You are on page 1of 16

Environmental Microbiology (2017) 19(3), 878–893 doi:10.1111/1462-2920.

13617

Minireview

Glycerol metabolism and transport in yeast and fungi:


established knowledge and ambiguities

Mathias Klein,1 Steve Swinnen,2 S. cerevisiae laboratory strains. This mini-review


Johan M. Thevelein,2,3,4 Elke Nevoigt1* summarizes what is known so far about the gene
1
Department of Life Sciences and Chemistry, Jacobs products and pathways involved in glycerol
University Bremen gGmbH, Campus Ring 1, Bremen, metabolism and transport in yeast and fungi as well
28759, Germany. as the regulation of these processes.
2
GlobalYeast NV, Kasteelpark Arenberg 31, Leuven-
Heverlee, 3001, Belgium.
3
Laboratory of Molecular Cell Biology, Institute of Botany Introduction
and Microbiology, KU Leuven, Leuven, Belgium.
4
Glycerol is an organic compound that is ubiquitous in
Department of Molecular Microbiology, VIB,
nature. It is therefore not surprising that organisms have
Kasteelpark Arenberg 31, 3001 Heverlee-Leuven,
developed ways to use this substance as a source of
Flanders, Belgium.
carbon and energy. Although it cannot be excluded that
trace amounts of glycerol in the environment could be of
geochemical origin (John Hallsworth, Personal Commu-
nication), glycerol from biotic origin likely accounts for
the largest fraction as it forms the backbone of phospho-
Summary lipids and triacylglycerols, which are common constitu-
ents of cell membranes and widespread storage lipids
There is huge variability among yeasts with regard to
respectively. Glycerol can be released from these con-
their efficiency in utilizing glycerol as the sole source
stituents into the environment through enzymatic degra-
of carbon and energy. Certain species show growth
dation by the action of microbial lipases. In modern
rates with glycerol comparable to those reached with
human society, glycerol is also generated in industrial
glucose as carbon source; others are virtually unable
processes such as the saponification and transesterifica-
to utilize glycerol, especially in synthetic medium.
tion of animal and vegetable fats and oils in soap and
Most of our current knowledge regarding glycerol
biodiesel production respectively. In fact, huge amounts
uptake and catabolic pathways has been gained from
of glycerol have been generated in recent years due to
studying laboratory strains of the model yeast
the immense growth of the biodiesel industry (Mattam
Saccharomyces cerevisiae. The growth of these
et al., 2013).
strains on glycerol is dependent on the presence of
Another biotic process of glycerol generation is its
medium supplements such as amino acids and
metabolic formation and excretion by certain (micro)or-
nucleobases. In contrast, there is only fragmentary
ganisms. Among them, mainly yeasts have been
knowledge about S. cerevisiae isolates able to grow
reported to produce relatively high quantities of glycerol
in synthetic glycerol medium without such
and, in the past, certain species have been used for the
supplements as well as about growth of non-
biotechnological production of glycerol (Wang et al.,
Saccharomyces yeast species on glycerol. Thus,
2001). The main reason why cells produce and accumu-
more research is required to understand why certain
late glycerol intracellularly is its protective properties
strains and species show superior growth
against stress, particularly hyperosmotic and thermal
performance on glycerol compared with common
stress. Moreover, the electron-dependent formation of
glycerol from the central metabolic intermediate dihy-
Received 6 September, 2016; accepted 16 November, 2016. *For
correspondence. E-mail e.nevoigt@jacobs-university.de; Tel. 149- droxyacetone phosphate (DHAP) might also serve as a
421-2003541; Fax 149-421-2003249. redox sink when regeneration of cytosolic NAD1 from
C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd.
V
This is an open access article under the terms of the Creative Commons Attribution-NonCommercial-NoDerivs License, which permits
use and distribution in any medium, provided the original work is properly cited, the use is non-commercial and no modifications or
adaptations are made.
Glycerol catabolism in yeast 879
NADH through electron transport to oxygen via the glycerol encouraged us to screen 52 different isolates
respiratory chain is insufficient. These two functions of including commonly used laboratory and industrial
glycerol formation have been very well studied in the strains as well as natural isolates in synthetic glycerol
yeast S. cerevisiae, and summarized in several reviews medium, thereby explicitly avoiding the addition of any of
(Nevoigt and Stahl, 1997; Bakker et al., 2001; Hohmann, the above-mentioned supplements (Swinnen et al.,
2002; 2015). Notably, the intracellular glycerol concen- 2013). The observed maximum specific growth rates
tration in certain fungi and algae can reach extremely (lmax) and lag phases (after shift from glucose to glycer-
high levels, up to 7–8 M. Such a high intracellular glyc- ol) confirmed the significant intra-species diversity with
erol concentration is a determinant of vigour and/or viru- regard to glycerol utilization. The strain with the fastest
lence of insect- and plant-pathogenic fungi (Hallsworth biomass accumulation (CBS 6412, combining a lmax of
and Magan, 1995; De Jong et al., 1997). In addition, the  0.10 h21 and a relatively short lag phase of  27 h)
insect haemolymph, in which entomopathogenic fungi was selected for thorough genetic analysis. A haploid
proliferate, can also contain glycerol at molar concentra- meiotic segregant of this isolate (CBS 6412-13A) with a
tions (Sformo et al., 2010). It has been demonstrated lmax of  0.13 h21 was used to unravel the (poly)genic
that such high glycerol concentrations might act as a basis of its better glycerol growth capacity in comparison
stress factor in yeasts and fungi, particularly at concen- with the glycerol-growth defective, well-established labo-
trations higher than 3–4 M (Cray et al., 2015; Stevenson ratory S. cerevisiae strain CEN.PK113-1A by genetic
et al., 2016). mapping (Swinnen et al., 2016). Three causative genetic
The ability to catabolize environmentally available determinants (UBR2CBS, SSK1CBS and GUT1CBS) were
glycerol as a carbon and energy source is widespread identified. Among these three genes, allele swapping of
from archeabacteria (Falb et al., 2008) to human cells UBR2 in the genetic background of CEN.PK113-1A had
(Hibuse et al., 2006). This minireview deals with glycerol the largest impact, establishing growth on glycerol in this
metabolism in fungal microorganisms with a particular strain with a lmax of  0.05 h21. Swapping of all three
focus on the catabolism of glycerol in yeasts. Our inter- superior alleles in the strain CEN.PK113-1A resulted in
est in this topic has been driven mainly by the goal of a lmax of  0.08 h21. While the function of GUT1 with
exploiting glycerol as a carbon source in biotechnologi- regard to glycerol utilization is obvious (GUT1 encodes
cal applications, in particular to valorize the large excess a glycerol kinase), the relevance of UBR2 and SSK1
of glycerol from biodiesel production. However, there are remains unknown. UBR2 encodes a ubiquitin ligase that
many experimental findings worth to be investigated in functions in the ubiquitin proteasome system of S. cere-
more detail from a fundamental point of view, particularly visiae (Finley et al., 2012). SSK1 encodes a protein that
with regard to the molecular basis of metabolic diversity. is part of a two-component system that senses hyperos-
motic stress and transduces the signal to the Hog1
Glycerol utilization by yeasts: high inter- and intra- MAPK cascade to increase intracellular glycerol levels
species diversity (Maeda et al., 1994). Further studies are necessary to
unravel the action mechanism of the superior alleles in
It is clear from the literature that different yeast species,
improving growth on glycerol.
as well as different strains from the same species, show Notably, two independent studies have reported that
high diversity with regard to growth on glycerol as the wild-type S. cerevisiae strains not able to grow in syn-
sole source of carbon (Kurtzman et al., 2011). However, thetic glycerol medium, can be evolved relatively easily
some of the reported differences might be caused by by repeated batch cultivations to obtain growth rates on
the growth medium used, instead of strain differences, glycerol of about 0.2 h21 (Ochoa-Estopier et al., 2010;
as already elaborated by Vasiliadis et al. (1987). In fact, Merico et al., 2011). These results underscore the inher-
it has been shown (at least for S. cerevisiae) that the ent potential of S. cerevisiae to utilize glycerol more effi-
addition of amino acids, nucleobases or complex supple- ciently. The exact mutations underlying the improved
ments such as peptone or yeast extract positively affects growth capacity of the evolved strains have recently
the capacity to grow on glycerol (Merico et al., 2011; been identified (Ho et al.; submitted). The results con-
Swinnen et al., 2013). Indeed, almost all remaining pre- firmed our above-mentioned finding that UBR2 and
vious studies with S. cerevisiae regarding growth on GUT1 are the most important targets for improving the
glycerol were conducted with supplemented media. As capacity of the strain CEN.PK113-1A for growth on glyc-
the laboratory strains used were for the most part auxo- erol. The latter study identified GUT1 alleles (when
trophic, the supplementation with amino acids and expressed in CEN.PK113-1A and combined with UBR2
nucleobases could not be circumvented. The ambiguity allele swapping) that resulted in an even higher growth
of the previously published data regarding the quantita- rate than the GUT1 allele previously identified in the
tive performance of S. cerevisiae strains for growth on strain CBS 6412-13A.
C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
880 M. Klein et al.
Literature surveys for studies on glycerol utilization in and Greig (2015). Still, when S. cerevisiae gets access
different yeast species have revealed that there are not to sugar-rich environments, it will overgrow other com-
much comparative data available. To the best of our peting microorganisms due to the fact that it liberates
knowledge, only one comprehensive study of 42 yeast the energy from sugar much faster via its fermentative
species in synthetic glycerol medium has been per- metabolism caused by the strong-Crabtree effect
formed (Lages et al., 1999). In this work, the yeasts (Pfeiffer et al., 2001). In addition, the rapid accumulation
Cyberlindnera jadinii (formerly known as Candida utilis of ethanol inhibits the growth of other microorganisms
and later as Pichia jadinii) and Pichia anomala showed present in the same niche. Hence, yeast species with a
the highest lmax on glycerol of 0.32 and 0.29 h21 strong Crabtree effect, such as S. cerevisiae, may not
respectively. In contrast, the S. cerevisiae strain ana- have developed or may have lost during evolution the
lysed by these authors showed a lmax of 0.11 h21. capability of efficiently utilizing glycerol. In fact, potential
Unfortunately, this study did not include yeast species competitors for the consumption of the fermentation
that are favoured in current industrial applications products ethanol and glycerol during the post-diauxic
employing glycerol-based cultivation media such as Yar- growth phase have already been outcompeted during
rowia (Candida) lipolytica, Komagataella (Pichia) pasto- the sugar-consuming phase. This supports the so-called
ris and Pachysolen tannophilus. For these three make-accumulate-consume strategy of S. cerevisiae
species, growth rates on glycerol in the range of 0.26– when growing in sugar-rich environments (Piskur et al.,
0.30 h21 have been reported in the literature (Papaniko- 2006).
laou et al., 2002; Mattanovich et al., 2009; Liu et al.,
2012). In order to directly compare these yeast species to Glycerol catabolic pathways
each other and to S. cerevisiae, we recently performed a
growth analysis in synthetic medium containing 6% Similar to bacteria, two pathways for the dissimilation of
(v/v) glycerol. One representative strain of Y. lipolytica, glycerol have been reported in yeasts (Fig. 1A). The
C. jadinii, P. tannophilus and K. pastoris was tested, and phosphorylative glycerol catabolic pathway, in which
the values for lmax observed under these conditions were L-glycerol 3-phosphate (G3P) is formed as the interme-
0.50, 0.42, 0.27 and 0.20 h21 respectively (Klein et al., diate, is widespread among fungal microorganisms (Fig.
2016b). Under the same conditions, the best performing 1Ai). This pathway, referred to here as the ‘catabolic
wild-type isolates of the species S. cerevisiae showed G3P pathway’, involves a glycerol kinase (GK) and an
growth rates on glycerol of up to 0.15 h21 (Swinnen et al., FAD-dependent glycerol 3-phosphate dehydrogenase
2013). located at the outer surface of the inner mitochondrial
One could assume that the ability of efficient glycerol membrane (FAD-dependent mG3PDH). The latter
utilization by certain yeast species is connected to their enzyme directly transfers the electrons via FADH2 to the
prevalence in habitats where significant quantities of respiratory chain. The second glycerol catabolic path-
glycerol are found. Such a connection is obvious for ole- way, referred to here as the ‘catabolic DHA pathway’,
aginous yeasts such as Y. lipolytica which thrive in envi- starts with the oxidation of glycerol to dihydroxyacetone
ronments rich in lipids. However, up to now there has (DHA) via an NAD1-dependent glycerol dehydrogenase
not been any dedicated work investigating this putative (GDH; EC 1.1.1.6; Table 1). DHA is subsequently phos-
correlation in a comprehensive way for yeast and fungal phorylated to DHAP via DHA kinase (DAK) (Fig. 1Aii).
species. In addition to its release in fat and lipid break- In order to understand the history of allocating a par-
down, glycerol can be produced in significant amounts ticular glycerol catabolic pathway to a certain yeast
by certain yeasts and fungi for osmoregulation and species or filamentous fungus, it has to be noted that
redox balancing, as already mentioned in the introduc- in early studies conclusions were solely based on the
tion. Glycerol of such microbial origin can, in theory, be measured in vitro activities of the pathways’ key
re-utilized by the glycerol-producer itself or, after excre- enzymes. The most comprehensive investigations have
tion, by other microorganisms present in the respective been delivered by Babel and Hoffmann (1982) and Tani
niche. We will discuss anabolism and catabolism of glyc- and Yamada (1987). Based on measuring GK as an
erol in S. cerevisiae in this context. It is well known that indicator for the G3P pathway and GDH (NAD1-depen-
S. cerevisiae outperforms any other microorganism in dent) for the DHA pathway, Tani and Yamada (1987)
sugar-rich media. However, such environments rich in divided the yeast species investigated into three
sugar, as found for instance in damaged fruits, are rare groups. The first group was assumed to exclusively
in nature and, moreover, not available all year in non- use the G3P pathway (e.g., Candida boidinii), the sec-
tropical climates. The question on the identity of the nat- ond group only the DHA pathway (e.g., Hansenula ofu-
ural niches of S. cerevisiae is still occupying researchers naensis), and the third group both pathways (e.g.,
intensively, as comprehensively summarized by Goddard Candida valida). Although these early conclusions may
C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
Glycerol catabolism in yeast 881

Fig. 1. Proposed pathways for glycerol catabolism (A) and anabolism (B) in yeasts via (i) L-glycerol 3-phosphate (G3P), (ii) dihydroxyacetone
(DHA) and (iii) glyceraldehyde (GA) as intermediate respectively. When known, the names of the S. cerevisiae genes allocated to the respec-
tive enzyme activities are indicated in italics. The asterisk (*) indicates those enzymes whose involvement in the pathway has been confirmed
by analysing deletion mutants of the respective genes in at least one fungal organism. Abbreviation: glycerolint, intracellular glycerol.

C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
V
882 M. Klein et al.

Table 1. Glycerol dehydrogenase activities detected in yeast and fungi

EC number Reaction pH used for in vitro Organism and gene References Experimental basis
according to predominantly enzyme assay name (if known) for the enzyme’s
BRENDA catalyzed physiological role

1.1.1.6 DHA 1 NADH 7.5–10.0 O. parapolymorpha Yamada-Onodera S. pombe gld1 deletion strain does
$ (glycerol oxidation) DL-1 gdh et al. (2002) not consume glycerol in the
Glycerol 1 NAD1 6.0 H. ofunaensis Yamada-Onodera presence of other carbon sources
(DHA reduction) S. pombe gld1 et al. (2006) (Matsuzawa et al., 2010)
Matsuzawa
et al. (2010)

1.1.1.156 DHA 1 NADPH 9.0–10.0 S. pombe Marshall et al. (1989) A. nidulans gldB
$ (glycerol oxidation) A. niger Schuurink et al. (1990) disruptant showed strongly
Glycerol 1 NADP1 6.0–7.0 A. nidulans gldB De Vries et al. (2003) reduced glycerol accumulation
(DHA reduction) A. oryzae Ruijter et al. (2004) during osmostress (De Vries
H. jecorina Liepins et al. (2006) et al., 2003)
(T. reesii) gld2

1.1.1.372/ D/L-Glyceraldehyde 9.5–10.0 N. crassa Viswanath-Reddy Assumed function in galacturonic


1.1.1.72* 1 NADPH Glycerol oxidation H. jecorina et al. (1977) acid catabolism due to transcrip-
$ The activity in the oxidizing reaction (T. reesii) gld1 Liepins et al. (2006) tional activation of A. niger gaaD
Glycerol 1 NADP1 with glycerol as substrate was A. niger gaaD Martens-Uzunova and by galacturonic acid (Martens-
Recommended name: below the detection limit (Liepins Schaap (2009) Uzunova and Schaap, 2009)
D/L-glyceraldehyde et al., 2006).
reductase 5.6–7.0
D/L-glyceraldehyde reduction

*A rationale behind having two different entries regarding this enzyme activity in BRENDA (www.brenda-enzymes.org) has not become obvious to the authors of this minireview.

Environmental Microbiology, 19, 878–893


C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
Glycerol catabolism in yeast 883
correctly reflect the glycerol catabolic pathways that media containing glycerol as the sole carbon source (at
are active in vivo, the sole measurement of a single least in the medium and conditions used) and glycerol
enzymatic activity (or the presence of a homologous assimilation only occurred in the presence of other car-
gene) is, in principle, not sufficient to assess the func- bon sources, such as galactose and ethanol. In spite of
tionality of a metabolic pathway. More comprehensive this, the authors identified the gene gld1 whose product
evidence can be delivered by showing that the enzy- is homologous to bacterial GDH enzymes and whose
matic activity as well as the metabolic activity of glycer- deletion caused a reduction in NAD1-dependent GDH
ol production or the capacity to grow on glycerol is activity and prevented glycerol assimilation.
reduced, abolished or improved, by mutating, deleting An additional pathway proposed for glycerol catabo-
or overexpressing, respectively, specific pathway lism with D-glyceraldehyde (GA) as the intermediate
gene(s). In general, studies on the pathways of glycerol (here referred to as the ‘catabolic GA pathway’) is
catabolism using mutant strains have been conducted shown in Fig. 1Aiii. In the proposed pathway, glycerol is
extensively in the model yeast S. cerevisiae, but are first oxidized to GA by an NADP1-dependent GDH (EC
rare in other yeast species or filamentous fungi. 1.1.1.372 or 1.1.1.72; Table 1). Subsequently, two differ-
Such studies with regard to the G3P pathway showed ent pathways have been proposed through which GA
that the deletion of either GUT1 (encoding GK) or GUT2 could be channeled into central carbon metabolism
(encoding FAD-dependent mG3PDH) in different S. cer- (Fig. 1Aiii). In the first pathway, GA is phosphorylated to
evisiae strains led to the complete abolishment of glyceraldehyde 3-phosphate by a glyceraldehyde kinase,
growth on glycerol (Sprague and Cronan, 1977; Pavlik while in the second pathway, GA is converted into
et al., 1993; Swinnen et al., 2013), strongly suggesting D-glycerate via the action of an aldehyde dehydroge-
that this pathway is the main route for glycerol utilization
nase. D-glycerate can subsequently be phosphorylated
in S. cerevisiae. These experimental findings confirmed
by a glycerate 3-kinase. The existence of the GA path-
(at least for S. cerevisiae) the assumption of Gancedo
way (including the two sub-pathways downstream of
et al. (1968) that the G3P pathway is used in C. utilis
GA) in the filamentous fungus N. crassa has been sug-
(nowadays classified as C. jadinii) and S. cerevisiae
gested by Tom et al. (1978), who determined the activi-
based on enzyme activity measurements in vitro.
ties of eight enzymes possibly involved in glycerol
Several authors have stated that S. cerevisiae might
metabolism. In crude cell extracts of a wild-type strain
also use the DHA pathway (Fig. 1Aii) for glycerol catab-
and two mutant strains with improved growth on glycer-
olism. The possibility for a catabolic DHA pathway in
ol, a low but significant level of glyceraldehyde kinase
this species has been raised by the discovery of signifi-
activity was detected in the mutants, but not in the wild
cant DAK activity and by the identification of the respec-
type. Moreover, the level of glycerate 3-kinase was
tive genes (DAK1 and DAK2) in S. cerevisiae (Norbeck
about twofold higher in the mutants compared with the
and Blomberg, 1997; Molin et al., 2003). However, no in
wild type. As glycerol kinase activity is also detectable in
vitro NAD1-dependent GDH activity could ever be mea-
sured in cell extracts of this organism although attempts N. crassa indicating the presence of the classical G3P
have been made by several authors (Norbeck and Blom- pathway (Viswanath-Reddy et al., 1977), it has been
berg, 1997; Ford and Ellis, 2002; Nguyen and Nevoigt, suggested that the catabolic GA pathway, including the
2009). The latter results make the existence of a func- two sub-pathways for the second part, might represent
tional endogenous glycerol catabolic DHA pathway in another route for glycerol catabolism in this organism.
S. cerevisiae very unlikely. Still, S. cerevisiae expresses Although the GA catabolic pathway has not been
proteins such as Gcy1 and Ypr1 which are homologous described in yeasts so far, homologs of the enzyme cat-
to fungal proteins with NADP1-dependent GDH activity alyzing the first step in this pathway have been reported
of the type EC 1.1.1.372/1.1.1.72 (Table 1). However, in S. cerevisiae (Gcy1 and Ypr1) as already mentioned
the contribution of this enzyme class in a glycerol cata- above.
bolic pathway with DHA as an intermediate is very For the sake of completeness, it has to be mentioned
unlikely due to its substrate specificity, which will be fur- that another enzyme activity converting glycerol to glyc-
ther scrutinized in the section ‘The dubious endogenous eraldehyde has been described in the filamentous fungi
DHA pathway in S. cerevisiae’. Aspergillus japonicus (Uwajima et al., 1984) and Penicil-
With regard to the DHA pathway (Fig. 1Aii), there is lium sp. (Lin et al., 1996). This enzyme catalyzes the
only one single study providing experimental evidence oxidation of glycerol with the consumption of oxygen to
that this pathway would indeed be used for glycerol utili- form glyceraldehyde and hydrogen peroxide. Whether
zation in yeast (Matsuzawa et al., 2010). The studied this enzyme activity (preliminary BRENDA-supplied EC
yeast species was Schizosaccharomyces pombe and number 1.1.3.B4) is essential for glycerol catabolism in
the strain used in this study did notably not grow in these organisms has not been investigated yet.
C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
884 M. Klein et al.
Glycerol anabolic pathways abolish glycerol biosynthesis under these conditions (Fil-
linger et al., 2001). The importance of gldB for osmoreg-
The best-studied pathway for the production of glycerol
ulation in A. nidulans is also supported by the fact that
from other carbon sources starts with the NADH-
its expression is strongly induced under conditions of
dependent reduction of glycolytic DHAP to G3P by a
hyper-osmotic shock. Moreover, the NADP1-dependent
cytosolic glycerol 3-phosphate dehydrogenase
GDH in A. nidulans as well as its orthologs in other fun-
(cG3PDH), after which G3P is dephosphorylated to glyc-
gal species (Table 1) have indeed strong preference for
erol by a glycerol 3-phosphatase (GPP) (Fig. 1Bi). This
the reduction of DHA under physiological conditions
‘anabolic G3P pathway’ is the major (or exclusive) path-
(i.e., at neutral pH). Although GDH activity has been
way for glycerol production in S. cerevisiae during osmo-
clearly demonstrated, this anabolic DHA pathway also
regulation and anaerobic redox balancing (Nevoigt and
requires a ‘DHAPase’ activity, which has to the best of
Stahl, 1997). In S. cerevisiae, cG3PDH activity is
our knowledge not yet been identified in any yeast or fil-
encoded by the two isogenes GPD1 and GPD2, while
amentous fungus. One might postulate that the glycerol
GPP activity is encoded by the isogenes GPP1 and
3-phosphatases Gpp1 and Gpp2 can also accept DHAP
GPP2. Genes encoding NAD1-dependent cG3PDH
as a substrate. However, a genetic screen specifically
have also been characterized in several other yeast spe-
developed for identifying genes whose gene products
cies and filamentous fungi, such as Candida magnoliae
dephosphorylate DHAP in S. cerevisiae did not result in
(Lee et al., 2008), Candida glycerinogenes (Chen et al.,
any candidate gene (E. Boles, Personal Communica-
2008), Debaryomyces hansenii (Thome, 2004) and A.
tion). Up to now, there are only two organisms (Neisse-
nidulans (Fillinger et al., 2001). In addition, cG3PDH
ria meningitidis and Plasmodium falciparum) listed in
activity using NADPH instead of NADH as the cofactor
BRENDA (www.brenda-enzymes.org) that exhibit sugar
(EC 1.1.1.94) has been identified in certain yeast spe-
phosphatase activity (EC 3.1.3.23) with DHAP as sub-
cies, such as Candida versatilis (Watanabe et al., 2008).
strate (Lee and Sowokinos, 1967; Guggisberg et al.,
A second pathway for glycerol production using DHA
2014).
as an intermediate has been proposed in filamentous
In Fig. 1Biii, we propose a third theoretical pathway for
fungi (referred to here as the ‘anabolic DHA pathway’,
glycerol formation. The existence of this pathway has not
Fig. 1Bii). The first step in this pathway is assumed to
been indicated by any molecular data so far nor even
be catalyzed by a so far uncharacterized enzyme that
been suggested in the literature. However, enzymes
dephosphorylates DHAP to DHA (see below). The sec-
converting D-glyceraldehyde to glycerol (EC-number
ond step is catalyzed by an NADP1-dependent GDH of
1.1.1.372 or 1.1.1.72; Table 1) have been described in fila-
the type EC 1.1.1.156 (Table 1). In contrast to the
mentous fungi such as N. crassa (Viswanath-Reddy et al.,
above-mentioned dehydrogenases belonging to EC
1977), H. jecorina (Liepins et al., 2006) and A. niger
1.1.1.372 or 1.1.1.72, enzymes from EC 1.1.1.156 are
(Martens-Uzunova and Schaap, 2009). As this enzymatic
able to reduce DHA in vitro but have only very low activi-
activity was induced in A. niger in the presence of galac-
ty (if at all) with glyceraldehyde. This anabolic DHA path-
turonic acid, a role of these enzymes in galacturonic acid
way has been postulated based on the fact that NADP-
dependent GDH activities converting DHA to glycerol catabolism has been suggested (Martens-Uzunova and
(EC 1.1.1.156) have been identified in S. pombe (Mar- Schaap, 2009). However, in principle the enzyme could
shall et al., 1989), A. nidulans (Redkar et al., 1995; De also be involved in an anabolic GA pathway for glycerol
Vries et al., 2003), A. niger (Schuurink et al., 1990), A. formation provided that an enzyme is present with sugar
oryzae (Ruijter et al., 2004) and H. jecorina (Liepins phosphatase activity, catalyzing the first step of the path-
et al., 2006) as also summarized in Table 1. To our way (EC 3.1.3.23 converting glyceraldehyde 3-phosphate
knowledge, studies using mutant strains for allocation of to GA).
this anabolic DHA pathway have only been conducted in
The dubious endogenous ‘DHA pathway’ in
A. nidulans (Fillinger et al., 2001; De Vries et al., 2003).
S. cerevisiae
These studies suggested that the anabolic DHA pathway
is the major pathway responsible for glycerol accumula- The first considerations concerning the possible exis-
tion during osmoregulation in this organism, while the tence of a catabolic DHA pathway in S. cerevisiae date
alternative anabolic G3P pathway did not seem to play a back to the study of Norbeck and Blomberg (1997).
significant role. In fact, the disruption of gldB encoding These authors analysed salt-induced changes in gene
NADPH-dependent GDH led to significantly reduced expression in S. cerevisiae at the proteome level and
intracellular glycerol levels and a severe growth defect in detected a significant increase in the abundance of
medium containing 1 M NaCl (De Vries et al., 2003), Dak1 (EC 2.7.1.29). Inspired by this finding, the authors
while the deletion of gfdA encoding cG3PDH did not attempted to identify the corresponding GDH but were
C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
Glycerol catabolism in yeast 885
unable to detect any in vitro GDH activity although differ- imposed conditions (Dak1/2 abolishment/Gcy1 overex-
ent pH values, substrates (DHA and glycerol) and co- pression) and that it is not relevant in wild-type cells. To
factors (NADH, NADPH, NAD1 and NADP1) were our knowledge, the major physiological functions for
tested. Gcy1 and Ypr1 gene products in S. cerevisiae have not
At the time when Norbeck and Blomberg (1997) con- yet been identified.
ducted their study, no protein sequence of an NAD1- Although there does not seem to be convincing evi-
dependent GDH was available, which would have dence for the existence of a native catabolic DHA pathway
allowed a search for genes encoding homologous pro- in S. cerevisiae, it is possible to engineer S. cerevisiae
teins in the S. cerevisiae genomic DNA sequence that strains for growth on glycerol using exclusively a synthetic
had just become publically available at that time. DHA pathway. We have shown recently that the native cat-
Instead, a search was conducted using a partial protein abolic G3P pathway of S. cerevisiae can be replaced by
sequence of a commercially available NADP1-depen- an artificial catabolic DHA pathway (Klein et al., 2016a).
dent GDH of A. niger (allocated to EC 1.1.1.72). The This was achieved by simultaneous expression of
authors identified the S. cerevisiae genes YPR1 and the NAD1-dependent GDH from O. parapolymorpha
GCY1 as encoding proteins with a sequence showing (EC 1.1.1.6; Table 1) and overexpression of the endoge-
the highest homology to the peptide fragment of the nous DAK1 gene in a gut1 deletion mutant in several
NADP1-dependent A. niger enzyme. The observation S. cerevisiae genetic backgrounds. Surprisingly, this path-
that Gcy1 expression is strongly increased upon osmo- way replacement together with the expression of a heter-
stress prompted (Blomberg, 2000) to speculate that ologous glycerol facilitator (see section ‘Glycerol transport
Gcy1 might function as a GDH for conversion of glycerol mechanisms’) even abolished the growth deficit in syn-
to DHA in a glycerol catabolic pathway. In addition, he thetic glycerol medium of several of the S. cerevisiae
proposed that increased activity of a catabolic DHA strains used, indicating that the growth deficit may be due
pathway during osmostress would lead to an ATP futile to insufficient activity or aberrant regulation of the native
cycle that might help to avoid substrate accelerated cell G3P pathway. Our results demonstrate that S. cerevisiae
death. However, so far no experimental proof has been naturally contains all prerequisites to utilize glycerol via
provided supporting this hypothesis. the DHA pathway (such as the ability to re-oxidize the sur-
The S. cerevisiae gene YPR1, of which the product plus of cytosolic NAD1 efficiently). Obviously, generation
shows 65% sequence similarity to that of the GCY1 of highly efficient utilization pathways for glycerol has not
gene, has been expressed in Escherichia coli to charac- been a priority during the evolutionary development of
terize the enzyme in more detail. It was demonstrated S. cerevisiae as already speculated in the section
that Ypr1 uses NADPH as a cofactor and preferably ‘Glycerol utilization by yeasts: high inter- and intra-species
reduces D/L-glyceraldehyde to glycerol. The activity of diversity’.
Ypr1 using DHA as a substrate was about 100-fold low-
er. Ypr1 also oxidized glycerol, however, this activity Glycerol transport mechanisms
was about 4,000-times lower than the activity in the
reducing direction with D/L-glyceraldehyde as substrate In order to be metabolized, glycerol must first enter the
(Ford and Ellis, 2002). The gene GCY1 has also been cell. Among fungi, the molecular details of the glycerol
expressed in E. coli and the activity of the gene product uptake mechanisms have been studied in greatest detail
tested with several substrates. D/L-glyceraldehyde was in the yeast S. cerevisiae. Only a few studies have been
the substrate with the highest kcat and NADPH was conducted in non-Saccharomyces yeasts and in filamen-
shown to be the co-factor (Chang et al., 2007). These tous fungi as elaborated below. The question of which
results suggest that neither Ypr1 nor Gcy1 significantly type of transport proteins are responsible for glycerol
contribute to in vivo DHA formation from glycerol in uptake in S. cerevisiae during growth on glycerol has
S. cerevisiae. been a matter of dispute for a long time, as comprehen-
It has to be mentioned that Jung et al. (2012) have sively summarized by Neves et al. (2004). In fact, pas-
shown that, at least under microaerobic conditions, a sive and facilitated diffusion as well as active transport
small amount of DHA is accumulated in a dak1 dak2 have all been considered in the last decades. We will
double deletion strain of S. cerevisiae. This DHA accu- summarize below the experimental findings in favour or
mulation further increased when GCY1 was overex- against the different hypotheses, which culminated in
pressed, indicating that Gcy1 has indeed a weak the currently accepted model of active transport.
glycerol oxidizing activity. However, it cannot be exclud- It is now generally accepted that during growth of
ed that the glycerol oxidation activity apparently present S. cerevisiae on glycerol the uptake of glycerol is solely
under these conditions is just a moonlighting activity that driven by a glycerol/H1-symporter encoded by STL1
has become significant only because of the artificially (Ferreira et al., 2005). The existence of such an active
C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
886 M. Klein et al.
transport system in S. cerevisiae had already been pre- additional deletion of GUP1 in a gpd1 gpd2 deletion
dicted by Sutherland et al. (1997) based on the demon- mutant compromises the rescuing effect of glycerol sup-
stration of simultaneous uptake of glycerol and protons plementation in hyperosmotic media suggesting a role of
accompanied by intracellular glycerol accumulation. Fer- Gup1 in glycerol import (Holst et al., 2000). However,
reira et al. (2005) finally revealed the crucial function of even before the glycerol/H1-symporter encoded by
Stl1 by showing that the deletion of STL1 completely STL1 had been identified, studies in which the expres-
abolished active glycerol transport as well as growth on sion patterns of GUP1 and GUP2 were compared with
glycerol. These results were originally obtained in auxo- glycerol uptake activity failed to support the concept that
trophic laboratory strains, but the same deletion in the these genes encode glycerol transporters in S. cerevi-
wild-type S. cerevisiae isolate CBS 6412-13A also siae (Oliveira and Lucas, 2004). While GUP1 and GUP2
resulted in inability to grow in synthetic glycerol medium expression was found to be constitutive, glycerol uptake
(Swinnen et al., 2013). activity is clearly affected by carbon catabolite repres-
Similar active glycerol uptake systems using H1- but sion (see below) as well as during growth under salt
also Na1-symport have been described for halotolerant stress. Moreover, mutants carrying deletions of gup1
yeast species such as Debaryomyces hansenii (Lucas and gup2 still show active glycerol uptake by glycerol/
et al., 1990), Pichia sorbitophila (Lages and Lucas, H1 symport under salt stress (Neves et al., 2004). In
1995) and Zygosaccharomyces rouxii (Van Zyl et al., addition, deletion of GUP1 in the wild-type isolate CBS
1990). In these highly osmotolerant species, the active 6412-13A did not impair the strain’s natural ability to
glycerol uptake has been studied mainly in the context grow in synthetic glycerol medium (our unpublished
of establishing and maintaining a glycerol gradient results). Therefore, the actual function of Gup1 and
across the cytoplasmic membrane in the presence of Gup2 in glycerol uptake remains to be resolved. As
high salt concentrations. It seems that uptake of glycerol GUP1 was later shown to encode an O-acyltransferase
in response to osmotic stress also occurs in S. cerevi- involved in remodelling of GPI anchors (Bosson et al.,
siae since transcription of STL1 is transiently induced by 2006) the gene product might affect glycerol uptake
osmotic shock during exponential growth in glucose- indirectly.
based media (Posas et al., 2000; Rep et al., 2000; Yale It is well-known that S. cerevisiae expresses a glycerol
and Bohnert, 2001; Ferreira et al., 2005). However, this facilitator, Fps1, a member of the major intrinsic protein
osmostress-induced active glycerol uptake seems to be (MIP) family of channel proteins (Luyten et al., 1995).
replaced subsequently by an increase in cellular glycerol This protein shows high sequence similarity to the previ-
production (and its intracellular retention due to closing ously described and extensively studied glycerol facilita-
of the Fps1 channel preventing glycerol efflux) as Stl1 tor GlpF from E. coli (Heller et al., 1980; Sweet et al.,
abundance and transport activity started to decrease 1990; Maurel et al., 1994). Notably, channel-mediated
gradually 1.5 h after the shift to hyperosmotic medium. diffusion via GlpF is the sole glycerol uptake system in
Interestingly, downregulation of STL1 expression and E. coli. The constitutively expressed glycerol facilitator,
degradation of Stl1 protein did not occur in a gpd1 gpd2 Fps1, in S. cerevisiae was shown to control glycerol
double deletion mutant (Ferreira et al., 2005). Such a efflux as a function of medium osmolarity and was initial-
strain cannot synthesize any glycerol and is therefore ly assumed also to contribute significantly to glycerol
sensitive to hyperosmotic shock. However, growth in the uptake in this organism. Early experimental studies dem-
presence of increased salt concentrations can be res- onstrated a saturable glycerol uptake component and
cued in this strain by supplementation of the growth thus seemed to confirm this hypothesis (Luyten et al.,
medium with small amounts of glycerol, which apparent- 1995; Sutherland et al., 1997). However, the results later
ly allows the persisting Stl1 activity to compensate at turned out to be caused by an artefact since the respec-
least partially for the absent glycerol anabolic pathway. tive studies used glucose-grown S. cerevisiae cells or
Prior to the demonstration that Stl1 is the main media- cells harvested during the diauxic shift (Oliveira et al.,
tor of glycerol uptake in S. cerevisiae, GUP1 and its 2003). Such cells contain considerable Gut1 activity
paralog GUP2 had been suggested to encode proteins (Grauslund et al., 1999; Oliveira et al., 2003) that inter-
responsible for glycerol active transport (Holst et al., fered with the experimental set-up causing the saturable
2000). GUP1 and GUP2 encode membrane proteins glycerol uptake kinetics. Indeed, Oliveira et al. (2003)
with multiple predicted transmembrane domains. These demonstrated that the latter was abolished by deletion
authors showed that the GUP1 gene product is essential of GUT1. These results are also consistent with the find-
for efficient growth in media containing glycerol as the ing of Luyten et al. (1995) that growth on glycerol as
sole carbon source and found that it also seems to play sole carbon source was unaffected in fps1D mutants.
an important role in the alleviation of the osmotic-stress Further studies confirmed that the main function of
induced cell death of gpd1 gpd2 mutants. In fact, Fps1 in S. cerevisiae is the control of glycerol export
C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
Glycerol catabolism in yeast 887
during osmoregulation rather than glycerol uptake et al. (2013) showed that FPS2 from the P. tannophilus
(Tamas et al., 1999). Oliveira et al. (2003) even claimed strain CBS 4044 (PtFPS2) was among the most strongly
that S. cerevisiae Fps1 should not be considered a glyc- upregulated genes upon switch from glucose- to glycerol-
erol facilitator with regard to glycerol uptake at all as it based medium. Interestingly, expression of the predicted
does not exhibit uptake properties (Vmax values) that glycerol facilitator was able to suppress the growth defect
would be in agreement with the definition of a facilitated on glycerol of a S. cerevisiae stl1 deletion mutant, which
diffusion type of transport. was in remarkable contrast to overexpression of the
Oliveira et al. (2003) also questioned the hitherto gen- endogenous S. cerevisiae FPS1. These results imply dif-
erally accepted concept that glycerol is able to cross the ferent functions for the two partially conserved facilitator
lipid bilayer of the plasma membrane by passive diffu- proteins with respect to glycerol transport (Liu et al.,
sion solely driven by its concentration gradient (Gancedo 2013). The apparently different physiological function of
et al., 1968; Heredia et al., 1968; Romano, 1986). the ScFps1 and PtFps2 orthologs is also reflected at the
Based on their experimental data, they hypothesized molecular level. The overall sequence similarity between
that the S. cerevisiae plasma membrane is almost the two proteins is 32% but the sequence similarity is
impermeable to glycerol and that any detectable passive largely restricted to the core section of ScFps1 with its six
diffusion into the cell is actually mediated via the Fps1 putative trans-membrane domains (TMDs). ScFps1 has
channel. Transport components different from active 669 amino acid residues, which makes it much longer
transport can anyway only be very low since stl1 dele- than PtFps2 (323 residues). Tamas et al. (1999) demon-
tion mutants are unable to grow in synthetic glycerol strated that the domains at the N- and C-terminus are
medium as mentioned above. Theoretical considerations important for controlling the specific ScFps1 function of
based on comparison of the glycerol olive oil/water parti- osmoregulated glycerol export. More specifically, the N-
tion coefficient (logP) with that of other molecules also terminus was found to control the closing of the channel
support the view that glycerol is unlikely to show signifi- thereby restricting glycerol efflux through ScFps1 during
cant diffusion through biological membranes (Oliveira osmotic stress. This property would be dispensable in
et al., 2003). Still, the same authors claim that some case the channel was solely used during glycerol
passive diffusion via the lipid bilayer cannot be assimilation.
completely excluded as cytoplasmic membranes are Overexpression in the S. cerevisiae strain CBS 6412-
highly flexible and dynamic systems. 13A of PtFPS2 and similar FPS1 homologs from differ-
As described in the section ‘Glycerol utilization ent non-conventional yeast species (C. jadinii, P. pastoris
by yeasts: high inter-species diversity’ several non- and Y. lipolytica) with superior growth performance on
conventional yeast species have been reported to show glycerol compared with S. cerevisiae, resulted in a sig-
growth rates on glycerol significantly higher than that of nificantly improved lmax from  0.13 h21 (for wild-type
S. cerevisiae. The molecular basis for the more efficient CBS 6412-13A) to  0.18 h21 (for the engineered
glycerol utilization in these yeast species as compared to strains) (Klein et al., 2016b). The high growth rates were
S. cerevisiae is not clear, although Gancedo et al. (1968) retained even after deletion of the endogenous STL1
provided indications that glycerol uptake might constitute demonstrating that the heterologous facilitators are able
a major limitation in S. cerevisiae. This study demonstrat- to fully replace the active transport system without loss
ed that glycerol uptake (described as membrane perme- of the superior growth capacity. Similarly, overexpression
ability) in the S. cerevisiae strain studied was  105 fold of FPS1 from C. jadinii (CjFPS1) in a reverse-
lower as compared with that in C. jadinii. In several yeast engineered S. cerevisiae strain of CEN.PK113-1A, car-
genome sequencing projects, homologs of STL1 have rying allele replacements for UBR2, GUT1 and SSK1
been identified, such as in Y. lipolytica (Dujon et al., 2004), from CBS 6412-13A, further improved the strain’s
P. tannophilus as well as K. pastoris, the latter even pos- growth performance on glycerol (Swinnen et al., 2016).
sessing four distinct glycerol/H1-symporters (Mattanovich However, the glycerol facilitators from the aforemen-
et al., 2009). Blast searches with FPS1 and genome tioned non-conventional yeast species should also be
sequence comparisons also resulted in the identification analysed by functional studies in the original organisms
of several homologs in various yeast species and the gen- in order to learn more about their actual contribution to
eration of unrooted phylogenetic trees showed that glycerol uptake in these species. A more detailed analy-
many of them, such as Fps1 homologs in P. tannophilus, sis of the different transporters in the original species
K. pastoris and Y. lipolytica, cluster in a branch separate might provide more insight into the molecular basis of
from S. cerevisiae Fps1 (ScFps1) (Liu et al., 2013). In con- the superior glycerol uptake capability of many non-
trast to S. cerevisiae the contribution of Stl1 and Fps1 conventional yeast species in comparison to S. cerevi-
homologs to glycerol utilization in other yeast species siae. In particular, the fact that many of these yeast spe-
have not yet been comprehensively studied. At least, Liu cies carry several STL1 and/or FPS1 isogenes and the
C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
888 M. Klein et al.
different molecular structure of the Fps homologs (see expression of genes encoding proteins related to mito-
above) stands in sharp contrast with the genetic consti- chondrial function and energy metabolism as well as
tution of S. cerevisiae in this respect. that of genes encoding enzymes involved in gluconeo-
The improvement of growth on glycerol achieved by genesis, the TCA and glyoxalate cycles, and in carbohy-
the replacement of S. cerevisiae’s native glycerol/H1- drate storage was highly up-regulated. Interestingly, the
symport system by a heterologous glycerol facilitator expression of genes encoding products associated with
confirmed that glycerol uptake is a the rate-limiting step stress response was also found to be up-regulated. In
for growth on glycerol in this yeast species. However, contrast, the expression of genes encoding proteins
many yeast species still show even higher growth rates required for biosynthesis, such as transcription by the
on glycerol than the engineered S. cerevisiae variants RNA polymerases I and III, DNA replication and ribo-
implying that other factors still restrict more efficient some biogenesis and assembly, was significantly down-
glycerol utilization in S. cerevisiae. Such a rate- regulated reflecting the generally lower growth rate of S.
controlling factor could for example reside in an ineffi- cerevisiae on respiratory carbon sources (DeRisi et al.,
cient glycerol catabolic pathway as elaborated in the 1997; Brauer et al., 2005; Roberts and Hudson, 2006).
section ‘The dubious catabolic DHA pathway in S. cere- However, one has to emphasize that only Roberts and
visiae’. In any case, the alleviation of glycerol transport Hudson (2006) investigated the transcriptome of cells
limitation now accomplished in S. cerevisiae paves the exponentially growing on either glycerol or ethanol, while
way for identification of further rate-controlling steps in the other two studies solely focused on analysing sam-
glycerol utilization and ultimately for further improvement ples around the diauxic shift after glucose depletion.
of the capacity of S. cerevisiae to grow on glycerol. The direct comparison of the results allowed the identifi-
cation of carbon source specific regulatory patterns
Carbon source regulation of glycerol transport and (Roberts and Hudson, 2006). For example, abundance
catabolic pathways of STL1 transcripts encoding the glycerol/H1-symporter
for glycerol uptake was strongly increased after glucose
The preferred carbon source of many unicellular organ-
depletion and reached a level that was several hundred-
isms is glucose and its presence often excludes the utili-
fold higher during continuous growth on glycerol as com-
zation of alternative carbon sources such as glycerol. pared with glucose. Likewise, the GUT1 and GUT2
This repressive effect often mediated by multiple inter- genes encoding the enzymes catalyzing the initial glyc-
linked regulatory interactions and signalling pathways is erol catabolic G3P pathway, showed a strong transcrip-
generally referred to as carbon catabolite repression. tional up-regulation in glycerol-based medium.
The shift to an alternative carbon source is character- Surprisingly, derepression of these genes was also
ized by a transition period with delayed cellular growth observed to a certain extent in medium containing etha-
(designated as diauxic shift) and the underlying molecu- nol as sole carbon source (Roberts and Hudson, 2006).
lar remodelling has been most thoroughly investigated in Interestingly, the expression of GCY1 also showed a
the yeast S. cerevisiae. Being a Crabtree-positive yeast strong up-regulation in medium containing glycerol, but
species (exhibiting fermentative metabolism even under not ethanol, as sole source of carbon. As discussed in
aerobic conditions as soon as the glucose concentration the section ‘The dubious endogenous “DHA pathway” in
exceeds a certain threshold concentration), S. cerevisiae S. cerevisiae’, Gcy1 is a homolog of the fungal NADP1-
naturally displays an extreme case of such a diauxic dependent GDH but its real function in S. cerevisiae’s
shift, when it switches from fermentative growth on glu- physiology has remained rather unclear.
cose to full respiratory metabolism of the previously pro- With regard to the above-mentioned transcriptome
duced fermentation products, mainly ethanol and to a studies, one has to keep in mind, that all three were per-
smaller extent glycerol. It is evident that such a drastic formed in media containing complex medium compo-
shift from utilizing a fermentable to a respiratory carbon nents and that data for exponentially growing
source requires a dramatic reprogramming of the cellular S. cerevisiae cells in synthetic media are still lacking.
machinery. In S. cerevisiae, the latter is mainly (but not This evidently (at least for glycerol) can be attributed to
exclusively) exerted at the transcriptional level as dem- the fact that (as elaborated above) commonly used labo-
onstrated by many studies of this phenomenon, includ- ratory strains of S. cerevisiae do not grow in synthetic
ing microarray-based transcriptome analyses at a global glycerol medium at all.
scale (DeRisi et al., 1997; Brauer et al., 2005; Roberts The molecular basis for carbon catabolite repression
and Hudson, 2006). Evaluation of the respective data has been studied in great detail in S. cerevisiae, and the
revealed many adaptations to respiratory growth that are reader is referred to a number of excellent reviews in
common to both carbon sources glycerol and ethanol. €ller, 2003; Turcotte et al.,
this field for more details (Schu
For instance all three studies showed that the 2010; Kayikci and Nielsen, 2015). Although it is clear
C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
Glycerol catabolism in yeast 889
that glucose depletion has a major impact on the initia- from sugar and intracellular accumulation of glycerol is
tion of the regulatory cascades resulting in growth on increased during hyperosmotic stress (Nevoigt and Stahl,
alternative respiratory carbon sources, the information 1997; Hohmann, 2002; 2015). Therefore, during growth
specifically related to the utilization of glycerol is rather on glycerol as the sole carbon source, catabolic and ana-
fragmentary. For example, it has been shown that the bolic pathways should be more or less strictly regulated
transcriptional activator Cat8 seems to affect the level of in order to avoid futile cycling. Indeed, transcriptome pro-
STL1 derepression during the diauxic shift (Haurie et al., filing of S. cerevisiae cells continuously growing on glycer-
2001). Cat8 is essential for growth on non-fermentable ol revealed a reduction of transcript levels for genes
carbon sources (Hedges et al., 1995) and derepresses involved in glycerol anabolic pathways, such as GPD2,
target genes by binding to their upstream carbon source GPP1 and GPP2 (Roberts and Hudson, 2006). Interest-
responsive elements (CSREs) (Young et al., 2003; Roth ingly, transcription of GPD1 was found to be slightly up-
et al., 2004). Another transcription factor important for regulated on glycerol- as compared with glucose-based
glycerol utilization is Adr1 that (together with Ino2 and medium probably reflecting an increased activity of the
Ino4) is required for the derepression of GUT1 (Graus- G3P shuttle (see Larsson et al., 1998) under these
lund et al., 1999). The derepression of GUT2 is depen- conditions.
dent on the activity of the protein kinase Snf1 as well as
on the heteromeric protein complex Hap2-Hap5, which Concluding remarks
activates the transcription of many nuclear genes encod-
Most of our knowledge on glycerol metabolism in yeasts
ing proteins with a mitochondrial function (Grauslund
and filamentous fungi has been derived originally from
and Ronnow, 2000). Both GUT1 and GUT2 are
research on laboratory strains of the yeast S. cerevisiae.
repressed by the negative regulator Opi1 during growth
Although this has led to elucidation of the glycerol
on glucose (Grauslund et al., 1999; Grauslund and Ron-
uptake mechanisms and the backbone of all major path-
now, 2000). Several studies have identified other factors
ways of glycerol metabolism, there are still some con-
with a crucial function for growth of S. cerevisiae on
spicuous gaps left in our knowledge of the genes
glycerol, such as Rsf1 (Lu et al., 2003; Roberts and
encoding enzymes supposed to be active in specific
Hudson, 2009) and Rsf2 (Zms1) (Lu et al., 2005). Dele-
steps of these pathways. Detailed insight has also been
tion mutants of RSF1 showed an imbalanced expression
gained for the regulation of glycerol metabolism in
of genes encoding enzymes for glycerol catabolic and
S. cerevisiae, in particular in the mechanism of carbon
anabolic pathways, decreased transcription of HAP4
catabolite repression for the utilization of non-
(whose gene product is part of the aforementioned
fermentable carbon sources such as glycerol. Progress
Hap2-Hap5 protein complex) as well as an increased
has also been made in understanding the striking differ-
transcript level of genes whose products function during
ences between S. cerevisiae strains in their capacity to
stress responses (Roberts and Hudson, 2009).
use glycerol as a sole source of carbon and this may
In comparison to S. cerevisiae, even less is known
lead to the generation of industrial yeast strains with
about the regulation of glycerol utilization in other yeast
greatly improved capacity to utilize crude glycerol as a
species. It would be particularly interesting to under-
carbon source. In spite of this, other yeast species still
stand the respective aspects of regulation in those yeast
display much higher growth rates on glycerol and more
species that exhibit a superior growth capability on glyc-
insight is needed into the molecular basis of this superi-
erol. The fact that, for example, Y. lipolytica consumes
or growth capacity in order to be able to use it as a tool
and actually prefers glycerol even if glucose is present
for further improvement of glycerol catabolic capacity in
(Workman et al., 2013) implies regulatory mechanisms
S. cerevisiae.
in this species which are distinct from common carbon
catabolite repression. Future investigations of such regu-
latory principles might deliver more detailed explanations References
for the superior glycerol utilization of certain non- Babel, W., and Hofmann, K.H. (1982) The relation between
conventional yeast species. the assimilation of methanol and glycerol in yeasts. Arch
In many fungal species, enzymes and transporters for Microbiol 132: 179–184.
a glycerol catabolic pathway often coexist with those Bakker, B.M., Overkamp, K.M., van Maris, A.J., Kotter, P.,
required for the formation of glycerol. For example, glyc- Luttik, M.A., van Dijken, J.P., and Pronk, J.T. (2001) Stoichi-
ometry and compartmentation of NADH metabolism in Sac-
erol is a usual by-product of sugar catabolism in S. cere-
charomyces cerevisiae. FEMS Microbiol Rev 25: 15–37.
visiae where it serves as the major redox sink for Blomberg, A. (2000) Metabolic surprises in Saccharomyces
reducing equivalents produced in biosynthetic pathways cerevisiae during adaptation to saline conditions: ques-
(biomass production) particularly under anaerobic condi- tions, some answers and a model. FEMS Microbiol Lett
tions (Bakker et al., 2001). Moreover, glycerol production 182: 1–8.

C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
890 M. Klein et al.
Bosson, R., Jaquenoud, M., and Conzelmann, A. (2006) glycerol-3-phosphate dehydrogenase gene, GUT2, from
GUP1 of Saccharomyces cerevisiae encodes an O- Saccharomyces cerevisiae. Can J Microbiol 46: 1096–
acyltransferase involved in remodeling of the GPI anchor. 1100.
Mol Biol Cell 17: 2636–2645. Grauslund, M., Lopes, J.M., and Ronnow, B. (1999) Expres-
Brauer, M.J., Saldanha, A.J., Dolinski, K., and Botstein, D. sion of GUT1, which encodes glycerol kinase in Saccha-
(2005) Homeostatic adjustment and metabolic remodeling in romyces cerevisiae, is controlled by the positive
glucose-limited yeast cultures. Mol Biol Cell 16: 2503–2517. regulators Adr1p, Ino2p and Ino4p and the negative regu-
Chang, Q., Griest, T.A., Harter, T.M., and Petrash, J.M. lator Opi1p in a carbon source-dependent fashion.
(2007) Functional studies of aldo-keto reductases in Sac- Nucleic Acids Res 27: 4391–4398.
charomyces cerevisiae. Biochim Biophys Acta 1773: Guggisberg, A.M., Park, J., Edwards, R.L., Kelly, M.L.,
321–329. Hodge, D.M., Tolia, N.H., and Odom, A.R. (2014) A sugar
Chen, X., Fang, H., Rao, Z., Shen, W., Zhuge, B., Wang, phosphatase regulates the methylerythritol phosphate
Z., and Zhuge, J. (2008) Cloning and characterization of (MEP) pathway in malaria parasites. Nat Commun 5:
a NAD1-dependent glycerol-3-phosphate dehydrogenase 4467.
gene from Candida glycerinogenes, an industrial glycerol Hallsworth, J.E., and Magan, N. (1995) Manipulation of
producer. FEMS Yeast Res 8: 725–734. intracellular glycerol and erythritol enhances germination
Cray, J.A., Stevenson, A., Ball, P., Bankar, S.B., Eleutherio, of conidia at low water availability. Microbiology 141 (Pt
E.C., Ezeji, T.C., et al. (2015) Chaotropicity: a key factor 5): 1109–1115.
in product tolerance of biofuel-producing microorganisms. Haurie, V., Perrot, M., Mini, T., Jeno, P., Sagliocco, F., and
Curr Opin Biotechnol 33: 228–259. Boucherie, H. (2001) The transcriptional activator Cat8p
De Jong, J.C., McCormack, B.J., Smirnoff, N., and Talbot, provides a major contribution to the reprogramming of
N.J. (1997) Glycerol generates turgor in rice blast. Nature carbon metabolism during the diauxic shift in Saccharo-
389: 244–244. myces cerevisiae. J Biol Chem 276: 76–85.
De Vries, R.P., Flitter, S.J., van de Vondervoort, P.J., Hedges, D., Proft, M., and Entian, K.D. (1995) CAT8, a new
Chaveroche, M.K., Fontaine, T., Fillinger, S., et al. (2003) zinc cluster-encoding gene necessary for derepression of
Glycerol dehydrogenase, encoded by gldB is essential for gluconeogenic enzymes in the yeast Saccharomyces cer-
osmotolerance in Aspergillus nidulans. Mol Microbiol 49: evisiae. Mol Cell Biol 15: 1915–1922.
131–141. Heller, K.B., Lin, E.C., and Wilson, T.H. (1980) Substrate
DeRisi, J.L., Iyer, V.R., and Brown, P.O. (1997) Exploring specificity and transport properties of the glycerol facilita-
the metabolic and genetic control of gene expression on tor of Escherichia coli. J Bacteriol 144: 274–278.
a genomic scale. Science 278: 680–686. Heredia, C.F., Sols, A., and DelaFuente, G. (1968) Specific-
Dujon, B., Sherman, D., Fischer, G., Durrens, P., ity of the constitutive hexose transport in yeast. Eur J
Casaregola, S., Lafontaine, I., et al. (2004) Genome evo- Biochem 5: 321–329.
lution in yeasts. Nature 430: 35–44. Hibuse, T., Maeda, N., Nagasawa, A., and Funahashi, T.
Falb, M., Muller, K., Konigsmaier, L., Oberwinkler, T., Horn, (2006) Aquaporins and glycerol metabolism. Biochim Bio-
P., von Gronau, S., et al. (2008) Metabolism of halophilic phys Acta 1758: 1004–1011.
archaea. Extremophiles 12: 177–196. Hohmann, S. (2002) Osmotic stress signaling and osmoa-
Ferreira, C., van Voorst, F., Martins, A., Neves, L., Oliveira, daptation in yeasts. Microbiol Mol Biol Rev 66: 300–372.
R., Kielland-Brandt, M.C., et al. (2005) A member of the Hohmann, S. (2015) An integrated view on a eukaryotic
sugar transporter family, Stl1p is the glycerol/H1 sym- osmoregulation system. Curr Genet 61: 373–382.
porter in Saccharomyces cerevisiae. Mol Biol Cell 16: Holst, B., Lunde, C., Lages, F., Oliveira, R., Lucas, C., and
2068–2076. Kielland-Brandt, M.C. (2000) GUP1 and its close homo-
Fillinger, S., Ruijter, G., Tamas, M.J., Visser, J., Thevelein, logue GUP2, encoding multimembrane-spanning proteins
J.M., and D’enfert, C. (2001) Molecular and physiological involved in active glycerol uptake in Saccharomyces cere-
characterization of the NAD-dependent glycerol 3- visiae. Mol Microbiol 37: 108–124.
phosphate dehydrogenase in the filamentous fungus Jung, J.Y., Kim, T.Y., Ng, C.Y., and Oh, M.K. (2012) Charac-
Aspergillus nidulans. Mol Microbiol 39: 145–157. terization of GCY1 in Saccharomyces cerevisiae by meta-
Finley, D., Ulrich, H.D., Sommer, T., and Kaiser, P. (2012) bolic profiling. J Appl Microbiol 113: 1468–1478.
The ubiquitin-proteasome system of Saccharomyces cer- Kayikci, O., and Nielsen, J. (2015) Glucose repression in
evisiae. Genetics 192: 319–360. Saccharomyces cerevisiae. FEMS Yeast Res 15: 68.
Ford, G., and Ellis, E.M. (2002) Characterization of Ypr1p Klein, M., Carrillo, M., Xiberras, J., Islam, Z.U., Swinnen,
from Saccharomyces cerevisiae as a 2-methylbutyraldehyde S., and Nevoigt, E. (2016a) Towards the exploitation of
reductase. Yeast 19: 1087–1096. glycerol’s high reducing power in Saccharomyces cerevi-
Gancedo, C., Gancedo, J.M., and Sols, A. (1968) Glycerol siae-based bioprocesses. Metab Eng 38: 464–472.
metabolism in yeasts. Pathways of utilization and produc- Klein, M., Islam, Z., Knudsen, P.B., Carrillo, M., Swinnen,
tion. Eur J Biochem 5: 165–172. S., Workman, M., and Nevoigt, E. (2016b) The expres-
Goddard, M.R., and Greig, D. (2015) Saccharomyces cere- sion of glycerol facilitators from various yeast species
visiae: a nomadic yeast with no niche? FEMS Yeast Res improves growth on glycerol of Saccharomyces cerevi-
15: 9. siae. Metab Eng Commun 3: 252–257.
Grauslund, M., and Ronnow, B. (2000) Carbon source- Kurtzman, C.P., Fell, J.W., and Boekhout, T. (eds) (2011)
dependent transcriptional regulation of the mitochondrial The Yeasts - a Taxonomic Study. New York: Elsevier.

C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
Glycerol catabolism in yeast 891
Lages, F., and Lucas, C. (1995) Characterization of a glyc- Matsuzawa, T., Ohashi, T., Hosomi, A., Tanaka, N., Tohda,
erol/H1 symport in the halotolerant yeast Pichia sorbito- H., and Takegawa, K. (2010) The gld1 gene encoding
phila. Yeast 11: 111–119. glycerol dehydrogenase is required for glycerol metabo-
Lages, F., Silva-Graca, M., and Lucas, C. (1999) Active lism in Schizosaccharomyces pombe. Appl Microbiol Bio-
glycerol uptake is a mechanism underlying halotolerance technol 87: 715–727.
in yeasts: a study of 42 species. Microbiology 145 (Pt 9): Mattam, A.J., Clomburg, J.M., Gonzalez, R., and Yazdani,
2577–2585. S.S. (2013) Fermentation of glycerol and production of
Larsson, C., Pahlman, I.L., Ansell, R., Rigoulet, M., Adler, valuable chemical and biofuel molecules. Biotechnol Lett
L., and Gustafsson, L. (1998) The importance of the glyc- 35: 831–842.
erol 3-phosphate shuttle during aerobic growth of Sac- Mattanovich, D., Graf, A., Stadlmann, J., Dragosits, M.,
charomyces cerevisiae. Yeast 14: 347–357. Redl, A., Maurer, M., et al. (2009) Genome, secretome
Lee, Y.P., and Sowokinos, J.R. (1967) Sugar phosphate and glucose transport highlight unique features of the
phosphohydrolase. I. Substrate specificity, intracellular protein production host Pichia pastoris. Microb Cell Fact
localization, and purification from Neisseria meningitidis. 8: 29.
J Biol Chem 242: 2264–2271. Maurel, C., Reizer, J., Schroeder, J.I., Chrispeels, M.J., and
Lee, D.H., Kim, M.D., Ryu, Y.W., and Seo, J.H. (2008) Cloning Saier, M.H. Jr. (1994) Functional characterization of the
and characterization of CmGPD1, the Candida magnoliae Escherichia coli glycerol facilitator, GlpF, in Xenopus
homologue of glycerol-3-phosphate dehydrogenase. FEMS oocytes. J Biol Chem 269: 11869–11872.
Yeast Res 8: 1324–1333. Merico, A., Ragni, E., Galafassi, S., Popolo, L., and
Liepins, J., Kuorelahti, S., Penttila, M., and Richard, P. Compagno, C. (2011) Generation of an evolved Saccha-
(2006) Enzymes for the NADPH-dependent reduction of romyces cerevisiae strain with a high freeze tolerance
dihydroxyacetone and D-glyceraldehyde and L- and an improved ability to grow on glycerol. J Ind Micro-
glyceraldehyde in the mould Hypocrea jecorina. Febs J biol Biotechnol 38: 1037–1044.
273: 4229–4235. Molin, M., Norbeck, J., and Blomberg, A. (2003) Dihydroxy-
Lin, S.F., Chiou, C.M., and Tsai, Y.C. (1996) Purification acetone kinases in Saccharomyces cerevisiae are
and characterization of a glycerol oxidase from Penicilli- involved in detoxification of dihydroxyacetone. J Biol
um sp. TS-622. Enzyme Microb Technol 18: 383–387. Chem 278: 1415–1423.
Liu, X., Jensen, P.R., and Workman, M. (2012) Bioconver- Neves, L., Lages, F., and Lucas, C. (2004) New insights on
sion of crude glycerol feedstocks into ethanol by Pachy- glycerol transport in Saccharomyces cerevisiae. FEBS
solen tannophilus. Bioresour Technol 104: 579–586. Lett 565: 160–162.
Liu, X., Mortensen, U.H., and Workman, M. (2013) Expres- Nevoigt, E. (2008) Progress in metabolic engineering of
sion and functional studies of genes involved in transport Saccharomyces cerevisiae. Microbiol Mol Biol Rev 72:
and metabolism of glycerol in Pachysolen tannophilus. 379–412.
Microb Cell Fact 12: 27. Nevoigt, E., and Stahl, U. (1997) Osmoregulation and glyc-
Lu, L., Roberts, G., Simon, K., Yu, J., and Hudson, A.P. erol metabolism in the yeast Saccharomyces cerevisiae.
(2003) Rsf1p, a protein required for respiratory growth of FEMS Microbiol Rev 21: 231–241.
Saccharomyces cerevisiae. Curr Genet 43: 263–272. Nguyen, H.T., and Nevoigt, E. (2009) Engineering of Sac-
Lu, L., Roberts, G.G., Oszust, C., and Hudson, A.P. (2005) charomyces cerevisiae for the production of dihydroxyac-
The YJR127C/ZMS1 gene product is involved in glycerol- etone (DHA) from sugars: a proof of concept. Metab Eng
based respiratory growth of the yeast Saccharomyces 11: 335–346.
cerevisiae. Curr Genet 48: 235–246. Norbeck, J., and Blomberg, A. (1997) Metabolic and regula-
Lucas, C., Da Costa, M., and van Uden, N. (1990) Osmo- tory changes associated with growth of Saccharomyces
regulatory Active Sodium-Glycerol Co-transport in the cerevisiae in 1.4 M NaCl. Evidence for osmotic induction
Halotolerant Yeast Debaryomyces hansenii. Yeast 6: of glycerol dissimilation via the dihydroxyacetone path-
187–191. way. J Biol Chem 272: 5544–5554.
Luyten, K., Albertyn, J., Skibbe, W.F., Prior, B.A., Ramos, Ochoa-Estopier, A., Lesage, J., Gorret, N., and Guillouet,
J., Thevelein, J.M., and Hohmann, S. (1995) Fps1, a S.E. (2010) Kinetic analysis of a Saccharomyces cerevi-
yeast member of the MIP family of channel proteins, is a siae strain adapted for improved growth on glycerol:
facilitator for glycerol uptake and efflux and is inactive implications for the development of yeast bioprocesses
under osmotic stress. Embo J 14: 1360–1371. on glycerol. Bioresour Technol 102: 1521–1577.
Maeda, T., Wurgler-Murphy, S.M., and Saito, H. (1994) A Oliveira, R., and Lucas, C. (2004) Expression studies of
two-component system that regulates an osmosensing GUP1 and GUP2, genes involved in glycerol active trans-
MAP kinase cascade in yeast. Nature 369: 242–245. port in Saccharomyces cerevisiae, using semi-
Marshall, J.H., Kong, Y.C., Sloan, J., and May, J.W. (1989) quantitative RT-PCR. Curr Genet 46: 140–146.
Purification and properties of glycerol:NADP1 2- Oliveira, R., Lages, F., Silva-Graca, M., and Lucas, C.
oxidoreductase from Schizosaccharomyces pombe. (2003) Fps1p channel is the mediator of the major part of
J Gen Microbiol 135: 697–701. glycerol passive diffusion in Saccharomyces cerevisiae:
Martens-Uzunova, E.S., and Schaap, P.J. (2009) Assess- artefacts and re-definitions. Biochim Biophys Acta 1613:
ment of the pectin degrading enzyme network of Asper- 57–71.
gillus niger by functional genomics. Fungal Genet Biol 46 Papanikolaou, S., Muniglia, L., Chevalot, I., Aggelis, G., and
Suppl 1: S170–S179. Marc, I. (2002) Yarrowia lipolytica as a potential producer

C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
892 M. Klein et al.
of citric acid from raw glycerol. J Appl Microbiol 92: 737– Sprague, G.F., and Cronan, J.E. (1977) Isolation and char-
744. acterization of Saccharomyces cerevisiae mutants defec-
Pavlik, P., Simon, M., Schuster, T., and Ruis, H. (1993) tive in glycerol catabolism. J Bacteriol 129: 1335–1342.
The glycerol kinase (GUT1) gene of Saccharomyces Stevenson, A., Hamill, P.G., Medina, A., Kminek, G.,
cerevisiae: cloning and characterization. Curr Genet 24: Rummel, J.D., Dijksterhuis, J., et al. (2016) Glycerol
21–25. enhances fungal germination at the water-activity limit for
Pfeiffer, T., Schuster, S., and Bonhoeffer, S. (2001) Cooper- life. Environ Microbiol. doi: 10.1111/1462-2920.13530.
ation and competition in the evolution of ATP-producing [Epub ahead of print]
pathways. Science 292: 504–507. Sutherland, F.C., Lages, F., Lucas, C., Luyten, K., Albertyn,
Piskur, J., Rozpedowska, E., Polakova, S., Merico, A., and J., Hohmann, S., et al. (1997) Characteristics of Fps1-
Compagno, C. (2006) How did Saccharomyces evolve to dependent and -independent glycerol transport in Sac-
become a good brewer?. Trends Genet 22: 183–186. charomyces cerevisiae. J Bacteriol 179: 7790–7795.
Posas, F., Chambers, J.R., Heyman, J.A., Hoeffler, J.P., de Sweet, G., Gandor, C., Voegele, R., Wittekindt, N., Beuerle,
Nadal, E., and Arino, J. (2000) The transcriptional J., Truniger, V., et al. (1990) Glycerol facilitator of Escheri-
response of yeast to saline stress. J Biol Chem 275: chia coli: cloning of glpF and identification of the glpF
17249–17255. product. J Bacteriol 172: 424–430.
Redkar, R.J., Locy, R.D., and Singh, N.K. (1995) Biosyn- Swinnen, S., Klein, M., Carrillo, M., McInnes, J., Nguyen,
thetic pathways of glycerol accumulation under salt stress H.T., and Nevoigt, E. (2013) Re-evaluation of glycerol uti-
in Aspergillus nidulans. Exp Mycol 19: 241–246. lization in Saccharomyces cerevisiae: characterization of
Rep, M., Krantz, M., Thevelein, J.M., and Hohmann, S. an isolate that grows on glycerol without supporting sup-
(2000) The transcriptional response of Saccharomyces plements. Biotechnol Biofuels 6: 157.
cerevisiae to osmotic shock. Hot1p and Msn2p/Msn4p Swinnen, S., Ho, P.W., Klein, M., and Nevoigt, E. (2016)
are required for the induction of subsets of high osmolari- Genetic determinants for enhanced glycerol growth of
ty glycerol pathway-dependent genes. J Biol Chem 275: Saccharomyces cerevisiae. Metab Eng 36: 68–79.
8290–8300. Tamas, M.J., Luyten, K., Sutherland, F.C., Hernandez, A.,
Roberts, G.G., and Hudson, A.P. (2006) Transcriptome pro- Albertyn, J., Valadi, H., et al. (1999) Fps1p controls the accu-
filing of Saccharomyces cerevisiae during a transition mulation and release of the compatible solute glycerol in
from fermentative to glycerol-based respiratory growth yeast osmoregulation. Mol Microbiol 31: 1087–1104.
reveals extensive metabolic and structural remodeling. Tani, T., and Yamada, K. (1987) Glycerol metabolism in
Mol Genet Genom 276: 170–186. methylotrophic yeasts. Agric Biol Chem 51: 1927–1933.
Roberts, G.G., 3rd, and Hudson, A.P. (2009) Rsf1p is Thome, P.E. (2004) Isolation of a GPD gene from Debaryo-
required for an efficient metabolic shift from fermentative myces hansenii encoding a glycerol 3-phosphate dehy-
to glycerol-based respiratory growth in S. cerevisiae. drogenase (NAD1). Yeast 21: 119–126.
Yeast 26: 95–110. Tom, G.D., Viswanath-Reddy, M., and Howe, H.B. Jr.
Romano, A.H. (1986) Microbial sugar transport systems (1978) Effect of carbon source on enzymes involved in
and their importance in biotechnology. Trends Biotechnol glycerol metabolism in Neurospora crassa. Arch Microbiol
4: 207–213. 117: 259–263.
Roth, S., Kumme, J., and Schuller, H.J. (2004) Transcrip- Turcotte, B., Liang, X.B., Robert, F., and Soontorngun, N.
tional activators Cat8 and Sip4 discriminate between (2010) Transcriptional regulation of nonfermentable carbon
sequence variants of the carbon source-responsive pro- utilization in budding yeast. FEMS Yeast Res 10: 2–13.
moter element in the yeast Saccharomyces cerevisiae. Uwajima, T., Shimizu, Y., and Terada, O. (1984) Glycerol
Curr Genet 45: 121–128. oxidase, a novel copper hemoprotein from Aspergillus
Ruijter, G.J., Visser, J., and Rinzema, A. (2004) Polyol japonicus. Molecular and catalytic properties of the
accumulation by Aspergillus oryzae at low water activity enzyme and its application to the analysis of serum trigly-
in solid-state fermentation. Microbiology 150: 1095– cerides. J Biol Chem 259: 2748–2753.
1101. Van Zyl, P.J., Kilian, S.G., and Prior, B.A. (1990) The role of
Schu €ller, H.J. (2003) Transcriptional control of nonfermenta- an active transport mechanism in glycerol accumulation
tive metabolism in the yeast Saccharomyces cerevisiae. during osmoregulation by Zygosaccharomyces rouxii.
Curr Genet 43: 139–160. Appl Microbiol Biotechnol 34: 231–235.
Schuurink, R., Busink, R., Hondmann, D.H., Witteveen, Vasiliadis, G.E., Sloan, J., Marshall, J.H., and May, J.W.
C.F., and Visser, J. (1990) Purification and properties of (1987) Glycerol and dihydroxyacetone metabolizing
NADP(1)-dependent glycerol dehydrogenases from enzymes in fission yeasts of the genus Schizosaccharo-
Aspergillus nidulans and A. niger. J Gen Microbiol 136: myces. Arch Microbiol 147: 263–267.
1043–1050. Viswanath-Reddy, M., Bennett, S.N., and Branch Howe,
Sformo, T., Walters, K., Jeannet, K., Wowk, B., Fahy, H.J. (1977) Characterization of glycerol nonutilizing and
G.M., Barnes, B.M., and Duman, J.G. (2010) Deep protoperithecial mutants of Neurospora. Mol Gen Genet
supercooling, vitrification and limited survival to 153: 29–38.
2100{degrees}C in the Alaskan beetle Cucujus clavipes Wang, Z.X., Zhuge, J., Fang, H., and Prior, B.A. (2001)
puniceus (Coleoptera: Cucujidae) larvae. J Exp Biol Glycerol production by microbial fermentation: a review.
213: 502–509. Biotechnol Adv 19: 201–223.

C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893
Glycerol catabolism in yeast 893
Watanabe, Y., Nagayama, K., and Tamai, Y. (2008) Expression dehydrogenase from a methylotrophic yeast, Hansenula
of glycerol 3-phosphate dehydrogenase gene (CvGPD1) in polymorpha DL-1, and its gene cloning. Acta Biotechnol
salt-tolerant yeast Candida versatilis is stimulated by high 22: 337–353.
concentrations of NaCl. Yeast 25: 107–116. Yamada-Onodera, K., Nakajima, A., and Tani, Y. (2006)
Workman, M., Holt, P., and Thykaer, J. (2013) Comparing cellular Purification, characterization, and gene cloning of glycerol
performance of Yarrowia lipolytica during growth on glucose dehydrogenase from Hansenula ofunaensis, and its
and glycerol in submerged cultivations. AMB Express 3: 58. expression for production of optically active diol. J Biosci
Yale, J., and Bohnert, H.J. (2001) Transcript expression in Bioeng 102: 545–551.
Saccharomyces cerevisiae at high salinity. J Biol Chem Young, E.T., Dombek, K.M., Tachibana, C., and Ideker, T.
276: 15996–16007. (2003) Multiple pathways are co-regulated by the protein
Yamada-Onodera, K., Yamamoto, H., Emoto, E., Kawahara, kinase Snf1 and the transcription factors Adr1 and Cat8.
N., and Tani, Y. (2002) Characterisation of glycerol J Biol Chem 278: 26146–26158.

C 2016 The Authors. Environmental Microbiology published by Society for Applied Microbiology and John Wiley & Sons Ltd,
V
Environmental Microbiology, 19, 878–893

You might also like