (Alper et al., 2006).

Frontiers in Microbiology | www.frontiersin.org 5 December 2020 | Volume 11 | Article 597828


Du et al. Low-Alcohol-Production Yeast

TABLE 2 | Fermentation characteristics of point-mutated strains.

Strains Residual sugar (g/L) Ethanol (g/L) Ethanol yield (g/g sugar) Changes in ethanol yield

409-46 0.56 ± 0.91a 53.09 ± 0.36b 0.348 ± 0.005ac 8.10%


409-56 1.10 ± 1.67a 52.23 ± 1.09b 0.357 ± 0.018a 11.00%
409-118 1.04 ± 0.01a 51.42 ± 1.76ab 0.325 ± 0.005b 1.10%
409-195 1.03 ± 0.02a 50.19 ± 0.49ab 0.332 ± 0.003bc 3.10%
409-205 1.03 ± 0.02a 53.34 ± 3.19b 0.355 ± 0.013a 10.40%
YS59-pY16 1.02 ± 0.02a 48.43 ± 0.24a 0.322 ± 0.003b 0
a,b,c Different letters indicate significant differences according to the Duncan test (p < 0.05). Data are mean ± SD of independent triplicate.

and Supplementary Table 4); this gene encodes hexokinase consumption under conditions of excess glucose, even in the
1, which is responsible for catalyzing the phosphorylation of presence of oxygen (also referred to as ‘overflow metabolism’)
glucose in the first irreversible step of glycolysis (Rodríguez et al., (de Deken, 1966). In the present study, we observed indications
2001). The downregulation of HXK1 also might contribute to the that the Crabtree effect of the YS59-409 strain might be
reduced glycolysis metabolism in the mutant; this observation is disturbed. On the one hand, the attenuated glycolysis flux
supported by a previous study that the repression of hexokinase in the YS59-409 strain, in combination with the higher CO2
activity resulted in reduced glycolysis flux (Tan et al., 2016). production and the lower ethanol formation rate, appeared
So, the downregulation of HXTs and HXK1 in strain YS59- to suggest that the TCA cycle was enhanced. This deduction
409 probably lead to a reduction in glycolysis flux compared could be supported by an earlier study showing that when
to the control strain, thus well explaining the reduced yield of glucose uptake rates were reduced, then the CO2 /ethanol ratio
ethanol. Varela et al. (2018) have reported that the low ethanol increased more than 50% and the net flux through the TCA
strain produced in their study exhibited reduced glycolysis cycle increased significantly (Heyland et al., 2009). That is to
activity, which agrees well with the present study. What’s more, say, the overflow metabolism of the YS59-409 strain might
a reduction in pyruvate content (Figure 3 and Supplementary shift toward respiratory metabolism. Our specific experimental
Table 6) provides further evidence of the weaker glycolysis flux conditions (the headspace of the 24-well plates and test tubes)
(Otterstedt et al., 2004). We hypothesize that the glycolysis flux could provide a micro-oxygen environment that can support
of the YS59-409 strain is reduced by mutation of the SPT15 gene, the respiration of yeast to some extent, as also noted by a
thus resulting in a weakened ethanol fermentation pathway of the previous study (Heyland et al., 2009). Respiratory metabolism
YS59-409 strain. plays an important role in terms of producing energy in the
form of ATP in aerobic growing cells. Intriguingly, we found
The Low-Ethanol-Producing Strain YS59-409 that the energetic metabolism of YS59-409 may be enhanced
Exhibited a Disturbance in the Crabtree Effect compared to the control strain. As is well-known, the histidine
Saccharomyces cerevisiae is a Crabtree-positive yeast, which and nucleotide biosynthetic pathways are connected (Ljungdahl
generally utilizes the ethanol fermentation pathway for glucose and Daignan-Fornier, 2012). The upregulation of genes related
to histidine metabolism (such as HIS1, HIS2, HIS5, and HIS7),
accompanied by the increased histidine content in YS59-409
TABLE 3 | KEGG pathway enrichment analysis (p-value < 0.05). (Figure 3), provided strong evidence for an enhancement in the
synthesis of histidine. And KEGG pathway enrichment analysis
Pathway P-value Number of Gene match indicated that purine synthesis metabolism was reinforced in
genes (genome match)a
the mutant strain (Table 3 and Supplementary Tables 4, 5).
Translation So, with the viewpoint that the de novo purine pathway feeds
Ribosome 1.00E-71 135 19.77 (4.33) into the histidine pathway and branches to allow ATP synthesis
Transcription (Ljungdahl and Daignan-Fornier, 2012), we can speculate that
RNA polymerase 5.87E-05 14 2.05 (0.69) ATP production in the YS59-409 strain is boosted. In addition,
Carbohydrate metabolism the upregulation of the ADE4 gene in the first step of the
Starch and sucrose metabolism 6.27E-10 22 3.22 (0.84) ATP formation pathway (Ljungdahl and Daignan-Fornier, 2012)
Galactose metabolism 6.11E-04 13 1.90 (0.74) provided supplementary evidence to the enhanced ATP synthesis.
Nucleotide metabolism What’s more, high ATP yields may result in excess biomass
Pyrimidine metabolism 1.28E-04 24 3.51 (1.65) formation at the expense of product yield (de Kok et al., 2012),
Purine metabolism 3.23E-02 24 3.51 (2.41) which further supports the fact that YS59-409 showed higher
Amino acid metabolism biomasses and lower ethanol production. Therefore, we can
Histidine metabolism 4.08E-03 6 0.88 (0.26) infer that the characteristic of Crabtree effect of the mutant
Arginine and proline metabolism 4.58E-02 6 0.88 (0.41) strain have been changed. On the other hand, some Crabtree-
a Thepercentage of DEGs involved in individual pathway account for all DEGs with negative yeast, such as Hanseniaspora uvarum and Metschnikowia
pathway annotation (683) and all genes with pathway annotation (4184). pulcherrima, possess different respiro-fermentative regulatory

Frontiers in Microbiology | www.frontiersin.org 6 December 2020 | Volume 11 | Article 597828


Du et al. Low-Alcohol-Production Yeast

FIGURE 3 | Changes in genes and metabolites of mutant strain YS59-409. Blue and green represent up- and down-regulation, respectively. The values in brackets
represent the fold change of expression level of mutant strain compared to control strain.

mechanisms than S. cerevisiae (Quiros et al., 2014). Researchers


have used these non-conventional species to reduce ethanol
production by partial controlling the aeration of grape juice
(Quiros et al., 2014). However, non-conventional yeasts cannot
generally complete alcoholic fermentation. We created a new
mutant strain (YS59-409), which might not only have similar
ethanol-producing properties as non-conventional yeasts, but
also is capable of finishing fermentation alone. Thus, these results
provide a new concept for the creation of new low-ethanol-
production strains.

NAD+ /NADH Homeostasis Was Disturbed in Strain


YS59-409
Sugar fermentation in S. cerevisiae is a redox neutral process
that is influenced by NAD+ /NADH balance, in which glycerol
plays important roles (Goold et al., 2017). We found those key
genes in the synthesis pathway of glycerol, GPD1, encoding
glycerol-3-phosphate dehydrogenase, along with GPP1 and
FIGURE 4 | CO2 weight loss (g). Black asterisk and black triangle represent
GPP2, encoding glycerol-3-phosphate phosphatase (Nevoigt knockout strains (YS59-1rgi1 and YS59-1rgi2, respectively) and black
and Stahl, 1997), were overexpressed in the low-ethanol- square represents control strain (YS59).
producing strain (Figure 3 and Supplementary Table 4).
Simultaneously, GUT1 and GUT2, genes that encode for glycerol
kinase for glycerol catabolism (Nevoigt and Stahl, 1997) were
both downregulated (Figure 3 and Supplementary Table 4). YS59-409 strain (Table 1). Higher production of glycerol is
Changes in the expression of those genes related to glycerol likely due to the need to balance cytosolic NADH produced
metabolism demonstrated the high production of glycerol; this and consumed. What is noticeable is that the de novo
was confirmed by the increased glycerol concentration in the pathway of NAD+ (namely, the kynurenine pathway) might

Frontiers in Microbiology | www.frontiersin.org 7 December 2020 | Volume 11 | Article 597828


Du et al. Low-Alcohol-Production Yeast

TABLE 4 | Fermentation characteristics of knockout strains.

Strains Residual sugar (g/L) Ethanol (g/L) Ethanol yield (g/g Sugar) Change in ethanol yield

YS59 1.96 ± 0.68a 52.04 ± 1.43c 0.351 ± 0.008c 0


YS59-1rgi1 0.99 ± 0.19a 40.59 ± 0.96a 0.272 ± 0.006a −22.0%
YS59-1rgi2 1.62 ± 0.40a 43.43 ± 0.77b 0.292 ± 0.004b −16.5%
a,b,c Values
followed by different letters within the same column are significantly different using Duncan’s multiple range test at the level of 0.05. Data are mean ± SD of
independent triplicate.

have been disturbed. In yeast, NAD+ can be synthesized de glycerol and energy metabolism observed in YS59-409. Then, we
novo from tryptophan (Panozzoa et al., 2002). We found found that only RGI2 exerted influence over ethanol production.
that tryptophan content was increased in the mutant strain The RGI2 gene possesses very little data concerning protein
(Figure 3 and Supplementary Table 6); this was consistent function or biological processes involved for itself, which was
with the fact that genes involved in the synthesis of tryptophan found to be significantly down-regulated by 41.9-fold in this
were also increased (Figure 3 and Supplementary Table 4). study (Supplementary Table 4). Noteworthily, the homologous
Moreover, as the key participator in the de novo synthesis of gene of RGI2, RGI1 (Yer067w), was significantly down-regulated
NAD+ from L-tryptophan (Panozzoa et al., 2002), kynurenine by 4.3-fold (Supplementary Table 4), which shared 70% identity
was found to be the increased metabolite with the highest with the sequence of RGI2 (Domitrovic et al., 2010). RGI1 gene
fold-change in this study (Figure 3 and Supplementary is reported to be regulated transcriptionally by SPT15 under
Table 6). Previous studies have shown that the de novo conditions involving ethanol stress (Yang et al., 2011). Previous
pathway plays only a minor role if a functional salvage studies show that Rgi1p and Rgi2p proteins most likely belong
pathway is present (Sporty et al., 2009); and one of the to the same complex and/or operate in the same pathway, and
key requirements in the formation of kynurenine is that the that these proteins are involved in the control of energetic
kynurenine pathway needs oxygen (Panozzoa et al., 2002). metabolism, particularly under respiratory growth conditions
Therefore, the increased production of kynurenine and L- (Domitrovic et al., 2010). Therefore, the effects of those two
tryptophan in the mutant strain probably shows the activation genes on ethanol metabolism were conducted in the present
of the de novo pathway for the synthesis of NAD+ under our study. We performed single gene knock-outs for RGI1 and
experimental conditions. The strengthening trend of NAD+ RGI2 genes in strain YS59 using Cre/loxP-mediated technology
level might further explain the enhanced synthesis of ATP in order to evaluate their effects on ethanol production in
synthesis for the reason that ATP synthesis and redox potential S. cerevisiae, generating two new strains: YS59-1rgi1 and YS59-
are directly proportional to the intracellular concentration of 1rgi2 (Supplementary Table 1). As follow, we compared the
NAD+ (Gonzalez Esquivel et al., 2017). Collectively, those data performance of strain YS59-1rgi1 and strain YS59-1rgi2 in
indicate that NAD+ /NADH equilibrium might be disturbed Triple M media (250-mL flasks containing 150-mL of media).
in the mutant strain. Certainly, the NAD+ /NADH balance of Compared with the control strain YS59, the weight loss of
the YS59-409 strain depends on a range of factors, including CO2 was both reduced in the two knockout strains (Figure 4),
biomass formation, respiration, ATP production, and the thus illustrating their reduced fermentation capacity; this was
generation of some metabolites, such as ethanol, glycerol, further supported by the fact that these strains showed a
and amino acids. Additionally, researchers have deployed 22.0 and 16.5% reduction in ethanol productivities, respectively
methods to intentionally perturb the NAD+ /NADH balance (Table 4). The results implied that perturbation of the RGI1
to reduce ethanol production (Goold et al., 2017), which or RGI2 gene could elicit alterations in ethanol metabolism,
also emphasizes the importance of NAD+ /NADH balance for which showed the positive effects of RGI1/2 deletion with regards
ethanol metabolism. to weakening the ethanol-yielding ability of yeast. However,
these two knockout strains may have a different low-yield
The Effects of Deleting the RGI1/2 Gene ethanol mechanism than the YS59-409 strain, because they
on the Ethanol-Production Capacity of had different CO2 production models; this possibility requires
further investigation.
Yeast
To further identify key genes in the regulatory network of our
new mutant strain, we constructed a protein-protein interaction CONCLUSION
(PPI) network using the STRING database V113 and Cytoscape
tools; the core gene module was then excavated using the In this study, we applied gTME technology to engineer a low-
MCODE plugin in Cytoscape. Nine genes were identified ethanol-production strain of S. cerevisiae achieved by mutating
(Supplementary Table 7); of these, AQY3 (Yfl054c) and RGI2 the transcription factor SPT15 to change gene expression in
(Yil057c) were shown to be associated with glycerol and energy a global level. We provide a novel insight into the use of
metabolism, which was corresponded with the changes in gTME technology to modulate ethanol metabolism, which could
not only facilitate the construction of a low-ethanol-production
3
https://string-db.org/ strain for the wine industry, but also, enhance our understanding

Frontiers in Microbiology | www.frontiersin.org 8 December 2020 | Volume 11 | Article 597828


Du et al. Low-Alcohol-Production Yeast

of the mechanisms underlying reduced ethanol production by DATA AVAILABILITY STATEMENT


yeast. We created a new strain of S. cerevisiae, YS59-409, with
weakened ethanol production capacity. The ethanol-production The datasets generated for this study can be found in
capacity of this strain was reduced by 34.9% compared online repositories. The names of the repository/repositories
to the control strain, which was caused by comprehensive and accession number(s) can be found in the article/
changes associated with the regulation of transcription and Supplementary Material.
metabolism. Sequence analysis was performed on the mutated
SPT15 gene, demonstrating that the five mutation sites may
work collectively, or at least partly, to create the specific AUTHOR CONTRIBUTIONS
characteristics of YS59-409, including a higher CO2 release,
biomass, and glycerol formation. The integration of RNA-Seq QD: experimental operation and writing – original draft. YL:
and metabolomics analysis showed that the specific phenotype supervision. YS: data analysis. YQ: experimental design and
of the new mutant strain featured changes in ribosome supervision. All authors edited and approved the final version
biogenesis, nucleotide metabolism, glycolysis flux, the Crabtree of the manuscript.
effect, NAD+ /NADH homeostasis, and energy metabolism.
Remarkably, we also found that RGI1 and RGI2 genes,
which play key roles in energy metabolism, were significantly FUNDING
down-regulated; it is possible that this was linked with low
ethanol metabolism, although this needs to be investigated This project was supported by National Key Research and
further in future research. Although public attitudes toward Development Project (2019YFD1002500), National Natural
the use of GMOs (Genetically Modified Organisms) in wines Science Foundation of China (31301541 and 31960470), the
are often less than positive, this study demonstrated that it Fundamental Research Funds for the Central Universities
is possible to reduce ethanol yields in yeast using gTME (2452020177), and China Agriculture Research System (grant
technology, while also reprogramming the metabolism of this no. CARS-29-jg-3).
new mutant strain. Currently, we can get some knowledge
from this study, which could be used to direct strategies
for generating wine yeast with weakened ethanol production SUPPLEMENTARY MATERIAL
capacity using other approaches, such as adaptive evolution.
In summary, this study highlights the potential to use gTME The Supplementary Material for this article can be found
technology to reduce the ethanol content of yeast for the wine- online at: https://www.frontiersin.org/articles/10.3389/fmicb.
making industry. 2020.597828/full#supplementary-material

REFERENCES composition under climate change conditions. Oeno One 51, 205. doi: 10.20870/
oeno-one.2016.0.0.1584
Alper, H., Moxley, J., Nevoigt, E., Fink, G. R., and Stephanopoulos, G. (2006). Domitrovic, T., Kozlov, G., Freire, J. C., Masuda, C. A., Almeida, M., Montero-
Engineering yeast transcription machinery for improved ethanol tolerance and Lomeli, M., et al. (2010). Structural and functional study of Yer067w, a new
production. Science 314, 1565–1568. doi: 10.1126/science.1131969 protein involved in yeast metabolism control and drug resistance. PLoS One
Alper, H., and Stephanopoulos, G. (2007). Global transcription machinery 5:e11163. doi: 10.1371/journal.pone.0011163
engineering: A new approach for improving cellular phenotype. Metab. Eng. Ehsani, M., Fernandez, M. R., Biosca, J. A., Julien, A., and Dequin, S. (2009).
9, 258–267. doi: 10.1016/j.ymben.2006.12.002 Engineering of 2,3-butanediol dehydrogenase to reduce acetoin formation by
Arndt, K. M., Ricupero, S. L., Eisenmann, D. M., and Winston, F. (1992). glycerol-overproducing, low-alcohol Saccharomyces cerevisiae. Appl. Environ.
Biochemical and genetic characterization of a yeast TFIID mutant that alters Microbiol. 75, 3196–3205. doi: 10.1128/AEM.02157-08
transcription in vivo and DNA binding in vitro. Molecular and Cellular Biology Eisenmann, D. M., Dollard, C., and Winston, F. (1989). SPT15, the gene encoding
12, 2372–2382. doi: 10.1128/mcb.12.5.2372 the yeast TATA binding factor TFIID, is required for normal transcription
Backhus, L. E., Derisi, J., Brown, P. O., and Bisson, L. F. (2001). Functional genomic initiation in vivo. Cell 58, 1183–1191. doi: 10.1016/0092-8674(89)90
analysis of a commercial wine strain of Saccharomyces cerevisiae under differing 516-3
nitrogen conditions. FEMS Yeast Res. 1, 111–125. doi: 10.1111/j.1567-1364. Gietz, R. D., and Woods, R. A. (2001). Genetic transformation of yeast.
2001.tb00022.x Biotechniques 30, 816–831. doi: 10.2144/01304rv02
Cambon, B., Monteil, V., Remize, F., Camarasa, C., and Dequin, S. (2006). Effects Gonzalez Esquivel, D., Ramirez-Ortega, D., Pineda, B., Castro, N., Rios, C., and
of GPD1 overexpression in Saccharomyces cerevisiae commercial wine yeast Perez de la Cruz, V. (2017). Kynurenine pathway metabolites and enzymes
strains lacking ALD6 genes. Appl. Environ. Microbiol. 72, 4688–4694. doi: 10. involved in redox reactions. Neuropharmacology 112, 331–345. doi: 10.1016/
1128/AEM.02975-05 j.neuropharm.2016.03.013
de Deken, R. H. (1966). The crabtree effect: A regulatory system in yeast. J. Gen. Goold, H. D., Kroukamp, H., Williams, T. C., Paulsen, I. T., Varela, C., and
Microbiol. 44, 149–156. doi: 10.1099/00221287-44-2-149 Pretorius, I. S. (2017). Yeast’s balancing act between ethanol and glycerol
de Kok, S., Kozak, B. U., Pronk, J. T., and van Maris, A. J. (2012). Energy coupling production in low-alcohol wines. Microb. Biotechnol. 10, 264–278. doi: 10.1111/
in Saccharomyces cerevisiae: Selected opportunities for metabolic engineering. 1751-7915.12488
FEMS Yeast Res. 12, 387–397. doi: 10.1111/j.1567-1364.2012.00799.x Heyland, J., Fu, J., and Blank, L. M. (2009). Correlation between TCA cycle flux
Dequin, S., Escudier, J. L., Bely, M., Noble, J., Albertin, W., Masneuf-Pomarède, and glucose uptake rate during respiro-fermentative growth of Saccharomyces
I., et al. (2017). How to adapt winemaking practices to modified grape cerevisiae. Microbiology 155, 3827–3837. doi: 10.1099/mic.0.030213-0

Frontiers in Microbiology | www.frontiersin.org 9 December 2020 | Volume 11 | Article 597828


Du et al. Low-Alcohol-Production Yeast

Hranilovic, A., Gambetta, J. M., Jeffery, D. W., Grbin, P. R., and Jiranek, V. Sporty, J., Lin, S. J., Kato, M., Ognibene, T., Stewart, B., Turteltaub, K., et al. (2009).
(2020). Lower-alcohol wines produced by Metschnikowia pulcherrima and Quantitation of NAD+ biosynthesis from the salvage pathway in Saccharomyces
Saccharomyces cerevisiae co-fermentations: The effect of sequential inoculation cerevisiae. Yeast 26, 363–369. doi: 10.1002/yea.1671
timing. Int. J. Food Microbiol. 329, 108651. doi: 10.1016/j.ijfoodmicro.2020. Tan, S. Z., Manchester, S., and Prather, K. L. (2016). Controlling central carbon
108651 metabolism for improved pathway yields in Saccharomyces cerevisiae. ACS
Hu, R., Lin, L., Liu, T., Ouyang, P., He, B., and Liu, S. (2008). Reducing sugar Synth. Biol. 5, 116–124. doi: 10.1021/acssynbio.5b00164
content in hemicellulose hydrolysate by DNS method: A revisit. J Biobased Tarazona, S., Garcia-Alcalde, F., Dopazo, J., Ferrer, A., and Conesa, A. (2011).
Mater. Bio. 2, 156–161. doi: 10.1166/jbmb.2008.306 Differential expression in RNA-seq: A matter of depth. Genome Res. 21, 2213–
Li, Y., Zhang, Y., Liu, M., Qin, Y., and Liu, Y. (2019). Saccharomyces cerevisiae 2223. doi: 10.1101/gr.124321.111
isolates with extreme hydrogen sulfide production showed different oxidative Tilloy, V., Cadière, A., Ehsani, M., and Dequin, S. (2015). Reducing alcohol
stress resistances responses during wine fermentation by RNA sequencing levels in wines through rational and evolutionary engineering of Saccharomyces
analysis. Food Microbiol. 79, 147–155. doi: 10.1016/j.fm.2018.10.021 cerevisiae. Int. J. Food Microbiol. 213, 49–58. doi: 10.1016/j.ijfoodmicro.2015.06.
Liu, N., Wang, D., Wang, Z. Y., He, X. P., and Zhang, B. (2007). Genetic basis 027
of flocculation phenotype conversion in Saccharomyces cerevisiae. FEMS Yeast Tilloy, V., Ortiz-Julien, A., and Dequin, S. (2014). Reduction of ethanol yield
Res. 7, 1362–1370. doi: 10.1111/j.1567-1364.2007.00294.x and improvement of glycerol formation by adaptive evolution of the wine
Ljungdahl, P. O., and Daignan-Fornier, B. (2012). Regulation of amino acid, yeast Saccharomyces cerevisiae under hyperosmotic conditions. Appl. Environ.
nucleotide, and phosphate metabolism in Saccharomyces cerevisiae. Genetics Microbiol. 80, 2623–2632. doi: 10.1128/AEM.03710-13
190, 885–929. doi: 10.1534/genetics.111.133306 Trapnell, C., Roberts, A., Goff, L., Pertea, G., Kim, D., Kelley, D. R., et al. (2012).
Nevoigt, E., and Stahl, U. (1996). ). Reduced pyruvate decarboxylase and Differential gene and transcript expression analysis of RNA-seq experiments
increased glycerol-3-phosphate dehydiogenase [NAD+] levels enhance glycerol with TopHat and Cufflinks. Nat. Protoc. 7, 562–578. doi: 10.1038/nprot.2012.
production in Saccharomyces cerevisiae. Yeast 12, 1331–1337. doi: 10.1002/ 016
(SICI)1097-0061(199610)12:13<1331::AID-YEA28<3.0.CO;2-0 Varela, C., Dry, P. R., Kutyna, D. R., Francis, I. L., Henschke, P. A., Curtin, C. D.,
Nevoigt, E., and Stahl, U. (1997). Osmoregulation and glycerol metabolism in the et al. (2015). Strategies for reducing alcohol concentration in wine. Aust. J.
yeast Saccharomyces cerevisiae. FEMS Microbiol. Rev. 21, 231–241. doi: 10.1016/ Grape Wine R. 21, 670–679. doi: 10.1111/ajgw.12187
S0168-6445(97)00058-2 Varela, C., Kutyna, D. R., Solomon, M. R., Black, C. A., Borneman, A., Henschke,
Otterstedt, K., Larsson, C., Bill, R. M., Stahlberg, A., Boles, E., Hohmann, S., et al. P. A., et al. (2012). Evaluation of gene modification strategies for the
(2004). Switching the mode of metabolism in the yeast Saccharomyces cerevisiae. development of low-alcohol-wine yeasts. Appl. Environ. Microbiol. 78, 6068–
EMBO Rep. 5, 532–537. doi: 10.1038/sj.embor.7400132 6077. doi: 10.1128/AEM.01279-12
Özcan, S., and Johnston, M. (1999). Function and regulation of yeast hexose Varela, C., Schmidt, S. A., Borneman, A. R., Pang, C. N. I., Kromerx, J. O., Khan,
transporters. Microbiol. Mol. Biol. Rev. 63, 554–569. A., et al. (2018). Systems-based approaches enable identification of gene targets
Panozzoa, C., Nawara, M., Suskia, C., and Kucharczyka, R. (2002). Aerobic and which improve the flavour profile of low-ethanol wine yeast strains. Metab. Eng.
anaerobic NAD+ metabolism in Saccharomyces cerevisiae. FEBS Lett. 517, 49, 178–191. doi: 10.1016/j.ymben.2018.08.006
97–102. doi: 10.1016/S0014-5793(02)02585-1 Varela, J., and Varela, C. (2018). Microbiological strategies to produce beer and
Quiros, M., Rojas, V., Gonzalez, R., and Morales, P. (2014). Selection of non- wine with reduced ethanol concentration. Curr. Opin. Biotechnol. 56, 88–96.
Saccharomyces yeast strains for reducing alcohol levels in wine by sugar doi: 10.1016/j.copbio.2018.10.003
respiration. Int. J. Food Microbiol. 181, 85–91. doi: 10.1016/j.ijfoodmicro.2014. Yang, J., Bae, J. Y., Lee, Y. M., Kwon, H., Moon, H. Y., Kang, H. A., et al.
04.024 (2011). Construction of Saccharomyces cerevisiae strains with enhanced ethanol
Rodríguez, A., De La Cera, T., Herrero, P., and Moreno, F. (2001). The hexokinase tolerance by mutagenesis of the TATA-binding protein gene and identification
2 protein regulates the expression of the GLK1, HXK1 and HXK2 genes of of novel genes associated with ethanol tolerance. Biotechnol. Bioeng. 108,
Saccharomyces cerevisiae. Biochem. J. 355, 625–631. doi: 10.1042/bj3550625 1776–1787. doi: 10.1002/bit.23141
Rolle, L., Englezos, V., Torchio, F., Cravero, F., Ríosegade, S., Rantsiou, K., et al.
(2017). Alcohol reduction in red wines by technological and microbiological Conflict of Interest: The authors declare that the research was conducted in the
approaches: A comparative study. Aust. J. Grape Wine R. 24, 62–74. doi: 10. absence of any commercial or financial relationships that could be construed as a
1111/ajgw.12301 potential conflict of interest.
Seong, Y. J., Park, H., Yang, J., Kim, S. J., Choi, W., Kim, K. H., et al. (2017).
Expression of a mutated SPT15 gene in Saccharomyces cerevisiae enhances Copyright © 2020 Du, Liu, Song and Qin. This is an open-access article distributed
both cell growth and ethanol production in microaerobic batch, fed-batch, and under the terms of the Creative Commons Attribution License (CC BY). The
simultaneous saccharification and fermentations. Appl. Microbiol. Biotechnol. use, distribution or reproduction in other forums is permitted, provided the
101, 3567–3575. doi: 10.1007/s00253-017-8139-2 original author(s) and the copyright owner(s) are credited and that the original
Spiropoulos, A., Tanaka, J., Flerianos, I., and Bisson, L. (2000). Characterization publication in this journal is cited, in accordance with accepted academic practice.
of hydrogen sulfide formation in commercial and natural wine isolates of No use, distribution or reproduction is permitted which does not comply with
saccharomyces. Am. J. Enol. Viticult. 51, 233–248. these terms.

Frontiers in Microbiology | www.frontiersin.org 10 December 2020 | Volume 11 | Article 597828


biomolecules

Review
Biotechnological Approaches to Lowering the Ethanol Yield
during Wine Fermentation
Ramon Gonzalez 1 , Andrea M. Guindal 1 , Jordi Tronchoni 2 and Pilar Morales 1, *

1 Instituto de Ciencias de la Vid y del Vino (CSIC, Gobierno de la Rioja, Universidad de La Rioja),
26007 La Rioja, Spain; rgonzalez@icvv.es (R.G.); andrea.martin@icvv.es (A.M.G.)
2 Faculty of Health Sciences, Valencian International University (VIU), 46002 Valencia, Spain;
jtronchoni@universidadviu.com
* Correspondence: pilar.morales@icvv.es

Abstract: One of the most prominent consequences of global climate warming for the wine industry
is a clear increase of the sugar content in grapes, and thus the alcohol level in wines. Among the
several approaches to address this important issue, this review focuses on biotechnological solutions,
mostly relying on the selection and improvement of wine yeast strains for reduced ethanol yields.
Other possibilities are also presented. Researchers are resorting to both S. cerevisiae and alternative
wine yeast species for the lowering of alcohol yields. In addition to the use of selected strains under
more or less standard fermentation conditions, aerobic fermentation is increasingly being explored for
this purpose. Genetic improvement is also playing a role in the development of biotechnological tools
to counter the increase in the wine alcohol levels. The use of recombinant wine yeasts is restricted to
research, but its contribution to the advancement of the field is still relevant. Furthermore, genetic

 improvement by non-GMO approaches is providing some interesting results, and will probably
Citation: Gonzalez, R.; Guindal, result in the development of commercial yeast strains with a lower alcohol yield in the near future.
A.M.; Tronchoni, J.; Morales, P. The optimization of fermentation processes using natural isolates is, anyway, the most probable
Biotechnological Approaches to source of advancement in the short term for the production of wines with lower alcohol contents.
Lowering the Ethanol Yield during
Wine Fermentation. Biomolecules 2021, Keywords: wine; low-alcohol; fermentation; global warming
11, 1569. https://doi.org/10.3390/
biom11111569

Academic Editors: Encarna 1. Introduction


Gómez-Plaza and Rocio Gil-Muñoz
Commercial wines have been experiencing a steady increase in alcohol levels since the
1980s. On average, the gain in alcohol has been about 1% (v/v) every ten years, currently
Received: 26 September 2021
Accepted: 20 October 2021
reaching a total increase of 3–4% (v/v). There are two main reasons behind this trend.
Published: 22 October 2021
From one side, global climate warming affects, in different ways, the different physiological
processes involved in grape berry ripening. Both sugar accumulation and phenolic and
Publisher’s Note: MDPI stays neutral
aromatic maturity develop faster, but the impact is greater for sugars. This imbalance
with regard to jurisdictional claims in
prevents early harvesting from being an adequate solution to this problem, as it would lead
published maps and institutional affil- to organoleptic defects associated with unripe berries. On the other hand, the consumer
iations. demand for tasty, structured, and full-bodied wines has pushed the market towards
winemaking styles that require the harvest of very ripe grapes. As a result, quality grapes
today tend to be richer in sugar content, leading to the observed increase in wine alcohol
levels when all this sugar is fermented by yeasts [1].
Copyright: © 2021 by the authors.
Excess alcohol in wines is becoming a problem for the industry from several points
Licensee MDPI, Basel, Switzerland.
of view. Wine balance is a delicate equilibrium between sweetness, acidity, bitterness,
This article is an open access article
and aroma compounds, with alcohol as a vehicle for the perception of the latter. Ethanol
distributed under the terms and surplus can lead to a sensory imbalance by promoting bitterness and masking fruit per-
conditions of the Creative Commons ceptions, and increasing the perception of sweetness, astringency and hotness. Besides
Attribution (CC BY) license (https:// this, sticking to moderate alcohol consumption becomes more difficult, and can therefore
creativecommons.org/licenses/by/ lead to increased risks for consumer health and road safety. Taxation based on alcohol
4.0/). levels in several importing countries is an additional incentive to look for wines with lower

Biomolecules 2021, 11, 1569. https://doi.org/10.3390/biom11111569 https://www.mdpi.com/journal/biomolecules


Biomolecules 2021, 11, 1569 2 of 16

ethanol contents. Taken together, all of these factors might contribute to discourage wine
consumption. Furthermore, from a bioprocess point of view, high levels of ethanol could
lead to sluggish or stuck fermentation processes, with significant economic consequences.
In recent years, there has been a collective effort by researchers, engineers, and wine-
makers to develop approaches to limit the ethanol content of wines [2]. This community
is targeting almost every stage of the production cycle, including, among other examples,
grapevine clonal selection, vineyard management, winemaking practices adapted to unripe
grapes, the use of yeast strains with a lower ethanol yield (often recombinant) or metabolic
inhibitors, and partial dealcoholisation by physical means. Unfortunately, straightforward
solutions, such as early harvesting or post-fermentation treatments, often have a negative
impact on the wine quality (e.g., a green character and altered aromatic profile). Apart from
sweet wines, most wines on the market are dry wines, which should contain very low
levels of residual sugar. Solutions that result in less alcohol but an increase in residual
sugars (i.e., stopping fermentation before completion) will drastically change the wine
style, and will not be considered in this review. Here, we will present biotechnological
solutions related to the fermentation process, mainly involving the selection and improve-
ment of wine yeast strains for reduced ethanol yield, but also the use of some enzymes or
metabolic inhibitors.
In this context, a pivotal concept is alcohol yield, which is the amount of ethanol
produced per unit of sugar consumed. Most microbiological approaches to the reduction
of the alcohol level in wines will target the ethanol yield of S. cerevisiae or alternative
wine yeast species. Associated with the lowering of the alcohol yield is the concept of
alternative carbon sinks (carbon-containing metabolic end products other than ethanol).
Under standard fermentation conditions, about 60% of the hexose carbon consumed by
S. cerevisiae ends up in the form of ethanol, 30% goes to carbon dioxide, and the remainder
goes to biomass and less abundant metabolites. This led to the rule of thumb which states
that for every 17 g/L of sugar in the grape must, an additional 1% v/v of alcohol wine is to
be expected in the wine. Reducing the alcohol yield requires the diversion of the metabolic
flux from ethanol to alternative carbon sinks.
However, alcoholic fermentation is the metabolic hallmark of Saccharomyces cerevisiae,
the yeast species dominating most wine fermentation processes in the industry, either spon-
taneously or because it is used as a starter culture. This species always shows a high ethanol
yield during fermentation, with very little biological diversity [3,4]. As an example, a sur-
vey of 72 S. cerevisiae strains from various geographical origins and ecological niches found
great phenotypic diversity for other traits, but almost no differences in ethanol yield [4].
Below, we describe the different efforts made to overcome this biotechnological hurdle.

2. Genetic Improvement of S. cerevisiae


2.1. Genetic Engineering
Since the 1990s, it has been a constant in wine biotechnology that product improvement
and problem solving tend to be addressed, in the first instance, by the genetic engineering
of wine yeasts. However, the wine industry is not the ideal breeding ground for genetically
modified organisms due to restrictions to GMOs on food production in most countries,
and their negative perception by wine consumers [5]. The only two recombinant wine
yeasts that have been commercialised so far do not seem to have become bestsellers in the
markets they were introduced in [5,6]. Therefore, none of the recombinant strains described
in this section were intended for direct commercialization. Nevertheless, their study
provided useful information for the improvement of yeast strains by alternative approaches.
Indeed, in several instances, the reduction in alcohol levels appeared as a side-effect of
genetic manipulations which were not intended for that purpose. As mentioned above,
when it comes to reducing the alcohol level, a key issue to guide genetic engineering is
the choice of carbon sinks. Figure 1 shows a summary of the main enzymatic reactions
involved in alcoholic fermentation, including those referred to on the following lines.
Biomolecules 2021, 11, 1569 3 of 16

Alcohol dehydrogenase (ADH) catalyses the final step in the production of ethanol
during alcoholic fermentation. It is therefore a rational deletion target for limiting ethanol
production. However, cell metabolism is complex and depends on a set of biochemical
reactions that must be balanced in several ways, especially regarding the redox balance.
For this reason, simple logics often lead to unwanted results. For example, alcohol dehy-
drogenase deletion leads to a reduction in ethanol production and an increase in that of
glycerol, but these recombinant strains are unable to grow under anaerobic conditions [7].
This genetic modification also results in higher levels of acetic acid and acetaldehyde,
far beyond levels suitable for commercial wines.
In addition to the modulation of the expression of endogenous genes by shutting
down or upregulating their expression, genetic engineering allows for the heterologous
expression of genes from other biological species. In 1994, Dequin and Barre [8] expressed
lactate dehydrogenase from Lactobacillus casei in S. cerevisiae to divert some of the pyruvate
formed in glycolysis to lactate formation. The engineered yeast strain simultaneously
performed alcoholic and lactic fermentations. The redox balance was not affected, as the
formation of lactic acid is compensated by the disappearance of ethanol in equimolecular
amounts, for the same consumption of NADH. Thus, a reduction of 1% ethanol leads to an
increase in 15 g/L lactic acid. The increased lactate content of these wines contributed to a
higher total acidity. This feature could be useful to compensate for low acidity, which is a
problem affecting some grape varieties and growing regions, and is also related to global
warming. However, this also sets a limit to this approach, as excessive acidity would make
the wines unpalatable.
Glycerol contributes positively to the viscosity, body, and sweetness of wine. It has
been a preferred carbon sink for the genetic engineering of wine yeasts. The overexpression
of GPD1, coding for glycerol 3-P dehydrogenase, results in a significant increase in glycerol
(plus 28 g/L) and a reduction in ethanol (minus 15 g/L), and also in the appearance of
excess acetate and acetoin to restore the intracellular redox balance during fermentation [9].
Researchers identified Ald6 as the aldehyde dehydrogenase isoenzyme responsible for most
of the acetate production of S. cerevisiae [10]. However, the knockout of ALD6 resulted in an
additional increase in acetoin production which, at high concentrations, negatively affects
wine aroma, providing buttery notes [11]. Following this line of research, the problem of
acetoin overproduction was tackled by overexpressing BDH1, which codes for butanediol
dehydrogenase [12]. This enzyme catalyses the conversion of acetoin to 2,3-butanediol,
a compound with no apparent sensory impact. Using this recombinant strain, wines show
a significant reduction in ethanol, and above 26 g/L of glycerol content. The number of
different genetic modifications required to achieve this objective is in itself an indication of
the difficulty of using this as a general approach to developing wine yeast strains with a
low ethanol yield.
The different adjustments that were required to develop the above-described recom-
binant yeasts respond to metabolic constraints concerning the redox balance and the
availability of an adequate pool of cofactors. Therefore, some authors have tried to act
directly on the reactions of the redox system. By cloning an NADH oxidase for the regen-
eration of the reducing power, and with controlled oxygenation, not only is a decrease in
ethanol obtained but also a significant increase in the levels of acetaldehyde, acetic acid,
and acetoin [13].
There are two enzymatic steps from pyruvate to ethanol during alcoholic fermentation
(Figure 1). The final one catalysed by alcohol dehydrogenase has been discussed above.
The first, from pyruvate to acetaldehyde, catalysed by pyruvate decarboxylase (PDC),
was also taken as an improvement target by some researchers. In the absence of PDC
activity, the Crabtree effect is abolished, but this activity is required for acetyl-CoA synthesis
in the cytoplasm and NAD+ regeneration. It is therefore essential for S. cerevisiae growing
on glucose [14]. The deletion of PDC1 resulted in a fourfold reduction of the PDC activity,
with an increase in pyruvate levels, but no reduction of ethanol levels [10]. The deletion of
PDC2, coding for a transcription factor involved in PDC gene expression, results in 19%
Biomolecules 2021, 11, 1569 4 of 16

PDC activity, with a clear reduction in the ethanol content, combined with an increase
in glycerol and pyruvate release [15]. This effect is enhanced by increased glycerol-3-
phosphate dehydrogenase activity, through the overexpression of GPD1, without the
over-production of acetic acid and acetaldehyde associated with GPD1 over-expression
in the original strain. Cuello et al. [16] replaced one of the native PDC2 alleles in different
diploid wine yeast strains by a truncated PDC2 allele. These engineered strains showed
a 2% ABV reduction without any increase in acetate production. An alternative way to
produce cytosolic acetyl CoA in pdc- strains was developed by substituting Acs1 and Acs2
(acetyl coA synthetase) with a heterologous pyruvate carboxylase and a transacetylase [17].
This strain produces glycerol, succinate, pyruvate and acetate from glucose, but not ethanol.
The additional deletion of gpp1 and gpp2 (glycerol 3p phosphatase) avoids the production
of these metabolites, including glycerol, and increases the growth rate.
Crabtree-negative strains of S. cerevisiae were also constructed by substituting all of the
active hexose transport genes with a single chimeric one [18]. This strain does not produce
ethanol as long as oxygen is available for respiration. However, its growth kinetics are
extremely slow. Further possibilities of using the respiratory metabolism of yeast to reduce
the ethanol yield during wine fermentation will be examined below in a specific section.
Triose-phosphate isomerase, encoded by TPI1, is the enzyme catalysing the inter-
conversion between the two products resulting from the breakdown of fructose 1,6-
bisphosphate: dihydroxyacetone phosphate and glyceraldehyde 3-phosphate (Figure 1).
The topology of the pathway suggests that TPI1 deletion would result in equimolar levels
of glycerol and ethanol precursors. However, in practice, such deletion strains are unable
to grow on glucose due to a redox imbalance [19]. In order to improve industrial glyc-
erol yields, additional genetic modifications have been introduced in strains deleted for
TPI1 [20]. These authors deleted ADH1 (coding for alcohol dehydrogenase) and TPI1 in a
yeast strain while overexpressing GPD1 (coding for glycerol-3-phosphate dehydrogenase),
FPS1 (coding for a glycerol transporter) and ALD3 (coding for an aldehyde dehydrogenase).
These strains produce very low levels of ethanol (about 25% of the parent strain), and their
glycerol levels are multiplied by 30. However, their growth rates were strongly reduced.
Another alternative carbon sink is yeast biomass. Despite the portion of carbon going
to biomass being quantitatively small during wine fermentation, some authors have tar-
geted reserve sugars to divert the carbon flux from ethanol. A moderate overexpression of
TPS1, the gene coding for trehalose-6-phosphate synthase, involved in trehalose biosynthe-
sis, resulted in a clear reduction in the ethanol yield, without an apparent impact on the
redox balance, according to the profile of the byproducts analysed [21].
Global transcription machinery engineering (gTME) technology has recently been
employed to construct a low-alcohol yeast strain. One strain with a reduced ethanol yield
was selected from a library of mutants in SPT15 (coding for a transcription factor) generated
by random directed mutagenesis [22]. The pleiotropic effect of this mutation downregulates
hexose transport and ethanol production, and upregulates glycerol production. Acetic acid
is not increased, but the mutant shows slower growth kinetics than the parent strain.
The treatment of must with glucose oxidase is an alternative method to decrease the
alcohol content of wine that will be discussed in the corresponding section. Based on
this method, Malherbe et al. [23] cloned a glucose oxidase gene from Aspergillus niger in a
laboratory strain of S. cerevisiae. Wines produced with the recombinant strain contained 2%
less ethanol. The recombinant strain showed antimicrobial activity against acetic and lactic
acid bacteria, probably due to the H2 O2 released by the oxidation reaction.
The CRISPR/Cas9 (Clustered Regularly Interspaced Short Palindromic Repeats/
CRISPR associated protein 9) genetic modification system has already been implemented
in S. cerevisiae, including wine yeasts [24,25]. The basic CRISPR/Cas9 system consists of a
guide RNA molecule (gRNA) and an endonuclease (Cas9). The endonuclease mediates a
double strand break guided by the hybridization of gRNA with genomic DNA. This triggers
the DNA repair mechanisms of the host cell. The system can be designed to target the
nuclease to the gene of interest, and to perform more or less large insertions and deletions.
Biomolecules 2021, 11, 1569 5 of 16

An advantage of this system is that, unlike other genetic engineering techniques, it can
easily be tuned to leave no trace of the ancillary genetic material. In addition, it is less
dependent on traditional transformation markers, and makes easier the manipulation
of diploid loci. Nonetheless, microorganisms obtained with this system still fall under
European Union GMO regulations. The genes GPD1 and ATF1 have been cloned in a
haploid wine strain using this technique [26], with no differences in ethanol production
between the strains, contrary to the results obtained in previous works [9,15]. Other works
applying this technique for the reduction of wine alcohol content will certainly appear soon.

2.2. Random Mutagenesis


Genetic improvement by random mutagenesis involves the artificial increase in
mutation rates by treatment with physical or chemical agents (UV radiation, X-rays,
ethyl methane sulfonate, nitrosoguanidine) coupled with a suitable selection strategy [27].
To be effective, the intensity of the treatments must be lethal for most of the cells in the
population. However, given the random nature of the process, the number of cells that
must remain viable in order to capture the expected mutations is often still too high for
strain-by-strain analysis. It is then critical to establish appropriate selection criteria that
link selectable or detectable phenotypes to technologically relevant features of the mutant
strains. While this seems like an easy task for some improvement outcomes—for example,
those related to the robustness of yeast strains—improvements in wine quality, including
the reduction of alcohol levels, require indirect and non-obvious selection criteria.
An improved strain of S. bayanus was developed by random mutagenesis, and was
marketed as a low ethanol producer (Oenoferm® LA-HOG from Erbslöh, Geisenheim,
Germany). According to commercial information, the genetic improvement process in-
volved two stages of chemical mutagenesis. The first one, with selection under hypersaline
conditions, was intended to recover strains with an improved HOG (High Osmolarity
Glycerol) response. In the second, the selection pressure was established with pyrazole,
an inhibitor of alcohol dehydrogenase. The commercial strain was selected among the
40 strains recovered from this last step. During the fermentation of the Pinot noir grape
juice, this strain produced 12.9 g/L glycerol and 94.6 g/L ethanol, compared to the 6.4 g/L
and 102.4 g/L, respectively, produced by the parent strain.

2.3. Adaptive Laboratory Evolution


Adaptive laboratory evolution (also known as experimental evolution) consists of
recreating in the laboratory, in an accelerated way, the natural selection process, promot-
ing the selection of the mutations that are most useful to our purpose. This technique
shares several features with random mutagenesis [27], including some limitations, such as
that it cannot be easily targeted to specific genes, that dominant mutations are favoured,
and the difficulties of connecting readily selectable phenotypes to technological traits. How-
ever, it also offers advantages, as the strains obtained in this way do not fall under GMO
regulations. In this case, mutations appear spontaneously in the population during each
round of duplication (although it can eventually be combined with random mutagenesis).
Yeasts are grown in a selective environment in which the desired metabolic changes would
favour their growth. Once such a mutation appears, and after several generations under
the same selection conditions, many individuals in the population will harbour the desired
mutation. In addition, new variants can arise from the original one, and in this way the
favourable mutations for the intended technological objective can be accumulated in the
evolving lineage. The random nature of spontaneous mutations, and the fact that many
of them can be added over time during experimental evolution, gives some advantages
to this approach over genetic engineering. As indicated above, the challenge of this tech-
nology is to anticipate the metabolic pathways to be altered in order to reach the expected
technological output, and the growth conditions that will give advantage to the yeast cells
improved in that way.
Biomolecules 2021, 11, 1569 6 of 16

In the context of alcohol level reduction, adaptive laboratory evolution has been
used to channel carbon flux towards the pentose phosphate pathway (PPP; Figure 1).
The selective pressure was established using gluconate as the only carbon source [28].
Gluconate is a substrate poorly assimilated by S. cerevisiae and metabolized by the PPP.
The rationale for targeting PPP was that, compared to glycolysis, an additional CO2
molecule is released for each glucose molecule entering the pathway (meaning that this
carbon will not contribute to ethanol production). The selected strain showed a 1.5-fold
increase in flux through the PPP compared to the ancestral strain, corresponding to a
reduction of 2 g/L in ethanol levels. A fivefold increase in flux through the PPP would
have been necessary to reduce the ethanol level of wine by approximately 1% (vol/vol).
However, selected evolved strains did not show a reduced ethanol yield, but an increased
production of esters [29], which can be useful, but not for the original purpose.
Other authors have used experimental evolution to push carbon flux towards glyc-
erol production. The process was based on a long-standing strategy for the industrial
production of glycerol based on sulphite addition [30]. The reaction between sulphite
and acetaldehyde decreases the availability of acetaldehyde for ethanol production and
NADH oxidation, and cells must compensate the redox balance by increasing glycerol
production. Kutyna et al. [31] used this principle to drive the evolution of a laboratory
strain, with sodium sulphite as a selective agent, at an alkaline pH. Strains evolved in this
way did show a relative increase in glycerol yields, but there was a very limited impact on
ethanol yield. In addition, genetic analysis showed that the sulphite tolerance and glycerol
yield were not linked in the evolved strains.
Glycerol is the major osmolyte in yeasts [32]. Tilloy et al. [33] targeted glycerol
production to reduce the ethanol yield from a wine yeast strain. In this case, the selective
pressure was osmotic stress and the HOG intracellular signal transduction pathway. Strains
evolved in a hypersaline medium were selected and further improved by sporulation and
back-crossing. The resulting strains showed reduced ethanol yield, increased glycerol
yield, and did not produce acetic acid. However, the metabolic analysis revealed that,
contrary to expectations, the overproduction of glycerol was not due to the activation
of the HOG pathway. Compared to recombinant strategies which were also based on
glycerol overproduction, the authors found the evolved strains did not show the drawbacks
of GPD1-overexpressing strains. At the pilot scale, these strains produced wines with
0.4% (v/v) to 1.3% (v/v) less alcohol than the control (for Syrah and Grenache grapes,
respectively). Grenache wines were tasted, and no organoleptic defects were detected.
This strain has been commercialized.
amounts, for the same consumption of NADH. Thus, a reduction of 1% ethanol leads to
an increase in 15 g/L lactic acid. The increased lactate content of these wines contributed
to a higher total acidity. This feature could be useful to compensate for low acidity, which
is a problem affecting some grape varieties and growing regions, and is also related to
global warming. However, this also sets a limit to this approach, as excessive acidity
Biomolecules 2021, 11, 1569 7 of 16
would make the wines unpalatable.

Figure
Figure 1.
1. Diagram
Diagram of
of the
the alcoholic
alcoholic fermentation
fermentation pathway
pathway with
with indications
indications of
of the
the steps
steps affected
affected by
by
the
thedifferent
differentgenetic
geneticmodifications
modificationsassayed
assayedtoto
lower
lowerethanol yields.
ethanol References
yields. References in black: genetic
in black: en-
genetic
gineering.
engineering.References in yellow:
References adaptive
in yellow: evolution.
adaptive Metabolites
evolution. outside
Metabolites of the
outside of central ethanol
the central fer-
ethanol
mentation
fermentationpathway areare
pathway indicated in purple.
indicated in purple.

3. Alternative Wine Yeast Species


Microbial ecology studies of grape juice and wine fermentation clearly indicate that,
most often, S. cerevisiae is a minor species at the beginning of the fermentation process.
In contrast, it is almost invariably the dominant one at the end of spontaneous fermenta-
tions. Therefore, this species is considered particularly well adapted to the process, and has
been selected for use in inoculated fermentations. The other yeast species present during
the initial stages of fermentation were traditionally considered to be spoilage microorgan-
isms, and therefore undesirable. Over time, after inoculation with S. cerevisiae became
common practice, non-Saccharomyces wine yeast species have been found to harbour in-
teresting characteristics, such as their contribution to the aroma and complexity of the
wine, and their ability as microbiological control agents [34]. Several commercial strains
of non-Saccharomyces yeast are currently available for use in wine fermentation, including
Torulaspora delbrueckii, Lachancea thermotolerans and Metschnikowia pulcherrima, among the
most common yeast species [35]. In addition to their features related to the sensorial quality
of wines, one potentially interesting feature of some non-Saccharomyces species is their
lower ethanol yield compared to S. cerevisiae. However, only during the last decade has the
possibility of using alternative yeast species to reduce alcohol content been considered [36].
A summary of the results described below, using different non-Saccharomyces yeast species
to obtain lower alcohol wines, is shown in Table 1.
Biomolecules 2021, 11, 1569 8 of 16

Table 1. Fermentation conditions and the results obtained for alcohol level reduction using strains of different non-
Saccharomyces wine yeast species on natural grape must.

Time for S. Ethanol


Sugar Content Sensory
Single Species Conditions cerevisiae Reduction Reference
(g/L) Analysis c
Addition (% ABV)
Saccharomyces uvarum 171 a standard Negative No. Pure culture 0.5 [37]
240 b standard Negativec No. Pure culture 1.7 [38]
210–240 a standard No c at 50% sugars 0.8–0.9 [39]
Metschnikowia pulcherrima 230–240 a standard No c at 50% sugars 0.9–1.6 [40]
210–240 a standard No c at 50% sugars 1.0–1.1 [39]
240 b standard No different c 0 h (1/10) 1.0 [38]
230 a standard No c 3 days 0.8–1.25 [41]
264 standard No different c 3 days 1.0 [42]
20 VVH discont
260 a No 0h 3.7 [43]
48 h
212 a DO = 20% 3 days Negative c 72 h 3.6 [44]
220 a 3 VVH 72 h No c 72 h 1.5 [45]
Starmerella bacillaris 244 standard Different 24 h 0.5–0.6 [46]
(Candida zemplinina) 250 standard No 48 h 0.5 [47]
DO = 20% 2–3
212 a Negative c 69–52 h 0.6 [44]
days
Starmerella bombicola 218 1.2 VVH O2 72 h No c 72 h 1.5 [48]
Hanseniaspora uvarum 230–236 standard Different c 7 days 0.8–1.1 [49]
238 standard Positive c 48 h 0.9 [50]
Hanseniaspora opuntiae 230–236 standard Different c 7 days 0.6–1.3 [49]
Hanseniaspora vineae 252 a 20 VVH 24 h No c 24 h 2.5 [51]
Torulaspora delbrueckii 223 standard Positive c 4 days 0.5 [52]
195 a standard No different c No. Pure culture 0.5 [53]
220 a 1.5–3 VVH 72 h No c 72 h 0.9–1.0 [45]
Candida oleophila 206 a DO = 20% 5 days Negative c 120 h 1.4 [44]
Pichia kudriavzevii 230–236 standard Different c 7 days 0.4–0.6 [49]
Pichia guilliermondii 206 a DO = 20% 5 days Negative c 120 h 3.6 [44]
212 a DO = 20% 3 days Negative c 72 h 1.8 [44]
Pichia kluyveri 212 a DO = 20% 3 days Negative c 72 h 2.7 [44]
Lachancea thermotolerans 222 standard Higher acidity c 48 h 0.7 [54]
220 a standard Negative c 6 days 1.2 [55]
Meyerozyma guilliermondii 230 a standard No c 3 days 1.2 [41]
Zygosaccharomyces bailii 220 a 3 VVH 72 h No c 72 h 1.2 [45]
Two Species
M. pulcherrima/S. uvarum 210–240 a standard No c at 50% sugars 1.8 [39]
M. pulcherrima/S. bayanus 195 a standard Positive c 96 h 0.9 [53]
Immobilized
Hanseniaspora osmophila 202 standard No c 72 h 1.0 [56]
Hanseniaspora uvarum 202 standard No c 72 h 1.2 [56]
Starmerella bombicola 202 standard No c 72 h 1.6 [56]
Metschnikowia pulcherrima 202 standard No c 72 h 1.5 [56]
204 a 1.2 VVH 72 h No c 72 h 1.4 [57]
Schizosaccharomyces
240 a standard No c 0 h/48 h 1.7/2.4 [58]
japonicus
a Heat treated or filter sterilized; b DMDC treated; c Volatile Compound Analysis performed; DO: dissolved oxygen; VVH: volume per

volume per hour.

The cryophilic species Saccharomyces uvarum has been described as a low ethanol,
low acetate, and high glycerol producer yeast [37,59,60]. In experimental fermentations,
the ethanol contents of wines produced with two strains of S. uvarum were 0.5 and 0.3%
ABV lower than S. cerevisiae wine, and their hybrids showed ethanol contents which were
intermediate between the parent species [37]. An important reduction (1.7% ABV) was
obtained with a strain of S uvarum in Merlot wine, but it showed predominantly negative
attributes in a sensory analysis, with aromas of meat and barnyard [38]. Another cryophilic
species, Saccharomyces kudriavzevii, has been described as a high glycerol producer [61].
Although S. kudriavzevii is not found in winemaking, S. cerevisiae × S. kudriavzevii hybrids
are often isolated from cold-climate wine environments. However, the ethanol yield of
these hybrids has been found to be higher than that of S. cerevisiae [62].
Biomolecules 2021, 11, 1569 9 of 16

In 2011, Magyar and Toth [63] studied the ethanol yield of several strains of S. cerevisiae,
S. uvarum/bayanus, Candida zemplinina (currently Starmerella bacillaris) and Candida stellata.
All of the other species showed a lower ethanol yield than S. cerevisiae, in particular
C. zemplinina isolates. In addition, this species is fructophilic and shows a high glycerol yield.
A great number of strains of this species have been explored for ethanol production [64,65].
Wines made with the selected strains contained, at best, 0.6% ABV less than the control
wine [46,47].
Gobbi et al. (2014) [66] assessed the ethanol yield of 32 strains belonging to 8 dif-
ferent species, compared to a control strain of S. cerevisiae, and found that the species
Hanseniaspora uvarum showed the lowest yield, followed by Zygosaccharomyces sapae and
Zygosaccharomyces bisporus. The authors highlight the variability among strains of the
same species, and the correlation found between ethanol yield and secondary metabolite
production. Mestre et al. (2017) [67] also selected a strain of H. uvarum among 117 strains
belonging to 32 different species. Malbec wine produced with the selected strain [50] had a
0.9% drop in ethanol in compared to that made from S. cerevisiae, and a good qualification
in a sensory analysis.
Contreras et al. (2014) [40] studied the ethanol yield and sugar consumption of
50 yeast strains, with the aim of selecting suitable strains to lower the alcohol content of
wine. The authors highlighted the potential of strains of the species C. stellata, Schizosaccha-
romyces malidevorans and, mainly, M. pulcherrima. A selected strain of M. pulcherrima was
used for sequential fermentations with S. cerevisiae, and allowed for an ethanol reduction
of 1.6% (v/v) for Shiraz must and 0.9% (v/v) for Chardonnay must. An increase in ethyl
acetate and 2-methyl propanol was observed, which can negatively impact the wine quality.
More recently, they fermented Merlot must with M. pulcherrima in a sequential inocu-
lation with S. cerevisiae [38]. Compared to wine inoculated with S. cerevisiae, wine from
spontaneous fermentation contained 0.7% (v/v) less ethanol, and wine fermented by M. pul-
cherrima contained 1% (v/v) less. In the sensory analysis, spontaneously fermented and
M. pulcherrima wines were indistinguishable, and very similar to the control. Coinoculation
with M. pulcherrima and S. uvarum, in sequential inoculation with S. cerevisiae, allowed a
stronger reduction (1.7% v/v) than using these strains separately (0.8–1.0% v/v) [39].
Rossouw and Bauer (2016) [49] studied the ethanol yield of 91 yeast strains, including
some isolates of S. cerevisiae and two commercial strains as control strains. The authors
highlight the variability in the rate and extent of sugar utilisation between strains of the
same species, except for S. cerevisiae species, and the differences in aroma compounds
produced by strains affecting the wine aroma. Wines made with selected strains of Hanse-
niaspora opuntiae, Hanseniaspora uvarum, and Pichia kudriavzevii, in sequential inoculation
with S. cerevisiae, contained between 0.4 and 1.5% ABV less than the control wines.
There are currently several strains of T. delbrueckii for winemaking in the market which
have been selected for their contribution to the aromatic complexity of wine. Regarding
the ability to lower the alcohol content, there was a 0.5% v/v reduction in the ethanol
content compared to the S. cerevisiae control wine [52]. Puškaš et al. (2019) [53] studied
the effectiveness of one commercial strain of T. delbrueckii and one strain of M. pulcherrima,
in co-inoculation with S. cerevisiae and/or S. bayanus. In pasteurised must, a reduction
of 0.9% (v/v) in ethanol concentration was achieved with Metschnikowia and 0.5% with
Torulaspora, but this effect was lost in natural must, probably due to the low dominance of
the selected strains.
L. thermotolerans has also been introduced in the market because its positive effects
on wine aroma, total acidity, and volatile acidity [68]. Wines obtained with the sequential
inoculation of L. thermotolerans and S. cerevisiae contained 1.2% less ethanol than the control
wine, but suffered a decrement in aroma quality [55].
Some authors are also exploring yeast cell immobilization to improve effectiveness
and process control (Table 1). Canonico et al. (2016) [56] studied the use of immobilised
alternative yeasts, together with free S. cerevisiae in sequential inoculation. Using Starmerella
bombicola, H. uvarum, Hanseniaspora osmophila and M. pulcherrima, ethanol reduction values
Biomolecules 2021, 11, 1569 10 of 16

from 1.0 to 1.6% (v/v) were achieved. Changes in the volatile compounds were also
observed. It was predicted that these could potentially affect the sensory properties of
the wines, either positively or negatively, depending on the case, but the study did not
include a sensory evaluation. Finally, the authors indicated that this technology makes
the process more expensive. Domizio et al. (2018) [58] obtained a 1.7–2.4% reduction of
ethanol levels with immobilized cells of Schizosaccharomyces japonicus in co-inoculation and
sequential inoculation with free S. cerevisiae. The levels of acetoin, ethyl acetate, and acetate
esters were higher in both wines, and acetic acid and glycerol were higher in wine made by
sequential inoculation.
Several of the studies described above show some promise for non-Saccharomyces
yeast used in sequential or co-inoculation with S. cerevisiae for the reduction of ethanol
yields during wine fermentation, and hence the alcohol content of wines. However,
even under laboratory controlled experimental conditions, the reduction values were
generally moderate (Table 1), and difficulties associated to scale-up were often reported.
Fairly different reduction values are published by different laboratories, which is not
surprising because the authors did not follow standardized conditions (the optimization
of fermentation conditions is usually a secondary objective of their work), and different
strains were used in each case. According to the available information, when developing
protocols aiming at alcohol level reduction, attention should also be paid to the impact of
strains and fermentation conditions on the sensory properties of the wines. In addition,
the use of non-Saccharomyces species for alcohol reduction faces technological challenges
common to other oenological applications of these strains. These include difficulties for
the industrial production of most of them as active dry yeast, a generally low tolerance to
sulphiting agents, low imposition, or considerations about compatibility with S. cerevisiae
starters (which will almost always be necessary, in sequential or simultaneous inoculation
with non-Saccharomyces, to ensure that fermentation is completed).

4. Aerobic Fermentation
4.1. Non-Saccharomyces Species
Glycerol has been the target in many attempts to reduce ethanol yields by genetic
improvement (see above). However, reducing the alcohol content by 1% (v/v), by diverting
all excess carbon to glycerol, can result in more than 15 g/L extra glycerol in the wine.
This sets a technological limit to this strategy, as a further increase in the glycerol yield
would lead to values well above the commercially acceptable levels. However, almost
every alternative carbon sink can have a detrimental sensory impact if produced in excess,
even for compounds with positive sensory attributes.
In this context, some authors, including our research group [69,70], proposed CO2
generated by the aerobic respiration of yeast as a neutral carbon sink. By respiration,
all of the carbon present in sugar is converted to CO2 , in a process that requires molecular
oxygen. However, S. cerevisiae is a Crabtree-positive yeast. This implies that, except
under very small concentrations of glucose, it will mostly ferment sugars and produce
ethanol, even in the presence of oxygen. Therefore, the proposal was made to perform a
multi-starter fermentation with S. cerevisiae and alternative yeasts with higher respiration
capacities [70]. The appropriate yeast would be inoculated into the must, and during the
first hours, oxygen would be provided to the fermentation tank. After some time, the forced
aeration would be stopped, and the traditional process would follow. S. cerevisiae could be
inoculated from the beginning (co-inoculation), or when the aeration is stopped (sequential
inoculation). The more sugar is respired during the aeration process, the lower the ethanol
yield of the overall process will be.
The availability of suitable yeast strains was a prerequisite to reach this goal. To be
useful, these strains must be able to respire in the stressful conditions of grape must,
with no negative impact on the wine sensory properties. In a screening of 65 strains from
28 yeast species, Quirós et al. (2014) [71] found an interesting result for some S. cerevisiae
strains. Despite the Crabtree effect, the residual respiration activity of some of them was
Biomolecules 2021, 11, 1569 11 of 16

enough to allow for a substantial reduction in the ethanol yield according to the measured
respiration quotient (RQ) values. However, S. cerevisiae strains showed a boost in their
acetic acid production when provided with air. In contrast, several non-Saccharomyces
strains showed a combination of metabolic features suggesting that they might be suited
for use in the reduction of alcohol levels. Namely, some strains showed a moderate-to-high
sugar consumption rate, very low acetic acid production, and a low RQ (which means
a low expected ethanol yield). Indeed, one strain of M. pulcherrima was used in a co-
inoculation with S. cerevisiae in laboratory-scale bioreactors for the fermentation of natural
grape must (260 g/L sugar content). Several inoculation and aeration rates were assayed.
The alcohol level reduction reached 3.6% (v/v) with a somehow too-high volatile acidity,
or 2.2% (v/v) with acceptable volatile acidity levels [43]. Some environmental factors that
can be controlled during wine fermentation—such as nutrients, temperature, and dissolved
oxygen levels—will certainly affect both the ethanol and acetic acid yield by different
yeast strains. Interestingly, strains of M. pulcherrima and Candida sake show low acetic acid
production almost independently of the fermentation conditions [72].
Fermentation conditions leading to lower ethanol yields based on the use of non-
Saccharomyces strains under aerobic conditions have been studied by these and other authors,
using T. delbrueckii, M. pulcherrima, Zygosaccharomyces bailii, C. zemplinina, M. pulcherrima,
Pichia guilliermondii, Pichia kluyveri, Candida oleophila, Starmerella bombicola, or Hanseniaspora
vineae [44,45,48,51,57,73,74]. The authors have worked with different aeration times and air
flows, free or immobilized cells, and different varieties of white grapes (Table 1). As a trend,
increased aeration leads to better results in terms of reduced alcohol levels, but also to a
greater impact on the volatile compounds (including volatile acidity) and a lower rating
from sensory panels. It can be concluded that the results are promising, but the process
needs to be further developed, including the criteria for the selection of non-Saccharomyces
strains for this application, in order to develop useful protocols for the industry.

4.2. S. cerevisiae Strains Improved for Aerobic Fermentation


As mentioned above, one major drawback for the use of S. cerevisiae under the aerobic
growth conditions required for respiration was that strains of this species produced un-
acceptable acetic acid levels under these conditions. In order to overcome this problem,
the researchers turned to other wine yeast species. Although this strategy is yielding some
promising results (see the previous section), and though the use of non-Saccharomyces
species could bring some other benefits to the wine, it also hinders the fermentation man-
agement in the winery. The use of a single yeast starter that will at the same time be able
to reduce alcohol levels, and drive fermentation to the end, while not increasing volatile
acidity, would be an optimal alternative.
A wine strain of S. cerevisiae improved for lower acetic acid production was expected
to be able to meet all of these requirements. Indeed, recombinant wine yeast strains partly
defective in carbon catabolite repression (CCR) due to REG1 gene knockout were shown to
produce reduced levels of acetic acid under anaerobic conditions [75]. Interestingly, reg1
mutant strains can grow on non-preferred carbon sources (glycerol or galactose, for ex-
ample) in the presence of the non-metabolizable glucose analogue 2-deoxyglucose (2DG).
In contrast, strains with an intact CCR will be inhibited under these growth conditions.
Although this link between CCR and acetic acid production under aerobic conditions is
not yet clarified, it is possible to take a practical advantage of it to design experimental
evolution strategies to develop S. cerevisiae wine yeast strains which are better suited for
alcohol level reduction. This was recently performed by Guindal et al. [76] using galactose
as a carbon source in the presence of growing amounts of 2DG for around 100 genera-
tions. Several evolved strains from different wine yeast genetic backgrounds were indeed
derepressed for glucose, and had been indirectly selected for lower acetic acid produc-
tion under aerobic conditions. As a trend, the glycerol yields were also increased for the
selected evolved strains. However, some strains had to be discarded due to impaired
fermentation kinetics. Indeed, there is always a risk of unwanted change during adaptive
Biomolecules 2021, 11, 1569 12 of 16

laboratory evolution. This must be managed by the careful design of experimental condi-
tions and a thorough characterization of the derived strains before they can by promoted
to industrial applications.
On the other side, in the framework of the European project CoolWine (ERA Co-
BioTech), researchers are working on the development of rational strategies for adaptive
laboratory evolution aiming at alcohol level reduction. For S. cerevisiae, the aim is to develop
new strategies to reduce the above-mentioned problem of excess acetic acid production in
the presence of oxygen, while for non-Saccharomyces strains, the aim is to improve their
survival and competitiveness in the context of winemaking.
Although the general trend for S. cerevisiae strains is the production of high volatile
acidity under aerated wine fermentation conditions, we recently found rare isolates showing
acceptable acetic acid yields under aerobic conditions [77]. These strains show promise
to be used for alcohol level reduction by sugar respiration, without resourcing to non-
Saccharomyces yeast or the requirement for genetic improvement. However, genetic improve-
ment will still be necessary to ensure enough genetic diversity for industrial applications
(a single strain cannot fulfil all industry requirements), and to take advantage of the features
of commercial strains that have previously shown optimal features for winemaking.

5. Enzymatic Treatment of Grape Must


One conceptually simple approach for the reduction of the ethanol content of wines
is to remove excess sugar from the grape juice before fermentation. Biotechnologically,
this goal can be achieved using enzymes, and the enzymatic activity proposed for this
application is glucose oxidase. This enzyme catalyses the oxidation of glucose to gluconic
acid. The reaction involves molecular oxygen as a co-substrate, and releases hydrogen
peroxide as a by-product.
The use of glucose oxidase during the pre-fermentation stages to produce low-alcohol
wines was first proposed by Villetaz (1986) [78], and was optimised by this and other
authors (as reviewed in [79]). However, the use of glucose oxidase is not among the
procedures included in the International Code of Oenological Practices [80].
The procedure proposed by Pickering et al. [81] involved an initial grape must deacid-
ification to accommodate the pH to that which is optimal for the enzyme used, as well as
aeration during the enzymatic treatment. Catalase was added with glucose oxidase to re-
move the hydrogen peroxide formed in the reaction. The removal of the hydrogen peroxide
improves the yeast survival and the activity of glucose oxidase enzyme itself. By treating
Riesling grape juice for 10 h under these conditions, these authors produced wines with an
alcohol content reduced by 3.7% (v/v) [82]. However, gluconic acid negatively affected the
taste of the wine, fruity aromas were perceived with less intensity, and their persistence
was also reduced [83]. A second deacidification or a sweetening were proposed to alleviate
the effect of gluconic acid [81].
Röcker et al. [84] assayed several methods of deacidification following glucose oxidase
treatment. However, the wines were more acidic than the control, and lost part of their
fruitiness. Other authors have used encapsulated enzymes in order to improve glucose
oxidase activity without prior pH correction [85,86]. Wines obtained from juices treated
for two days under these conditions showed a 0.68% (v/v) reduction in their ethanol
content [85].

6. Metabolic Inhibitors
Depending on their mode of action, sublethal concentrations of some metabolic
inhibitors might be expected to alter the ethanol yield of S. cerevisiae. This is the case of
furfural, an aromatic aldehyde resulting from sugar dehydration. It can be found in wines
at concentrations below the perception limit (20 mg/L). Furfural is toxic for S. cerevisiae,
and the detoxification mechanism involves its transformation to furfuryl alcohol by alcohol
dehydrogenase, the same enzyme catalysing the final step of alcoholic fermentation from
acetaldehyde to ethanol. The use of furfural has been proposed to reduce the ethanol yield
Biomolecules 2021, 11, 1569 13 of 16

during wine fermentation [87]. However, the suggested dose, 50 mg/L, is clearly above
the perception limit, for a modest reduction in ethanol content of 0.6% (v/v). The same
research group suggested other metabolic inhibitors to be used for the same purpose [88],
but no experimental results have been published yet.

7. Conclusions
The metabolic versatility of oenological yeasts, considering both non-Saccharomyces
and Saccharomyces species, and our growing knowledge about it, can be exploited to
address the problem of the rising alcohol levels observed over the last few decades in
wine. Research groups have approached the problem from various perspectives: the
exploration of natural genetic diversity, focusing mainly on non-Saccharomyces yeasts;
the genetic improvement of S. cerevisiae by genetic engineering, random mutagenesis,
or experimental evolution; or the modification of the environmental conditions (metabolic
inhibitors, aerobiosis). A combination of aerobic conditions with yeast genetic improvement
(of both Saccharomyces and non-Saccharomyces yeast strains) is currently one of the most
promising options. However, considering the metabolic diversity of wine yeast species and
the impact of oxygen on yeast metabolism, any technological improvement developed in
the laboratories must be validated by the sensory analysis of the wines produced at the
pilot or industrial scale.

Author Contributions: P.M. and R.G. conceived the idea, wrote the first draft, and supervised the
writing. A.M.G. and J.T. contributed to the writing of several sections. All authors have read and
agreed to the published version of the manuscript.
Funding: The work of the Microwine group is currently funded by the Spanish Government through
grants PID2019-105159RB-I00 and PCI2018-092949 (co-funded by ERA-CoBioTech). JT is funded by
FGCSIC by the COMFUTURO program. AMG’s predoctoral contract is funded by Consejería de
Desarrollo Económico e Innovación de la Comunidad Autónoma de La Rioja.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Mira de Orduña, R. Climate change associated effects on grape and wine quality and production. Food Res. Int. 2010, 43,
1844–1855. [CrossRef]
2. Goold, H.D.; Kroukamp, H.; Williams, T.C.; Paulsen, I.T.; Varela, C.; Pretorius, I.S. Yeast’s balancing act between ethanol and
glycerol production in low-alcohol wines. Microb. Biotechnol. 2017, 10, 264–278. [CrossRef]
3. Palacios, A.; Raginel, F.; Ortiz-Julien, A. Can the selection Saccharomyces cerevisiae yeast lead to variations in the final alcohol
degree of wines? Aust. N. Zeal. Grapegrow. Winemak. 2007, 527, 71–75.
4. Camarasa, C.; Sanchez, I.; Brial, P.; Bigey, F.; Dequin, S. Phenotypic landscape of Saccharomyces cerevisiae during wine fermentation:
Evidence for origin-dependent metabolic traits. PLoS ONE 2011, 6, e25147. [CrossRef]
5. Cebollero, E.; González-Ramos, D.; Tabera, L.; Gonzalez, R. Transgenic wine yeast technology comes of age: Is it time for
transgenic wine? Biotechnol. Lett. 2007, 29, 191–200. [CrossRef]
6. Gonzalez, R.; Tronchoni, J.; Quirós, M.; Morales, P. Genetic improvement and genetically modified microorganisms. In Wine Safety,
Consumer Preference, and Human Health; Moreno-Arribas, M.V., Bartolomé Sualdea, B., Eds.; Springer International Publishing:
Cham, Switzerland, 2016; Chapter 4; pp. 71–96. ISBN1 978-3-319-24512-6. ISBN2 978-3-319-24514-0. [CrossRef]
7. Drewke, C.; Thielen, J.; Ciriacy, M. Ethanol formation in adh0 mutants reveals the existence of a novel acetaldehyde-reducing
activity in Saccharomyces cerevisiae. J. Bacteriol. 1990, 172, 3909–3917. [CrossRef] [PubMed]
8. Dequin, S.; Barre, P. Mixed lactic acid-alcoholic fermentation by Saccharomyces cerevisiae expressing the Lactobacillus casei
L(+)-LDH. Nat. Biotechnol. 1994, 12, 173–177. [CrossRef] [PubMed]
9. Michnick, S.; Roustan, J.L.; Remize, F.; Barre, P.; Dequin, S. Modulation of glycerol and ethanol yields during alcoholic fermentation
in Saccharomyces cerevisiae strains overexpressed or disrupted for GPD1 encoding glycerol 3-phosphate dehydrogenase. Yeast
1997, 13, 783–793. [CrossRef]
Biomolecules 2021, 11, 1569 14 of 16

10. Remize, F.; Andrieu, E.; Dequin, S. Engineering of the pyruvate dehydrogenase bypass in Saccharomyces cerevisiae: Role of the
cytosolic Mg2+ and mitochondrial K+ acetaldehyde dehydrogenases Ald6p and Ald4p in acetate formation during alcoholic
fermentation. Appl. Environ. Microbiol. 2000, 66, 3151–3159. [CrossRef] [PubMed]
11. Cambon, B.; Monteil, V.; Remize, F.; Camarasa, C.; Dequin, S. Effects of GPD1 overexpression in Saccharomyces cerevisiae
commercial wine yeast strains lacking ALD6 genes. Appl. Environ. Microbiol. 2006, 72, 4688–4694. [CrossRef] [PubMed]
12. Ehsani, M.; Fernández, M.R.; Biosca, J.A.; Julien, A.; Dequin, S. Engineering of 2,3-butanediol dehydrogenase to reduce acetoin
formation by glycerol-overproducing, low-alcohol Saccharomyces cerevisiae. Appl. Environ. Microbiol. 2009, 75, 3196–3205.
[CrossRef]
13. Heux, S.; Sablayrolles, J.-M.; Cachon, R.; Dequin, S. Engineering a Saccharomyces cerevisiae wine yeast that exhibits reduced ethanol
production during fermentation under controlled microoxygenation conditions. Appl. Environ. Microbiol. 2006, 72, 5822–5828.
[CrossRef] [PubMed]
14. Flikweert, M.T.; van der Zanden, L.; Janssen, W.M.T.M.; Steensma, H.Y.; van Dijken, J.P.; Pronk, J.T. Pyruvate decarboxylase:
An indispensable enzyme for growth of Saccharomyces cerevisiae on glucose. Yeast 1996, 12, 247–257. [CrossRef]
15. Nevoigt, E.; Stahl, U. Reduced pyruvate decarboxylase and increased glycerol-3-phosphate dehydrogenase [NAD+] levels
enhance glycerol production in Saccharomyces cerevisiae. Yeast 1996, 12, 1331–1337. [CrossRef]
16. Cuello, R.A.; Montero, K.J.F.; Mercado, L.A.; Combina, M.; Ciklic, I.F. Construction of low-ethanol–wine yeasts through partial
deletion of the Saccharomyces cerevisiae PDC2 gene. AMB Expr. 2017, 7, 67. [CrossRef]
17. Dai, Z.; Huang, M.; Chen, Y.; Siewers, V.; Nielsen, J. Global rewiring of cellular metabolism renders Saccharomyces cerevisiae
Crabtree negative. Nat. Commun. 2018, 9, 3059. [CrossRef]
18. Otterstedt, K.; Larsson, C.; Bill, R.M.; Stahlberg, A.; Boles, E.; Hohmann, S.; Gustafsson, L. Switching the mode of metabolism in
the yeast Saccharomyces cerevisiae. EMBO Rep. 2004, 5, 532–537. [CrossRef]
19. Compagno, C.; Brambilla, L.; Capitanio, D.; Boschi, F.; Ranzi, B.M.; Porro, D. Alterations of the glucose metabolism in a triose
phosphate isomerase-negative Saccharomyces cerevisiae mutant. Yeast 2001, 18, 663–670. [CrossRef] [PubMed]
20. Cordier, H.; Mendes, F.; Vasconcelos, I.; Francois, J.M. A metabolic and genomic study of engineered Saccharomyces cerevisiae
strains for high glycerol production. Metab. Eng. 2007, 9, 364–378. [CrossRef] [PubMed]
21. Rossouw, D.; Heyns, E.H.; Setati, M.E.; Bosch, S.; Bauer, F.F. Adjustment of trehalose metabolism in wine Saccharomyces cerevisiae
strains to modify ethanol yields. Appl. Environ. Microbiol. 2013, 79, 5197–5207. [CrossRef] [PubMed]
22. Du, Q.; Liu, Y.; Song, Y.; Qin, Y. Creation of a low-alcohol-production yeast by a mutated SPT15 transcription regulator triggers
transcriptional and metabolic changes during wine fermentation. Front. Microbiol. 2020, 11, 597828. [CrossRef]
23. Malherbe, D.F.; du Toit, M.; Cordero Otero, R.R.; van Rensburg, P.; Pretorius, I.S. Expression of the Aspergillus niger glucose
oxidase gene in Saccharomyces cerevisiae and its potential applications in wine production. Appl. Microbiol. Biotechnol. 2003, 61,
502–511. [CrossRef] [PubMed]
24. Mans, R.; van Rossum, H.M.; Wijsman, M.; Backx, A.; Kuijpers, N.G.; van den Broek, M.; Daran-Lapujade, P.; Pronk, J.; van
Maris, A.J.A.; Daran, J.M.G. CRISPR/Cas9: A molecular Swiss army knife for simultaneous introduction of multiple genetic
modifications in Saccharomyces cerevisiae. FEMS Yeast Res. 2015, 15, fov004. [CrossRef] [PubMed]
25. Vigentini, I.; Gebbia, M.; Belotti, A.; Foschino, R.; Roth, F.P. CRISPR/Cas9 system as a valuable genome editing tool for wine
yeasts with application to decrease urea production. Front. Microbiol. 2017, 8, 1–11. [CrossRef] [PubMed]
26. Van Wyk, N.; Kroukamp, H.; Espinosa, M.I.; von Wallbrunn, C.; Wendland, J.; Pretorius, I.S. Blending wine yeast phenotypes
with the aid of CRISPR DNA editing technologies. Int. J. Food Microbiol. 2020, 324, 108615. [CrossRef] [PubMed]
27. Gonzalez, R.; Morales, P. Truth in wine yeast. Microb. Biotechnol. 2021, 13848. [CrossRef]
28. Cadière, A.; Ortiz-Julien, A.; Camarasa, C.; Dequin, S. Evolutionary engineered Saccharomyces cerevisiae wine yeast strains with
increased in vivo flux through the pentose phosphate pathway. Metab. Eng. 2011, 13, 263–271. [CrossRef]
29. Cadière, A.; Aguera, E.; Caillé, S.; Ortiz-Julien, A.; Dequin, S. Pilot-scale evaluation the enological traits of a novel, aromatic wine
yeast strain obtained by adaptive evolution. Food Microbiol. 2012, 32, 332–337. [CrossRef]
30. Taherzadeh, M.J.; Adler, L.; Liden, G. Strategies for enhancing fermentative production of glycerol—A review. Enzym. Microb.
Technol. 2002, 31, 53–66. [CrossRef]
31. Kutyna, D.R.; Varela, C.; Stanley, G.A.; Borneman, A.R.; Henschke, P.A.; Chambers, P.J. Adaptive evolution of Saccharomyces
cerevisiae to generate strains with enhanced glycerol production. Appl. Microbiol. Biotechnol. 2012, 93, 1175–1184. [CrossRef]
32. Blomberg, A.; Adler, L. Physiology of osmotolerance in fungi. Adv. Microb. Physiol. 1992, 33, 145–212. [CrossRef] [PubMed]
33. Tilloy, V.; Ortiz-Julien, A.; Dequin, S. Reduction of ethanol yield and improvement of glycerol formation by adaptive evolution of
the wine yeast Saccharomyces cerevisiae under hyperosmotic conditions. Appl. Environ. Microbiol. 2014, 80, 2623–2632. [CrossRef]
[PubMed]
34. Mas, A.; Guillamón, J.M.; Beltran, G. Editorial: Non-conventional yeast in the wine industry. Front. Microbiol. 2016, 7, 1494.
[CrossRef] [PubMed]
35. Vejarano, R.; Gil-Calderón, A. Commercially available non-Saccharomyces yeasts for winemaking: Current market, advantages
over Saccharomyces, biocompatibility, and safety. Fermentation 2021, 7, 171. [CrossRef]
36. Ciani, M.; Morales, P.; Comitini, F.; Tronchoni, J.; Canonico, L.; Curiel, J.A.; Oro, L.; Rodrigues, A.J.; Gonzalez, R. Non-conventional
yeast species for lowering ethanol content of wines. Front. Microbiol. 2016, 7, 642. [CrossRef]
Biomolecules 2021, 11, 1569 15 of 16

37. Coloretti, F.; Zambonelli, C.; Tini, V. Characterization of flocculent Saccharomyces interspecific hybrids for the production of
sparkling wines. Food Microbiol. 2006, 23, 672–676. [CrossRef]
38. Varela, C.; Barker, A.; Tran, T.; Borneman, A.; Curtin, C. Sensory profile and volatile aroma composition of reduced alcohol Merlot
wines fermented with Metschnikowia pulcherrima and Saccharomyces Uvarum. Int. J. Food Microbiol. 2017, 252, 1–9. [CrossRef]
39. Varela, C.; Sengler, F.; Solomon, M.; Curtin, C. Volatile flavour profile of reduced alcohol wines fermented with the non-
conventional yeast species Metschnikowia pulcherrima and Saccharomyces Uvarum. Food Chem. 2016, 209, 57–64. [CrossRef]
[PubMed]
40. Contreras, A.; Hidalgo, C.; Schmidt, S.; Henschke, P.A.; Curtin, C.; Varela, C. Evaluation of non-Saccharomyces yeasts for the
reduction of alcohol content in wine. Appl. Environ. Microbiol. 2014, 80, 1670–1678. [CrossRef]
41. García, M.; Esteve-Zarzoso, B.; Cabellos, J.M.; Arroyo, T. Sequential non-Saccharomyces and Saccharomyces cerevisiae fermentations
to reduce the alcohol content in wine. Fermentation 2020, 6, 60. [CrossRef]
42. Aplin, J.J.; Paup, V.D.; Ross, C.F.; Edwards, C.G. Chemical and sensory profiles of Merlot wines produced by sequential inoculation
of Metschnikowia pulcherrima or Meyerozyma guilliermondii. Fermentation 2021, 7, 126. [CrossRef]
43. Morales, P.; Rojas, V.; Quirós, M.; Gonzalez, R. The impact of oxygen on the final alcohol content of wine fermented by a mixed
starter culture. Appl. Microbiol. Biotechnol. 2015, 99, 3993–4003. [CrossRef]
44. Röcker, J.; Strub, S.; Ebert, K.; Grossmann, M. Usage of different aerobic non-Saccharomyces yeasts and experimental conditions as
a tool for reducing the potential ethanol content in wines. Eur. Food Res. Technol. 2016, 242, 2051–2070. [CrossRef]
45. Canonico, L.; Solomon, M.; Comitini, F.; Ciani, M.; Varela, C. Volatile profile of reduced alcohol wines fermented with selected
non-Saccharomyces yeasts under different aeration conditions. Food Microbiol. 2019, 84, 103247. [CrossRef]
46. Giaramida, P.; Ponticello, G.; Di Maio, S.; Squadrito, M.; Genna, G.; Barone, E.; Scacco, A.; Amore, G.; di Stefano, R.; Oliva, D.
Candida zemplinina for production of wines with less alcohol and more glycerol. S. Afr. J. Enol. Vitic. 2013, 34, 204–211. [CrossRef]
47. Englezos, V.; Rantsiou, K.; Cravero, F.; Torchio, F.; Ortiz-Julien, A.; Gerbi, V.; Rolle, L.; Cocolin, L. Starmerella bacillaris and
Saccharomyces cerevisiae mixed fermentations to reduce ethanol content in wine. Appl. Microbiol. Biotechnol. 2016, 100, 5515–5526.
[CrossRef]
48. Canonico, L.; Galli, E.; Agarbati, A.; Comitini, F.; Ciani, M. Starmerella bombicola and Saccharomyces cerevisiae in wine sequential
fermentation in aeration condition: Evaluation of ethanol reduction and analytical profile. Foods 2021, 10, 1047. [CrossRef]
49. Rossouw, D.; Bauer, F.F. Exploring the phenotypic space of non-Saccharomyces wine yeast biodiversity. Food Microbiol. 2016, 55,
32–46. [CrossRef]
50. Mestre, M.V.; Maturano, Y.P.; Gallardo, C.; Combina, M.; Mercado, L.; Toro, M.E.; Carrau, F.; Vazquez, F.; Dellacassa, E. Impact
on sensory and aromatic profile of low ethanol Malbec wines fermented by sequential culture of Hanseniaspora uvarum and
Saccharomyces cerevisiae native yeasts. Fermentation 2019, 5, 65. [CrossRef]
51. Yan, G.; Zhang, B.; Joseph, L.; Waterhouse, A.L. Effects of initial oxygenation on chemical and aromatic composition of wine in
mixed starters of Hanseniaspora vineae and Saccharomyces cerevisiae. Food Microbiol. 2020, 90, 103460. [CrossRef] [PubMed]
52. Belda, I.; Ruiz, J.; Beisert, B.; Navascués, E.; Marquina, D.; Calderón, F.; Rauhut, D.; Benito, S.; Santos, A. Influence of Torulaspora
delbrueckii in varietal thiol (3-SH and 4-MSP) release in wine sequential fermentations. (2018). Int. J. Food Microbiol. 2017, 257,
183–191. [CrossRef] [PubMed]
53. Puškaš, V.S.; Miljić, U.D.; Djuran, J.J.; Vučurović, M.M. The aptitude of commercial yeast strains for lowering the ethanol content
of wine. Food Sci. Nutr. 2019, 8, 1489–1498. [CrossRef] [PubMed]
54. Gobbi, M.; Comitini, F.; Domizio, P.; Romani, C.; Lencioni, L.; Mannazzu, I.; Ciani, M. Lachancea thermotolerans and Saccharomyces
cerevisiae in simultaneous and sequential co-fermentation: A strategy to enhance acidity and improve the overall quality of wine.
Food Microbiol. 2013, 33, 271–281. [CrossRef]
55. Del Fresno, J.M.; Morata, A.; Loira, I.; Bañuelos, M.A.; Escott, C.; ·Benito, S.; González Chamorro, C.; Suárez-Lepe, J.A. Use of
non-Saccharomyces in single-culture, mixed and sequential fermentation to improve red wine quality. Eur. Food Res. Technol. 2017,
243, 2175–2185. [CrossRef]
56. Canonico, L.; Comitini, F.; Oro, L.; Ciani, M. Sequential fermentation with selected immobilized non-Saccharomyces yeast for
reduction of ethanol content in wine. Front. Microbiol. 2016, 7, 278. [CrossRef]
57. Canonico, L.; Comitini, F.; Ciani, M. Metschnikowia pulcherrima selected strain for ethanol reduction in wine: Influence of cell
immobilization and aeration condition. Foods 2019, 8, 378. [CrossRef]
58. Domizio, P.; Lencioni, L.; Calamai, L.; Portaro, L.; Bisson, L.F. Evaluation of the yeast Schizosaccharomyces japonicus for use in wine
production. Am. J. Enol. Vitic. 2018, 69, 266–277. [CrossRef]
59. Castellari, L.; Ferruzzi, M.; Magrini, A.; Giudici, P.; Passarelli, P.; Zambonelli, C. Unbalanced wine fermentation by cryotolerant
vs. Non cryotolerant Saccharomyces strains. Vitis 1994, 33, 49–52.
60. Giudici, P.; Zambonelli, C.; Passarelli, P.; Castellari, L. Improvement of wine composition with cryotolerant Saccharomyces strains.
Am. J. Enol. Vitic. 1995, 46, 143–147.
61. Arroyo-Lopez, F.N.; Perez-Torrado, R.; Querol, A.; Barrio, E. Modulation of the glycerol and ethanol syntheses in the yeast
Saccharomyces kudriavzevii differs from that exhibited by Saccharomyces cerevisiae and their hybrid. Food Microbiol. 2010, 27, 628–637.
[CrossRef]
62. Gangl, H.; Batusic, M.; Tscheik, G.; Tiefenbrunner, W.; Hack, C.; Lopandic, K. Exceptional fermentation characteristics of natural
hybrids from Saccharomyces cerevisiae and S. kudriavzevii. New Biotechnol. 2009, 25, 244–251. [CrossRef]
Biomolecules 2021, 11, 1569 16 of 16

63. Magyar, I.; Toth, T. Comparative evaluation of some oenological properties in wine strains of Candida stellata, Candida zemplinina,
Saccharomyces uvarum and Saccharomyces cerevisiae. Food Microbiol. 2011, 28, 94–100. [CrossRef] [PubMed]
64. Di Maio, S.; Genna, G.; Gandolfo, V.; Amore, G.; Ciaccio, M.; Oliva, D. Presence of Candida zemplinina in Sicilian musts and
selection of a strain for wine mixed fermentations. S. Afr. J. Enol. Vitic. 2012, 33, 80–87. [CrossRef]
65. Englezos, V.; Rantsiou, K.; Torchio, F.; Rolle, L.; Gerbi, V.; Cocolin, L. Exploitation of the non-Saccharomyces yeast Starmerella
bacillaris (synonym Candida zemplinina) in wine fermentation: Physiological and molecular characterizations. Int. J. Food Microbiol.
2015, 199, 33–40. [CrossRef] [PubMed]
66. Gobbi, M.; De Vero, L.; Solieri, L.; Comitini, F.; Oro, L.; Giudici, P.; Ciani, M. Fermentative aptitude of non-Saccharomyces wine
yeast for reduction in the ethanol content in wine. Eur. Food Res. Technol. 2014, 239, 41–48. [CrossRef]
67. Mestre, M.V.; Maturano, Y.P.; Combina, M.; Mercado, L.A.; Toro, M.E.; Vazquez, F. Selection of non-Saccharomyces yeasts to be
used in grape musts with high alcoholic potential: A strategy to obtain wines with reduced ethanol content. FEMS Yeast Res.
2017, 17, fox010. [CrossRef]
68. Morata, A.; Loira, I.; Tesfaye, W.; Bañuelos, M.A.; González, C.; Suárez Lepe, J.A. Lachancea thermotolerans applications in wine
technology. Fermentation 2018, 4, 53. [CrossRef]
69. Erten, H.; Campbell, I. The production of low-alcohol wines by aerobic yeasts. J. Inst. Brew. 2001, 107, 207–215. [CrossRef]
70. Gonzalez, R.; Quirós, M.; Morales, P. Yeast respiration of sugars by non-Saccharomyces yeast species: A promising and barely
explored approach to lowering alcohol content of wines. Trends Food Sci. Technol. 2013, 29, 55–61. [CrossRef]
71. Quirós, M.; Rojas, V.; Gonzalez, R.; Morales, P. Selection of non-Saccharomyces yeast strains for reducing alcohol levels in wine by
sugar respiration. Int. J. Food Microbiol. 2014, 181, 85–91. [CrossRef]
72. Rodrigues, A.J.; Raimbaud, T.; Gonzalez, R.; Morales, P. Environmental factors influencing the efficacy of different yeast strains
for alcohol level reduction in wine by respiration. LWT Food Sci. Technol. 2016, 65, 1038–1043. [CrossRef]
73. Tronchoni, J.; Curiel, J.A.; Sáenz-Navajas, M.P.; Morales, P.; de-la-Fuente-Blanco, A.; Fernández-Zurbano, P.; Ferreira, V.;
Gonzalez, R. Aroma profiling of an aerated fermentation of natural grape must with selected yeast strains at pilot scale. Food
Microbiol. 2018, 70, 14–223. [CrossRef]
74. Contreras, A.; Hidalgo, C.; Schmidt, S.; Henschke, P.A.; Curtin, C.; Varela, C. The application of non-Saccharomyces yeast in
fermentations with limited aeration as a strategy for the production of wine with reduced alcohol content. Int. J. Food Microbiol.
2015, 205, 7–15. [CrossRef]
75. Curiel, J.A.; Salvadó, Z.; Tronchoni, J.; Morales, P.; Rodrigues, A.J.; Quirós, M.; Gonzalez, R. Identification of target genes to
control acetate yield during aerobic fermentation with Saccharomyces cerevisiae. Microb. Cell Fact. 2016, 15, 156. [CrossRef]
[PubMed]
76. Guindal, A.M.; Morales, P.; Gonzalez, R.; Tronchoni, J. Adaptive Laboratory Evolution to reduce acetic acid yield of Saccharomyces
cerevisiae wine yeast strains under aerobic conditions. In Proceedings of the 30th International Conference on Yeast Genetics and
Molecular Biology (Virtual Conference), Vienna, Austria, 23–27 August 2021; Abstract B40. p. 317.
77. Tronchoni, J.; Gonzalez, R.; Guindal, A.M.; Calleja, E.; Morales, P. Exploring the suitability of Saccharomyces cerevisiae strains for
winemaking under aerobic conditions. Food Microbiol. 2022, 101, 103893. [CrossRef] [PubMed]
78. Villettaz, J.-C. Method for Production of a Low Alcoholic Wine and Agent for Performance of the Method. European Patent No.
EP0194043A1, 10 September 1986.
79. Pickering, G.J. Low- and Reduced-alcohol Wine: A Review. J. Wine Res. 2000, 11, 129–144. [CrossRef]
80. International Code of Oenological Practices. Available online: https://www.oiv.int/en/technical-standards-and-documents/
oenological-practices/international-code-of-oenological-practices (accessed on 24 September 2021).
81. Pickering, G.J.; Heatherbell, D.A.; Barnes, M.F. Optimising glucose conversion in the production of reduced alcohol wines from
glucose oxidase treated musts. Food Res. Int. 1999, 31, 685–692. [CrossRef]
82. Pickering, G.J.; Heatherbell, D.A.; Barnes, M.F. The production of reduced-alcohol wine using glucose oxidase treated juice. Part I.
Composition. Am. J. Enol. Vitic. 1999, 50, 291–298.
83. Pickering, G.J.; Heatherbell, D.A.; Barnes, M.F. The production of reduced-alcohol wine using glucose oxidase-treated juice.
Part III. Sensory. Am. J. Enol. Vitic. 1999, 50, 307–316.
84. Röcker, J.; Schmitt, M.; Pasch, L.; Ebert, K.; Grossmann, M. The use of glucose oxidase and catalase for the enzymatic reduction of
the potential ethanol content in wine. Food Chem. 2016, 210, 660–670. [CrossRef]
85. Biyela, B.N.E.; du Toit, W.J.; Divol, B.; Malherbe, D.F.; Van Rensburg, P. The production of reduced-alcohol wines using Gluzyme
Mono 10.000 BG-treated grape juice. S. Afr. J. Enol. Vitic. 2009, 30, 124–132. [CrossRef]
86. Ruiz, E.; Busto, M.D.; Ramos-Gómez, S.; Palacios, D.; Pilar-Izquierdo, M.C.; Ortega, N. Encapsulation of glucose oxidase in
alginate hollow beads to reduce the fermentable sugars in simulated musts. Food Biosci. 2018, 24, 67–72. [CrossRef]
87. Abalos, D.; Vejarano, R.; Morata, A.; González, C.; Suárez-Lepe, J.A. The use of furfural as a metabolic inhibitor for reducing the
alcohol content of model wines. Eur. Food Res. Technol. 2001, 232, 663–669. [CrossRef]
88. Vejarano, R.; Morata, A.; Loira, I.; González, M.C.; Suárez-Lepe, J.A. Theoretical considerations about usage of metabolic inhibitors
as possible alternative to reduce alcohol content of wines from hot areas. Eur. Food Res. Technol. 2013, 237, 281–290. [CrossRef]
| |
Received: 17 July 2019    Revised: 18 December 2019    Accepted: 27 December 2019

DOI: 10.1002/fsn3.1433

ORIGINAL RESEARCH

The aptitude of commercial yeast strains for lowering the


ethanol content of wine

Vladimir S. Puškaš | Uroš D. Miljić  | Jovana J. Djuran | Vesna M. Vučurović

Faculty of Technology Novi Sad, University


of Novi Sad, Novi Sad, Serbia Abstract
The high alcohol content in wine usually has a negative impact on its sensory prop-
Correspondence
Uroš D. Miljić, Faculty of Technology Novi erties, but can also affect the general health of the consumers. The possibility to
Sad, University of Novi Sad, Blvd. cara reduce ethanol production in wines during fermentation involves the use of different
Lazara 1, Novi Sad 21000, Serbia.
Email: urosmiljic@yahoo.com yeast strains characterized by the increased production of fermentation by-products
(glycerol, 2,3-butanediol, etc.) from the available sugar. The activity of these strains
Funding information
Ministry of Education, Science and should not impair the sensory properties of the wine. In general, the use of geneti-
Technological Development of the Republic cally and evolutionarily (non-GM) engineered Saccharomyces cerevisiae strains is still
of Serbia, Grant/Award Number: Project
TR-31002 not close enough to commercial application, and therefore, it is unavailable for wine
producers. Thus, the aim of this study was to examine the possibility of reducing
the production of ethanol in wines using different selected yeast strains (S. cerevi-
siae, Saccharomyces bayanus, Torulaspora delbrueckii, and Metschnikowia pulcherrima)
available at the market. The application of individual yeast and sequential inoculation
for wine alcoholic fermentation was examined. The achieved effects were evaluated
by determining the content of ethanol, as well as fermentation by-products (glyc-
erol and volatile acids) and aromatic components in wine samples. Depending on the
strain/s used, a decrease in ethanol content of up to 0.9% v/v was recorded in com-
parison with fermentation by S. cerevisiae alone. The sensory analysis of produced
wine showed significant differences in taste and flavor. The results of the experiment
conducted at the laboratory level and with the use of sterile must were compared to
the ones from the scale-up experiment in real vinification conditions. The observed
differences in the alcohol content of produced wines were significantly lower.

KEYWORDS

commercial yeast strains, global warming, lower-alcohol wines, sequential inoculation

1 |  I NTRO D U C TI O N wine-growing regions (Goold et al., 2017). A wide range of factors sig-
nificantly affects sugar accumulation in the grape such as warm climate
Climate change, new market trends, and technologies' progress now- conditions combined with the lengthy maturation periods used to sat-
adays strongly affect the wine industry. Over recent decades, the av- isfy the consumer demand for rich and ripe fruit flavor in the wine.
erage alcohol content of table wines has increased by about 2% (v/v), The high alcohol content in wine has several negative consequences.
due to the high sugar content of the grapes currently used in the most One of the major issues of higher alcohol content in wines is its effect

This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium,
provided the original work is properly cited.
© 2020 The Authors. Food Science & Nutrition published by Wiley Periodicals, Inc.

Food Sci Nutr. 2020;8:1489–1498.  |


www.foodscience-nutrition.com     1489
|
1490       PUŠKAŠ et al.

on the sensory properties, which increases the perception of the heat appropriate metabolic characteristics of different yeast strains indi-
and alters the perception of wine aroma complexity (Goldner, Zamora, vidually or in mixed cultures might serve as an efficient mechanism
Di Leo Lira, Gianninoto, & Bandoni, 2009). Also, excessive alcohol for reducing ethanol production in wines (Ciani & Comitini, 2015).
intake through the consumption of wines with higher ethanol levels The application of non-Saccharomyces species for decreased alcohol
is often associated with undesirable implications on human health. production is possible through both aerobic (respiration) and anaer-
Furthermore, higher alcohol in wine may increase costs in countries obic (fermentation) metabolism (Ciani et al., 2016).
where taxes are levied according to alcohol concentration. Thus, the The aim of this research was to examine the possibility of reduc-
combination of quality, health, and economic issues associated with ing ethanol production in wines using different selected yeast strains
high-alcohol wines has created significant interest in the development (Saccharomyces cerevisiae, Saccharomyces bayanus, Torulaspora del-
of technologies for the production of reduced ethanol wines. brueckii, and Metschnikowia pulcherrima) which are currently available
Different approaches to reduce alcohol levels in wines have been at the market. The experiments implied a series of wine fermentations
proposed at all stages of the winemaking process. These mainly fit carried out both by the activity of chosen strains individually and in the
into four basic groups as viticultural, prefermentation, fermentation, form of mixed cultures applied through sequential inoculation.
and postfermentation strategies (Varela et al., 2015). Viticulture
strategies, as promising but long-term techniques, are based on the
selection of new grape varieties with low sugar accumulation, viticul- 2 | M ATE R I A L S A N D M E TH O DS
tural practices adapted to unripe grapes, and different agronomical
methods (Olego et al., 2016). On the other hand, postfermentation 2.1 | Inoculation strategies for the fermentation of
strategies such as reverse osmosis, nanofiltration, and distillation experimental wines
represent a short-term perspective dependent on the current EU
and OIV regulations. Moreover, these procedures may increase pro- In order to investigate the possibility of reducing the production of
duction costs and also can compromise the wine organoleptic quality ethanol in wine using different yeasts as producing microorganisms,
due to the elimination of volatile compounds (Schmidtke, Blackman, the following commercial strains were used:
& Agboola, 2012). Considering possible approaches, the application
of yeast strains characterized by lower sugar-to-ethanol transforma- 1. Saccharomyces cerevisiae Oenoferm Bouquet (Erbslöh, Germany),
tion rates has been imposed as an attractive way to deal with the shortened CER.
problem of high-alcohol wines (Kutyna, Varela, Henschke, Chambers, 2. Saccharomyces bayanus LittoLevure CHA (Erbslöh, Germany),
& Stanley, 2010). Lower ethanol-producing yeast strains could be shortened BAY.
isolated and characterized from spontaneous wine alcoholic fermen- 3. Torulaspora delbrueckii Oenoferm Wild and Pure (Erbslöh,
tations or obtained through the application of adaptive evolution Germany), shortened TOR.
(development of the low-alcohol variants of existing Saccharomyces 4. Metschnikowia pulcherrima FLAVIA MP346 (Lallemand, France),
cerevisiae strains) and genetic modification techniques (GMO ap- shortened MET.
proaches) (Ozturk & Anli, 2014; Varela et al., 2015). Due to poor
consumer acceptance of GMO foods and beverages, there is a need The yeast inoculation was carried out according to the plan shown
to investigate and develop the non-GMO approaches for the gener- in Table 1, for both laboratory and scale-up experiments. Production
ation of wine yeasts that produce less ethanol (Kutyna et al., 2010).
Nonconventional yeasts such as Kloeckera, Pichia, Candida,
Metschnikowia, Schizosaccharomyces, and Torulaspora species are TA B L E 1   Inoculation plan the fermentation of experimental
wines
among the main representatives of grape natural microbiota. In gen-
eral, their pronounced sensitivity to antimicrobial agents (e.g., SO2) Experimental
and higher alcohol contents prevent the complete transformation of sample Plan of inoculation with selected strains
grape sugars into ethanol during alcoholic fermentation. Therefore, CER Saccharomyces cerevisiae
their application in co-inoculation or sequential inoculation with BAY Saccharomyces bayanus
S. cerevisiae is increasingly getting popular especially regarding their MET+CER Metschnikowia pulcherrima and then 48 hr after
potential positive effects on wine flavor (Ciani et al., 2016). On the the first inoculation of S. cerevisiae
other hand, species, such as Saccharomyces bayanus, are associ- MET+BAY Metschnikowia pulcherrima and then 48 hr after
ated with spontaneous fermentation of must and have been shown the first inoculation of S. bayanus
to be of oenological interest (González, Barrio, Gafner, & Querol, TOR Torulaspora delbrueckii
2006). The use of mixed cultures of selected Saccharomyces and TOR+BAY Torulaspora delbrueckii and then 48 hr after the
non-Saccharomyces yeasts for wine fermentation can result in the first inoculation of S. bayanus
formation of higher amounts of undesired compounds (e.g., volatile MET+BAY+CER Metschnikowia pulcherrima, then 48 hr after the
phenols and ethyl acetate) which can affect both structure and the first inoculation of S. bayanus, and then 96 hr
after the first inoculation of S. cerevisiae
aromatic profile of the wines. Therefore, ensuring the expression of
PUŠKAŠ et al. |
      1491

of experimental wines was carried out under identical conditions for using official OIV methods (OIV, 2016). The content of yeast assimila-
all samples (laboratory and scale-up level) and was done in triplicate. ble nitrogen in the must was determined by formol titration method
(Zoecklein, Fugelsang, Gump, & Nury, 1999). The content of glycerol
was determined by a commercial enzyme test (Megazyme, CO).
2.2 | Laboratory-scale experiment Concentration of methanol, ethanol, and higher alcohols was de-
termined by gas chromatographic analysis using gas chromatograph
Wines were produced from the Serbian white grape variety Sila (Vitis Agilent 7890A equipped with flame ionization detector (FID) and a
vinifera L.), from grapes originating from the experimental vineyards of split/splitless injector, while a capillary column HP-INNOWax (poly-
the Faculty of Agriculture, University of Novi Sad, located in Sremski ethylene glycol; 30  m  ×  250  µm i.d. with 0.25  µm film thickness)
Karlovci. Grapes were harvested at the stage of technological maturity was used. The GC-FID conditions were previously described (Miljić,
(optimal ratio between sugar and acids, phenolic and aromatic maturity Puškaš, Vučurović, & Muzalevski, 2017).
also ensured), at the end of September 2016. The processing of grapes
included crushing and destemming (Zambelli Gamma 30), followed by
pressing in classical vertical basket press (capacity 100 kg). The sugar 2.5 | Wine sensory evaluation
content in the must was 19.5%, total acidity 5.1 g/L (as tartaric acid),
and assimilable nitrogen 225 mg/L. The must was sulfited by the addi- Buxbaum model of positive ranking, described in the paper of Amerine
tion of potassium metabisulfite (0.1 g/L) in order to prevent oxidation. and Roessler (1983), was applied for the evaluation of sensory prop-
The clarification of must was carried out by classic precipitation with erties of experimental wines. Sensory evaluation was performed by a
the addition of the Trenolin FastFlow pectolytic enzyme (Erbslöh) in panel of five qualified testers (officially certified tasters authorized for
the amount of 5 ml/hl. The must was pasteurized (Grant instruments, wine sensory analysis by Serbian Ministry of Agriculture). Four sen-
75°C during 15 min) to neutralize the influence of autochthonous sorial experiences rated up making a maximum of 20 points (up to 2
yeasts on the experimental fermentations. Fermentation was carried points for color, up to 2 points for clearness, up to 4 points for aroma,
out in 10-liter glass vessels. Must samples were inoculated with se- and up to 12 points for overall flavor). Overall flavor implies both taste
lected yeasts and rehydrated according to the manufacturer's instruc- and aroma components evaluated by retronasal olfaction. The minor
tions, and in the amounts suggested. The yeast nutrient VitaFerm Ultra unit of the scale is 0.1. The better the parameter is rated, the higher
(Erbslöh) was added after one-third of sugar fermented (sugar content the mark is given. The bottom part of the evaluation sheet contained
120 g/L), in the amount of 0.2 g/L. The fermentation temperature was the space for tasters' comments since they were asked to highlight the
maintained at 16–18°C. During alcoholic fermentation, several param- most dominant aroma descriptors of the assessed wines.
eters were evaluated in duplicate: sugars, glycerol, volatile acidity, and Wines were presented to panelists in ISO standard wine glasses,
ethanol. After the end of fermentation, the wines were racked, sulfited in isolated booths, and under daylight-type lighting. All wine sam-
(0.08 g/L potassium metabisulfite), and stored in 2-L glass bottles. The ples were evaluated by the panel (seven trial sets, Table 1) with ran-
analysis of the major volatile compounds and the sensory analysis of domized presentation order. Three sessions were employed in three
produced wines were carried out after 1 month. consecutive days where the panelist evaluated all seven individual
wines each day in two sets. Each replicate presented on three con-
secutive days of tasting was poured from a separate bottle.
2.3 | Scale-up experiment

The experiments conducted at laboratory level and with the use of 2.6 | Statistical analysis
pasteurized must were followed by the scale-up trials. The goal was to
evaluate the previously obtained results in real winemaking conditions Statistical analysis in the present study was performed using Statistica
which do not imply the pasteurization of the must. For this purpose, ex- 12.0 (StatSoft). The statistical difference between mean values of pa-
periments were carried out in 200-liter stainless still tanks. Grape pro- rameters was estimated by analyses of variance (ANOVA), at the 95%
cessing and wine fermentation were conducted in an identical way as confidence level. Values detected as significantly different by the use of
in the case of the laboratory experiment. The only difference was the Duncan multiple range test were marked with different letters (a, b, c …).
use of fresh unpasteurized grape must. After the end of fermentation,
the content of ethanol, glycerol, and volatile acids was determined.
3 | R E S U LT S A N D D I S CU S S I O N

2.4 | Analyses 3.1 | The influence of different yeast strains’


metabolic activity on ethanol production
Total sugars, total acidity (expressed as tartaric acid), volatile acidity (ex-
pressed as acetic acid), and ethanol content (hydrostatic balance Densi Different selected yeast strains were used in wine fermentation tri-
Alcomat; Gibertini) of the must and produced wines were determined als, and the effect of their metabolic activity on the content of most
|
1492       PUŠKAŠ et al.

important fermentation products was assessed. At the same time, S. cerevisiae, while sequential inoculation of grape must with M. pul-
the activity of autochthonous microbiota was suppressed by pas- cherrima, S. bayanus, and S. cerevisiae resulted in the production of
teurization of the must. the lowest amount of this compound in the test wines (11.01% v/v).
Sugar consumption profiles during the fermentation of must inoc- A significant reduction in the production of ethanol is also registered
ulated according to the defined inoculation plan are shown in Figure 1. in the experimental wines produced by the activity of M. pulcherrima
A different rate of sugar consumption between individual and sequen- and S. bayanus (MET+BAY) and T. delbrueckii with the final ethanol
tial inoculation regime was observed. The use of individual yeasts concentration of 11.30% v/v and 11.40% v/v, respectively (Figure 2).
S. cerevisiae and S. bayanus resulted in the shortest time of fermen- The highest rate of ethanol production (fermentative power) was
tation, that is, the highest rate of sugar consumption. Also, the sug- observed in wines produced only with S. cerevisiae, where approx-
ars were almost consumed completely in 6 days of fermentation by imately 70% of the ethanol content formed within the first 3 days
these individual yeast strains. The activity of M. pulcherrima in the first of fermentation. Fermentation of grape must inoculated only with
48–72 hr of fermentation, before the inoculation with S. cerevisiae and S. bayanus showed the lower fermentative power (about 50% of the
S. bayanus, resulted in a slight decrease in sugar content. After sequen- total ethanol) in the first 3 days compared with fermentation using a
tial inoculation, M. pulcherrima ferments (MET+CER, MET+BAY, and pure culture of S. cerevisiae.
MET+CER+BAY) completed fermentation in 8–9  days. T. delbrueckii In the grape must fermented with M. pulcherrima, the ethanol con-
showed similar fermentation activity as M. pulcherrima with the slower centration in the first 3 days of fermentation was only 1.30%–1.70%
sugar consumption in the first 3 days of fermentation. These results v/v, while it is assumed that the rest of the formed alcohol is produced
may be the consequence of a higher sensitivity to SO2 of M. pulcher- mainly by the activity of S. bayanus and S. cerevisiae after sequential
rima and T. delbrueckii. Higher sensitivity to SO2 increases the time inoculation. The activity of T. delbrueckii resulted in an ethanol concen-
necessary for yeasts to adapt to the environmental conditions and to tration of 0.8%–0.9% v/v in the early days of fermentation. The high-
start the alcoholic fermentation, which is associated with the reduced est ethanol production rate with a pure culture of T. delbrueckii was
consumption of sugars. In the end, it is important to emphasize that all achieved between the third and sixth days of fermentation.
treatments produced wines that had 0.5–1.0 g/L of reducing sugars. Previous studies (Canonico, Comitini, Oro, & Ciani, 2016;
Figures 2‒4 show changes in the content of the observed com- Contreras, Curtin, & Varela, 2015; Contreras et al., 2014; Gobbi et
ponents (concentration of ethanol and glycerol and volatile acidity al., 2013; Varela, Barker, Tran, Borneman, & Curtin, 2017) have re-
value) during the alcoholic fermentation of experimental wines. ported a significant reduction in ethanol yield (0.3%–1.7% v/v) when
Through the comparison of the final values of observed parameters, using non-Saccharomyces and S. cerevisiae strains in co-inoculated
the efficiency of applied yeast strains in the production of wines or sequential cultures. The use of M. pulcherrima co-inoculated with
with reduced ethanol concentration was assessed. S. cerevisiae in the pilot-scale production of Merlot wines resulted
The highest content of ethanol (11.92% v/v) was determined in in 1.0% v/v lower ethanol content compared to wines fermented by
the experimental wine produced only using a commercial strain of S. cerevisiae alone (Varela et al., 2017).
Contreras et al. (2014) have reported that sequential inoculation of
the selected strain of M. pulcherrima AWRI1149 and S. cerevisiae, after
3  days, resulted in the production of wines with a lower concentra-
tion of ethanol in relation to wine produced only by the application
of S. cerevisiae (in Chardonnay, the decrease was 0.9% v/v, while in
Shiraz, this reduction was 1.60% v/v). Canonico et al. (2016) investi-
gated the use of immobilized selected strains of non-Saccharomyces
yeasts (Starmerella bombicola, M. pulcherrima, Hanseniaspora osmophila,
and Hanseniaspora uvarum) to start fermentation, followed by inocula-
tion of free S. cerevisiae cells. The sequential inoculations of S. bombi-
cola- and M. pulcherrima-immobilized cells and S. cerevisiae-free cells
showed the best reductions in the final ethanol content (1.6% and
1.4% v/v, respectively). Also, lower ethanol production was recorded
in ferments obtained by joint activity of different Saccharomyces spe-
cies. Sequential inoculation of Saccharomyces uvarum (AWRI 2846) and
S. cerevisiae resulted in ethanol reduction of 0.8% v/v and an increase in
F I G U R E 1   Sugar consumption profiles during the alcoholic glycerol content for 6.4 g/L (Contreras, Curtin, et al., 2015). Moreover,
fermentation of pasteurized must inoculated with different yeast wines fermented with S. uvarum in the study of Varela et al. (2017)
strains in laboratory conditions. Marks describing each trial set
had a 1.7% v/v lower ethanol concentration than S. cerevisiae ferment.
are given in the figure legend: Saccharomyces cerevisiae—CER;
Previous works also showed that the application of non-Saccharomy-
Saccharomyces bayanus—BAY; Torulaspora delbrueckii—TOR;
Metschnikowia pulcherrima—MET; a detailed inoculation plan is ces yeast in fermentations with controlled aeration can be an efficient
given in Table 1 approach for the production of wines with decreased alcohol content
PUŠKAŠ et al. |
      1493

F I G U R E 2   Changes in ethanol content during alcoholic fermentations carried out by different commercial yeast strains used in
laboratory conditions. Marks describing each trial set are given in the figure legend: Saccharomyces cerevisiae—CER; Saccharomyces bayanus—
BAY; Torulaspora delbrueckii—TOR; Metschnikowia pulcherrima—MET; a detailed inoculation plan is given in Table 1.a,b,c different letters for the
results in every specific day of alcoholic fermentation indicate significant differences between values (p < .05)

F I G U R E 3   Changes in glycerol content during alcoholic fermentations carried out by different commercial yeast strains used in
laboratory conditions. Marks describing each trial set are given in the figure legend: Saccharomyces cerevisiae—CER; Saccharomyces bayanus—
BAY; Torulaspora delbrueckii—TOR; Metschnikowia pulcherrima—MET; a detailed inoculation plan is given in Table 1. a,b,c different letters for
the results in every specific day of alcoholic fermentation indicate significant differences between values (p < .05)

(Canonico, Comitini, & Ciani, 2019; Canonico, Solomon, Comitini, Ciani, relatively high value (6.7 g/L) of glycerol. The lowest production of
& Varela, 2019; Contreras, Hidalgo, et al., 2015). The following exper- this compound (5.7 g/L) was established in the wine produced only
iments generally implied controlled aeration (1–10 ml L-1 min-1) during with S. cerevisiae. Similar results for glycerol levels were reported by
the first 24–72 hr of fermentation. Under these conditions, M. pulcher- the study of Canonico et al. (2016). From the obtained results, it can
rima, T. delbrueckii, and Z. bailii strains produced wines with 0.9%–2.0% be noticed that S. bayanus, among the tested yeasts, was the most
v/v lower ethanol content compared to S. cerevisiae wines. efficient glycerol producer. The individual inoculation with both
Glycerol is nonvolatile three-hydroxy alcohol which indirectly S. cerevisiae and S. bayanus led to an intensive glycerol production in
contributes to the sensory character of a wine. At high concentra- the first 3 days of fermentation which resulted in formation of more
tion, it contributes significantly to the sweetness, body, and fullness than 70% of total glycerol amount in this period. Available literature
of wines. For these reasons, glycerol production is one of the desir- (Ribéreau-Gayon, Dubourdieu, Donèche, & Lonvaud, 2006) points
able features during grape must fermentation. Furthermore, of all out that the production of glycerol is associated with fermentation
the microbiological strategies to reduce ethanol yield, glycerol over- of the first 50 g/L of sugar from must, and these data relate primarily
production was proven to be the most effective (Varela et al., 2012). on S. cerevisiae yeast activity. From the obtained results, it can be no-
Among experimental wines produced in this study, the highest glyc- ticed that the production of glycerol in the ferments with T. delbruec-
erol content (6.99 g/L) was obtained during sequential fermentation kii during the first 3 days was minimal. T. delbrueckii was shown to be
with T. delbrueckii and S. bayanus (Figure 3). Sequential inoculation a lower producer of secondary by-products of fermentation (Ciani et
of strain S. bayanus, M. pulcherrima, and S. cerevisiae also resulted in al., 2016). Glycerol production by M. pulcherrima in the first 3 days of
|
1494       PUŠKAŠ et al.

F I G U R E 4   Changes in volatile acidity during alcoholic fermentations carried out by different commercial yeast strains used in laboratory
conditions. Marks describing each trial set are given in the figure legend: Saccharomyces cerevisiae—CER; Saccharomyces bayanus—BAY;
Torulaspora delbrueckii—TOR; Metschnikowia pulcherrima—MET; A detailed inoculation plan is given in Table 1. a,b,c different letters for the
results in every specific day of alcoholic fermentation indicate significant differences between values (p < .05)

fermentation was also low (1.3–1.75 g/L), compared to the individual volatile compounds are formed by yeast during alcoholic fermenta-
fermentations by S. cerevisiae and S. bayanus. Positive linear correla- tion, and they significantly impact the aroma and overall quality of
tion between the production of ethanol and glycerol was recorded in wines. The most important volatile compounds synthesized by wine
fermentations started by S. cerevisiae, S. bayanus, and M. pulcherrima yeast include higher alcohols, acetate esters, ethyl esters, and alde-
(0.991, 0.918, and 0.933, respectively). On the other hand, signifi- hydes among others. The influence of commercial selected yeast
cantly lower degree of linear correlation (.516) was determined be- strains used individual or sequential inoculation, on the composition
tween ethanol and glycerol content in T. delbrueckii ferments. of the aromatic compounds in experimental wines, is shown in Table 2.
Volatile acidity (VA) is a parameter which significantly deter- Metabolic activity of commercial selected yeast strains during
mines the quality of wine, and, in general, it represents the content fermentation caused significant differences in the content of aro-
of acetic acid as a by-product of alcoholic fermentation (more than matic compounds in produced wines. As shown, ten higher alcohols
80% of VA originates from acetic acid). If adequate vinification and two aldehydes were detected in the analyzed wine samples;
practices had been employed, the resulting values of this param- meanwhile, the compounds such as 2-propanol, 1-butanol, 2-buta-
eter are below the ones that can negatively impact the organo- nol, 1-pentanol, and 1-hexanol were not detected. The produced
leptic properties of the wine. According to Zoecklein et al. (1999), wines were characterized by relatively uniform content of ethyl ac-
the sensory detection threshold of volatile acidity in the wine is in etate, 2-methyl-1-propanol, that is, isobutanol, octanol, and deca-
the range of 0.7–1.1 g/L. Analysis of produced experimental wines nol. The use of commercial non-Saccharomyces strains did not lead
(Figure 4) showed the highest volatile acidity in wines obtained to the increase in the ethyl acetate levels, and the amounts deter-
by the activity of T. delbrueckii (0.44 g/L), but similar values were mined (about 50 mg/L) were far below the threshold associated with
determined also for the other ferments (ranging 0.30–0.40  g/L). a negative impact on wine sensory characteristics (above 150 mg/L).
It should be emphasized that higher glycerol production in some The content of acetaldehyde was the highest in the MET+CER and
ferments (TOR+BAY and MET+BAY+CER) was not followed by the MET+BAY+CER samples (15–20  mg/L); however, even the highest
increase in volatile acidity. This is very important, especially from levels determined were significantly lower than the ones associ-
the organoleptic point of view. The changes of the volatile acidity ated with oxidative faults (“over-ripe bruised apples,” “sherry,” and
recorded during fermentations with individual or mixed yeast cul- “nut-like” characters). At lower levels (below 50 mg/L), acetaldehyde
tures clearly indicate that the most intensive production of these can contribute pleasant fruity aromas to a wine. As for comparison,
compounds is in the first 3 days in all fermentations. Varela et al. (2017) and Canonico et al. (2016) reported that wines
produced with M. pulcherrima showed higher concentrations of ethyl
acetate, total esters, total higher alcohols, and total sulfur compounds
3.2 | The impact of used yeast strains on the compared to S. cerevisiae ferments. The wine obtained by S. bayanus
composition of volatile compounds in wines activity had a significantly higher concentration of 1-propanol (sam-
ples BAY and TOR+BAY) and 1-heptanol (sample BAY) compared
The aroma is one of the main characteristics that determine the qual- to the rest of the experimental wines. Also, the wines BAY and
ity and value of a wine. The aroma of wine is a unique mixture of TOR+BAY had significantly higher amounts of benzaldehyde (76.4
volatile compounds originating from grape and secondary products and 87.3 mg/L, respectively), in comparison with other experimen-
formed during the wine fermentation and aging. A large number of tal wines. The highest contents of 3-methyl-1-butanol (24–28 mg/L)
PUŠKAŠ et al. |
      1495

TA B L E 2   Concentration of the major volatile compounds in wines produced in laboratory-scale experiment

Volatile compound
(mg/L) CER BAY MET+CER MET+BAY TOR TOR+BAY MET+BAY+CER
a bc c a b a
Acetaldehyde 6.5 ± 0.3 13 ± 1.2 18 ± 0.9 7.4 ± 0.7 10.7 ± 0.8 7 ± 0 15 ± 0.5c
Ethyl acetate 49 ± 1.5a 49 ± 0.7a 48.2 ± 2.3a 49.4 ± 1.9a 46.1 ± 1.1a 49.8 ± 1.5a 47.5 ± 1.1a
2-Butanol nd nd nd nd nd nd nd
2-Propanol nd nd nd nd nd nd nd
1-Propanol 0.7 ± 0.1a 4.5 ± 0.4b 1.0 ± 0.2a 1.1 ± 0a 0.4 ± 0.1a 3.7 ± 0.3b 0.9 ± 0a
a a a a a a
2-Methyl−1-propanol 11.7 ± 0.8 11.6 ± 0.7 12.2 ± 1.2 11.6 ± 1.4 10.9 ± 0.3 11.5 ± 0.7 11.4 ± 0.8a
1-Butanol nd nd nd nd nd nd nd
b b c b ab a
3-Methyl−1-butanol 26.4 ± 2.4 23.9 ± 1.3 28.1 ± 0.7 22.8 ± 1 20.6 ± 1.5 18.9 ± 0.6 24 ± 1.5b
1-Pentanol nd nd nd nd nd nd nd
1-Hexanol nd nd nd nd nd nd nd
1-Heptanol 10.7 ± 0.1a 23 ± 1.9d 19.5 ± 1.0 c 14.5 ± 0.5b 19.7 ± 1c 9 ± 0.8a 16.4 ± 0.4bc
a a a a a a
Furfural 0.6 ± 0.2 0.7 ± 0.2 0.8 ± 0.3 0.6 ± 0.2 1.1 ± 0.3 0.9 ± 0.3 0.4 ± 0.1a
Benzaldehyde 48.5 ± 2.1a 76.4 ± 1d 49.5 ± 0.7a 56.8 ± 1.4b 61.2 ± 1.5c 87.3 ± 1.9e 51.4 ± 0.9a
a a a a a a
1-Octanol 22 ± 1.2 23.2 ± 1 21.7 ± 1.5 21.5 ± 0.8 21.2 ± 0.3 23.5 ± 1.2 20.5 ± 0.8a
1-Nonanol 5.2 ± 0.4a nd nd 5.4 ± 0.7a 5.1 ± 0.4a nd nd
a a a a a a
1-Decanol 6.2 ± 0.2 5.9 ± 0.6 5.7 ± 0 6.3 ± 0.8 6.7 ± 0.2 5.9 ± 0.7 5.4 ± 0.2a
Benzyl alcohol 10.5 ± 0.6c 8 ± 0.4b 11.2 ± 0.8c 9.1 ± 2b 8 ± 0.7b 5.8 ± 0.3a 9.7 ± 0.8bc

Note: Saccharomyces cerevisiae—CER; Saccharomyces bayanus—BAY; Torulaspora delbrueckii—TOR; Metschnikowia pulcherrima—MET. Different letters
in the same row indicate significant differences between values (p < .05). nd, not detected.

TA B L E 3   Sensory evaluation of wines produced in laboratory-scale experiment

Clearness (max 2 Overall flavor (max 12 Total (max


  Color (max 2 points) points) Aroma (max 4 points) points) 20 points)

CER 2.0 ± 0.0a 2.0 ± 0.0a 3.2 ± 0.2a 11.3 ± 0.2ab 18.5 ± 0.2a


BAY 2.0 ± 0.0a 2.0 ± 0.0a 3.6 ± 0.1b 11.4 ± 0.0 b 19.0 ± 0.1bc
a a a a
MET+CER 2.0 ± 0.0 2.0 ± 0.0 3.2 ± 0.1 11.2 ± 0.1 18.4 ± 0.1a
MET+BAY 2.0 ± 0.0a 2.0 ± 0.0a 3.7 ± 0.0 b 11.5 ± 0.1bc 19.2 ± 0.1c
a a a b
TOR 2.0 ± 0.0 2.0 ± 0.0 3.4 ± 0.1 11.4 ± 0.2 18.8 ± 0.2b
TOR+BAY 2.0 ± 0.0a 2.0 ± 0.0a 3.7 ± 0.1b 11.6 ± 0.0 c 19.3 ± 0.1c
a a a a
MET+BAY+CER 2.0 ± 0.0 2.0 ± 0.0 3.4 ± 0.2 11.0 ± 0.2 18.4 ± 0.2a

Note: The data represent the average values of the scores given by five members of the panel. Different letters in the same column indicate
significant differences between values (p < .05).

and benzyl alcohol (9.7–11.2 mg/L) were detected in wines produced experimental wines produced using different commercially available
by the activity of S. cerevisiae, in both individual and sequential in- yeast strains in laboratory conditions. On the other hand, the color
oculations (samples CER, MET+CER, and MET+BAY+CER). The pres- differences could be considered insignificant. In terms of aroma,
ence of furfural was detected in small amounts in all wine samples. the best-evaluated wines were produced by the activity of yeasts
One of the main sources of furfural in wines is toasted oak wood. MET+BAY and TOR+BAY. The lowest marks were given to MET+CER
Considering the fact that wines in this study were not in contact with and MET+BAY+CER wines, which can be due to the higher content
wood, the possible explanation for furfural occurrence is the pas- of acetaldehyde compared to other wines. Wine MET+BAY was
teurization of must before inoculation of yeast strains. characterized by citrus flavors, while in MET+CER, melon and ba-
nana flavor dominated. Wines MET+BAY and TOR+BAY had fuller,
more complex, and rounded flavors compared to other experimental
3.3 | Wine sensory analysis wines. These properties could partly be impacted by the higher con-
tents of 1-propanal, 1-heptanol (flower tones), and benzaldehyde (al-
The results of the sensory analysis (Table 3) indicate that a dif- mond flavor) compared to other wines. The use of M. pulcherrima and
ference in the aroma, as well as taste and flavor, was recorded in T. delbrueckii in co-inoculation with S. bayanus commercial strains
|
1496       PUŠKAŠ et al.

gave the complexity to the sensory profiles of produced wines which


was confirmed by the highest taste and overall flavor marks, as well
as the highest overall marks (19.2–19.3). Varela et al. (2017) used a
consensus-based descriptive methodology and demonstrated that
apart from being able to produce wine with reduced alcohol concen-
tration, M. pulcherrima in co-inoculation with S. cerevisiae gave wines
with similar sensory profiles (not differing significantly in any attrib-
ute and with high scores given for desirable sensory descriptors) as
uninoculated and control S. cerevisiae ferments.

3.4 | Scale-up experiment

Evaluation of the results obtained in the laboratory experiments


was done by simulating the same vinification conditions in the
scale-up experiment. The aim was to assess the activity of ap-
plied yeast strains in a real must sample and in the presence of
autochthonous microbiota. The final concentration of ethanol and
glycerol, and value of volatile acidity in these wines are shown
in Figure 5. The aptitude of commercial yeast strains for lower-
ing the ethanol content of wine was lower compared to ferments
based on sterile must use. The competitiveness between inocu-
lated non-Saccharomyces yeasts and yeasts from autochthonous
microbiota resulted in a slower start of fermentation than in the
laboratory experiment (results not shown). Moreover, the content
of ethanol was in a pretty narrow range 11.4%–11.8% v/v, while
the content of glycerol was 3.78–4.89  g/L which is significantly
lower than that obtained in wines from laboratory-scale experi-
ment (5.7–6.99 g/L). The value of volatile acidity was very simi-
lar in both experimental sets (0.3–0.45  g/L). Similar differences
among the results of experiments with model and real must sam-
ples were reported (Ciani & Ferraro, 1998; Ferraro, Fatichenti, &
Ciani, 2000). The use of Starmerella bombicola (formerly C. stellata)
and S. cerevisiae resulted in a high production of glycerol, succinic
acid, and different by-products (interactions involving acetalde- F I G U R E 5   Final concentrations of glycerol, volatile acids, and
ethanol in wines produced by different commercial yeast strains
hyde and acetoin) with a consequent reduction of final ethanol
used in scale-up conditions. Marks describing each trial in the figure
amount. The reduction in ethanol content in these experiments
legend: Saccharomyces cerevisiae—CER; Saccharomyces bayanus—
varied from 0.64% v/v at pilot scale in natural grape juice to 1.60% BAY; Torulaspora delbrueckii—TOR; Metschnikowia pulcherrima—MET;
v/v at laboratory scale using synthetic grape juice. Furthermore, a detailed inoculation plan is given in Table 1. a,b,c different letters
different factors affecting the metabolism of M. pulcherrima indicate significant differences between values (p < .05)
AWRI1149 during fermentation of nonsterile Shiraz must were
evaluated (Contreras, Curtin, et al., 2015). Among different in-
oculation regimes which were applied, only initial inoculation with 4 | CO N C LU S I O N
1 × 10 6 cells/mL of M. pulcherrima AWRI1149, followed by S. cer-
evisiae after 50% sugar consumption, leads to a significant ethanol In summary, this work evaluated the potential of commercial
concentration reduction. Canonico et al. (2016) showed that appli- selected yeast strain application for purpose of wines with de-
cation of immobilized selected strains of M. pulcherrima, followed creased alcohol content production. Sequential inoculation of the
by inoculation of free S. cerevisiae cells, resulted in a decrease in must with M. pulcherrima, S. bayanus, and S. cerevisiae resulted in
ethanol content for 1.3% v/v when synthetic grape juice was used, the production of wines with the lowest ethanol content among
while in the case of natural grape juice the reduction was 1% v/v experimental samples (decrease of 0.9% v/v compared to the con-
(in both trials, the beads of immobilized yeast were removed after trol wine). Significant differences in the content of certain aro-
72 hr from inoculation). matic compounds, as well as in taste and flavor, were also found
PUŠKAŠ et al. |
      1497

in produced wines. The experiment in real conditions showed that and Biotechnology, 99(4), 1885–1895. https​ ://doi.org/10.1007/
s00253-014-6193-6
used commercially available Saccharomyces and non-Saccharomy-
Contreras, A., Hidalgo, C., Henschke, P. A., Chambers, P. J., Curtin, C.,
ces were not effective enough in lowering ethanol concentration & Varela, C. (2014). Evaluation of non-Saccharomyces yeasts for
in wines due to interactions and competitiveness with yeasts from the reduction of alcohol content in wine. Applied and Environmental
autochthonous microbiota. Already evident problems of wines Microbiology, 80(5), 1670–1678. https​
://doi.org/10.1128/
aem.03780-13
with very high alcohol content will force wine producers in near
Contreras, A., Hidalgo, C., Schmidt, S., Henschke, P. A., Curtin, C., &
future to find more efficient ways of directing alcoholic fermenta- Varela, C. (2015). The application of non-Saccharomyces yeast in
tion to the production of wines with lower alcohol content and not fermentations with limited aeration as a strategy for the produc-
worsened sensory properties. The application of Saccharomyces tion of wine with reduced alcohol content. International Journal
and non-Saccharomyces strains obtained by adaptive evolution of Food Microbiology, 205, 7–15. https​ ://doi.org/10.1016/j.ijfoo​
dmicro.2015.03.027
principles (non-GM) for this purpose will need further investiga-
Ferraro, L., Fatichenti, F., & Ciani, M. (2000). Pilot scale vinification
tion and commercialization. process using immobilized Candida stellata cells and Saccharomyces
cerevisiae. Process Biochemistry, 35(10), 1125–1129. https​ ://doi.
AC K N OW L E D G M E N T org/10.1016/S0032-9592(00)00148-5
Gobbi, M., Comitini, F., Domizio, P., Romani, C., Lencioni, L., Mannazzu, I.,
Financial support from the Ministry of Education, Science and
& Ciani, M. (2013). Lachancea thermotolerans and Saccharomyces cerevi-
Technological Development of the Republic of Serbia (Project TR- siae in simultaneous and sequential co-fermentation: A strategy to en-
31002) is greatly appreciated. hance acidity and improve the overall quality of wine. Food Microbiology,
33(2), 271–281. https​://doi.org/10.1016/j.fm.2012.10.004
Goldner, M. C., Zamora, M. C., Di Leo Lira, P., Gianninoto, H., &
C O N FL I C T O F I N T E R E S T
Bandoni, A. (2009). Effect of ethanol level in the perception of
The authors declare that they have no conflicts of interest. aroma attributes and the detection of volatile compounds in
red wine. Journal of Sensory Studies, 24(2), 243–257. https​://doi.
E T H I C A L A P P R OVA L org/10.1111/j.1745-459X.2009.00208.x
González, S. S., Barrio, E., Gafner, J., & Querol, A. (2006). Natural hybrids from
This study does not involve any human or animal testing.
Saccharomyces cerevisiae, Saccharomyces bayanus and Saccharomyces
kudriavzevii in wine fermentations. FEMS Yeast Research, 6(8), 1221–
ORCID 1234. https​://doi.org/10.1111/j.1567-1364.2006.00126.x
Uroš D. Miljić  https://orcid.org/0000-0002-1947-8122 Goold, H. D., Kroukamp, H., Williams, T. C., Paulsen, I. T., Varela, C., &
Pretorius, I. S. (2017). Yeast's balancing act between ethanol and
glycerol production in low-alcohol wines. Microbial Biotechnology,
REFERENCES 10(2), 264–278. https​://doi.org/10.1111/1751-7915.12488​
Amerine, M. F., & Roessler, A. B. (1983). Wines, their sensory evaluation. Kutyna, D. R., Varela, C., Henschke, P. A., Chambers, P. J., & Stanley, G.
New York, NY: Freeman W.H. A. (2010). Microbiological approaches to lowering ethanol concen-
Canonico, L., Comitini, F., & Ciani, M. (2019). Metschnikowia pulcherrima tration in wine. Trends in Food Science & Technology, 21(6), 293–302.
selected strain for ethanol reduction in wine: Influence of cell im- https​://doi.org/10.1016/j.tifs.2010.03.004
mobilization and aeration condition. Foods, 8(9), 378. https​://doi. Miljić, U., Puškaš, V., Vučurović, V., & Muzalevski, A. (2017).
org/10.3390/foods​8 090378 Fermentation characteristics and aromatic profile of plum wines
Canonico, L., Comitini, F., Oro, L., & Ciani, M. (2016). Sequential fermen- produced with indigenous microbiota and pure cultures of se-
tation with selected immobilized non-Saccharomyces yeast for reduc- lected yeast. Journal of Food Science, 82(6), 1443–1450. https​://doi.
tion of ethanol content in wine. Frontiers in Microbiology, 7(278), 1–11. org/10.1111/1750-3841.13736​
https​://doi.org/10.3389/fmicb.2016.00278​ Olego, M. Á., Álvarez-Pérez, J. M., Quiroga, M. J., Rebeca Cobos, M.-
Canonico, L., Solomon, M., Comitini, F., Ciani, M., & Varela, C. (2019). S.-G., Medina, J. E., González-García, S., … Garzón-Jimeno, J. E.
Volatile profile of reduced alcohol wines fermented with se- (2016). Viticultural and biotechnological strategies to reduce alcohol
lected non-Saccharomyces yeasts under different aeration condi- content in red wines. In I. L. Antonio Morata (Ed.), Grape and wine
tions. Food Microbiology, 84, 103247. https​://doi.org/10.1016/j. biotechnology (pp. 361–380). Rijeka, Croatia: InTechOpen.
fm.2019.103247 Ozturk, B., & Anli, E. (2014). Different techniques for reducing alcohol
Ciani, M., & Comitini, F. (2015). Yeast interactions in multi-starter wine levels in wine: A review. BIO Web of Conferences, 3, 02012. https​://
fermentation. Current Opinion in Food Science, 1, 1–6. https​://doi. doi.org/10.1051/bioco​nf/20140​3 02012
org/10.1016/j.cofs.2014.07.001 Ribéreau-Gayon, P., Dubourdieu, D., Donèche, B., & Lonvaud, A. (2006).
Ciani, M., & Ferraro, L. (1998). Combined use of immobilized Candida The microbiology of wine and vinifications (Vol 1). Chichester, UK: John
stellata cells and Saccharomyces cerevisiae to improve the quality of Wiley & Sons Ltd.
wines. Journal of Applied Microbiology, 85(2), 247–254. https​://doi. Schmidtke, L. M., Blackman, J. W., & Agboola, S. O. (2012). Production
org/10.1046/j.1365-2672.1998.00485.x technologies for reduced alcoholic wines. Journal of Food Science,
Ciani, M., Morales, P., Comitini, F., Tronchoni, J., Canonico, L., Curiel, J. 77(1), R25–R41.https​://doi.org/10.1111/j.1750-3841.2011.02448.x
A., … Gonzalez, R. (2016). Non-conventional yeast species for lower- Varela, C., Barker, A., Tran, T., Borneman, A., & Curtin, C. (2017). Sensory
ing ethanol content of wines. Frontiers in Microbiology, 7(642), 1–13. profile and volatile aroma composition of reduced alcohol Merlot
https​://doi.org/10.3389/fmicb.2016.00642​ wines fermented with Metschnikowia pulcherrima and Saccharomyces
Contreras, A., Curtin, C., & Varela, C. (2015). Yeast population dynam- uvarum. International Journal of Food Microbiology, 252, 1–9. https​://
ics reveal a potential ‘collaboration’ between Metschnikowia pul- doi.org/10.1016/j.ijfoo​dmicro.2017.04.002
cherrima and Saccharomyces uvarum for the production of reduced Varela, C., Dry, P. R., Kutyna, D. R., Francis, I. L., Henschke, P. A., Curtin,
alcohol wines during Shiraz fermentation. Applied Microbiology C. D., & Chambers, P. J. (2015). Strategies for reducing alcohol
|
1498       PUŠKAŠ et al.

concentration in wine. Australian Journal of Grape and Wine Research,


21(S1), 670–679.https​://doi.org/10.1111/ajgw.12187​ How to cite this article: Puškaš VS, Miljić UD, Djuran JJ,
Varela, C., Kutyna, D. R., Solomon, M. R., Black, C. A., Borneman, A., Vučurović VM. The aptitude of commercial yeast strains for
Henschke, P. A., … Chambers, P. J. (2012). Evaluation of gene modi-
lowering the ethanol content of wine. Food Sci Nutr.
fication strategies for the development of low-alcohol-wine yeasts.
Applied and Environmental Microbiology, 78(17), 6068–6077. https​:// 2020;8:1489–1498. https​://doi.org/10.1002/fsn3.1433
doi.org/10.1128/aem.01279-12
Zoecklein, B. W., Fugelsang, K. C., Gump, B. H., & Nury, F. (1999). Wine
analysis and production. New York, NY: Kluwer Academic.
Available online at www.sciencedirect.com

ScienceDirect

Microbiological strategies to produce beer and wine


with reduced ethanol concentration
Javier Varela1 and Cristian Varela2

Changes in consumer preferences, government policies and as either biological or physical processes [reviewed by
environmental conditions have driven research efforts towards Branyik et al. 5]. Biological processes are based on
producing alcoholic beverages with reduced alcohol content, limiting alcohol formation, while physical methods
namely wine and beer. While the strategies available to involve ethanol removal from regular beer. Biological
accomplish this goal vary for wine and beer, a common processes cover arrested or limited fermentation, the use
approach relies on the use of yeast strains which are less of non-conventional yeast species and continuous fer-
efficient at producing ethanol. Here we discuss current mentation, whereas physical methods include: vacuum
research on the isolation and/or generation of yeast strains able rectification and evaporation, spinning cone column
to produce beer or wine with reduced ethanol concentration. distillation, osmotic distillation, dialysis and reverse
Particular consideration is given to the impact of ‘low-ethanol’ osmosis [1,5].
yeasts on volatile composition and sensory profile of beer and
wine. Strategies aimed at reducing wine alcohol concentration
include: viticultural practices, such as decreasing the leaf
Addresses area to fruit weight ratio in order to lower sugar concen-
1
School of Microbiology/Centre for Synthetic Biology and tration in the berry; pre-fermentation and winemaking
Biotechnology/Environmental Research Institute/APC Microbiome practices, including dilution and sugar removal from the
Institute, University College Cork, Cork T12 YN60, Ireland
2
The Australian Wine Research Institute, P.O. Box 197, Glen Osmond,
grape must; microbiological strategies such as the use of
Adelaide, SA 5064, Australia yeasts that are less efficient at ethanol production; and
post-fermentation practices and processing technologies,
Corresponding author: Varela, Cristian (cristian.varela@awri.com.au) including the physical removal of alcohol after fermenta-
tion [reviewed by Varela et al. 6].
Current Opinion in Biotechnology 2019, 56:88–96
This review comes from a themed issue on Food biotechnology Yeast diversity during brewing and
Edited by Rute Neves and Herwig Bachmann winemaking
The first beers and wines were the products of spontane-
ous fermentations; microorganisms present on the surface
of grains and fruits, or unintentionally introduced by
https://doi.org/10.1016/j.copbio.2018.10.003 human action, were likely responsible for their produc-
0958-1669/ã 2018 Elsevier Ltd. All rights reserved. tion [7]. Spontaneous fermentation involves the sequen-
tial action of several yeast species, which not only gener-
ate ethanol but also produce numerous secondary
metabolites that shape the aroma and flavor of beer
and wine [7,8].

Introduction Two Saccharomyces species are associated with the pro-


Health concerns related to alcohol consumption, no-alco- duction of beer. Saccharomyces carlbergensis is employed in
hol policies advocated for some consumers (e.g. drivers, ale beer production, while Saccharomyces pastorianus (syn.
pregnant women) and growing market demand, have S. carlbergensis), which is an interspecies hybrid between
focused research attention on reducing alcohol concentra- S. cerevisiae and Saccharomyces eubayanus [9], is used for the
tion in beer and wine [1,2]. Additional reasons for this production of lager beers. Inoculation is performed with
focus, particularly in wine, include the effect of high pure starter cultures which ensures a dominant yeast
alcohol levels on flavor perception, sometimes impacting population from the beginning of fermentation. Although
negatively on wine sensory properties [3,4], and greater most beers are produced by inoculation, certain beer
taxes for wines with high alcohol concentration. styles (e.g. Belgium lambic beer and American Coolship
Ales) rely on spontaneous fermentation. Several yeast
Reduced-alcohol beers are often classified according to species have been found during these latter fermenta-
their alcohol content, with low-alcohol beers and alco- tions including Meyerozyma guilliermondii, Debaryomyces
hol-free beers containing <1.2% v/v and <0.5% v/v of spp., Pichia spp., Wickerhamomyces anomalus, Brettanomyces
alcohol, respectively. These beers can be produced anomalus, Brettanomyces custersii, Brettanomyces bruxellensis,
through several strategies which, are broadly categorized Candida krusei, Cryptococcus keutzingii, Saccharomyces

Current Opinion in Biotechnology 2019, 56:88–96 www.sciencedirect.com


Microbiological strategies for reducing alcohol concentration Varela and Varela 89

pastorianus, Naumovia castellii, Kazachstania servazzii and target complete wine fermentation, therefore minimising
Rhodotorula mucilaginosa [10,11]. residual sugar and any potential impact on wine
sweetness.
Until the 1980s wine was largely produced by sponta-
neous fermentation. This process is characterized by a In wine, gene modification technologies (GM
rich microbial diversity, with more than 40 of the approaches) have been used to partially divert carbon
1500 known yeast species isolated from fermenting metabolism away from ethanol formation, by redirect-
grape must [8]. Fermentation usually commences by ing carbon to other end points, while attempting to
the action of different non-Saccharomyces yeasts until maintain yeast fermentation performance and wine
indigenous S. cerevisiae yeasts dominate. The micro- quality [reviewed by Varela et al. [6] and Kutyna
bial communities responsible for this originate in the et al. [27]]. Therefore, researchers not only have to
vineyard and the winery [12]. While several species find suitable end-point metabolites, they also have to
live on the surfaces of vines, grapes and vineyard soil determine whether or not the introduced gene mod-
[13,14], others reside in the winery environment with ifications affect the formation of other metabolites
different yeast species isolated from equipment, soil which will in turn affect wine quality. The most rele-
and air [15–19]. vant GM strategies to decrease ethanol production in S.
cerevisiae are summarized in Figure 1, while gene targets
Until recently, non-Saccharomyces yeasts were regarded as and their location in the yeast metabolic network are
responsible for microbial-related problems during wine detailed in [6]. Successful strategies proven to
production, which resulted in the almost global use of decrease ethanol production include: increasing glyc-
pure S. cerevisiae yeast starter cultures. However, over the erol production, diverting carbon to the tricarboxylic
past 5–10 years thanks to numerous reports describing the acids (TCA) cycle, relieving glucose repression,
potential of these so-called non-conventional yeasts to increasing trehalose production, increasing lactic acid
introduce desirable aromas and flavors to wine during production and expressing genes from other species.
fermentation, their role in the production of fermented Although not yet tested, increasing the production of
beverages has been revised [8,20–22]. This has generated complex carbohydrates, such as glycogen and cellulose,
a growing interest in isolating and characterizing non- could also provide an alternative to decrease ethanol
conventional yeasts for development of starter cultures production in yeast.
that increase flavor diversity in beer and wine. Given the
enormous diversity of non-conventional yeasts, these GM approaches have also been used to generate S.
species are also being studied to determine their ability cerevisiae strains able to produce beer with reduced alco-
to produce beer and wine with reduced ethanol hol concentration (Figure 1). Strains lacking the genes
concentration. FUM1, and KGD1/KGD2, encoding fumarase and alpha-
ketoglutarate dehydrogenase, respectively, produced
Generating S. cerevisiae yeasts with reduced beer with an alcohol concentration below 0.5% v/v [28
ethanol production ]. In these strains, carbon was directed towards the
S. cerevisiae is the main yeast used for wine fermentation. formation of lactic acid, a compound that effectively
Depending on the environmental conditions S. cerevisiae masks wort-like off-flavors. However, ‘side effects’ were
can either respire or ferment sugar. If oxygen is available observed, including increased concentration of the off-
and sugar concentration is low, respiration takes place, flavor diacetyl and depleted production of ethyl acetate
whereas under anaerobic conditions fermentation is and isoamyl acetate, two key beer flavor molecules.
favored (Crabtree effect). Aeration is a common practice Deletion of genes encoding alcohol dehydrogenase
in commercial wine production, performed through (ADH), the last enzymatic step in ethanol formation,
‘pump-overs’ or macro-oxygenation techniques. has also been applied to successfully reduce beer alcohol
Although air addition can lower ethanol production by concentration (US Patent No. 4814188). Strains lacking
S. cerevisiae, the amount of air required to achieve this also ADH genes divert carbon towards the formation of glyc-
increases the concentration of acetic acid, which can erol, and unfortunately also to the production of acetal-
negatively affect wine aroma [23,24]. There are numerous dehyde, an off-flavor associated with ‘bruised apple’
strains of this species commercially available, however sensory descriptors.
they all generate similar ethanol yields, which translates
to comparable wine ethanol concentrations [25,26]. GM approaches have proven very effective at reducing
Although there are no published studies describing the wine ethanol concentration with some engineered strains
ethanol productivity of S. cerevisiae brewing strains, it is producing 1.5–2.5 % v/v less ethanol than their parent
safe to assume that ethanol yields are comparable. Thus, strains. Although most GM strategies have reported an
initial efforts to produce ‘low-ethanol’ yeasts focused effect on the formation of volatile compounds in addition
primarily on generating new S. cerevisiae strains. It is to reduced alcohol concentration, none have described
key to highlight that microbial strategies described here the sensory profile of the resulting wines. Thus, it is not

www.sciencedirect.com Current Opinion in Biotechnology 2019, 56:88–96


90 Food biotechnology

Figure 1

Increasing glycerol production


GPD overexpression
Sugar Ethanol PDC deletion
ADH deletion
TPI1 deletion
FPS1 modification
Diverting carbon to TCA cycle

Sugar Ethanol MDH2 overexpression


FRD1 overexpression
Relieving glucose repression
HXT1/7 chimera
Other carbon HXK2 deletion
metabolites MIG1 deletion
Increasing trehalose production
TPS1 overexpression
Increasing lactic acid production
FUM1 deletion
KGD1/KGD2 deletion
LDH overexpression
Expression of genes from other species
GOX expression
NoxE expression
Increasing complex carbohydrates
Glycogen
Cellulose
Formate production through
formate lyase expression
Current Opinion in Biotechnology

GM strategies to decrease ethanol production in S. cerevisiae. The different genetic modifications aiming to divert carbon away from ethanol
production are shown. Several of these strategies have been tested and proven successful in reducing ethanol in wine and beer (ticked boxes)
while others have not been evaluated yet (empty boxes). Gene targets and their location in the yeast metabolic network are detailed in Ref. [6].

possible to completely assess the success of these engi- pressure on producing enough hops to cope with the
neered strains. market demand. This GM strain represents an interesting
alternative for brewing in a more controlled and sustain-
Considering the negative public perception toward geneti- able manner.
cally modified organisms (GMOs) in food and beverage
production it is not surprising that only one article [29] has Adaptive evolution offers another potential strategy to
been published in this area in recent years. Perhaps the generate yeast strains that produce reduced amounts of
possibility of using CRISPR-Cas9 strategies to generate ethanol. This approach relies on applying a selection
yeast strains, which could be potentially regarded as non- pressure that favors a metabolic diversion away from
GM [30,31] could revitalize research on this topic. Yeast ethanol production. Several reagents can be used to shift
strains which, in addition to lower ethanol concentration, yeast carbon metabolism and potentially generate ‘low
generate products with added value, may be attractive for alcohol’ strains [reviewed by Kutyna et al. 27]. Although
the fermented beverage industries. For example, expres- different agents/conditions have been tested [35,36], only
sing the enzyme pyruvate-formate lyase would divert yeast one adaptively evolved strain has been shown to decrease
carbon metabolism towards the formation of formate [32], ethanol levels significantly and become a commercial
which then could be cleaved by formate hydrogenlyase product [37]. After 200 generations under hyperosmotic
resulting in the production of carbon dioxide and hydrogen stress caused by potassium chloride, an evolved strain
[33]. Potentially, hydrogen can be then stored and sold or showing increased glycerol production and a reduced
used to provide energy in-house. ethanol yield was isolated. This strain was then subjected
to two rounds of spore mating generating a hybrid strain
Recently, GM approaches have been used to generate a able to produce Shiraz wines with 1.3% v/v less ethanol,
yeast strain able to produce the molecules that impart hop and increased acidity and glycerol concentration than the
flavor in beer [34]. Hops are an expensive ingredient in parental wine strain [37]. These findings suggest that the
beer and the increasing preference for beers with high generation of ‘low-ethanol’ strains by adaptive evolution
hoppy flavors (IPA and Pale Ale varieties) has put is not an easy task, and that promising strain isolates may

Current Opinion in Biotechnology 2019, 56:88–96 www.sciencedirect.com


Microbiological strategies for reducing alcohol concentration Varela and Varela 91

need further rounds of improvement to create a commer- Saccharomyces hybrids has not been shown, a natural
cial yeast product. S. cerevisiae x S. kudriavzevii hybrid has been described
which, relative to S. cerevisiae, produced more glycerol
Use of non-conventional Saccharomyces and potentially less ethanol during wine fermentation
yeasts to reduce ethanol concentration at low temperature (14  C) [41].
Members of the Saccharomyces clade, which includes
Saccharomyces kudriavzevii, Saccharomyces cariocanus, Sac- Similarly, hybridization has been used to improve the
charomyces mikatae, Saccharomyces eubayanus, Saccharomyces fermentation performance, stress tolerance and flavor
uvarum and Saccharomyces paradoxus, can provide an addi- profile of brewing strains [reviewed by Krogerus et al.
tional source of novel wine yeast strains with reduced 47]. However, a set of S. cerevisiae x S. eubayanus hybrids
ethanol production. generated to increase aroma in lager beers showed vari-
able ethanol production, with some strains producing up
S. uvarum has been shown to produced Malvasia delle to 6% v/v ethanol and others only 3% v/v [48]. This
Lipari wines with lower alcohol concentration (0.7% v/v suggest that generating hybrids for low-ethanol beer
less) and volatile acidity (acetic acid) than wines fermented production is a viable option.
with S. cerevisiae [38]. S. uvarum wines showed increased
positive sensory attributes and decreased negative attri-
butes compared to S. cerevisiae wines [38]. A study by Varela Non-Saccharomyces species exhibiting
et al. [39] also showed the ability of S. uvarum to produce decreased ethanol yield
wines with reduced ethanol concentration; Merlot wines Despite an initial bad reputation, non-Saccharomyces
fermented with this yeast had 1.7% v/v less ethanol than S. yeasts are now considered to contribute positively during
cerevisiae wines and increased glycerol concentration [39]. the production of alcoholic beverages [7,8,49]. These
However, S. uvarum wines were characterized by a sensory species can be used for controlling spoilage microorgan-
profile mostly dominated by unusual and negative sensory isms, to conduct malolactic fermentation, which is the
attributes that would conventionally be considered off- transformation of L-malic acid to L-lactic acid intended
flavors. Nonetheless, these reports demonstrate the poten- mainly to deacidify wines, to improve wine clarification
tial of S. uvarum to produce wines with decreased alcohol and filtration, to shape beer and wine sensory profile, and
concentration, but they highlight the importance of sensory to reduce the ethanol concentration in fermented bev-
analysis of such wines. erages [7,50].

Although not found during wine fermentation S. kudriav- Table 1 lists recent studies reporting the use of non-
zevii presents several traits relevant for winemaking [40], Saccharomyces yeast species to produce wine with reduced
including the potential to reduce ethanol concentration in ethanol concentration and indicates their impact on wine
wine [41]. S. kudriavzevii is not able to compete well with S. composition and/or wine sensory properties. While non-
cerevisiae during fermentation and it usually requires par- Saccharomyces yeasts are generally able to ferment wort to
ticular environmental conditions, such as low temperature produce beer, they are not capable of completing wine
and aeration, to grow competitively and improve its per- fermentation on their own. Instead they have to be
formance [40,42]. Mixed fermentations inoculated with S. accompanied by a S. cerevisiae wine strain, which is added
kudriavzevii and S. cerevisiae and supplied with limited sequentially (inoculating first with the non-Saccharomyces
aeration, produced wines with 1% v/v lower ethanol con- yeast followed by S. cerevisiae) or simultaneously [7,50].
centration than wines fermented with S. cerevisiae alone and Thus, different inoculation regimes have been trialled to
increased glycerol concentration [42]. Although S. kudriav- reduced ethanol concentration and shape flavor profile
zevii has shown potential to reduce wine ethanol concen- [51]. Additionally, several reports have explored the oxi-
tration at lab scale, it has not been yet tested in grape juice dative metabolism observed in some non-Saccharomyces
or at pilot-scale. species, which can respire sugar at high sugar concentra-
tions and therefore, ‘burn off’ carbon that would other-
A strategy to eliminate unwanted traits in Saccharomyces wise go into ethanol formation [23,52–54]. While aeration
yeasts, such as production of off-flavor compounds or can enhance the growth and the persistence of non-
poor performance, is the generation of interspecific Saccharomyces yeasts during wine fermentation [55], it
yeast hybrids. These can be generated from crosses can also negatively impact wine sensory profile increasing
between S. cerevisiae and other members of the Saccha- the concentration of some off-flavors [53,54]. The pro-
romyces clade [21]. This approach has been used to duction of some of these off-flavors is related to amino
generate robust hybrids with novel sensory attributes acid metabolism and it has been suggested that supple-
[43,44], to increase cryotolerance and reduce volatile mentation with carefully formulated nitrogen sources
acidity [45], and to produce hybrids with enhanced could be used to reduce their formation [54]. In fact,
flavor diversity and reduced hydrogen sulfide produc- nitrogen sources can affect the fermentation performance
tion [46]. Although the generation of ‘low-ethanol’ of mixed yeast cultures, stimulate nitrogen consumption

www.sciencedirect.com Current Opinion in Biotechnology 2019, 56:88–96


92 Food biotechnology

Table 1

Summary of recent studies evaluating the use of non-Saccharomyces yeasts to lower wine alcohol concentration and their impact on
wine composition and/or wine sensory properties

Non-Saccharomyces Scale and Grape variety Ethanol Impact on chemical composition References
yeast a conditions of reduction b or sensory attributes c
fermentation
Candida sake Lab-scale Tempranillo 2.4% v/v Increased sorbitol, low production of ethyl [62]
esters. No sensory analysis reported
Candida stellata Medium-scale Chardonnay 0.7% v/v Increased glycerol concentration, decreased [63]
‘floral’ and ‘fruity’, increased ‘ethyl acetate’
sensory descriptors
Candida stellata Pilot-scale Trebbiano 0.6% v/v Increased glycerol and succinic acid content, [64]
immobilized cells decreased esters and higher alcohols. No
sensory reported
Candida zemplinina Pilot-scale Riesling 0.8% v/v Decreased ‘purity’ and increased ‘oxidation’ [53]
aeration and ‘solvent’ sensory attributes
Hanseniaspora opuntiae Medium-scale Pinotage 0.6% v/v Increased ‘hazelnut’, ‘coffee’, ‘caramel’, [65]
‘cherry’ and ‘acetone’ sensory descriptors
Hanseniaspora uvarum Medium-scale Pinotage 0.8% v/v Increased ‘hazelnut’, ‘coffee’, ‘caramel’, [65]
‘cherry’ and ‘acetone’ sensory descriptors
Lachancea Lab-scale Tempranillo 1.2% v/v Reduced ‘aromatic intensity’ and ‘aromatic [51]
thermotolerans quality’, increased ‘herbaceous’ character
Lachancea Industrial-scale Sangiovese 0.7% v/v Increased ‘spicy’ and ‘acidity’ sensory [66]
thermotolerans attributes
Metschnikowia Lab-scale Chardonnay 0.9% v/v Increased concentration of esters and higher [67]
pulcherrima alcohols, decreased volatile acids content. No
sensory reported
Metschnikowia Lab-scale Shiraz 1.6% v/v Increased concentration of higher alcohols, [67]
pulcherrima decreased volatile acids content. No sensory
reported
Metschnikowia Lab-scale Verdicchio 1.4% v/v Increased geraniol and acetaldehyde content. [68]
pulcherrima immobilized cells No sensory reported
Metschnikowia Lab-scale Malvasia/Viura 2.2% v/v No volatile composition or sensory analysis [52]
pulcherrima aeration blend reported
Metschnikowia Pilot-scale Merlot 1.0% v/v Decreased ‘brown tint’ and increased ‘red fruit [39]
pulcherrima aroma’ sensory attributes
Metschnikowia Pilot-scale Riesling 3.8% v/v Decreased ‘purity’ and ‘fruity’, and increased [53]
pulcherrima aeration ‘vinegar’, ‘oxidation’ and ‘solvent’ sensory
attributes
Metschnikowia Pilot-scale Viura-Malvası́a 0.8% v/v Increased glycerol content, decreased ‘tropical [54]
pulcherrima Aeration fruit’ and ‘white fruits’, and increased
‘oxidation’ sensory attributes
Pichia guilliermodii Pilot-scale Riesling 2.0% v/v Decreased ‘purity’ and ‘fruity’, and increased [53]
aeration ‘vinegar’ and ‘reductive’ sensory attributes
Pichia kluyveri Pilot-scale Riesling 3.0% v/v Decreased ‘purity’ and ‘fruity’, and increased [53]
aeration ‘vinegar’, ‘oxidation’ and ‘solvent’ sensory
attributes
Schizosaccharomyces Medium-scale Airen 0.7% v/v Decreased malic acid concentration, increased [69]
pombe ‘bitterness’ and decreased ‘acidity’ sensory
attributes
Starmerella bacillaris Lab-scale Barbera 0.7% v/v Increased glycerol content. No volatile [70]
composition or sensory analysis reported
Starmerella bacillaris Pilot- scale Barbera 0.5% v/v Increased glycerol content. No volatile [70]
composition or sensory analysis reported
Starmerella bacillaris Pilot- scale Barbera 0.3% v/v Reduced concentration of higher alcohols. No [71]
sensory analysis reported
Starmerella bombicola Lab-scale Verdicchio 1.6% v/v Increased ethyl acetate and isoamyl acetate [68]
immobilized cells concentration. No sensory reported
Torulaspora delbrueckii Pilot-scale Airen 0.3% v/v Increased ‘aromatic intensity’ and ‘fruity’ [72]
sensory attributes
Torulaspora delbrueckii Pilot-scale Viura-Malvası́a 0.5% v/v Increased glycerol content, decreased ‘tropical [54]
Aeration fruit’ and ‘white fruits’, and increased
‘reduction’ and ‘dried fruits’ sensory attributes
a
Depending on the study non-Saccharomyces yeasts were used in single, sequential or simultaneous inoculation.
b
Ethanol concentration compared to control wines fermented with S. cerevisiae.
c
Chemical composition or sensory attributes compared to control wines fermented with S. cerevisiae.

Current Opinion in Biotechnology 2019, 56:88–96 www.sciencedirect.com


Microbiological strategies for reducing alcohol concentration Varela and Varela 93

Table 2

Summary of recent studies evaluating the use of non-Saccharomyces yeasts to decrease ethanol concentration in beer and the impact on
beer sensory properties

Non-Saccharomyces yeast Maltose Ethanol concentration Impact on chemical References


utilizationa composition or sensory attributes
Candida tropicalis – 0.2 % v/v Low content of aroma compounds [73]
Candida shehatae – <0.5 % v/v ‘Sweet’, no ‘wort-like flavors’ were Patent CN102220198B
identified
Pichia kluyveri – 0.1 % v/v ‘Fruity’. High concentrations of esters Patent US9580675B2
Saccharomycodes ludwigii – 0.5–1.4 % v/v High ester concentration, low diacetyl [1]
Torulaspora delbruenckii – 2.66 % v/v High levels of phenethyl acetate, ethyl [20]
hexanoate and
ethyl octaonate. ‘Fruity’, ‘Hoppy’
Torulaspora delbruenckii v 0.9–4.0 % v/v Slight ‘wort-like’ flavor. [74]
‘Honey’ and ‘pear-like’ flavors
Williopsis saturnus – 1.7% v/v High levels of terpenes and terpernoids [75]
Zygosaccharomyces rouxii – 0.9–3.3 % v/v High diacetyl levels [1]
a
Variable maltose utilization is indicated as ‘v’ and refers to the use of maltose positive and negative strains in the same study.

Figure 2

Engineered and adaptively Non-conventional Novel Sacchromyces


evolved novel S. cerevisiae yeast species hybrids

Screening for low-ethenol


producing strains
Ethanol concentration (% v/v)

16

12

0
Sc

Y1

Y2

Y3

Y4

Y5

Y6

Y7
LA

LA

LA

LA

LA

LA

LA

Assessing sensory
properties

Fruity Fruity

Worty Sweet Vegetal Floral

Beer sensory Wine sensory


analysis analysis
Current Opinion in Biotechnology

The process of selecting low-alcohol yeasts. Novel Saccharomyces cerevisiae strains generated by adaptive evolution or gene modification; non-
conventional Saccharomyces and non-Saccharomyces yeast species; and new Saccharomyces hybrids are screened for ethanol production in
wort or grape must. Strains showing reduced ethanol formation are then evaluated for beer or wine sensory profile.

www.sciencedirect.com Current Opinion in Biotechnology 2019, 56:88–96


94 Food biotechnology

in S. cerevisiae and alter the competitiveness of non- communities during fermentation, enhancing our under-
Saccharomyces yeasts [56,57]. standing of microbial ecology and population dynamics
[10,61]. Applying these metagenomics approaches not only
Several non-Saccharomyces species have been tested for the to samples taken during alcoholic fermentation, but also to
production of low-alcohol and alcohol-free beers (Table 2) environmental samples promises to increase our knowledge
[58]. The rationale was to use yeasts species that are on microbial diversity. In turn, systematic characterization of
unable to utilize maltose and/or maltotriose, the main this diversity could lead to the isolation or development of
sugars present in wort. Thus, ethanol production is limited novel ‘low-ethanol’ yeasts for the production of beer and
by the amount of other sugars, including glucose, fructose wine with reduced ethanol concentration.
and sucrose, which accounts for less than 20% of the
carbohydrates in wort. Beers produced this way contain Conflict of interest statement
high amounts of maltose, but this does not translate into a Nothing declared.
higher calorie content compared to standard beers. In fact,
since the calorific value of ethanol is higher than carbo- Acknowledgements
hydrates, low-alcohol beers tend to have a lower calorie The authors would like to thank Dr Paul Cambers for reviewing the
manuscript. The AWRI, a member of the Wine Innovation Cluster in
content [59]. Regardless, high amounts of maltose often Adelaide, is supported by Australia’s grape growers and winemakers through
result in higher susceptibility to microbial contamination their investment body Wine Australia with matching funds from the
and in unpleasant ‘wort-like’ flavors. The latter can be Australian Government. J.V works on the CHASSY project, which has
received funding from the European Union’s Horizon 2020 Framework
overcome by using yeasts that produce high amounts of Programme for Research and Innovation – Grant Agreement No. 720824.
esters that mask ‘wort-like’ flavors. An example of this is
the use of Saccharomycodes ludwigii to produce alcohol-free References and recommended reading
beer [1], which is now commercially available. Beer Papers of particular interest, published within the period of review,
have been highlighted as
produced with this yeast contains very low alcohol con-
centration (<0.1% v/v) and shows high amounts of esters  of special interest
that contribute to a ‘fruity’ character. Similarly, beer  of outstanding interest
produced with Pichia kluyveri contain reduced alcohol
1. De Francesco G, Turchetti B, Sileoni V, Marconi O, Perretti G:
concentration and increased concentration of esters (US  Screening of new strains of Saccharomycodes ludwigii and
Patent No. 9580675B2). Although all of the examples Zygosaccharomyces rouxii to produce low-alcohol beer. J Inst
Brewing 2015, 121:113-121.
listed in Table 2 have used single inoculation with This article describes the evaluation of non-conventional yeast
non-Saccharomyces species, sequential fermentations with strains for producing low-alcohol beer. Beer samples produced by
these strains were analysed for ethanol content and main volatile
S. cerevisiae have also been investigated. Beers produced profile
using this strategy contained a more complex flavor profile
2. MacAvoy MG: Wine — harmful or healthy? What is being
compared to single fermentations with S. cerevisiae and, in considered in Australia and New Zealand?. Fourteenth
some cases, lower alcohol content [60]. Australian Wine Industry Technical Conference. 2010:28-31:.
Edited by Inc AWITC.

Concluding remarks 3. Athes V, Pena y Lillo M, Bernard C, Perez-Correa R, Souchon I:


Comparison of experimental methods for measuring infinite
Several yeast-based strategies have been shown to be effec- dilution volatilities of aroma compounds in water/ethanol
tive in producing beer and/or wine with reduced ethanol mixtures. J Agric Food Chem 2004, 52:2021-2027.
concentration. Regardless of the approach, ‘low-ethanol’ 4. Guth H, Sies A: Flavour of wines: towards and understanding by
yeast strains should always be evaluated for sensory impacts reconstitution experiments and an analysis of ethanol’s effect on
odour activity of key compounds. Eleventh Australian Wine Industry
they have, ideally, at pilot-scale to determine their real Technical Conference. 2001:128-139:. Edited by Inc AWITC.
contribution to the flavor characteristics of the final product 5. Branyik T, Silva DP, Baszczynski M, Lehnert R, Silva J: A review of
(Figure 2). Reduction in alcohol content has been widely methods of low alcohol and alcohol-free beer production. J
studied in wine but production of low-alcohol and alcohol- Food Eng 2012, 108:493-506.
free beers has received less attention. This difference can be 6. Varela C, Dry PR, Kutyna DR, Francis IL, Henschke PA, Curtin CD,
 Chambers PJ: Strategies for reducing alcohol concentration in
explained by the different priorities of the wine and beer wine. Aust J Grape Wine Res 2015, 21:670-679.
industries. While the wine industry sees production of This review summarises research aimed at reducing wine alcohol con-
reduced-alcohol wine as a necessity, the beer industry seems centration, including viticultural practices, pre-fermentation and wine-
making practices, microbiological strategies, and post-fermentation
to focus mainly in producing beers with new sensory attri- approaches and processing technologies. It also covers the effect of
butes (e.g. new flavors). Nevertheless, the expansion of the these practices on wine flavour.
beer industry into new markets (e.g. beer production in 7. Varela C: The impact of non-Saccharomyces yeasts in the
 production of alcoholic beverages. Appl Microbiol Biotechnol
countries where alcohol consumption is banned) could moti- 2016, 100:9861-9874.
vate researchers to look for new strategies to reduce alcohol This review covers the impacts of non-Saccharomyces yeasts on volatile
content in beer. composition and sensory profile of beer, wine, spirits and other fermented
beverages.
8. Jolly NP, Varela C, Pretorius IS: Not your ordinary yeast: non-
Recent advances in next-generation sequencing techniques Saccharomyces yeasts in wine production uncovered. FEMS
have enabled to investigate at great depth microbial Yeast Res 2014, 14:215-237.

Current Opinion in Biotechnology 2019, 56:88–96 www.sciencedirect.com


Microbiological strategies for reducing alcohol concentration Varela and Varela 95

9. Baker E, Wang B, Bellora N, Peris D, Hulfachor AB, Koshalek JA, This review describes biological-based approaches that influence etha-
Adams M, Libkind D, Hittinger CT: The genome sequence of nol yield during fermentation, most of which have attempted to partially
Saccharomyces eubayanus and the domestication of lager- redirect yeast metabolism away from ethanol production.
brewing yeasts. Mol Biol Evol 2015, 32:2818-2831.
28. Selecký R, mogrovi9 cová D, Sulo P: Beer with reduced ethanol
10. Bokulich NA, Bamforth CW, Mills DA: Brewhouse-resident  content produced using Saccharomyces cerevisiae yeasts
microbiota are responsible for multi-stage fermentation of deficient in various tricarboxylic acid cycle enzymes. J Inst
American coolship ale. PLoS One 2012, 7 e35507. Brewing 2008, 114:97-101.
This article examines the potential of yeast strains lacking genes encod-
11. Spitaels F, Wieme AD, Janssens M, Aerts M, Daniel HM, Van ing for enzymes of the TCA cycle to produce low-alcohol beer. Strains
Landschoot A, De Vuyst L, Vandamme P: The microbial diversity deficient in fumarase and a-ketoglutarate dehydrogenase made non-
of traditional spontaneously fermented lambic beer. PLoS One alcoholic beers with a reduced alcohol concentration. Low ethanol con-
2014, 9. tent was compensated by the considerable increase of organic acids.
12. Varela C, Borneman AR: Yeasts of vineyards and wineries. Yeast 29. Cuello RA, Montero KJF, Mercado LA, Combina M, Ciklic IF:
2017, 34:111-128. Construction of low-ethanol-wine yeasts through partial
deletion of the Saccharomyces cerevisiae PDC2 gene. AMB
13. Morrison-Whittle P, Goddard MR: Quantifying the relative roles Expr 2017, 7.
of selective and neutral processes in defining eukaryotic
microbial communities. ISME J 2015, 9:2003-2011. 30. Ishii T, Araki M: Consumer acceptance of food crops developed
by genome editing. Plant Cell Rep 2016, 35:1507-1518.
14. Taylor MW, Tsai P, Anfang N, Ross HA, Goddard MR:
Pyrosequencing reveals regional differences in fruit- 31. Stovicek V, Holkenbrink C, Borodina I: CRISPR/cas system for
associated fungal communities. Environ Microbiol 2014, yeast genome engineering: advances and applications. FEMS
16:2848-2858. Yeast Res 2017, 17:16.
15. Bokulich NA, Ohta M, Richardson PM, Mills DA: Monitoring 32. Waks Z, Silver PA: Engineering a synthetic dual-organism
seasonal changes in winery-resident microbiota. PLoS One system for hydrogen production. Appl Environ Microbiol 2009,
2013, 8 e66437. 75:1867-1875.
16. Grangeteau C, Gerhards D, von Wallbrunn C, Alexandre N, 33. Nandi R, Sengupta S: Microbial production of hydrogen: an
Rousseaux S, Guilloux-Benatier M: Persistence of two non- overview. Crit Rev Microbiol 1998, 24:61-84.
Saccharomyces yeasts (Hanseniaspora and Starmerella) in
the cellar. Front Microbiol 2016, 7:11. 34. Denby CM, Li RA, Vu VT, Costello Z, Lin W, Chan LJG, Williams J,
Donaldson B, Bamforth CW, Petzold CJ et al.: Industrial brewing
17. Ocon E, Gutierrez AR, Garijo P, Lopez R, Santamaria P: Presence yeast engineered for the production of primary flavor
of non-Saccharomyces yeasts in cellar equipment and grape determinants in hopped beer. Nat Commun 2018, 9:965.
juice during harvest time. Food Microbiol 2010, 27:1023-1027.
35. Cadiere A, Ortiz-Julien A, Camarasa C, Dequin S: Evolutionary
18. Pérez-Martı́n F, Seseña S, Fernández-González M, Arévalo M, engineered Saccharomyces cerevisiae wine yeast strains with
Palop ML: Microbial communities in air and wine of a winery at increased in vivo flux through the pentose phosphate
two consecutive vintages. Int J Food Microbiol 2014, 190:44-53. pathway. Metab Eng 2011, 13:263-271.
19. Sabate J, Cano J, Esteve-Zarzoso B, Guillamon JM: Isolation and 36. Kutyna DR, Varela C, Stanley GA, Borneman AR, Henschke PA,
identification of yeasts associated with vineyard and winery by Chambers PJ: Adaptive evolution of Saccharomyces cerevisiae
RFLP analysis of ribosomal genes and mitochondrial DNA. to generate strains with enhanced glycerol production. Appl
Microbiol Res 2002, 157:267-274. Microbiol Biotechnol 2012, 93:1175-1184.
20. Canonico L, Agarbati A, Comitini F, Ciani M: Torulaspora 37. Tilloy V, Ortiz-Julien A, Dequin S: Reduction of ethanol yield and
delbrueckii in the brewing process: a new approach to improvement of glycerol formation by adaptive evolution of
enhance bioflavour and to reduce ethanol content. Food the wine yeast Saccharomyces cerevisiae under hyperosmotic
Microbiol 2016, 56:45-51. conditions. Appl Environ Microbiol 2014, 80:2623-2632.
21. Chambers PJ, Borneman AR, Varela C, Cordente AG, Bellon JR, 38. Muratore G, Asmundo CN, Lanza CM, Caggia C, Licciardello F,
Tran TMT, Henschke PA, Curtin CD: Ongoing domestication of Restuccia C: Influence of Saccharomyces uvarum on volatile
wine yeast: past, present and future. Aust J Grape Wine Res acidity, aromatic and sensory profile of Malvasia delle Lipari
2015, 21:642-650. wine. Food Technol Biotechnol 2007, 45:101-106.
22. Ciani M, Comitini F, Mannazzu I, Domizio P: Controlled mixed 39. Varela C, Barker A, Tran T, Borneman A, Curtin C: Sensory profile
culture fermentation: a new perspective on the use of non- and volatile aroma composition of reduced alcohol Merlot
Saccharomyces yeasts in winemaking. FEMS Yeast Res 2010, wines fermented with Metschnikowia pulcherrima and
10:123-133. Saccharomyces uvarum. Int J Food Microbiol 2017, 252:1-9.
23. Contreras A, Hidalgo C, Schmidt S, Henschke PA, Curtin C, 40. Perez-Torrado R, Barrio E, Querol A: Alternative yeasts for
Varela C: The application of non-Saccharomyces yeast in winemaking: Saccharomyces non-cerevisiae and its hybrids.
fermentations with limited aeration as a strategy for the Crit Rev Food Sci Nutr 2017:1-11.
production of wine with reduced alcohol content. Int J Food
Microbiol 2015, 205:15. 41. Arroyo-Lopez FN, Perez-Torrado R, Querol A, Barrio E:
Modulation of the glycerol and ethanol syntheses in the yeast
24. Quiros M, Rojas V, Gonzalez R, Morales P: Selection of non- Saccharomyces kudriavzevii differs from that exhibited by
Saccharomyces yeast strains for reducing alcohol levels in Saccharomyces cerevisiae and their hybrid. Food Microbiol
wine by sugar respiration. Int J Food Microbiol 2014, 181:85-91. 2010, 27:628-637.
25. Palacios A, Raginel F, Ortiz-Julien A: Can the selection of 42. Alonso-del-Real J, Contreras-Ruiz A, Castiglioni GL, Barrio E,
Saccharomyces cerevisiae yeast lead to variations in the final Querol A: The use of mixed populations of Saccharomyces
alcohol degree of wines? Aust N Z Grapegrow Winemak 2007, cerevisiae and S. kudriavzevii to reduce ethanol content in
527:71-75. wine: limited aeration, inoculum proportions, and sequential
inoculation. Front Microbiol 2017, 8.
26. Varela C, Kutyna D, Henschke PA, Chambers PJ, Herderich MJ,
Pretorius IS: Taking control of alcohol. Aust N Z Wine Ind J 2008, 43. Bellon JR, Eglinton JM, Siebert TE, Pollnitz AP, Rose L, de Barros
23:41-43. Lopes M, Chambers PJ: Newly generated interspecific wine
yeast hybrids introduce flavour and aroma diversity to wines.
27. Kutyna DR, Varela C, Henschke PA, Chambers PJ, Stanley GA: Appl Microbiol Biotechnol 2011, 91:603-612.
 Microbiological approaches to lowering ethanol
concentration in wine. Trends Food Sci Technol 2010, 21:293- 44. Bellon JR, Schmid F, Capone DL, Dunn BL, Chambers PJ:
302. Introducing a new breed of wine yeast: interspecific

www.sciencedirect.com Current Opinion in Biotechnology 2019, 56:88–96


96 Food biotechnology

hybridisation between a commercial Saccharomyces This article reviews reported brewing trials which have used pure non-
cerevisiae wine yeast and Saccharomyces mikatae. PLoS One Saccharomyces starter cultures to produce normal gravity beers, low-
2013, 8 e62053. alcohol and alcohol-free beers. The article focuses on the production of
secondary metabolites and the potential use of non-Saccharomyces
45. Bellon JR, Yang F, Day MP, Inglis DL, Chambers PJ: Designing species in breweries.
and creating Saccharomyces interspecific hybrids for
improved, industry relevant, phenotypes. Appl Microbiol 59. Bamforth CW: Beer, carbohydrates and diet. J Inst Brewing
Biotechnol 2015, 99:8597-8609. 2005, 111:259-264.
46. Bizaj E, Cordente AG, Bellon JR, Raspor P, Curtin CD, Pretorius IS: 60. Holt S, Mukherjee V, Lievens B, Verstrepen KJ, Thevelein JM:
A breeding strategy to harness flavor diversity of Bioflavoring by non-conventional yeasts in sequential beer
Saccharomyces interspecific hybrids and minimize hydrogen fermentations. Food Microbiol 2018, 72:55-66.
sulfide production. Fems Yeast Res 2012, 12:456-465.
61. Bokulich NA, Mills DA: Microbial terroir of wine: deep insight
47. Krogerus K, Magalhães F, Vidgren V, Gibson B: Novel brewing into site-specific winery and vineyard microbial communities.
 yeast hybrids: creation and application. Appl Microbiol Am J Enol Viticult 2012, 63:435a.
Biotechnol 2017, 101:65-78.
This article describes the use of hybridisation as a strain development tool 62. Ballester-Tomas L, Prieto JA, Gil JV, Baeza M, Randez-Gil F: The
for brewery applications. Creation and application of yeast hybrids for Antarctic yeast Candida sake: understanding cold metabolism
beer productions is also discussed. impact on wine. Int J Food Microbiol 2017, 245:59-65.

48. Mertens S, Steensels J, Saels V, De Rouck G, Aerts G, 63. Soden A, Francis IL, Oakey H, Henschke PA: Effects of co-
 Verstrepen KJ: A large set of newly created interspecific fermentation with Candida stellata and Saccharomyces
Saccharomyces hybrids increases aromatic diversity in lager cerevisiae on the aroma and composition of Chardonnay wine.
beers. Appl Environ Microbiol 2015, 81:8202-8214. Aust J Grape Wine Res 2000, 6:21-30.
This article evaluates the potential of novel yeast hybrids for the produc- 64. Ferraro L, Fatichenti F, Ciani M: Pilot scale vinification process
tion of beer. S. cerevisiae x S. eubayanus hybrids showed broad temn- using immobilized Candida stellata cells and Saccharomyces
perature tolerance and desirable flavour profile, with one hybrid perform- cerevisiae. Process Biochem 2000, 35:1125-1129.
ing succesfully in pilot-scale trials.
65. Rossouw D, Bauer FF: Exploring the phenotypic space of non-
49. Gschaedler A: Contribution of non-conventional yeasts in Saccharomyces wine yeast biodiversity. Food Microbiol 2016,
alcoholic beverages. Curr Opin Food Sci 2017, 13:73-77. 55:32-46.
50. Ciani M, Morales P, Comitini F, Tronchoni J, Canonico L, Curiel JA, 66. Gobbi M, Comitini F, Domizio P, Romani C, Lencioni L, Mannazzu I,
 Oro L, Rodrigues AJ, Gonzalez R: Non-conventional yeast Ciani M: Lachancea thermotolerans and Saccharomyces
species for lowering ethanol content of wines. Front Microbiol cerevisiae in simultaneous and sequential co-fermentation: a
2016, 7. strategy to enhance acidity and improve the overall quality of
This review describes the use of non-Saccharomyces yeasts to produce wine. Food Microbiol 2013, 33:271-281.
wines with lower alcohol content than those from pure Saccharomyces
starters. It also discusses the selection of suitable yeast strains, identi- 67. Contreras A, Hidalgo C, Henschke PA, Chambers PJ, Curtin C,
fication of key environmental parameters influencing ethanol yield, and Varela C: Evaluation of non-Saccharomyces yeasts for the
managing mixed fermentations. reduction of alcohol content in wine. Appl Environ Microbiol
2014, 80:1670-1678.
51. Del Fresno JM, Morata A, Loira I, Banuelos MA, Escott C, Benito S,
Chamorro CG, Suarez-Lepe JA: Use of non-Saccharomyces in 68. Canonico L, Comitini F, Oro L, Ciani M: Sequential
single-culture, mixed and sequential fermentation to improve fermentation with selected immobilized non-
red wine quality. Eur Food Res and Technol 2017, 243:2175- Saccharomyces yeast for reduction of ethanol content in
2185. wine. Front Microbiol 2016, 7.
52. Morales P, Rojas V, Quiros M, Gonzalez R: The impact of oxygen 69. Benito S, Palomero F, Morata A, Calderon F, Palmero D, Suarez-
on the final alcohol content of wine fermented by a mixed Lepe JA: Physiological features of Schizosaccharomyces
starter culture. Appl Microbiol Biotechnol 2015, 99:3993-4003. pombe of interest in making of white wines. Eur Food Res
Technol 2013, 236:29-36.
53. Rocker J, Strub S, Ebert K, Grossmann M: Usage of different
aerobic non-Saccharomyces yeasts and experimental 70. Englezos V, Rantsiou K, Cravero F, Torchio F, Ortiz-Julien A,
conditions as a tool for reducing the potential ethanol content Gerbi V, Rolle L, Cocolin L: Starmerella bacillaris and
in wines. Eur Food Res Technol 2016, 242:2051-2070. Saccharomyces cerevisiae mixed fermentations to reduce
ethanol content in wine. Appl Microbiol Biotechnol 2016,
54. Tronchoni J, Curiel JA, Saenz-Navajas MP, Morales P, de-la- 100:5515-5526.
Fuente-Blanco A, Fernandez-Zurbano P, Ferreira V, Gonzalez R:
Aroma profiling of an aerated fermentation of natural grape 71. Rolle L, Englezos V, Torchio F, Cravero F, Segade SR,
must with selected yeast strains at pilot scale. Food Microbiol Rantsiou K, Giacosa S, Gambuti A, Gerbi V, Cocolin L: Alcohol
2018, 70:214-223. reduction in red wines by technological and microbiological
approaches: a comparative study. Aust J Grape Wine Res
55. Shekhawat K, Bauer FF, Setati ME: Impact of oxygenation on the 2018, 24:62-74.
performance of three non-Saccharomyces yeasts in co-
fermentation with Saccharomyces cerevisiae. Appl Microbiol 72. Izquierdo Canas PM, Palacios Garcia AT, Garcia Romero E:
Biotechnol 2017, 101:2479-2491. Enhancement of flavour properties in wines using sequential
inoculations of non-Saccharomyces (Hansenula and
56. Curiel JA, Morales P, Gonzalez R, Tronchoni J: Different non- Torulaspora) and Saccharomyces yeast starter. Vitis 2011,
Saccharomyces yeast species stimulate nutrient 50:177-182.
consumption in S. cerevisiae mixed cultures. Front Microbiol
2017, 8. 73. N’Guessan FK, N’Dri DY, Camara F, Djè MK: Saccharomyces
cerevisiae and Candida tropicalis as starter cultures for the
57. Rollero S, Bloem A, Ortiz-Julien A, Camarasa C, Divol B: Altered alcoholic fermentation of tchapalo, a traditional sorghum
fermentation performances, growth, and metabolic footprints beer. World J Microbiol Biotechnol 2010, 26:693-699.
reveal competition for nutrients between yeast species
inoculated in synthetic grape juice-like medium. Front 74. Michel M, Kopecká J, Meier-Dörnberg T, Zarnkow M, Jacob F,
Microbiol 2018, 9. Hutzler M: Screening for new brewing yeasts in the non-
Saccharomyces sector with Torulaspora delbrueckii as model.
58. Michel M, Meier-Dörnberg T, Jacob F, Methner FJ, Wagner RS, Yeast 2016, 33:129-144.
 Hutzler M: Review: pure non-Saccharomyces starter cultures
for beer fermentation with a focus on secondary metabolites 75. Liu S-Q, Quek AYH: Evaluation of beer fermentation with a novel
and practical applications. J Inst Brewing 2016, 122:569-587. yeast Williopsis saturnus. Food Technol Biotechnol 2016, 54:403-412.

Current Opinion in Biotechnology 2019, 56:88–96 www.sciencedirect.com

Anda mungkin juga menyukai