Yeast Cell Wall and Tolerance To Stress
Yeast Cell Wall and Tolerance To Stress
Yeast Cell Wall and Tolerance To Stress
REVIEWED BY
yeasts
Javier Arroyo,
Complutense University of Madrid,
Spain Ricardo A. Ribeiro1,2,3 , Nuno Bourbon-Melo1 and
Katy Kao, Isabel Sá-Correia1,2,3*
San Jose State University, United States
Antonio D. Moreno, 1
Institute for Bioengineering and Biosciences, Instituto Superior Técnico, Universidade de Lisboa,
Centro de Investigaciones Energéticas, Lisbon, Portugal, 2 Department of Bioengineering, Instituto Superior Técnico, Universidade
Medioambientales y Tecnológicas, de Lisboa, Lisbon, Portugal, 3 Associate Laboratory i4HB—Institute for Health and Bioeconomy
Spain at Instituto Superior Técnico, Universidade de Lisboa, Lisbon, Portugal
*CORRESPONDENCE
Isabel Sá-Correia
[email protected]
SPECIALTY SECTION In industrial settings and processes, yeasts may face multiple adverse
This article was submitted to environmental conditions. These include exposure to non-optimal
Microbial Physiology and Metabolism,
a section of the journal temperatures or pH, osmotic stress, and deleterious concentrations of diverse
Frontiers in Microbiology inhibitory compounds. These toxic chemicals may result from the desired
RECEIVED 26 May 2022 accumulation of added-value bio-products, yeast metabolism, or be present
ACCEPTED 11 July 2022
PUBLISHED 28 July 2022
or derive from the pre-treatment of feedstocks, as in lignocellulosic biomass
hydrolysates. Adaptation and tolerance to industrially relevant stress factors
CITATION
Ribeiro RA, Bourbon-Melo N and involve highly complex and coordinated molecular mechanisms occurring
Sá-Correia I (2022) The cell wall in the yeast cell with repercussions on the performance and economy
and the response and tolerance
to stresses of biotechnological of bioprocesses, or on the microbiological stability and conservation of
relevance in yeasts. foods, beverages, and other goods. To sense, survive, and adapt to different
Front. Microbiol. 13:953479.
doi: 10.3389/fmicb.2022.953479
stresses, yeasts rely on a network of signaling pathways to modulate the
global transcriptional response and elicit coordinated changes in the cell.
COPYRIGHT
© 2022 Ribeiro, Bourbon-Melo and These pathways cooperate and tightly regulate the composition, organization
Sá-Correia. This is an open-access and biophysical properties of the cell wall. The intricacy of the underlying
article distributed under the terms of
the Creative Commons Attribution regulatory networks reflects the major role of the cell wall as the first
License (CC BY). The use, distribution line of defense against a wide range of environmental stresses. However,
or reproduction in other forums is
permitted, provided the original the involvement of cell wall in the adaptation and tolerance of yeasts
author(s) and the copyright owner(s) to multiple stresses of biotechnological relevance has not received the
are credited and that the original
publication in this journal is cited, in
deserved attention. This article provides an overview of the molecular
accordance with accepted academic mechanisms involved in fine-tuning cell wall physicochemical properties
practice. No use, distribution or
during the stress response of Saccharomyces cerevisiae and their implication
reproduction is permitted which does
not comply with these terms. in stress tolerance. The available information for non-conventional yeast
species is also included. These non-Saccharomyces species have recently
been on the focus of very active research to better explore or control
their biotechnological potential envisaging the transition to a sustainable
circular bioeconomy.
KEYWORDS
polysaccharide fraction, and alone represent 30–60% of the There are two major classes of CWPs, namely the
cell wall dry weight (Fleet, 1991). Approximately 30–60% of glycosylphosphatidylinositol CWPs (GPI-CWPs), and the
the cell wall dry biomass is composed by β-glucans, around alkali-sensitive linkage CWPs (ASL-CWPs), which include
85% of these are 1,3-β-glucans and the remaining are 1,6- proteins of the internal repeats (Pir) family (Klis, 2002; Klis
β-glucans (Aguilar-Uscanga and François, 2003; Orlean, 2012; et al., 2006; Figure 1). While GPI-CWPs are typically linked
Figure 1). The β1,3-glucan chains are composed of ≈1500 to β1,6-chains by a GPI anchor, ASL-CWPs are directly linked
glucose units and are assembled in coiled spring-like structure to β1,3-glucans through an alkali-labile bond (Lesage and
that confer elasticity and tensile strength to the cell wall Bussey, 2006). They also differ in their distribution throughout
(Klis, 2002). The β1,6-glucans chains are shorter than the the cell wall. While GPI-CWPs are found in the outer layer,
β1,3-glucan chains and, although quantitatively being a minor Pir-CWPs seem to be uniformly distributed throughout the
component of the wall, β1,6-glucans have a central role in cross- inner layer, which is consistent with their direct association with
linking β1,3-glucans together (Kollár et al., 1997; Orlean, 2012). β1,3-glucans (Kapteyn et al., 1999b; Klis et al., 2006; Figure 1).
Additionally, β1,6-glucans can be connected to mannoproteins Cell wall biogenesis involves cell wall polysaccharide
with a glycosylphosphatidylinositol (GPI) anchor and to chitin synthases and enzymes involved in cell wall remodeling,
(Kollár et al., 1995; Orlean, 2012). The reducing ends of β1,3- assembly and degradation (Klis et al., 2006; Free, 2013;
glucans chains can be linked to a side-branching β1,6-glucan Figure 2). The β1,3-glucans are synthetized as a linear polymer
on β1,3-glucans chains, forming a fibrillar structure that serves by the FKS family of genes (FKS1-3). The FKS1 and FKS2
as backbone and anchorage point for other constituents of the genes code for β1,3-glucan synthases and differ mostly on the
cell wall (Lesage and Bussey, 2006; Figure 1). The non-reducing expression pattern, whereas FKS3 remains poorly characterized
ends of 1,3-β-glucans are linked to the reducing ends of chitin (Garrett-Engele et al., 1995; Mazur et al., 1995; Igual et al.,
through a β-1,4 link (Kollár et al., 1995, 1997; Orlean, 2012). 1996; Zhao et al., 1998; Lesage and Bussey, 2006). While FKS1
Chitin is a linear polymer of β1,4-linked expression is prevalent under optimal growth, FKS2 expression
N-acetylglucosamine (GlcNAc), representing less than 2% is induced in response to different stresses such as glucose
of S. cerevisiae cell wall dry weight (Figure 1). Chitin can occur depletion, alternative carbon sources (e.g., acetate, glycerol,
both in the free form, or bound to β-glucans (Kollár et al., galactose), high extracellular Ca2+ or heat stress (Garrett-
1995, 1997; Orlean, 2012). Chitin is normally concentrated as a Engele et al., 1995; Mazur et al., 1995; Igual et al., 1996;
ring between the mother cell and the emerging bud and in the Zhao et al., 1998). Several genes influence β-1,6-glucan levels,
lateral walls of the mother cell after septation (Cabib, 2009). In in particular the KRE6 and SKN1 genes, encoding proteins
response to stress, chitin levels can increase to as much as 20% involved in β1,6-glucan synthesis, and the KRE9 and KHN1
of the cell wall dry weight (Popolo et al., 1997; Dallies et al., encoding CWPs presumably involved in β1,6-glucan cross-link
1998; Ram et al., 1998; Osmond et al., 1999; Valdivieso et al., to other components of the cell wall (Roemer et al., 1994; Lesage
2000; Magnelli et al., 2002; Orlean, 2012). Changes occurring and Bussey, 2006). Chitin is synthetized as a linear polymer by
in cell wall nanomechanical properties, such as cell surface the chitin synthases encoded by the CHS1-3 gene family. The
stiffness, are mostly dependent on the crosslinking between gene CHS3, encoding chitin synthase III, is by far responsible
β-glucans and chitin (Dague et al., 2010) and occur in response for the majority of chitin synthesis, whether it is during
to stresses of industrial relevance (Dague et al., 2010; Pillet et al., optimal growth conditions or during stress response when
2014; Schiavone et al., 2016; Ribeiro et al., 2021, 2022). increased chitin deposition in the cell wall occurs (Roncero,
The outer layer is composed by cell wall mannoproteins 2002; Klis et al., 2006; Lesage and Bussey, 2006; Orlean, 2012).
(Figure 1). These mannoproteins are heavily glycosylated, Also, CWPs mannosylation requires several genes encoding
modified with both N-and O-linked carbohydrates, commonly proteins with mannosyltransferase activity (Gonzalez et al.,
formed by mannose (Klis et al., 2006; Schiavone et al., 2014). 2009a; Orlean, 2012). The GAS family (GAS1-5), encodes β1,3-
The outer layer serves an important protective role in the cell glucanosyltransferases involved in β1,3-glucan branching and
by limiting the access of external aggressors, such as foreign elongation, in which GAS1 has a major role and is also required
enzymes, to the inner layer (Klis et al., 2006). The proteins for β1,6-glucan linkage with β1,3-glucan chains (Ram et al.,
that constitute this layer are involved in a wide range of 1998; Ragni et al., 2007; Orlean, 2012; Aimanianda et al.,
functions often related to cell-to-cell interactions (e.g., mating, 2017). Shortening of glucan chains is also required for cell wall
flocculation, biofilm formation, etc.) (Klis, 2002). Increased cell remodeling and involves BGL2 and EXG1-2 encoding a major
wall hydrophobicity influences flocculation leading to increased endoglucanase and major exoglucanases, respectively (Larriba
robustness to inhibitory chemical compounds (Kahar et al., et al., 1995; Lesage and Bussey, 2006; Orlean, 2012; Aimanianda
2022). This was associated with changes in the expression of et al., 2017). The transglycosylases GPI proteins encoded by the
MOT3 gene, encoding a transcription regulator that controls the CRH1-2 genes have a major role in the cross-linkage between
expression of a cell wall protein (CWPs) encoding gene, YGP1, chitin and glucans (Cabib et al., 2007, 2008; Cabib, 2009; Orlean,
that influences cell wall hydrophobicity (Kahar et al., 2022). 2012; Blanco et al., 2015). Endochitinases, encoded by CTS1-2,
FIGURE 1
Schematic representation of S. cerevisiae cell wall. The inner layer of the cell wall is mostly composed of β1,3-glucan chains branched with
β1,6-glucans, and chitin. The outer layer is composed of mannoproteins, most of which are linked to the inner wall by a GPI anchor, whereas
ASL-Pir proteins seem to be uniformly distributed throughout the inner layer. ASL-CWP, alkali-sensitive linkage cell wall protein. GPI-CWP,
glycosylphosphatidylinositol cell wall protein. Details in the main text.
are required for septation and cell separation (Kuranda and in response to environmental external stimuli that may damage
Robbins, 1991; Cabib et al., 1992; Orlean, 2012). the cell wall (Gustin et al., 1998; Levin, 2011; Sanz et al., 2018;
Saccharomyces cerevisiae cell wall appears to have many Figure 2). CWI pathway relies on a family of plasma membrane
organizational similarities with the walls of other ascomycetous sensors, namely Wsc1-3, Mid2, and Mtl1. Stimuli that may
yeasts and even of basidiomycetous yeasts (Jaafar and Zueco, impose a stress to the cell are detected by these sensors and
2004; De Groot et al., 2005, 2008; Patel and Free, 2019; Garcia- triggers a phosphorylation cascade that elicit transcriptional
Rubio et al., 2020). However, there are differences between the changes to enable the cell to adapt to stress conditions (Kock
cell walls of S. cerevisiae and other fungi. For instance, the et al., 2016; Sanz et al., 2018; Figure 2). For the Wsc1 sensor,
cell wall of S. cerevisiae does not contain α-glucans, chitosan, the signaling transduction to the downstream components of
or melanin, and its chitin content is relatively low, especially the pathway is influenced by the spatial distribution of these
when compared to other yeasts or filamentous fungi (e.g., sensors within the plasma membrane, requiring a functional
Candida albicans, Aspergillus niger, Paracoccidioides brasiliensis, extracellular cysteine-rich domain to form clusters in specific
Blastomyces dermatitidis) (Kanetsuna et al., 1969; Garcia-Rubio microdomains or rafts (Heinisch et al., 2010; Schöppner
et al., 2020; van Leeuwe et al., 2020). et al., 2022). Industrially relevant stresses reported to trigger
a response by the referred CWI pathway membrane sensors
are highlighted in Table 1. Together with phosphatidylinositol
4,5-biphosphate, PIP2 , which recruits the guanine nucleotide
Cell wall-related signaling exchange factors (GEFs) Rom1 and Rom2, the surface sensors
pathways induced by stress activate the Rho1 GTPase (Levin, 2011; Sanz et al., 2018). Rho1
regulates both β1,3-glucan synthases encoded by FKS1 and
Yeast cells sense and respond to environmental stresses FKS2, and β1,6-glucan synthase activities, and is considered a
through the induction and activity of different signaling key regulator of CWI signaling (Levin, 2011). Rho1 activates
pathways (Chen and Thorner, 2007). Depending on the type Pkc1 and sets in motion a series of phosphorylation events
of stress, specific pathways can be triggered and directly, or which sequentially activate Bck1, Mkk1/2, and the pathway’s
indirectly, elicit changes in the composition and architecture of MAPK, Slt2 (Levin, 2011; Figure 2). Finally, Slt2 activates
the cell wall (Levin, 2011; Figure 2). The CWI signaling pathway two transcription factors, Rlm1 and SBF (Swi4/Swi6 complex),
is one of the most well-described pathways, characterized for that coordinate the CWI transcriptional response (Levin, 2011;
its role in cell wall maintenance and homeostasis, particularly Orlean, 2012; Figure 2). This response mostly involves the
FIGURE 2
Schematic representation of the CWI, HOG, and calcineurin signaling pathways and interactions. The cell wall integrity pathway (CWI), in blue, is
the main pathway involved in the maintenance of yeast cell wall integrity. It is frequently activated by a signal starting at plasma membrane
sensors (Wsc1-3, Mid2, Mtl1), which then trigger a cascade culminating in the phosphorylation of Slt2. Phosphorylated Slt2 activates the
transcription factor Rlm1 and the SBF complex (Swi4/6), triggering a transcriptional response that elicit changes at cell wall level.
Phosphoinositide synthesis (in orange) plays an important role in the activation of the CWI pathway by recruiting Rom1/2 to the plasma
membrane, which activates the main regulator of the pathway, the GTPase Rho1. The CWI pathway receives input and interplays with other
pathways, including the high osmolarity glycerol (HOG) pathway (in green) and the calcineurin pathway (in pink), represented in the figure, but
also with the protein kinase A (PKA) and TOR signaling pathways. Details in the text.
activation of cell wall biogenesis genes. The Rlm1 transcription revealed for other yeast species (Donlin et al., 2014; Valiante
factor is responsible for the expression of most of the genes et al., 2015; Madrid et al., 2016; Assis et al., 2018).
induced in response to cell wall stress, such as the CHS3 gene, Changes occurring at the level of plasma membrane can also
and also regulates the expression of FKS1 (Jung and Levin, 1999; elicit responses at the level of the cell wall. Plasma membrane
Klis et al., 2006). Importantly, Rlm1 activates MLP1, coding for stretching appears to be the main factor in activating the
a Slt2 pseudo-kinase paralog, that, together with Slt2, activates CWI pathway in response to a number of stresses, including
the SBF complex for transcription of a subset of cell wall stress- high osmotic pressure and supra-optimal temperature (Kamada
activated genes. Among them are genes related with glucan et al., 1995; Hohmann, 2002b). This is consistent with the CWI
synthesis (FKS1, FKS2) or glucan elongation and branching pathway-activating sensors being located at the membrane and
(GAS1) (Igual et al., 1996; Baetz et al., 2001; Kim et al., 2008; the likely role of Wsc1 as a mechanosensor (Kock et al., 2015).
Levin, 2011; Harris et al., 2013)This brief description of the CWI Alterations of plasma membrane lipid composition have also
pathway as a simple linear cascade of events can be deceiving. In been shown to impact the CWI pathway, namely, a defective
fact, the CWI pathway receives lateral influences from cAMP- biosynthesis of the complex sphingolipid mannosylinositol
Protein Kinase A (PKA) signaling, TOR signaling, calcineurin phosphorylceramide (MIPC) leads to increased abundance of
signaling, the HOG pathway, and likely from others not yet the phosphorylated form of Slt2 and sensitivity to cell wall-
clarified (Fuchs and Mylonakis, 2009; Rodríguez-Peña et al., perturbing agents (Morimoto and Tani, 2015). This phenotype is
2010; García et al., 2017; Udom et al., 2019; Jiménez-Gutiérrez partly suppressed by the upregulation of ergosterol biosynthesis,
et al., 2020a,b; Figure 2). This complex network is what enables suggesting that MIPC and ergosterol are coordinately involved
the CWI pathway to be activated by numerous types of stressors, in maintaining the integrity of the cell wall (Tanaka and Tani,
ensuring an adequate response to each stress or group of 2018). Also, a reported crosstalk between plasma membrane
stresses (Figure 2). The existence of a complex interplay between ergosterol content (related with the level of expression of the
the CWI pathway and other signaling pathways is also being plasma membrane ergosterol transporter Pdr18), and cell wall
TABLE 1 Cell wall integrity membrane sensors implicated in the ethanol, acetic acid or lactic acid lead to oxidative stress
sensing of industrially relevant stresses.
that occurs when cellular defense mechanisms are unable to
CWI membrane Industrially relevant Bibliographic cope with existing ROS (Costa and Moradas-Ferreira, 2001;
sensor stress references Abbott et al., 2009; Teixeira et al., 2009; Morano et al., 2012;
Charoenbhakdi et al., 2016).
Wsc1 Hyper-osmolorarity García-Rodríguez et al.,
Heat 2005 The cell wall or, more broadly, the cell envelope, has
Acetic acid Rajavel et al., 1999 been associated with oxidative stress toxicity and tolerance.
Alkaline pH Mollapour et al., 2009
Diamide and H2 O2 Serrano et al., 2006;
This involvement has been unveiled by studies focused on
Kwon et al., 2016 S. cerevisiae exposure to pro-oxidant agents such as hydrogen
Vilella et al., 2005 peroxide, lipid hyperoxides, diamine, catecholamines, and
Wsc2 H2 O2 Vélez-Segarra et al., 2020 organic hydroperoxides, such as cumene hydroperoxide (CHP)
Wsc1/Wsc2 Impaired mannosylinositol Tanaka and Tani, 2018 and linoleic acid hydroperoxide (LoaOOH) (Staleva et al.,
phosphorylceramide
metabolism 2004; Vilella et al., 2005; Petkova et al., 2010; Taymaz-Nikerel
Mid2 Hyper-osmolorarity García-Rodríguez et al., et al., 2016). Oxidative stress resulting from exposure to
Heat 2005 these agents induces distinct responses in S. cerevisiae. For
Low pH – media acidified Rajavel et al., 1999
instance, a quantitative proteomics study reported differences
with a strong acid Claret et al., 2005
H2 O2 Jin et al., 2013 in the activation of the CWI pathway between hydrogen
peroxide, CHP and diamide (Pandey et al., 2020). Decreased
Mtl1 Diamide and H2 O2 Vilella et al., 2005; cell permeability, a property influenced by the thickness and
Jin et al., 2013
composition of the cell wall and plasma membrane, results in
increased resistance to pro-oxidant compounds by limiting their
biophysical properties suggested a coordinated response to diffusion into the cell (Sousa-Lopes et al., 2004). Furthermore,
counteract acetic acid deleterious effects, reinforcing the notion membrane lipid composition is a determinant of oxidative stress
that plasma membrane lipid composition influences cell wall resistance, with cells containing a higher level of saturated fatty
integrity during stress (Ribeiro et al., 2022). acids being more resistant than cells with a higher level of
For the main and better studied stresses encountered polyunsaturated fatty acids (Jamieson, 1998).
during industrial bioprocesses, this review provides, whenever Diamide induces the formation of disulfide bonds in the
possible, an integrative view of the pathways and responses that three-dimensional structure of CWPs, causing changes in the
cooperatively maintain cell wall integrity and, in so doing, helps morphology of the cell outer layer and the increase of cell
yeasts to resist to multiple stresses. wall thickness (De Souza Pereira and Geibel, 1999; Vilella
et al., 2005). CHP causes oxidative damage to the cell wall
periphery leading to the upregulation of genes related to cell
Yeast response and tolerance to wall biogenesis (HSP150), CWI pathway regulation (RHO1,
ROM2) and β1,6-glucan synthesis (KRE5, KRE6, KHN1) (Sha
industrially relevant stresses et al., 2013; Figure 3). The transcriptional reprogramming of
involving the cell wall yeast in response to oxidative stress is mainly regulated by
two oxidative stress-responsive transcription factors, Yap1 and
The cell wall in the response to Skn7, and the general stress transcription factors, Msn2/Msn4
oxidative stress (Auesukaree, 2017). Notably, Skn7 has an important role in the
regulation of genes involved in the maintenance of cell wall
All aerobically growing organisms suffer exposure to integrity, including genes coding for CWPs (Alloush et al., 2002;
oxidative stress, caused by reactive oxygen species (ROS) Li et al., 2002; Levin, 2011). The CWI pathway sensors Wsc1,
capable of damaging cellular DNA, lipids, carbohydrates, and Wsc2, Mtl1, and Mid2 play an important role in the sensing of
proteins, threatening cell integrity (Jiménez-Gutiérrez et al., oxidative stress induced by H2 O2 (Vilella et al., 2005; Jin et al.,
2020b). Consequently, mechanisms to protect cell components 2013; Vélez-Segarra et al., 2020). Surviving and overcoming
against ROS were evolved and the antioxidant defenses can oxidative stress induced by H2 O2 , as for diamine, requires Pkc1
be induced either by respiratory growth or in the presence of and the upstream element of the CWI pathway Rom2 (Vilella
pro-oxidants (Moradas-Ferreira and Costa, 2000). Yeasts have et al., 2005; Figure 3).
several inducible adaptive stress responses to oxidants regulated In Yarrowia lipolytica, a promising oleaginous yeast
at the transcriptional and posttranscriptional levels (Jamieson, (Ledesma-Amaro and Nicaud, 2016), oxidative stress induced
1998; Moradas-Ferreira and Costa, 2000). Stresses commonly by H2 O2 leads to changes in the morphology of the cell
arising during industrial fermentations, such as supra-optimal wall, with the formation of globular surface structures in the
temperatures or presence of inhibitory concentrations of cell wall surface (Arinbasarova et al., 2018). When exposed
FIGURE 3
Schematic model of the described S. cerevisiae adaptive responses common to multiple industrial relevant stresses involving the cell wall. The
cell wall integrity pathway (CWI) and other signaling pathways (HOG and Calcineurin pathway) that, in collaboration with the CWI pathway and
the upregulation of cell wall genes, lead to changes in the composition and nanomechanical properties of the cell wall in response to different
stresses are highlighted. Squares of different colors indicate the different industrial-relevant stresses for which data is available in the literature.
to soluble complexes of UO2, the mRNA levels from genes of cellular turgor (Davenport et al., 1995; Hohmann, 2002a).
involved in elongation of β1,3-glucan and chitin synthesis Although there are many similarities in how the yeast cell
were reduced in a tropical marine strain of Y. lipolytica, able responds to osmotic stress caused by different agents, thus
to immobilize in the cell surface uranium often associated allowing the description of a generalized stress response, there
with oxidative stress (Kolhe et al., 2020). These complexes, are also specific features for each osmolyte or salt stress
formed above pH 7.0, are prevalent in aquatic environments (Gonzalez et al., 2016).
such as rivers or sea water, being this strain considered Sudden exposure to increased osmolarity leads to cell
promising for the bioremediation of uranium-contaminated shrinkage due to water efflux and likely causes the release of
aquatic environments (Kolhe et al., 2020, 2021). plasma membrane material, as well as changes in cell wall
structure (Morris et al., 1983; Claro et al., 2007; Saldaña et al.,
2021). The decrease in cell volume is accompanied by changes
Osmotic stress in cell morphology and surface roughness (Adya et al., 2006;
Saldaña et al., 2021). Intracellularly, a compensatory response is
Hyper-osmotic stress is common in industrial bioprocesses triggered, recruiting water from the vacuole into the cytoplasm
(Mattanovich et al., 2004). Very high gravity fermentations are (Mager and Siderius, 2002). These changes occur immediately
frequently used to enhance product titres in specific sectors, after an osmotic upshift, likely caused by rapid biophysical forces
such as in first-generation bioethanol production by increasing (Ene et al., 2015). To survive such a shock, the properties of
initial sugar concentration and, consequently, osmotic pressure the cell wall have to be modified to have enough elasticity
(Deparis et al., 2017). In winemaking and baking, yeasts to respond to shifts in osmolarity, while maintaining enough
also have to grow in hyper-osmotic environments and their rigidity to preserve cell morphology and integrity (Ene et al.,
performance depends on the ability to adapt and respond 2015). The rapid changes of cellular volume and shape occurring
to this stress (Attfield and Kletsas, 2000; Noti et al., 2015). in response to hyperosmotic shock that are fully reversible when
Also, the increase of osmotic pressure induced by the addition cells are introduced in an isosmotic solution, are evidences of
of stressful concentrations of sugars or salts is beneficial for the remarkable plasticity of the cell wall (Saldaña et al., 2021).
food preservation (Gonzalez et al., 2016). In yeasts, the High The HOG pathway (Figures 2, 3) is a MAPK signal
Osmolarity Glycerol (HOG) and the CWI signaling pathways transduction system and the major pathway in the adaptation
are central to osmotic stress-induced response and control of yeast cells to increased osmolarity (Hohmann, 2002a). The
pathway relies on two osmolarity sensors, Sln1 and Sho1. During of ZrFPS1 (Pøibylová et al., 2007; Hou et al., 2013; Guo
an osmotic upshift, Sln1 is transiently inhibited, diminishing the et al., 2020). Among Z. rouxii upregulated genes under high
levels of phosphorylated Ssk1. Unphosphorylated Ssk1 can then osmolarity stress are FKS1(encoding a β1,3-glucan synthase),
interact with Ssk2/Ssk22, which in turn phosphorylates Pbs2. UTR2 (encoding a cell wall transglycosylase), KRE9 (encoding
The Sho1 osmosensor also transmits a signal to Pbs2 via the a glycoprotein involved in β-glucan assembly), CHS1 (encoding
MAPKKK Ste11 (Hohmann, 2009). The phosphorylated Pbs2 a chitin synthase) and KAR2 (encoding an ATPase involved
activates Hog1 and modulates the expression of several genes β1,6-glucan synthesis and also involved in the translocation of
(Hohmann, 2009), including the transcriptional activation proteins into the endoplasmic reticulum) (Guo et al., 2020). The
of GPD1 and GPP2 genes, encoding a glycerol 3-phosphate upregulation of the encoding genes helps to explain why the cell
dehydrogenase and a glycerol-1-phosphatase respectively, wall became thicker as the cell volume decreases, resulting in a
involved in glycerol production (Hohmann, 2009). In response smaller amplitude of cell size variation (Guo et al., 2020).
to hyper-osmotic stress, Hog1 phosphorylates Rgc2, a regulator In a Zygosaccharomyces mellis strain, isolated from honey
of the Fps1 glycerol channel, causing the dissociation of Rgc2 and tolerant to high-glucose concentrations (Liu et al., 2016,
from Fps1 and the consequent glycerol channel closure (Lee 2021), the genes KRE5 (involved in β1,6-glucan synthesis) and
et al., 2013). Together, the blockade of glycerol efflux by SLT2 (encoding a kinase of the CWI pathway), are upregulated
the closing of the Fps1 channel and the increase in glycerol under hyper-osmotic stress (Liu et al., 2021). This indicates that,
biosynthetic capacity lead to glycerol accumulation in the in Z. mellis, the maintenance of cell wall integrity under this
cytoplasm and effectively counteract the loss of turgor pressure stress is also important. In D. hansenii, the cell wall was also
caused by an osmotic upshift (Jiménez-Gutiérrez et al., 2020b). shown to play a critical role in osmosensing and genes involved
A coordinated interplay between the HOG pathway and the in CWP mannosylation (MNN1, PMT2, PSA1, and MNT1) are
CWI pathway allows yeast cells to better adapt to hyper-osmotic upregulated in response to hyper-osmotic stress (Thomé, 2007;
stress conditions (García-Rodríguez et al., 2005; Vaz et al., 2020; Gonzalez et al., 2009b).
Figure 3). The CWI pathway appears to be indirectly activated
by an increase in turgor pressure caused by the accumulation
of glycerol, leading to transient phosphorylation of Slt2 that Heat stress
depends on a functional HOG pathway (García-Rodríguez et al.,
2005). Nevertheless, the CWI sensors Mid2 and Wsc1 seem Another obstacle that can be faced by yeast cell factories
to be involved in the sensing of Hog1-driven accumulation of is temperature stress, in particular heat stress. Heat stress is
glycerol (García-Rodríguez et al., 2005). Further evidences of common during alcoholic fermentations to produce alcoholic
the interplay between the HOG pathway and the CWI pathway beverages or bioethanol due to their exothermic nature. If the
were reported (Vaz et al., 2020). Chemogenomic studies have temperature is not controlled, a significant rise in temperature
implicated genes related to β1,3-glucan synthesis (FKS1), β1,6- occurs (Parapouli et al., 2020; Walker and Basso, 2020).
glucan synthesis (KRE6) and cell wall mannosylation (MNN10, Given that in the presence of other stresses, in particular of
ANP1) as determinants of tolerance to high-glucose and sucrose ethanol or acetic acid stress, the optimum and the maximum
concentrations (Ando et al., 2006; Teixeira et al., 2010), known temperatures for growth decrease as the stress level increase,
to induce hyper-osmotic stress (Deparis et al., 2017). even temperatures close to the optimal range of growth
In Y. lipolytica, hyper-osmotic stress was also found to can become lethal temperatures, depending on the level of
lead to cell wall remodeling (Kubiak-Szymendera et al., 2022). stress (van Uden, 1985; Godinho et al., 2021). Despite the
YlHOG1 deletion impacts filamentous growth, cytokinesis, and importance of cold stress, the related literature is scarce in the
resistance to cell wall perturbing agents (Rzechonek et al., context of this review, so this part of the article is essentially
2018). A quantitative proteomic analysis has shown that, dedicated to heat stress.
among the upregulated proteins under hyper-osmotic stress is At supra-optimal temperatures in the absence of any other
Pil1, involved in formation of membrane-associated protein stress (higher than 35–37◦ C), S. cerevisiae activates the heat
complexes commonly referred as eisosomes and distributed shock response (HSR) and undergoes physiological changes
across the cell surface periphery, and UTR2 coding for a CWP, which include membrane and cell wall restructuring (Verghese
involved in glucan-chitin crosslinking in S. cerevisiae (Kubiak- et al., 2012; Schiavone et al., 2014). During heat stress, β1,6-
Szymendera et al., 2022). Other non-Saccharomyces species, glucan and chitin content increases by 20 and 100% respectively,
such as Zygosaccharomyces rouxii, Debaryomyces hansenii, and and β1,3-glucans levels decrease by 45% (Schiavone et al.,
Pichia sorbitophila are notable for their osmotolerance (Bubnová 2014; Figure 3). The increased β1,6-glucan synthesis under heat
et al., 2014). The spoilage yeast of high-sugar or high-salt stress is in accordance with a higher number of cross-linkages
foods Z. rouxii appears to react to hyperosmolarity similarly between this polymer and chitin, presumably compensating
to S. cerevisiae, accumulating intracellular glycerol through cell wall weakening during this stress (Kapteyn et al., 1999a;
the increased expression of ZrGPS1 and decreased expression Schiavone et al., 2014). Heat stress also induces changes in
the morphology of the yeast cell surface with the formation of often required to allow an adequate fermentative performance
circular structures in the surface of heat-stressed cells (Pillet (Lehnen et al., 2019; Thorwall et al., 2020). Two prominent
et al., 2014). The emergence of these structures is, apparently, examples of thermotolerant yeast species, shown to be able to
related to a dysfunction in the budding machinery, accompanied grow relatively well above 40◦ C, are Kluyveromyces marxianus
by a concomitant increase in chitin and cell wall stiffness, and Ogataea polymorpha (formerly Hansenula polymorpha)
regulated by the CWI pathway (Pillet et al., 2014; Figure 3) (Lehnen et al., 2019; Thorwall et al., 2020). In K. marxianus,
activated in response to heat stress (Torres et al., 1991; Lee high temperatures were shown to lead to upregulation of genes
and Levin, 1992; Irie et al., 1993). Together with Msn2/4, associated with changes in plasma membrane composition (Fu
Hsf1 is a transcription factor responsible for the bulk of the et al., 2019). Very few is reported concerning changes occurring
HSR (Imazu and Sakurai, 2005; Truman et al., 2007). Heat in K. marxianus cell wall under heat stress but a strain isolated
stress also promotes glycerol efflux by the opening of Fps1 from agave when grown at 42◦ C exhibit increased sensitivity
channels regulated by the CWI pathway (Dunayevich et al., to lyticase than when grown at 30◦ C, indicating that heat
2018). The resulting turgor loss stimulates the activation of stress may affect cell wall integrity (Castillo-Plata et al., 2021).
the HOG pathway, promoting glycerol production and the In Ogataea species, thermotolerance has been attributed to
re-establishment of turgor pressure (Dunayevich et al., 2018). a structural predisposition of the cell envelope, related with
Heat shock leads to the intracellular accumulation of trehalose membrane and cell wall composition, allowing a higher cell
in S. cerevisiae (Attfield, 1987; Hottiger et al., 1987). Since envelope stability (Lehnen et al., 2019). Among the candidate
trehalose and thermotolerance are closely related, trehalose was genes that contribute to heat tolerance in O. polymorpha is
suggested to act as a thermoprotectant (Hottiger et al., 1989). an ortholog of S. cerevisiae PSA1, coding for GPD-mannose
The intracellular accumulation of trehalose causes a decrease pyrophosphorylase, involved in cell wall biosynthesis (Seike
of the specific growth rate that may trigger the environmental et al., 2021). In the genus Ogataea, Mpk1 appears to be involved
stress response (ESR) and higher thermotolerance (Gibney in CWI signaling in response to heat stress, but, different
et al., 2015). The accumulation of trehalose increases cytosolic Ogataea strains exhibit different growth phenotypes when
osmolarity and turgor pressure, mimicking hypotonic stress, MPK1 is absent (Kim et al., 2018). Compared to O. polymorpha,
and causing plasma membrane stretching known to lead O. parapolymorpha has a thinner β-glucan and chitin layer
to activation of the CWI pathway (Kamada et al., 1995; with short mannan chains and the derived deletion mutant
Mensonides et al., 2005). Consistently, preventing trehalose mpk11 exhibit more severe growth defects during heat stress
synthesis by deletion of the TPS1 gene led to decreased and higher susceptibility to cell wall-perturbing agents, leading
activation of the CWI pathway upon exposure of yeast cells the authors to consider these differences related to cell wall
to heat stress (Mensonides et al., 2005). Interestingly, a recent structural differences (Kim et al., 2018).
study suggests that the UDP-Glucose Pyrophosphorylase Ugp1 The characterization of the yeast response to low
is required for heat stress response by influencing trehalose and temperatures is also important in industry-related bioprocesses
glucan content (Zhang and Xu, 2021). (Liszkowska and Berlowska, 2021), in particular in many wine
Although the mechanism behind heat-induced activation and beer fermentations, resulting in the retention of more
of the CWI pathway is not clear, the Wsc1 and Mid2 CWI volatile compounds that influence the sensory properties of
sensors appear to be involved in the activation of CWI pathway, the product (Liszkowska and Berlowska, 2021). Furthermore,
reinforcing the notion that this stress is ultimately transmitted cold-adapted spoilage yeasts can potentially impose health risks
to the cell surface (Verna et al., 1997; Rajavel et al., 1999; Martín to the consumers and economic burden in the Food industry,
et al., 2000; Levin, 2011; Verghese et al., 2012). In fact, Wsc1 being able to grow and proliferate at temperatures at which
sensors form clusters in the plasma membrane upon heat stress food products are refrigerated (Fleet, 2011). Among the genes
(the so-called Wsc1 sensosomes) enhancing the CWI pathway that are upregulated in response to cold temperatures are genes
signaling capability (Heinisch et al., 2010). The expression of the encoding cell wall mannoproteins (TIR1-2, TIR4, TIP1, and
CWI pathway SLT2 gene, the HSP150 gene encoding a protein DAN1) (Abramova et al., 2001; Sahara et al., 2002; Homma
required for cell stability and the CWP-encoding gene YGP1 et al., 2003; Schade et al., 2004; Murata et al., 2006; Abe, 2007;
are induced in response to heat stress (Causton et al., 2001; Figure 3). Increased resistance to the cell wall-perturbing
Varol et al., 2018; Figure 3). Also, FKS2, encoding a β1,3-glucan compound SDS in cold-shock-stress-induced cells has also
synthase, is upregulated during heat stress, being its expression been attributed, at least partially, to the upregulation of DAN1
regulated by both CWI and Calcineurin pathways (Zhao et al., (Abe, 2007). The CWI pathway WSC1, BCK1, and SLT2 genes
1998; Figure 3). encoding a CWI sensor, the MAPKKK Bck1 and the MAPK
The ability to withstand elevated temperatures while Slt2, respectively, have also been implicated in response and
maintaining high growth rates and ethanol productivity can be adaptation to cold temperature-induced-stress, being Slt2
beneficial for bioethanol production and could alleviate some phosphorylation partially dependent on the Wsc1 CWI sensor
of the costs associated with the cooling of the bioreactors (Córcoles-Sáez et al., 2012; Figure 3). Psychrophilic yeasts,
yeasts adapted to low temperatures, with an optimal growth mannose backbone present in CWPs, and LDB7, encoding a
performance at 15◦ C, are studied due to its biotechnological component of the chromatin structure remodeling complex,
potential, in particular for the production of cold-active involved in the regulation of the mannosylphosphate content
enzymes (Buzzini et al., 2012). Interestingly, the absence of the of mannoproteins (Kubota et al., 2004; Auesukaree et al., 2009;
SWI4 or SWI6 genes, encoding transcriptional activators of Teixeira et al., 2009). Moreover, the CWP-encoding genes PIR3,
the CWI pathway in Metschnikowia australis W7-5, leads to SED1, SPI1, and YGP1, as well as the GPI-CWPs-encoding
impaired growth at temperatures as low as 5◦ C, suggesting that genes of the TIR family (TIR1-3 and TIP1), were found to
these genes are determinants of tolerance to low temperature in be transcriptionally responsive to ethanol stress (Ogawa et al.,
this species (Wei et al., 2021). 2000; Rossignol et al., 2003; Wu et al., 2006; Udom et al.,
2019; Figure 3). The overrepresentation of genes related to the
mannoprotein-rich outer layer of the cell wall in response to
Ethanol – and methanol – induced ethanol-induced stress suggests that mannoproteins have a role
stresses in yeast adaptation to this stress, limiting ethanol’s access to
the plasma membrane and thus counteracting ethanol-induced
Ethanol toxicity is arguably the major environmental stress membrane permeabilization and subsequent deleterious effects.
limiting industrial titres and overall productivity in a wide Due to its amphiphilic nature, ethanol can bind to the exposed
range of industrial yeast fermentations (Gibson et al., 2007; proteins at the outer layer of the cell wall, likely altering its
Parapouli et al., 2020). Accumulation of high concentrations organization and increasing cell wall porosity (Zlotnik et al.,
of ethanol, whether during bioethanol production, wine 1984; De Nobel et al., 1990; Udom et al., 2019).
and brewing industries, or other alcoholic fermentations, Genes related to chitin and glucan synthesis, namely CHS1
often results in decreased fermentation yield or complete (chitin), FKS1 (β1,3-glucans), ACF2 (β1,3-glucans), and KRE6
cessation of yeast metabolic activity and fermentation arrest (β1,6-glucans), as well as genes involved in cell wall integrity
(Walker and Basso, 2020). (PUN1) and cell wall organization (SIT4), were also found
Due to its liposolubility, ethanol disrupts plasma to be involved in ethanol tolerance (Kubota et al., 2004;
membrane lipid organization, increasing its fluidity, non- Auesukaree et al., 2009; Teixeira et al., 2009; Mota et al.,
specific permeability, and compromising transmembrane 2021). The upregulation of FKS1 and also of SED1 and
electrochemical potential (van Uden, 1985). Among the SMI1 genes, involved in cell wall biosynthesis depends on the
adaptive responses is the alteration of plasma membrane Znf1 transcription factor, recently implicated in adaptation
composition (e.g., increase in ergosterol content and in to ethanol stress (Samakkarn et al., 2021). Consistent with a
unsaturated/saturated fatty acid ratio) leading to increased transcriptional response involving multiple components of the
lipid order and counteracting plasma membrane fluidisation cell wall under ethanol stress, elements of the CWI signaling
(Navarro-Tapia et al., 2018), together with and adaptive pathway, namely the membrane sensors Mid2 and Wsc1, the
response at the level of the cell wall (Ding et al., 2009; Stanley MAPKKK Bck1, and the MAPK Slt2, were found to be crucial
et al., 2010). Remarkably, the impairment of plasma membrane for maximum tolerance to ethanol stress (Fujita et al., 2006;
integrity was shown to influence the nanomechanical properties Auesukaree et al., 2009; Teixeira et al., 2009). Also, yeasts
of the cell wall during ethanol stress (Schiavone et al., 2016) deficient in either one of the two Slt2-activated SBF subunits,
(Figure 3). Specifically, ethanol-induced changes to the Swi4 and Swi6, reveal higher sensitivity to ethanol (Udom
plasma membrane, compromising the proper delivery of et al., 2019). The SBF complex controls the expression of
plasma membrane-anchored GPI-CWPs to the membrane FKS2 and CRH1, coding for a β1,3-glucan synthase and a
via the secretory pathway, affect their crosslinking to cell chitin transglycosylase involved in glucan-chitin crosslinking,
wall polysaccharides resulting in reduced cell wall stiffness respectively; these genes were shown to be upregulated during
(Schiavone et al., 2016; Figure 3). ethanol stress (Udom et al., 2019; Figure 3). Additionally, the
Several chemogenomic studies have implicated genes related HOG pathway also seems to contribute to the regulation of
to cell wall biosynthesis, cell wall remodeling, and CWI pathway the transcriptional response of cell wall genes during ethanol
as determinants of ethanol tolerance (Kubota et al., 2004; Fujita stress, since the dysfunction of the HOG pathway leads to
et al., 2006; van Voorst et al., 2006; Auesukaree et al., 2009; decreased expression of FKS2 and CRH1 (Udom et al., 2019).
Teixeira et al., 2009). Additionally, suitable supplementation Furthermore, the wall of yeast cells exposed to ethanol exhibits
of the fermentation medium with potassium, zinc or inositol higher resistance to lyticase and zymolyase treatment (Udom
was shown to improve tolerance to ethanol stress (Krause et al., 2019). In summary, the mechanism underlying ethanol
et al., 2007; Zhao et al., 2009; Xu et al., 2020). Genes tolerance involves a collaborative role of the CWI and HOG
reported as determinants of ethanol tolerance include MNN10, signaling pathway in the transcriptional regulation of cell
MNN11, ANP1, and HOC1, encoding four subunits of the wall genes leading to cell wall changes (Udom et al., 2019;
polymerase complex responsible for the elongation of the Figure 3).
Not many non-conventional yeasts can compete with pathway for activation of methanol-inducible genes (Ohsawa
S. cerevisiae when it comes to ethanol production and tolerance, et al., 2017). Additionally, differences in methanol metabolism,
but some have other interesting features that could make their vector transformation efficiency, growth and heterologous
use in the bioethanol industry appealing. K. marxianus, known protein production between different K. phaffii strains were
for its thermotolerance and ability to metabolize several carbon related with cell wall integrity (Zhang et al., 2020), providing
sources (Thorwall et al., 2020), has a fair number of studies another example of the important role of cell wall.
dedicated to understanding the response to ethanol stress (Diniz
et al., 2017; Alvim et al., 2019; Mo et al., 2019; da Silveira
et al., 2020), although not much is known about ethanol- Low- or high-pH-induced stress
induced changes to its cell wall. However, the adaptive evolution
under ethanol stress was found to lead to the improvement Cell wall composition and architecture and the CWI
of multiple pathways, including cell wall biogenesis, suggesting pathway were also implicated in yeast adaptive response to
that cell wall remodeling is part of a strategy to mitigate acid and alkaline stress conditions (Claret et al., 2005; Serrano
the toxic effects of ethanol, also in this species (Mo et al., et al., 2006; De Lucena et al., 2012). However, the mechanisms
2019). The increase in cell wall thickness also occurs in underlying yeast response and tolerance to acidic conditions
Saccharomyces boulardii under ethanol stress (Ramírez-Cota are complex and dependent not only on the pH value but
et al., 2021). S. boulardii is known for its probiotic capacities also on the nature of the acid used to adjust low pH. Strong
as a biotherapeutic agent in infections and medical disorders inorganic acids, such as sulphuric acid and hydrochloric acid,
(Hudson et al., 2016; Suvarna et al., 2018; Hossain et al., 2020) are fully dissociated at any external pH, while weak organic
and is a promising species to be used in certain crafts of acids dissociation depends on medium pH and their pKa, the
beer fermentation (Ramírez-Cota et al., 2021). In Issatchenkia toxic form being the liposoluble non dissociated form (Carmelo
orientalis, with potential to be used in winemaking, cell wall- et al., 1996, 1997; Booth and Stratford, 2003). Since the plasma
related genes GAS4 (β-1,3-glucanosyltransferase involved in membrane of unstressed cells is very poorly permeable to H+ ,
glucan elongation), FLO1 (CWP involved in adhesion events the effect of strong acids relies, essentially, on the concentration
important for flocculation) and IFF6 (GPI-CWP involved in cell of H+ /medium pH (Carmelo et al., 1996, 1997; Lucena et al.,
wall organization) were found to be up-regulated under ethanol 2020). For this reason, pH stress and stress induced by organic
stress (Miao et al., 2018). acid stress at low pH are discussed separately.
A recent chemogenomic analysis reported that, in
S. cerevisiae, methanol and ethanol, share genetic determinants Acid pH-induced stress
of tolerance involved in cell wall maintenance (Mota et al., In bioethanol production, yeast biomass is reused after being
2021). Methanol is a feedstock alternative to sugar-based raw washed between successive batches using inorganic acids. It is a
substrates for biorefinery processes and a toxic compound common procedure, frequently carried out at a pH below 3.0
commonly found in crude glycerol, a by-product of the as a means of eliminating contaminant bacteria from pitching
biodiesel industry, and in hydrolysates from pectin-rich yeast. This disinfection treatment, together with the presence
biomass residues (Yasokawa et al., 2010; Martins et al., 2020; of toxic metabolites and other stressful conditions occurring
Mota et al., 2021). Methanol and ethanol tolerance genes include during fermentation, may lead to the loss of cell viability and
FKS1 and SMI1 (β1,3-glucan synthesis), ROT2 (β1,6-glucan limit fermentation yield (Simpson and Hammond, 1989; De
synthesis), MNN11 (mannosylation of CWPs), and WSC1 (CWI Melo et al., 2010; Lucena et al., 2020). A scalable and economic
pathway membrane sensor). However, KRE6, CWH41 (β1,6- solution to control bacterial contamination during alcoholic
glucan synthesis) and GAS1 (β1,3-glucan chain elongation fermentation is to run the fermentations at low pH (<4.0),
and branching) were found to be required for maximum at which growth and viability of most bacteria are drastically
tolerance to methanol but not for equivalent inhibitory reduced (Narayanan et al., 2016). Thus, understanding how
concentrations of ethanol (Mota et al., 2021). This suggests that, yeast strains tolerate low pH set up with strong acids may enable
despite the similarities of these alcohols, equivalent inhibitory the improvement of ethanol yield and reduce production costs
concentrations might impact the cell wall differently and, as (Brosnan et al., 2000). It is likely that during fermentation,
such, elicit not fully overlapping remodeling responses (Mota organic acids (weak acids) may also play a role since they are
et al., 2021). The CWI pathway is implicated in methanol produced during yeast metabolism and may be already present
adaptation in Komagataella phaffii (formerly Pichia pastoris), in the fermentation medium (e.g., when lignocellulosic biomass
involving the upregulation of the SLT2 homolog encoding gene hydrolysates are used).
in K. phaffii (Zhang et al., 2020). The homolog of Wsc1 and In response to acid pH, the cell wall structure and
Wsc3 CWI sensors in the methylotrophic yeast K. phaffii were composition suffers alteration (Cabib et al., 1989; Kapteyn
implicated in sensing methanol, interacting upstream with et al., 2001) leading to deformation of surface morphology
the K. phaffii homolog of Rom2, a GEF of S. cerevisiae CWI (De Lucena et al., 2015). Cell wall chitin levels decrease at
growth pH values below 5.0, likely as a result of increased cooperation between Slt2 and Crz1 for the expression of FKS2
chitinase activity (Cabib et al., 1989; Figure 3). Many of the in response to cell wall injury (De Lucena et al., 2012; Figure 3).
genes found to be upregulated at low pH are related to cell This Ca2+ -dependent response is likely responsible for the
wall biogenesis, including FKS1 (β1,3-glucan synthase), GAS1 increment of CWI protein trafficking and their localisation at
(β-1,3-glucanosyltransferase involved in cell wall remodeling- cell surface to repair the structural changes caused by medium
elongation of (1 → 3)-β-D-glucan chains and branching), acidification (Lucena et al., 2020). Together, CWI, HOG, and
CHS1 (chitin synthase), NCW2 (GPI-protein involved in chitin- calcineurin signaling pathways ensure the post-translational
glucan assembly), KRE6 (glucosyl hydrolase required for β1,6- activation of the transcription factors needed to promote cell
glucan synthesis) and MNN9 (mannosyltransferase subunit wall maintenance and regeneration to survive acidic pH stress.
involved in wall protein mannosylation) (De Melo et al., 2010;
De Lucena et al., 2012; Figure 3). Interestingly, using QTL Alkaline pH stress
mapping to uncover the genetic basis of a bioethanol industrial The yeast S. cerevisiae proliferates better at acidic than
strain Pedra-2 (PE-2) tolerance, a prevalent non-synonymous at neutral or alkaline pH and medium alkalinisation has
mutation (A631G) in GAS1 was identified during growth at widespread effects in yeast physiology (Serra-Cardona et al.,
low pH induced by sulfuric acid exposure, reinforcing the 2015). An increase of pH from 4.0 to 6.0 leads to the decrease of
idea of the relevant role of this GPI-protein in yeast tolerance the relative proportion of β1,6-glucans in the β-glucan fraction
in acidic environments (Coradini et al., 2021). Together with of the cell wall (Aguilar-Uscanga and François, 2003) and cells
the up-regulation of genes related with 1,3-β-glucan synthesis, grown at pH 6.0 are more susceptible to zymolyase treatment
elongation, and anchoring, low pH stress leads to increased cell (Aguilar-Uscanga and François, 2003; Figure 3). The CWI
wall resistance to compounds with β1,3-glucanase activity and pathway is necessary for tolerance to alkaline pH, as shown
to the establishment of more alkali-sensitive linkages between by the strong alkali-sensitive phenotype of the bck11, slt21,
CWPs and the β1,3-glucan network (Kapteyn et al., 2001; swi41, and swi61 mutants (Serrano et al., 2006). The alkaline
De Melo et al., 2010; De Lucena et al., 2012, 2015; Lucena stress-mediated activation of Slt2 was also shown to depend
et al., 2020). This suggests that low pH established by a strong on the CWI Wsc1 membrane sensor (Serrano et al., 2006;
inorganic acid affects the β-glucan fraction of the cell wall. Kwon et al., 2016; Figure 3). Both FKS2 and GAS1 encoded
As for other stresses, response to acid pH imposed by proteins, involved in the synthesis and elongation of β1,3-
inorganic acids involves a crosstalk between different signaling glucans respectively, are required to resist alkaline stress, and
pathways. The CWI signaling pathway is involved in response likely play a major role in altering the ratio between different
to low pH in S. cerevisiae and has been proposed as the main types of glucans in the cell wall (Serrano et al., 2006). The Ca2+ -
mechanism for tolerance to acid pH (Claret et al., 2005; De dependent calcineurin response has also been implicated in the
Lucena et al., 2012; Figure 3). Its activation is mainly mediated regulation of cell wall synthesis during alkaline stress by the
by Mid2, but also by Wsc1, with the latter appearing to have upregulation of FKS2 expression via the calcineurin-activated
a more prominent role in activating the general response of transcription factor Crz1 (Viladevall et al., 2004; Figure 3).
the cell to this stress (Claret et al., 2005; De Lucena et al., Changes in cell wall composition were also implicated
2012). The proposed model is that cell wall injury due to acid in Y. lipolytica adaptation to high pH stress (Sekova et al.,
stress results in lower turgor and consequently mimics the 2019). In particular, the structural mannoprotein YlPir1 is
effect of hyper-osmotic shock, justifying the activation of the abundant in the cell wall in unstressed conditions but absent
HOG pathway (De Lucena et al., 2012; Figure 3). The HOG when Y. lipolytica cells are exposed to high pH stress (Sekova
pathway then appears to have a dual role. First, Hog1 activates et al., 2019). This readjustment is consistent with the fact that
the protein complex Msn2/4, which induces the expression of mannans, unlike other main polysaccharides of the cell wall, are
ESR genes, including RGD1, a major regulator of yeast survival prone to alkaline hydrolysis, and therefore unstable at high pH
at low pH stress, that encodes a protein implicated in the (Sekova et al., 2019).
activation of CWI pathway under acid stress (Claret et al., To summarize, there are major differences between the
2005). Second, the Hog1 kinase may help to establish a positive signaling responses elicited by acidic and alkaline pH stresses
feedback loop at the downstream module of the CWI pathway involving the cell wall. First, Mid2 appears to be the main sensor
by cooperating with the Slt2-activated Rlm1 transcription factor activating the CWI pathway in response to acidic pH, while
to increase the expression of SLT2 (Claret et al., 2005; De Wsc1 is the main sensor in alkaline pH (Claret et al., 2005; Fuchs
Lucena et al., 2012). Therefore, the HOG pathway can activate and Mylonakis, 2009; Serra-Cardona et al., 2015). Second, acidic
the CWI pathway while bypassing its membrane sensors. pH stress mostly leads to the transcription of Rlm1-dependent
The Ca2+ /calmodulin-dependent calcineurin pathway is also genes, while alkaline pH stress favors transcription of SBF-
involved in the response to acid pH, and interacts with the CWI dependent genes. Third, while the CWI pathway manages acidic
pathway by activating Cch1/Mid1 calcium channels by Slt2, stress in a Hog1-dependent manner, response to alkaline stress
the Crz1 transcription factor by Rho1-Skn7, and through the is Hog1-independent (Fuchs and Mylonakis, 2009).
Organic acid-induced stress yeast cell population to sudden exposure to a sub-lethal stress
induced by acetic acid (Ribeiro et al., 2021). Yeast cell wall
The response and tolerance of the yeast cell to the various resistance to lyticase activity was found to increase during
industrially relevant weak acids and the underlying toxicity acetic acid-induced growth latency, corresponding to the period
mechanisms are not fully shared by all the acids, with specific of yeast population adaptation to sudden exposure to acetic
mechanisms for a weak acid/group of weak acids (Mira et al., acid. This response was correlated with the increase of cell
2010c). In general, and broadly speaking, the higher the stiffness, assessed by atomic force microscopy (Ribeiro et al.,
lipophilicity of each of a weak acid is, the higher its toxicity. 2021; Figure 3). The increased content of cell wall β-glucans,
The straight medium chain weak acids (e.g., butyric, hexanoic, also assessed by fluorescence microscopy, and the slight increase
octanoic, and decanoic acids) and sorbic and benzoic acids are of the transcription level of the GAS1 gene encoding a β-1,3-
more lipophilic and toxic than the short-chain volatile fatty acids glucanosyltransferase that leads to elongation of β1,3-glucan
(VFA) formic, acetic, and propionic acids (Mira et al., 2010c; chains, were also implicated (Ribeiro et al., 2021; Figure 3).
Skoneczny, 2018). These observations reinforce the notion that the adaptive yeast
response to acetic acid stress involves a coordinate alteration of
Acetic acid stress the cell wall at the biophysical and molecular levels, essential to
Acetic acid is widely used as a food preservative in the limit the futile cycle associated to the re-entry of the toxic acid
food industry and is also a major inhibitory compound present form after the active expulsion of acetate from the cell interior
in lignocellulosic biomass hydrolysates limiting the use of this (Ribeiro et al., 2021).
low cost and abundant biomass (Palma et al., 2018; Cunha The adaptive genomic response to acetic acid in S. cerevisiae
et al., 2019). Acetic acid is also produced by yeast metabolic is mainly regulated by the Haa1 transcription factor involved
activity and can lead, together with ethanol and other yeast toxic in the direct, or indirect, transcriptional activation of
metabolites, to decreased ethanol yield and even fermentation approximately 80% of acetic acid-responsive genes and
arrest depending on the level of stress (Palma et al., 2018; likely involved in the response at the cell wall level (Mira
Cunha et al., 2019; Palma and Sá-Correia, 2019). Elucidation et al., 2010a, 2011). Haa1 increased expression or Haa1 specific
of the mechanisms underlying yeast adaptation and tolerance mutations lead to increased tolerance to acetic acid stress
to acetic acid is instrumental to pave way for strain and and to a lower intracellular accumulation of acetate (Tanaka
process optimisation in several important biotechnological and et al., 2012; Palma et al., 2018). Overexpression of HAA1
food industries. improves cell wall robustness in response to this weak acid,
When the external pH is below acetic acid pKa (below as suggested by the decreased susceptibility of the cell wall to
4.75 at 25◦ C) (Lide, 2003), the undissociated form of the lyticase activity mediated disruption (Cunha et al., 2018). Also,
acid (CH3 COOH) is able to passively diffuse through the YGP1 and SPI1, encoding CWPs, belong to the Haa1 regulon;
plasma membrane lipid bilayer (Casal et al., 1996; Mira et al., they are upregulated under acetic acid stress and contribute
2010c; Palma et al., 2018; Palma and Sá-Correia, 2019). Once to yeast tolerance (Simões et al., 2006; Sakihama et al., 2015;
in the near-neutral cytosol, acetic acid dissociates into the Figure 3). Increased mRNA levels from YGP1 were reported in
negatively charged acetate counterion, CH3 COO− , releasing cells overexpressing HAA1 (Tanaka et al., 2012). These results
protons, H+ . Being unable to cross the hydrophobic lipid suggest that not only the activation of acetate expulsion through
layer due to the electric charge, these ions accumulate in the efflux pumps is involved in acetic acid tolerance, as proposed
cytosol, resulting in decreased intracellular pH, increased turgor (Piper et al., 1998; Holyoak et al., 1999; Tenreiro et al., 2002;
pressure and oxidative stress, disrupting normal metabolism Fernandes et al., 2005; Kawahata et al., 2006; Mira et al., 2010b),
(Mira et al., 2010c; Palma et al., 2018; Palma and Sá-Correia, but a more efficient restriction of the diffusional entry of acetic
2019; Ribeiro et al., 2022). acid, partially dependent of the CWP Ygp1, can also be involved
Several genome-wide studies have sought to shed light into (Tanaka et al., 2012). The Znf1 transcription factor was also
the global mechanisms involved in the response and tolerance implicated in acetic acid tolerance and in the upregulation of
of S. cerevisiae to acetic acid (Kawahata et al., 2006; Abbott et al., the YGP1 gene (Songdech et al., 2020).
2007; Almeida et al., 2009; Mira et al., 2010a,b; Bajwa et al., Genes involved in CWP mannosylation (MNN2, MNN9,
2013; Longo et al., 2015). Increased cell wall impermeabilization MNN11, KTR4), chitin synthesis (CHS1, CHS5), β1,3-glucan
in adapted yeast cells can reduce the passive diffusion of the synthesis (FKS1, ROM2), and β1,6-glucan synthesis (KRE6)
weak acids into the cytosol, in this way restraining the futile were also reported as being required for maximum tolerance
cycle associated with the re-entry of the liposoluble acid form of S. cerevisiae to acetic acid stress (Mollapour et al., 2009;
after the active expulsion of its counter-ion from the cell interior Mira et al., 2010b). However, the mRNA levels from RLM1,
(Ullah et al., 2013). Recently, it was reported a coordinate encoding a major transcriptional regulator of the CWI pathway,
and comprehensive view on the time course of the alterations and from Rlm1-target genes, were found to decrease in cells
occurring at the level of the cell wall during adaptation of a exposed to acetic acid stress, suggesting that the CWI pathway
is not the major key player in acetic acid stress response response to lactic acid stress (Kawahata et al., 2006; Figure 3).
(Ribeiro et al., 2021). Genes encoding elements of the MAPK module of the CWI
Zygosaccharomyces bailii is a common food spoilage yeast pathway (BCK1, SLT2), as well as genes involved in glucan
capable of adapting and proliferating in the presence of synthesis (KRE1, KRE11) and remodeling (GAS1), have also
remarkably high concentrations of acetic acid (Palma et al., been implicated in lactic acid tolerance (Kawahata et al., 2006;
2018). During exposure to acetic acid, several genes involved Suzuki et al., 2012).
in the modulation of plasma membrane composition and cell The Haa1 transcription factor also plays an important role
wall architecture were found to be differentially expressed in tolerance to lactic acid and is involved in the control of the
(Antunes et al., 2018). Among those genes is the homologue expression of CWPs (Abbott et al., 2008; Mira et al., 2010c;
of S. cerevisiae YGP1, encoding a cell wall-related glycoprotein, Sugiyama et al., 2014). As found for acetic acid stress (Kim
whose upregulation in the presence of acetic acid was shown to et al., 2019), exposure to lactic acid leads to Haa1 translocation
depend on ZbHaa1 (Palma et al., 2017; Antunes et al., 2018), from the cytoplasm to the nucleus where gene transcription, in
showing a high degree of similarity between the responses particular of YGP1 and SPI1 occurs (Sugiyama et al., 2014). The
involving the cell wall in both yeast species. In a K. marxianus overexpression of these two genes, encoding CWP, likely confers
strain isolated from agave, lyticase assays showed that the a stronger protective effect against lactic acid-induced toxicity
addition of KCl/KOH leads to the increase of cell wall robustness (Sugiyama et al., 2014).
in cells grown in the presence of acetic acid (Castillo-Plata et al., Due to the high tolerance of Zygosaccharomyces parabailii
2021). However, further studies are necessary to elucidate the to high lactic acid concentrations at low pH, this species was
link between changes in cell wall during acetic acid adaptation proposed as a promising novel host for lactic acid production
and potassium homeostasis (Castillo-Plata et al., 2021). In (Ortiz-Merino et al., 2018). In Z. parabailii, several cell wall-
I. orientalis, the expression of MPG1 gene, coding for a GDP- related genes were found to be down-regulated in the presence
mannose pyrophosphorylase involved in cell wall synthesis, of lactic acid (Ortiz-Merino et al., 2018). Under the stress
was also found to be upregulated under acetic acid stress conditions tested, during exponential growth in the presence
(Li et al., 2021). of lactic acid, a slight decrease in glucans was reported in
S. cerevisiae and Z. bailii, and a slight decrease in mannans in
Lactic acid stress S. cerevisiae (Berterame et al., 2016; Kuanyshev et al., 2016).
Lactic acid is another important weak acid in the food
industry and also in the pharmaceutical and cosmetic industries. Other weak acids
Its industrial production is currently carried out by lactic Propionic acid is commonly used to preserve baked goods
acid bacteria (Van Maris et al., 2004; Sauer et al., 2010) but and dairy (Suhr and Nielsen, 2004). Transcriptomic and
bacteria are sensitive to low pH, requiring large amounts chemogenomic studies have hinted at a role for the cell wall in
of neutralizing agents to counteract the acidification of the yeast adaptation to this weak acid (Fernandes et al., 2005; Mira
fermentation media, thus compromising the recovery yield et al., 2009). Specifically, CWP1 (encoding a GPI-CWP), BAG7
of precipitated lactic acid (Altaf et al., 2007; Yen et al., (β1,3-glucan synthesis), and KNH1 (β1,6-glucan synthesis) are
2010). Yeasts typically fare better than bacteria in acidic required to resist propionic acid stress, all of which are regulated
environments, which has motivated attempts to produce by the transcription factor Rim101 (Mira et al., 2009), suggesting
lactic acid through the heterologous expression of lactate that cell wall remodeling during adaptation to propionic acid
dehydrogenases in yeast (Dequin and Barre, 1994; Porro stress may be dependent on RIM101 expression.
et al., 1999; Pacheco et al., 2012). The success of such Sorbic and benzoic acids are two other weak acids used to
strategy and its industrial application require knowledge of preserve foods and beverages. In both cases, the S. cerevisiae
the mechanisms behind yeast tolerance to lactic acid stress. cell wall has been implicated in stress tolerance (De Nobel
Exposure to lactic acid leads to a decrease in cell wall et al., 2001; Simões et al., 2006). Common to the responses
glucan content and, to a lesser extent, of the mannan to sorbic and benzoic acids is the induction of SPI1, a GPI-
content of S. cerevisiae cell wall (Berterame et al., 2016; CWP that likely leads to the decrease of cell wall porosity and,
Figure 3). Several genes involved in the synthesis of cell in turn, limits the access to plasma membrane, thus reducing
wall polysaccharides, cell wall remodeling, and synthesis of membrane damage, intracellular acidification, and viability loss
mannoproteins are transcriptionally responsive and/or are (De Nobel et al., 2001; Simões et al., 2006; Figure 3). As
determinants of tolerance to lactic acid stress (Kawahata et al., previously mentioned, SPI1 is also activated in response to other
2006; Abbott et al., 2007; Suzuki et al., 2012). Specifically, weak acids and ethanol stress (Ogawa et al., 2000; Simões et al.,
genes required for glucan remodeling (EXG1, GAS2, SCW10), 2006; Wu et al., 2006).
cross-linking between β-glucans and chitin (CRH1), chitin Formic acid in an inhibitory weak acid present
synthesis (CHS1), and genes encoding mannoproteins required in lignocellulosic hydrolysates negatively impacting
for cell wall stability (HSP150, PIR3, SED1), are up-regulated in lignocellulosic-based biorefining (Cunha et al., 2019).
Under formic acid stress, S. cerevisiae cells exhibited a the important role of CWP in decreasing cell surface porosity
deformed shape, with collapsed cell wall edges, indicating that and increasing cell wall stability when coping with stress. The
the cell wall was damaged, and an FTIR analysis suggested that upregulation of genes involved in β1,3-glucan synthesis was also
chitin structure was altered (Zeng et al., 2022). S. cerevisiae reported to be shared by several stresses (heat, ethanol, low
genes involved in β1,3-glucan synthesis (FKS1, ELO2), β1,6- and high pH and formic acid stress) and during hyper-osmotic
glucan synthesis (TRS65), chitin synthesis (CHS5), CWP stress in Z. rouxii (Zhao et al., 1998; Viladevall et al., 2004; De
mannosylation (PMT2), cell wall integrity (PUN1), and CWI Melo et al., 2010; De Lucena et al., 2012; Udom et al., 2019;
pathway regulation (ROM2) are important determinants of Guo et al., 2020; Samakkarn et al., 2021; Zeng et al., 2022).
formic acid tolerance while FKS3, encoding an FKS1-2 homolog, Genes involved in β1,3-glucan elongation and branching were
is upregulated in response to this acid stress (Henriques et al., also found to be upregulated in response to low pH, acetic
2017; Zeng et al., 2022; Figure 3). Moreover, the upregulation and lactic acids-induced stresses and under ethanol stress in
of EXG2 encoding a major exoglucanase, and PKC1, encoding a I. orientalis (Kawahata et al., 2006; De Melo et al., 2010; De
CWI pathway kinase, was reported in an industrial S. cerevisiae Lucena et al., 2012; Miao et al., 2018; Ribeiro et al., 2021).
strain (S6), engineered to ferment xylose, when grown in a Genes involved in chitin synthesis found to be upregulated
medium with glucose and xylose supplemented with formic acid were also found to be shared by the response to low pH and
(Li et al., 2020). lactic acid stress in S. cerevisiae, and to hyper-osmotic stress in
Z. rouxii (Kawahata et al., 2006; De Melo et al., 2010; De Lucena
et al., 2012; Guo et al., 2020). Genes involved in β-glucan-chitin
Adaptive responses to several crosslinking were also found to be upregulated under ethanol
industrially relevant stresses involving and lactic acid stresses in S. cerevisiae, and under hyper-osmotic
the cell wall stress in Z. rouxii (Kawahata et al., 2006; Udom et al., 2019;
Guo et al., 2020). The upregulation of genes involved in β1,6-
Some of the responses involving the modification cell glucan synthesis are shared by oxidative and low pH stress
wall metabolism and cell wall physicochemical properties in in S. cerevisiae, and by hyper-osmotic stress in Z. rouxii and
yeasts are shared by relevant industrial stresses (Figure 3). For Z. mellis (De Melo et al., 2010; De Lucena et al., 2012; Sha
example, the CWI pathway is implicated in oxidative-, osmotic-, et al., 2013; Guo et al., 2020; Liu et al., 2021). Genes involved in
heat and cold-, ethanol- and low and high pH- induced stresses CWI pathway were found to be upregulated in oxidative-, heat-
(Kamada et al., 1995; Hohmann, 2002a; Claret et al., 2005; and low pH-induced stress in S. cerevisiae, under hyper-osmotic
García-Rodríguez et al., 2005; Mensonides et al., 2005; Vilella stress in Z. mellis, and under methanol stress in K. phaffii (Claret
et al., 2005; Fujita et al., 2006; Serrano et al., 2006; Auesukaree et al., 2005; De Lucena et al., 2012; Sha et al., 2013; Varol et al.,
et al., 2009; Teixeira et al., 2009; Córcoles-Sáez et al., 2012; 2018; Zhang et al., 2020; Liu et al., 2021). Genes involved in CWP
De Lucena et al., 2012; Pillet et al., 2014; Kwon et al., 2016; mannosylation were found to be upregulated in S. cerevisiae
Udom et al., 2019; Vaz et al., 2020). Also, the coordinated under low pH stress and in D. hansenii under hyper-osmotic
regulation involving the CWI and the HOG pathways occurs stress (Thomé, 2007; Gonzalez et al., 2009b; De Melo et al.,
during osmotic, heat, ethanol and low pH stresses (Kamada 2010; De Lucena et al., 2012). Genes involved in β1,3-glucan
et al., 1995; Hohmann, 2002a; García-Rodríguez et al., 2005; hydrolysis were found to be upregulated under lactic acid stress
Mensonides et al., 2005; De Lucena et al., 2012; Dunayevich in S. cerevisiae (Kawahata et al., 2006).
et al., 2018; Udom et al., 2019; Vaz et al., 2020) and the Of all the industrially relevant stresses herein described,
Calcineurin pathway interacts with the CWI pathway during exposure to several stresses were found to decrease, at least,
high and low pH stresses (Viladevall et al., 2004; De Lucena one type of cell wall polysaccharide in S. cerevisiae. Specifically,
et al., 2012). The reported upregulation of genes encoding CWPs a decrease in the β-glucan content was reported for heat-
was also observed under several stress induced conditions (e.g., (β1,3-glucans), high pH- (β1,6-glucans), lactic acid- (β1,3 and
under ethanol stress in S. cerevisiae and I. orientalis and under β1,6-glucans) induced stresses, and a decrease in mannans
acetic acid stress in S. cerevisiae and Z. bailii (Ogawa et al., was reported during lactic acid stress (Aguilar-Uscanga and
2000; Rossignol et al., 2003; Simões et al., 2006; Wu et al., 2006; François, 2003; Schiavone et al., 2014; Berterame et al., 2016).
Sakihama et al., 2015; Palma et al., 2017; Antunes et al., 2018; A slight decrease in the β-glucan content (β1,3 and β1,6-
Miao et al., 2018; Udom et al., 2019) and under oxidative, high glucans) was also reported for Z. bailii during lactic acid stress
and low temperature, low pH, lactic, sorbic and benzoic acids- (Kuanyshev et al., 2016). A decrease in the chitin content was
induced stresses in S. cerevisiae) (Abramova et al., 2001; Causton reported for inorganic acid-induced low pH in S. cerevisiae
et al., 2001; De Nobel et al., 2001; Sahara et al., 2002; Homma (Cabib et al., 1989). Nevertheless, an increase of cell wall
et al., 2003; Schade et al., 2004; Murata et al., 2006; Simões polysaccharide content was also reported, in particular an
et al., 2006; Abe, 2007; Sha et al., 2013; Sugiyama et al., 2014; increase in chitin and β1,6-glucans under heat stress and an
Varol et al., 2018). This common response is consistent with increase in the β-glucans content under acetic acid stress in
S. cerevisiae (Schiavone et al., 2014; Ribeiro et al., 2021). It is (Kawahata et al., 2006; Mira et al., 2010b). The increased
noteworthy to take in consideration that the methods used for expression of the HAA1 gene, encoding a transcription factor
cell wall polysaccharides quantification were not the same in and a major determinant of acetic acid tolerance in S. cerevisiae,
different articles and both the levels of stress and the adaptation led to the improvement of cell wall robustness under acetic
phase of the cells examined (early response, cells adapted to the acid stress, as suggested by the decreased susceptibility of the
stress) could be different or not clearly reported. cell wall to lyticase activity mediated disruption (Cunha et al.,
2018). Through Haa1 amino acid sequence engineering, a single
amino acid exchange at position 135 (serine to phenylalanine)
was found to lead to the upregulation of genes of the Haa1-
Improvement of yeast tolerance to regulon, increasing acetic acid tolerance, in particular YGP1
multiple stresses involving cell wall (Swinnen et al., 2017). The deletion of ATG22, encoding a
engineering vacuolar membrane protein that mediates the efflux of amino
acids resulting from autophagic protein degradation, was found
A few successful examples of the alteration of the to delay programmed cell death (PDC) induced by acetic acid.
physicochemical properties of the cell wall either by the genetic The deletion of ATG22 contributes to the maintenance of cell
engineering of the yeast cell or by the adaptive laboratory wall integrity, by preventing the decrease in total cell wall
evolution (ALE) of yeast cells leading to the increase of tolerance polysaccharides induced by PDC caused by severe acetic acid
to stress(es) have been reported in the literature. stress, increasing the transcript levels of CWI pathway genes
As referred above, through chemogenomic analyses, it was (Yang et al., 2006; Hu et al., 2019). This suggested ATG22 as
demonstrated that the expression of genes related with chitin a potential target for genetic engineering strategies to improve
and glucan synthesis, namely CHS1 (chitin), FKS1 and ACF2 yeast cell wall robustness and tolerance to acetic acid and other
(β1,3-glucans), KRE6 (β1,6-glucans), and others involved in industrial stresses. Enhanced tolerance to both acetic and formic
cell wall integrity (PUN1) and cell wall organization (SIT4), acids at low pH (pH 2.4 or below) by expressing the GAS1 gene
are required for maximum tolerance to ethanol in S. cerevisiae of Issatchenkia orientalis in S. cerevisiae was reported (Wada
(Kubota et al., 2004; Auesukaree et al., 2009; Teixeira et al., et al., 2020). Also, the overexpression of GAS1 in S. cerevisiae
2009; Mota et al., 2021). Superior fermentation performance led to increased lactic acid productivity at low pH (Zhong et al.,
of lignocellulosic hydrolysates was reported for a recombinant 2021), reinforcing the importance of GAS1, encoding a beta-1,3-
S. cerevisiae WXY70 strain overexpressing the CCW12 gene, glucanosyltransferase, in this context (Baek et al., 2017).
encoding a cell wall mannoprotein, compared to the control As previously referred, chemogenomic studies
strain, and CCW12 expression was found to improve cell demonstrated that genes related with cell wall polysaccharide
wall stability and tolerance to the growth inhibitors present synthesis (FKS1, KRE6) and cell wall mannosylation (MNN10,
(Kong et al., 2021). The deletion of GAL6, encoding a cysteine ANP1) are determinants of tolerance to osmotic stress (high
aminopeptidase with homocysteine-thiolactonase activity, was glucose concentrations) in S. cerevisiae (Ando et al., 2006;
found to lead to improved growth and higher viability of Teixeira et al., 2010). In Pichia pastoris (now Komagataella
S. cerevisiae, in the presence of ethanol stress (Yazawa et al., phaffii), the deletion of YPS7, encoding a putative GPI-linked
2007). The gal61 cells showed increased resistance to zymolyase aspartyl protease, led to increased osmotic tolerance and this
activity indicating the occurrence of structural changes in the gene was proposed as a promising molecular target for the
cell wall (Yazawa et al., 2007). A marked increased tolerance to engineering of yeast robustness (Guan et al., 2012). This species
ethanol stress associated with transcription rewiring involving is used to produce heterologous proteins in the pharmaceutical
cell wall synthesis was also identified in the highly–ethanol and food industry.
tolerant strain K. marxianus FIM1 obtained by ALE in an In a laboratory adaptively evolved K. marxianus strain
ethanol supplemented medium (Mo et al., 2019). Among exhibiting a significantly improved tolerance to high
the identified changes at the transcriptional level were the temperatures, a single nucleotide polymorphism was found in
alterations related with the upregulation of cell wall metabolism the coding region of the exoglucanase gene EXGI, required for
involving, in particular, chitin synthesis (CHS3), β1,6-glucan cell wall remodeling (Ai et al., 2022).
synthesis (KNH1), glucanosyltransferases activity involved in
β1,3-glucan branching and elongation (GAS4) and the cell wall
integrity pathway (BCK1, MID2, WSC3) (Mo et al., 2019). Concluding remarks
Numerous genes encoding proteins required for cell
wall polysaccharides synthesis (FKS1, ROM2, KRE6, CHS1, The important role played by the cell wall in the adaptation
CHS5), cell wall remodeling (GAS1) and protein mannosylation and tolerance of yeasts to different stresses of biotechnological
(MNN2, MNN9, MNN11, KTR4) were found to be determinants interest emerges from this review article. Although several
of acetic acid stress tolerance by chemogenomic analyses evidences support this idea, the truth is that these evidences
References
Abbott, D. A., Knijnenburg, T. A., De Poorter, L. M. I., Reinders, M. J. T., Pronk, chemostat cultures of Saccharomyces cerevisiae. Appl. Env. Microbiol. 74, 5759–
J. T., and Van Maris, A. J. A. (2007). Generic and specific transcriptional responses 5768. doi: 10.1128/AEM.01030-08
to different weak organic acids in anaerobic chemostat cultures of Saccharomyces
Abe, F. (2007). Induction of DAN/TIR yeast cell wall
cerevisiae. FEMS Yeast Res. 7, 819–833. doi: 10.1111/j.1567-1364.2007.00242.x
mannoprotein genes in response to high hydrostatic pressure and low
Abbott, D., Suir, E., Duong, G., de Hulster, E., Pronk, J., and van Maris, A. temperature. FEBS Lett. 581, 4993–4998. doi: 10.1016/j.febslet.2007.
(2009). Catalase overexpression reduces lactic acid-induced oxidative stress in 09.039
Saccharomyces cerevisiae. Appl. Env. Microbiol. 75, 2320–2325. doi: 10.1128/AEM.
Abramova, N., Sertil, O., Mehta, S., and Lowry, C. V. (2001). Reciprocal
00009-09
regulation of anaerobic and aerobic cell wall mannoprotein gene expression in
Abbott, D., Suir, E., van Maris, A., and Pronk, J. (2008). Physiological and Saccharomyces cerevisiae. J. Bacteriol. 183, 2881–2887. doi: 10.1128/JB.183.9.2881-
transcriptional responses to high concentrations of lactic acid in anaerobic 2887.2001
Adya, A. K., Canetta, E., and Walker, G. M. (2006). Atomic force microscopic Bajwa, P. K., Ho, C. Y., Chan, C. K., Martin, V. J. J., Trevors, J. T., and Lee,
study of the infuence of physical stresses on Saccharomyces cerevisiae and H. (2013). Transcriptional profiling of Saccharomyces cerevisiae T2 cells upon
Schizosaccharomyces pombe. FEMS Yeast Res. 6, 120–128. doi: 10.1111/j.1567- exposure to hardwood spent sulphite liquor: comparison to acetic acid, furfural
1364.2005.00003.x and hydroxymethylfurfural. Antonie Leeuwenhoek Int. J. Gen. Mol. Microbiol. 103,
1281–1295. doi: 10.1007/s10482-013-9909-1
Aguilar-Uscanga, B., and François, J. M. (2003). A study of the yeast cell
wall composition and structure in response to growth conditions and mode of Bauer, F. F., and Pretorius, I. S. (2000). Yeast stress response and fermentation
cultivation. Lett. Appl. Microbiol. 37, 268–274. doi: 10.1046/j.1472-765X.2003. efficiency: how to survive the making of wine - a review. South African J. Enol.
01394.x Vitic. 21, 27–51. doi: 10.21548/21-1-3557
Ai, Y., Luo, T., Yu, Y., Zhou, J., and Lu, H. (2022). Downregulation Berterame, N. M., Porro, D., Ami, D., and Branduardi, P. (2016). Protein
of ammonium uptake improves the growth and tolerance of Kluyveromyces aggregation and membrane lipid modifications under lactic acid stress in wild type
marxianus at high temperature. Microbiol. Open 11, 1–13. doi: 10.1002/mbo3.1290 and OPI1 deleted Saccharomyces cerevisiae strains. Microb. Cell Fact. 15, 1–12.
doi: 10.1186/s12934-016-0438-2
Aimanianda, V., Simenel, C., Garnaud, C., Clavaud, C., Tada, R., Barbin, L., et al.
(2017). The dual activity responsible for the elongation and branching of β-(1,3)- Blanco, N., Sanz, A. B., Rodríguez-Peña, J. M., Nombela, C., Farkaš, V.,
glucan in the fungal cell wall. MBio 8, e619. doi: 10.1128/mBio.00619-17 Hurtado-Guerrero, R., et al. (2015). Structural and functional analysis of yeast
Crh1 and Crh2 transglycosylases. FEBS J. 282, 715–731. doi: 10.1111/febs.13176
Alloush, H. M., Edwards, T. A., Valle-Lisboa, V., and Wheals, A. E. (2002).
Cdc4 is involved in the transcriptional control of OCH1, a gene encoding α-1,6- Booth, I. R., and Stratford, M. (2003). “Acidulants and low pH,” in Food
mannosyltransferase in Saccharomyces cerevisiae. Yeast 19, 69–77. doi: 10.1002/ Preservatives, eds N. J. Russell and G. W. Gould (Boston, MA: Springer), 25–47.
yea.801 doi: 10.1007/978-0-387-30042-9_3
Almeida, B., Ohlmeier, S., Almeida, A. J., Madeo, F., Leão, C., Rodrigues, F., Borodina, I., and Nielsen, J. (2014). Advances in metabolic engineering of yeast
et al. (2009). Yeast protein expression profile during acetic acid-induced apoptosis Saccharomyces cerevisiae for production of chemicals. Biotechnol. J. 9, 609–620.
indicates causal involvement of the TOR pathway. Proteomics 9, 720–732. doi: doi: 10.1002/biot.201300445
10.1002/pmic.200700816 Bourbon-Melo, N., Palma, M., Rocha, M. P., Ferreira, A., Bronze, M. R.,
Altaf, M., Venkateshwar, M., Srijana, M., and Reddy, G. (2007). An economic Elias, H., et al. (2021). Use of Hanseniaspora guilliermondii and Hanseniaspora
approach for L-(+) lactic acid fermentation by Lactobacillus amylophilus GV6 opuntiae to enhance the aromatic profile of beer in mixed-culture fermentation
using inexpensive carbon and nitrogen sources. J. Appl. Microbiol. 103, 372–380. with Saccharomyces cerevisiae. Food Microbiol. 95:103678. doi: 10.1016/j.fm.2020.
doi: 10.1111/j.1365-2672.2006.03254.x 103678
Alvim, M. C. T., Vital, C. E., Barros, E., Vieira, N. M., da Silveira, Brosnan, M. P., Donnelly, D., James, T. C., and Bond, U. (2000). The stress
F. A., Balbino, T. R., et al. (2019). Ethanol stress responses of Kluyveromyces response is repressed during fermentation in brewery strains of yeast. J. Appl.
marxianus CCT 7735 revealed by proteomic and metabolomic analyses. Antonie Microbiol. 88, 746–755. doi: 10.1046/j.1365-2672.2000.01006.x
Leeuwenhoek Int. J. Gen. Mol. Microbiol. 112, 827–845. doi: 10.1007/s10482-018- Bubnová, M., Zemanciková, J., and Sychrová, H. (2014). Osmotolerant yeast
01214-y species differ in basic physiological parameters and in tolerance of non-osmotic
Ando, A., Tanaka, F., Murata, Y., Takagi, H., and Shima, J. (2006). Identification stresses. Yeast 31, 309–321.
and classification of genes required for tolerance to high-sucrose stress revealed by Buzzini, P., Branda, E., Goretti, M., and Turchetti, B. (2012). Psychrophilic
genome-wide screening of Saccharomyces cerevisiae. FEMS Yeast Res. 6, 249–267. yeasts from worldwide glacial habitats: diversity, adaptation strategies and
doi: 10.1111/j.1567-1364.2006.00035.x biotechnological potential. FEMS Microbiol. Ecol. 82, 217–241. doi: 10.1111/j.
Antunes, M., Palma, M., and Sá-correia, I. (2018). Transcriptional profiling of 1574-6941.2012.01348.x
Zygosaccharomyces bailii early response to acetic acid or copper stress mediated by Cabib, E. (2009). Two novel techniques for determination of polysaccharide
ZbHaa1. Sci. Rep. 2, 1–14. doi: 10.1038/s41598-018-32266-9 cross-links show that Crh1p and Crh2p attach chitin to both β(1-6)- and β(1-
Arinbasarova, A. Y., Machulin, A. V., Biryukova, E. N., Sorokin, V. V., 3)glucan in the Saccharomyces cerevisiae cell wall. Eukaryot. Cell 8, 1626–1636.
Medentsev, A. G., and Suzina, N. E. (2018). Structural changes in the cell envelope doi: 10.1128/EC.00228-09
of Yarrowia lipolytica yeast under stress conditions. Can. J. Microbiol. 64, 359–365. Cabib, E., Blanco, N., Grau, C., Rodríguez-Peña, J. M., and Arroyo, J. (2007).
doi: 10.1139/cjm-2018-0034 Crh1p and Crh2p are required for the cross-linking of chitin to beta(1-6)glucan in
Arnesen, J. A., Belmonte Del Ama, A., Jayachandran, S., Dahlin, J., the Saccharomyces cerevisiae cell wall. Mol. Microbiol. 63, 921–935. doi: 10.1111/j.
Rago, D., Andersen, A. J. C., et al. (2022). Engineering of Yarrowia 1365-2958.2006.05565.x
lipolytica for the production of plant triterpenoids: asiatic, madecassic, and Cabib, E., Farkas, V., Kosík, O., Blanco, N., Arroyo, J., and McPhie, P. (2008).
arjunolic acids. Metab. Eng. Commun. 14:e00197. doi: 10.1016/j.mec.2022.e Assembly of the yeast cell wall. Crh1p and Crh2p act as transglycosylases
00197 in vivo and in vitro. J. Biol. Chem. 283, 29859–29872. doi: 10.1074/jbc.M8042
Assis, L., Manfiolli, A., Mattos, E., Fabri, J., Malavazi, I., Jacobsen, I., et al. (2018). 74200
Protein kinase a and high-osmolarity glycerol response pathways cooperatively Cabib, E., Silverman, S. J., and Shaw, J. A. (1992). Chitinase and chitin synthase
control cell wall carbohydrate mobilization in Aspergillus fumigatus. Mol. Biol. 1: counterbalancing activities in cell separation of Saccharomyces cerevisiae. J. Gen.
Physiol. 9, 1–15. doi: 10.1128/mBio.01952-18 Microbiol. 138, 97–102. doi: 10.1099/00221287-138-1-97
Attfield, P. V. (1987). Trehalose accumulates in Saccharomyces cerevisiae during Cabib, E., Sburlati, A., Bowers, B., and Silverman, S. J. (1989). Chitin synthase
exposure to agents that induce heat shock response. FEBS Lett. 225, 259–263. 1, an auxiliary enzyme for chitin synthesis in Saccharomyces cerevisiae. J. Cell Biol.
doi: 10.1016/0014-5793(87)81170-5 108, 1665–1672. doi: 10.1083/jcb.108.5.1665
Attfield, P. V., and Kletsas, S. (2000). Hyperosmotic stress response by strains Carmelo, V., Bogaerts, P., and Sá-Correia, I. (1996). Activity of plasma
of bakers’ yeasts in high sugar concentration medium. Lett. Appl. Microbiol. 31, membrane H+-ATPase and expression of PMA1 and PMA2 genes in
323–327. doi: 10.1046/j.1472-765X.2000.00825.x Saccharomyces cerevisiae cells grown at optimal and low pH. Arch. Microbiol. 166,
Auesukaree, C. (2017). Molecular mechanisms of the yeast adaptive response 315–320. doi: 10.1007/s002030050389
and tolerance to stresses encountered during ethanol fermentation. J. Biosci. Carmelo, V., Santos, H., and Sá-Correia, I. (1997). Effect of extracellular
Bioeng. 124, 133–142. doi: 10.1016/j.jbiosc.2017.03.009 acidification on the activity of plasma membrane ATPase and on the cytosolic and
Auesukaree, C., Damnernsawad, A., Kruatrachue, M., Pokethitiyook, vacuolar pH of Saccharomyces cerevisiae. Biochim. Biophys. Acta - Biomembr. 1,
P., Boonchird, C., Kaneko, Y., et al. (2009). Genome-wide identification 63–70. doi: 10.1016/S0005-2736(96)00245-3
of genes involved in tolerance to various environmental stresses in Casal, M., Cardoso, H., and Leão, C. (1996). Mechanisms regulating the
Saccharomyces cerevisiae. J. Appl. Genet. 50, 301–310. doi: 10.1007/BF031 transport of acetic acid in Saccharomyces cerevisiae. Microbiology 142, 1385–1390.
95688 doi: 10.1099/13500872-142-6-1385
Baek, S.-H., Kwon, E. Y., Bae, S.-J., Cho, B.-R., Kim, S.-Y., and Hahn, J.-S. Castillo-Plata, A. K., Sigala, J. C., Lappe-Oliveras, P., and Le Borgne, S. (2021).
(2017). Improvement of d-lactic acid production in saccharomyces cerevisiae under KCl/KOH supplementation improves acetic acid tolerance and ethanol production
acidic conditions by evolutionary and rational metabolic engineering. Biotechnol. in a thermotolerant strain of Kluyveromyces marxianus isolated from henequen
J. 12:1700015. doi: 10.1002/biot.201700015 (agave fourcroydes). Rev. Mex. Ing. Química 21:Bio2567. doi: 10.24275/rmiq/
Bio2567
Baetz, K., Moffat, J., Haynes, J., Chang, M., and Andrews, B. (2001).
Transcriptional coregulation by the cell integrity mitogen-activated protein kinase Causton, H. C., Ren, B., Koh, S. S., Harbison, C. T., Kanin, E., Jennings, E. G.,
Slt2 and the cell cycle regulator Swi4. Mol. Cell. Biol. 21, 6515–6528. doi: 10.1128/ et al. (2001). Remodeling of yeast genome expression in response to environmental
MCB.21.19.6515-6528.2001 changes. Mol. Biol. Cell 12, 323–337. doi: 10.1091/mbc.12.2.323
Charoenbhakdi, S., Dokpikul, T., Burphan, T., Techo, T., and Auesukaree, De Nobel, H., Lawrie, L., Brul, S., Klis, F., Davis, M., Alloush, H., et al. (2001).
C. (2016). Vacuolar H+-ATPase protects Saccharomyces cerevisiae cells against Parallel and comparative analysis of the proteome and transcriptome of sorbic
ethanolinduced oxidative and cell wall stresses. Appl. Environ. Microbiol. 82, acid-stressed Saccharomyces cerevisiae. Yeast 18, 1413–1428. doi: 10.1002/yea.793
3121–3130. doi: 10.1128/AEM.00376-16
De Nobel, J. G., Klis, F. M., Priem, J., Munnik, T., and Van Den Ende, H. (1990).
Chen, R. E., and Thorner, J. (2007). Function and regulation in MAPK signaling The glucanase-soluble mannoproteins limit cell wall porosity in Saccharomyces
pathways: lessons learned from the yeast Saccharomyces cerevisiae. Biochim. cerevisiae. Yeast 6, 491–499. doi: 10.1002/yea.320060606
Biophys. Acta Mol. Cell Res. 1773, 1311–1340. doi: 10.1016/j.bbamcr.2007.05.003
De Souza Pereira, R., and Geibel, J. (1999). Direct observation of oxidative stress
Claret, S., Gatti, X., Doignon, F., Thoraval, D., and Crouzet, M. (2005). The on the cell wall of Saccharomyces cerevisiae strains with atomic force microscopy.
rgd1p rho GTPase-activating protein and the Mid2p cell wall sensor are required at Mol. Cell. Biochem. 201, 17–24. doi: 10.1023/a:1007007704657
low pH for protein kinase C pathway activation and cell survival in Saccharomyces
Deparis, Q., Claes, A., Foulquié-Moreno, M. R., and Thevelein, J. M. (2017).
cerevisiae. Eukaryot. Cell 4, 1375–1386. doi: 10.1128/EC.4.8.1375-1386.2005
Engineering tolerance to industrially relevant stress factors in yeast cell factories.
Claro, F. B., Rijsbrack, K., and Soares, E. V. (2007). Flocculation onset in FEMS Yeast Res. 17, 1–17. doi: 10.1093/femsyr/fox036
Saccharomyces cerevisiae: effect of ethanol, heat and osmotic stress. J. Appl.
Dequin, S., and Barre, P. (1994). Mixed lactic acid–alcoholic fermentation
Microbiol. 102, 693–700. doi: 10.1111/j.1365-2672.2006.03130.x
by Saccharomyes cerevisiae expressing the Lactobacillus casei L(+)–LDH. Nat.
Coradini, A. L. V., da Silveira Bezerra de Mello, F., Furlan, M., Maneira, C., Biotechnol. 12, 173–177. doi: 10.1038/nbt0294-173
Carazzolle, M. F., Pereira, G. A. G., et al. (2021). QTL mapping of a brazilian
Ding, J., Huang, X., Zhang, L., Zhao, N., Yang, D., and Zhang, K. (2009).
bioethanol strain links the cell wall protein-encoding gene GAS1 to low pH
Tolerance and stress response to ethanol in the yeast Saccharomyces cerevisiae.
tolerance in S. cerevisiae. Biotechnol. Biofuels 14:239. doi: 10.1186/s13068-021-
Appl. Microbiol. Biotechnol. 85, 253–263. doi: 10.1007/s00253-009-2223-1
02079-6
Diniz, R. H. S., Villada, J. C., Alvim, M. C. T., Vidigal, P. M. P., Vieira, N. M.,
Córcoles-Sáez, I., Ballester-Tomas, L., Torre-Ruiz, M. A., Prieto, J. A., and
Lamas-Maceiras, M., et al. (2017). Transcriptome analysis of the thermotolerant
Randez-Gil, F. (2012). Low temperature highlights the functional role of the cell
yeast Kluyveromyces marxianus CCT 7735 under ethanol stress. Appl. Microbiol.
wall integrity pathway in the regulation of growth in Saccharomyces cerevisiae.
Biotechnol. 101, 6969–6980. doi: 10.1007/s00253-017-8432-0
Biochem. J. 446, 477–488. doi: 10.1042/BJ20120634
Donlin, M. J., Upadhya, R., Gerik, K. J., Lam, W., Vanarendonk, L. G.,
Costa, V., and Moradas-Ferreira, P. (2001). Oxidative stress and signal
Specht, C. A., et al. (2014). Cross talk between the cell wall integrity and cyclic
transduction in Saccharomyces cerevisiae: insights into ageing, apoptosis and
AMP/protein kinase a pathways in Cryptococcus neoformans. MBio 5, 1–10. doi:
diseases. Mol. Aspects Med. 22, 217–246. doi: 10.1016/S0098-2997(01)00012-7
10.1128/mBio.01573-14
Cunha, J. T., Costa, C. E., Ferraz, L., Romaní, A., Johansson, B., Sá-Correia,
Dunayevich, P., Baltanás, R., Clemente, J. A., Couto, A., Sapochnik, D., Vasen,
I., et al. (2018). HAA1 and PRS3 overexpression boosts yeast tolerance towards
G., et al. (2018). Heat-stress triggers MAPK crosstalk to turn on the hyperosmotic
acetic acid improving xylose or glucose consumption: unravelling the underlying
response pathway. Sci. Rep. 8:33203. doi: 10.1038/s41598-018-33203-6
mechanisms. Appl. Microbiol. Biotechnol. 102, 4589–4600. doi: 10.1007/s00253-
018-8955-z Dupres, V., Dufrene, Y., and Heinisch, J. (2010). Measuring cell wall thickness
in living yeast cells using single molecular rulers. ACS Nano 4, 5498–5504. doi:
Cunha, J. T., Romaní, A., Costa, C. E., Sá-Correia, I., and Domingues, L. (2019).
10.1021/nn101598v
Molecular and physiological basis of Saccharomyces cerevisiae tolerance to adverse
lignocellulose-based process conditions. Appl. Microbiol. Biotechnol. 103, 159–175. Eardley, J., and Timson, D. J. (2020). Yeast cellular stress: impacts on bioethanol
doi: 10.1007/s00253-018-9478-3 production. Fermentation 6:109. doi: 10.3390/fermentation6040109
da Silveira, F. A., de Oliveira Soares, D. L., Bang, K. W., Balbino, T. R., de Moura Ene, I. V., Walker, L. A., Schiavone, M., Lee, K. K., Martin-Yken, H., Dague, E.,
Ferreira, M. A., Diniz, R. H. S., et al. (2020). Assessment of ethanol tolerance of et al. (2015). Cell wall remodeling enzymes modulate fungal cell wall elasticity and
Kluyveromyces marxianus CCT 7735 selected by adaptive laboratory evolution. osmotic stress resistance. MBio 6:e00986. doi: 10.1128/mBio.00986-15
Appl. Microbiol. Biotechnol. 104, 7483–7494. doi: 10.1007/s00253-020-10768-9
Estruch, F. (2000). Stress-controlled transcription factors, stress-induced genes
Dague, E., Bitar, R., Ranchon, H., Durand, F., and François, J. M. (2010). and stress tolerance in budding yeast. FEMS Microbiol. Rev. 24, 469–486. doi:
An atomic force microscopy analysis of yeast mutants defective in cell wall 10.1016/S0168-6445(00)00035-8
architecture. Yeast 27, 673–684. doi: 10.1002/yea.1801
Fathi, Z., Tramontin, L. R. R., Ebrahimipour, G., Borodina, I., and Darvishi, F.
Dallies, N., François, J., and Paquet, V. (1998). A new method for quantitative (2021). Metabolic engineering of Saccharomyces cerevisiae for production of β-
determination of polysaccharides in the yeast cell wall. application to the cell wall carotene from hydrophobic substrates. FEMS Yeast Res. 21:foaa068. doi: 10.1093/
defective mutants of Saccharomyces cerevisiae. Yeast 14, 1297–1306. doi: 10.1002/ femsyr/foaa068
(SICI)1097-0061(1998100)14.0.CO;2-L
Fernandes, A. R., Mira, N. P., Vargas, R. C., Canelhas, I., and Sá-Correia,
Davenport, K. R., Sohaskey, M., Kamada, Y., Levin, D. E., and Gustin, M. C. I. (2005). Saccharomyces cerevisiae adaptation to weak acids involves the
(1995). A second osmosensing signal transduction pathway in yeast. hypotonic transcription factor Haa1p and Haa1p-regulated genes. Biochem. Biophys. Res.
shock activates the PKC1 protein kinase-regulated cell integrity pathway. J. Biol. Commun. 337, 95–103. doi: 10.1016/j.bbrc.2005.09.010
Chem. 270, 30157–30161. doi: 10.1074/jbc.270.50.30157
Fleet, G. H. (1991). “Cell walls,” in The Yeasts, Vol. 4, ed. H. J. S. Rose (London:
De Groot, P. W. J., Kraneveld, E. A., Qing, Y. Y., Dekker, H. L., Groß, U., Academic Press), 199–277. doi: 10.1002/jobm.19720120411
Crielaard, W., et al. (2008). The cell wall of the human pathogen candida glabrata:
Fleet, G. H. (2011). Yeast spoilage of foods and beverages. Yeasts 2011, 53–63.
differential incorporation of novel adhesin-like wall proteins. Eukaryot. Cell 7,
doi: 10.1016/B978-0-444-52149-1.00005-7
1951–1964. doi: 10.1128/EC.00284-08
Free, S. J. (2013). “Chapter two - fungal cell wall organization and biosynthesis,”
De Groot, P. W. J., Ram, A. F., and Klis, F. M. (2005). Features and functions
in Advances in Genetics, eds T. Friedmann and J. C. Dunlap (Academic Press),
of covalently linked proteins in fungal cell walls. Fungal. Genet. Biol. 42, 657–675.
33–82. doi: 10.1016/B978-0-12-407677-8.00002-6
doi: 10.1016/j.fgb.2005.04.002
Fu, X., Li, P., Zhang, L., and Li, S. (2019). Understanding the stress responses
De Groot, P. W. J., Ruiz, C., de Aldana, C. R. V., Duenas, E., Cid, V. J., Del
of Kluyveromyces marxianus after an arrest during high-temperature ethanol
Rey, F., et al. (2001). A genomic approach for the identification and classification
fermentation based on integration of RNA-Seq and metabolite data. Appl.
of genes involved in cell wall formation and its regulation in Saccharomyces
Microbiol. Biotechnol. 103, 2715–2729. doi: 10.1007/s00253-019-09637-x
cerevisiae. Comput. Funct. Geno. 2, 124–142. doi: 10.1002/cfg.85
Fuchs, B. B., and Mylonakis, E. (2009). Our paths might cross: the role of the
De Lucena, R. M., Elsztein, C., de Barros Pita, W., de Souza, R. B., de Sá Leitão
fungal cell wall integrity pathway in stress response and cross talk with other stress
Paiva Júnior, S., and de Morais Junior, M. A. (2015). Transcriptomic response
response pathways. Eukaryot. Cell 8, 1616–1625. doi: 10.1128/EC.00193-09
of Saccharomyces cerevisiae for its adaptation to sulphuric acid-induced stress.
Antonie Van Leeuwenhoek 108, 1147–1160. doi: 10.1007/s10482-015-0568-2 Fujita, K., Matsuyama, A., Kobayashi, Y., and Iwahashi, H. (2006). The genome-
wide screening of yeast deletion mutants to identify the genes required for
De Lucena, R. M., Elsztein, C., Simões, D. A., and De Morais, M. A. (2012).
tolerance to ethanol and other alcohols. FEMS Yeast Res. 6, 744–750. doi: 10.1111/
Participation of CWI, HOG and Calcineurin pathways in the tolerance of
j.1567-1364.2006.00040.x
Saccharomyces cerevisiae to low pH by inorganic acid. J. Appl. Microbiol. 113,
629–640. doi: 10.1111/j.1365-2672.2012.05362.x García-Rodríguez, L. J., Valle, R., Durán, Á, and Roncero, C. (2005). Cell
integrity signaling activation in response to hyperosmotic shock in yeast. FEBS
De Melo, H. F., Bonini, B. M., Thevelein, J., Simões, D. A., and Morais,
Lett. 579, 6186–6190. doi: 10.1016/j.febslet.2005.10.001
M. A. (2010). Physiological and molecular analysis of the stress response of
Saccharomyces cerevisiae imposed by strong inorganic acid with implication to Garcia-Rubio, R., de Oliveira, H. C., Rivera, J., and Trevijano-Contador, N.
industrial fermentations. J. Appl. Microbiol. 109, 116–127. doi: 10.1111/j.1365- (2020). The fungal cell wall: candida, Cryptococcus, and Aspergillus Species. Front.
2672.2009.04633.x Microbiol. 10:1–13. doi: 10.3389/fmicb.2019.02993
García, R., Bravo, E., Diez-Muñiz, S., Nombela, C., Rodríguez-Peña, J. M., Homma, T., Iwahashi, H., and Komatsu, Y. (2003). Yeast gene expression during
and Arroyo, J. (2017). A novel connection between the cell wall integrity and growth at low temperature. Cryobiology 46, 230–237. doi: 10.1016/s0011-2240(03)
the PKA pathways regulates cell wall stress response in yeast. Sci. Rep. 7, 1–15. 00028-2
doi: 10.1038/s41598-017-06001-9
Hossain, M. N., Afrin, S., Humayun, S., Ahmed, M. M., and Saha, B. K. (2020).
Garrett-Engele, P., Moilanen, B., and Cyert, M. S. (1995). Calcineurin, the Identification and growth characterization of a novel strain of Saccharomyces
Ca2+/calmodulin-dependent protein phosphatase, is essential in yeast mutants boulardii isolated from soya paste. Front. Nutr. 7:27. doi: 10.3389/fnut.2020.00027
with cell integrity defects and in mutants that lack a functional vacuolar H(+)-
Hottiger, T., Boller, T., and Wiemken, A. (1989). Correlation of trenalose
ATPase. Mol. Cell. Biol. 15, 4103–4114. doi: 10.1128/MCB.15.8.4103
content and heat resistance in yeast mutants altered in the RAS/adenylate cyclase
Gibney, P. A., Schieler, A., Chen, J. C., Rabinowitz, J. D., and Botstein, D. pathway: is trehalose a thermoprotectant? FEBS Lett. 255, 431–434. doi: 10.1016/
(2015). Characterizing the in vivo role of trehalose in Saccharomyces cerevisiae 0014-5793(89)81139-1
using the AGT1 transporter. Proc. Natl. Acad. Sci. U.S.A. 112, 6116–6121. doi:
Hottiger, T., Schmutz, P., and Wiemken, A. (1987). Heat-induced accumulation
10.1073/pnas.1506289112
and futile cycling of trehalose in Saccharomyces cerevisiae. J. Bacteriol. 169,
Gibson, B. R., Lawrence, S. J., Leclaire, J. P. R., Powell, C. D., and Smart, K. A. 5518–5522. doi: 10.1128/jb.169.12.5518-5522.1987
(2007). Yeast responses to stresses associated with industrial brewery handling.
FEMS Microbiol. Rev. 31, 535–569. doi: 10.1111/j.1574-6976.2007.00076.x Hou, L., Wang, M., Wang, C., Wang, C., and Wang, H. (2013). Analysis of
salt-tolerance genes in Zygosaccharomyces rouxii. Appl. Biochem. Biotechnol. 170,
Godinho, C. P., Costa, R., and Sá-Correia, I. (2021). The ABC transporter Pdr18 1417–1425. doi: 10.1007/s12010-013-0283-2
is required for yeast thermotolerance due to its role in ergosterol transport and
plasma membrane properties. Environ. Microbiol. 23, 69–80. doi: 10.1111/1462- Hu, J., Dong, Y., Wang, W., Zhang, W., Lou, H., and Chen, Q. (2019). Deletion
2920.15253 of Atg22 gene contributes to reduce programmed cell death induced by acetic
acid stress in Saccharomyces cerevisiae. Biotechnol. Biofuels 12, 1–20. doi: 10.1186/
Gonzalez, M., Lipke, P. N., and Ovalle, R. B. (2009a). “Chapter 15 - GPI proteins s13068-019-1638-x
in biogenesis and structure of yeast cell walls,” in Glycosylphosphatidylinositol
(GPI) Anchoring of Proteins, eds A. K. Menon, T. Kinoshita, P. Orlean, and F. Hudson, L. E., McDermott, C. D., Stewart, T. P., Hudson, W. H., Rios, D.,
Tamanoi (New York, NY: Academic Press), 321–356. doi: 10.1016/S1874-6047(09) Fasken, M. B., et al. (2016). Characterization of the probiotic yeast Saccharomyces
26015-X boulardii in the healthy mucosal immune system. PLoS One 11:e0153351. doi:
10.1371/journal.pone.0153351
Gonzalez, M, Vásquez, A., Zuazaga, H., Sen, A., Olvera, H., Ortiz, S., et al.
(2009b). Genome-wide expression profiling of the osmoadaptation response of Igual, J. C., Johnson, A. L., and Johnston, L. H. (1996). Coordinated regulation
Debaryomyces hansenii. Yeast 26, 111–124. of gene expression by the cell cycle transcription factor Swi4 and the protein
kinase C MAP kinase pathway for yeast cell integrity. EMBO J. 34, 1049–1057.
Gonzalez, R., Morales, P., Tronchoni, J., Cordero-Bueso, G., Vaudano, E., doi: 10.1002/j.1460-2075.1996.tb00880.x
Quirós, M., et al. (2016). New genes involved in osmotic stress tolerance in
Saccharomyces cerevisiae. Front. Microbiol. 7:1–12. doi: 10.3389/fmicb.2016.01545 Imazu, H., and Sakurai, H. (2005). Saccharomyces cerevisiae heat shock
transcription factor regulates cell wall remodeling in response to heat shock.
Gschaedler, A. (2017). Contribution of non-conventional yeasts in alcoholic Eukaryot. Cell 4, 1050–1056. doi: 10.1128/EC.4.6.1050-1056.2005
beverages. Curr. Opin. Food Sci. 13, 73–77. doi: 10.1016/j.cofs.2017.02.004
Irie, K., Takase, M., Lee, K. S., Levin, D. E., Araki, H., Matsumoto, K., et al.
Guan, B., Lei, J., Su, S., Chen, F., Duan, Z., Chen, Y., et al. (2012). Absence of (1993). MKK1 and MKK2, which encode Saccharomyces cerevisiae mitogen-
Yps7p, a putative glycosylphosphatidylinositol-linked aspartyl protease in pichia activated protein kinase-kinase homologs, function in the pathway mediated by
pastoris, results in aberrant cell wall composition and increased osmotic stress protein kinase C. Mol. Cell. Biol. 13, 3076–3083. doi: 10.1128/mcb.13.5.3076
resistance. FEMS Yeast Res. 12, 969–979. doi: 10.1111/1567-1364.12002
Jaafar, L., and Zueco, J. (2004). Characterization of a
Guo, H., Qiu, Y., Wei, J., Niu, C., Zhang, Y., Yuan, Y., et al. (2020). Genomic glycosylphosphatidylinositol-bound cell-wall protein (GPI-CWP) in Yarrowia
insights into sugar adaptation in an extremophile yeast Zygosaccharomyces rouxii. lipolytica. Microbiology 150, 53–60. doi: 10.1099/mic.0.26430-0
Front. Microbiol. 10:1–10. doi: 10.3389/fmicb.2019.03157
Jamieson, D. J. (1998). Oxidative stress responses of the yeast Saccharomyces
Gustin, M. C., Albertyn, J., Alexander, M., and Davenport, K. (1998). MAP cerevisiae. Yeast 14, 1511–1527. doi: 10.1002/(SICI)1097-0061(199812)14:163.0.
kinase pathways in the yeast Saccharomyces cerevisiae. Microbiol. Mol. Biol. Rev. CO;2-S
62:1264. doi: 10.1128/mmbr.62.4.1264-1300.1998
Jiménez-Gutiérrez, E., Alegría-Carrasco, E., Alonso-Rodríguez, E., Fernández-
Hagen, I., Ecker, M., Lagorce, A., Francois, J. M., Sestak, S., Rachel, R., Acero, T., Molina, M., and Martín, H. (2020a). Rewiring the yeast cell wall integrity
et al. (2004). Sed1p and Srl1p are required to compensate for cell wall (CWI) pathway through a synthetic positive feedback circuit unveils a novel
instability in Saccharomyces cerevisiae mutants defective in multiple GPI-anchored role for the MAPKKK Ssk2 in CWI pathway activation. FEBS J. 22, 4881–4901.
mannoproteins. Mol. Microbiol. 52, 1413–1425. doi: 10.1111/j.1365-2958.2004. doi: 10.1111/febs.15288
04064.x
Jiménez-Gutiérrez, E., Alegría-Carrasco, E., Sellers-Moya, Á, Molina, M., and
Harris, M. R., Lee, D., Farmer, S., Lowndes, N. F., and de Bruin, R. A. M. (2013). Martín, H. (2020b). Not just the wall: the other ways to turn the yeast CWI pathway
Binding specificity of the G1/S transcriptional regulators in budding yeast. PLoS on. Int. Microbiol. 23, 107–119. doi: 10.1007/s10123-019-00092-2
One 8:e61059. doi: 10.1371/journal.pone.0061059
Jin, C., Parshin, A. V., Daly, I., Strich, R., and Cooper, K. F. (2013). The cell
Heinisch, J. J., Dupres, V., Wilk, S., Jendretzki, A., and Dufrêne, Y. F. wall sensors Mtl1, Wsc1, and Mid2 are required for stress-induced nuclear to
(2010). Single-molecule atomic force microscopy reveals clustering of the yeast cytoplasmic translocation of cyclin C and programmed cell death in yeast. Oxid.
plasma-membrane sensor Wsc1. PLoS One 5:e11104. doi: 10.1371/journal.pone.00 Med. Cell. Longev. 2013:320823. doi: 10.1155/2013/320823
11104
Jung, U. S., and Levin, D. E. (1999). Genome-wide analysis of gene expression
Henriques, S. F., Mira, N. P., and Sá-Correia, I. (2017). Genome-wide search regulated by the yeast cell wall integrity signalling pathway. Mol. Microbiol. 34,
for candidate genes for yeast robustness improvement against formic acid reveals 1049–1057. doi: 10.1046/j.1365-2958.1999.01667.x
novel susceptibility (Trk1 and positive regulators) and resistance (Haa1-regulon)
determinants. Biotechnol. Biofuels 10:96. doi: 10.1186/s13068-017-0781-5 Kahar, P., Itomi, A., Tsuboi, H., Ishizaki, M., Yasuda, M., Kihira, C., et al. (2022).
The flocculant Saccharomyces cerevisiae strain gains robustness via alteration of
Hohmann, S. (2002a). Osmotic adaptation in yeast-control of the yeast osmolyte the cell wall hydrophobicity. Metab. Eng. 72, 82–96. doi: 10.1016/j.ymben.2022.03.
system. Int. Rev. Cytol. 215, 149–187. doi: 10.1016/S0074-7696(02)15008-X 001
Hohmann, S. (2002b). Osmotic stress signaling and osmoadaptation in yeasts. Kamada, Y., Jung, U. S., Piotrowski, J., and Levin, D. E. (1995). The protein
Microbiol. Mol. Biol. Rev. 66, 300–372. doi: 10.1128/mmbr.66.2.300-372.2002 kinase C-activated MAP kinase pathway of Saccharomyces cerevisiae mediates a
Hohmann, S. (2009). Control of high osmolarity signalling in the yeast novel aspect of the heat shock response. Genes Dev. 9, 1559–1571. doi: 10.1101/
Saccharomyces cerevisiae. FEBS Lett. 583, 4025–4029. doi: 10.1016/j.febslet.2009. gad.9.13.1559
10.069
Kanetsuna, F., Carbonell, L. M., Moreno, R. E., and Rodriguez, J. (1969).
Holt, S., Mukherjee, V., Lievens, B., Verstrepen, K. J., and Thevelein, J. M. Cell wall composition of the yeast and mycelial forms of Paracoccidioides
(2018). Bioflavoring by non-conventional yeasts in sequential beer fermentations. brasiliensis. J. Bacteriol. 97, 1036–1041. doi: 10.1128/jb.97.3.1036-1041.
Food Microbiol. 72, 55–66. doi: 10.1016/j.fm.2017.11.008 1969
Holyoak, C. D., Bracey, D., Piper, P. W., Kuchler, K., and Coote, P. J. (1999). The Kapteyn, J. C., Ter Riet, B., Vink, E., Blad, S., De Nobel, H., Van Den Ende,
Saccharomyces cerevisiae weak acid-inducible ABC transporter Pdr12 transports H., et al. (2001). Low external ph induces HOG1-dependent changes in the
fluorescein and preservative anions from the cytosol by an energy-dependent organization of the Saccharomyces cerevisiae cell wall. Mol. Microbiol. 39, 469–479.
mechanism. J. Bacteriol. 181, 4644–4652. doi: 10.1128/JB.181.15.4644-4652.1999 doi: 10.1046/j.1365-2958.2001.02242.x
Kapteyn, J. C., Van Den Ende, H., and Klis, F. M. (1999a). The contribution of Kwak, S., and Jin, Y. S. (2017). Production of fuels and chemicals from xylose by
cell wall proteins to the organization of the yeast cell wall. Biochim. Biophys. Acta engineered Saccharomyces cerevisiae: a review and perspective. Microb. Cell Fact.
Gen. Subj. 1426, 373–383. doi: 10.1016/s0304-4165(98)00137-8 16, 1–15. doi: 10.1186/s12934-017-0694-9
Kapteyn, J. C., Van Egmond, P., Sievi, E., Van Den Ende, H., Makarow, M., and Kwon, Y., Chiang, J., Tran, G., Giaever, G., Nislow, C., Hahn, B.-S., et al. (2016).
Klis, F. M. (1999b). The contribution of the O-glycosylated protein Pir2p/Hsp150 Signaling pathways coordinating the alkaline pH response confer resistance to the
to the construction of the yeast cell wall in wild-type cells and beta 1,6-glucan- hevein-type plant antimicrobial peptide Pn-AMP1 in Saccharomyces cerevisiae.
deficient mutants. Mol. Microbiol. 31, 1835–1844. doi: 10.1046/j.1365-2958.1999. Planta 244, 1229–1240. doi: 10.1007/s00425-016-2579-2
01320.x
Larriba, G., Andaluz, E., Cueva, R., and Basco, R. D. (1995). Molecular biology
Kawahata, M., Masaki, K., Fujii, T., and Iefuji, H. (2006). Yeast genes involved of yeast exoglucanases. FEMS Microbiol. Lett. 125, 121–126. doi: 10.1111/j.1574-
in response to lactic acid and acetic acid: acidic conditions caused by the organic 6968.1995.tb07347.x
acids in Saccharomyces cerevisiae cultures induce expression of intracellular metal
Ledesma-Amaro, R., and Nicaud, J. M. (2016). Yarrowia lipolytica as a
metabolism genes regulated by Aft1p. FEMS Yeast Res. 6, 924–936. doi: 10.1111/j.
biotechnological chassis to produce usual and unusual fatty acids. Prog. Lipid Res.
1567-1364.2006.00089.x
61, 40–50. doi: 10.1016/j.plipres.2015.12.001
Kim, H., Thak, E. J., Yeon, J. Y., Sohn, M. J., Choo, J. H., Kim, J.-Y., et al.
Lee, J., Reiter, W., Dohnal, I., Gregori, C., Beese-Sims, S., Kuchler, K.,
(2018). Functional analysis of Mpk1-mediated cell wall integrity signaling pathway
et al. (2013). MAPK Hog1 closes the S. cerevisiae glycerol channel Fps1 by
in the thermotolerant methylotrophic yeast Hansenula polymorpha. J. Microbiol.
phosphorylating and displacing its positive regulators. Genes Dev. 27, 2590–2601.
56, 72–82. doi: 10.1007/s12275-018-7508-6
doi: 10.1101/gad.229310.113
Kim, K.-Y., Truman, A. W., and Levin, D. E. (2008). Yeast Mpk1
Lee, K. S., and Levin, D. E. (1992). Dominant mutations in a gene encoding
mitogen-activated protein kinase activates transcription through Swi4/Swi6 by
a putative protein kinase (BCK1) bypass the requirement for a Saccharomyces
a noncatalytic mechanism that requires upstream signal. Mol. Cell. Biol. 28,
cerevisiae protein kinase C homolog. Mol. Cell. Biol. 12, 172–182. doi: 10.1128/
2579–2589. doi: 10.1128/mcb.01795-07
mcb.12.1.172
Kim, M. S., Cho, K. H., Park, K. H., Jang, J., and Hahn, J.-S. (2019). Activation of
Lehnen, M., Ebert, B. E., and Blank, L. M. (2019). Elevated temperatures do
Haa1 and War1 transcription factors by differential binding of weak acid anions
not trigger a conserved metabolic network response among thermotolerant yeasts.
in Saccharomyces cerevisiae. Nucleic Acids Res. 47, 1211–1224. doi: 10.1093/nar/
BMC Microbiol. 19:1–11. doi: 10.1186/s12866-019-1453-3
gky1188
Lesage, G., and Bussey, H. (2006). Cell wall assembly in Saccharomyces
Klis, F. M. (2002). Dynamics of cell wall structure in Saccharomyces cerevisiae.
cerevisiae. Microbiol. Mol. Biol. Rev. 70, 317–343. doi: 10.1128/MMBR.00038-05
FEMS Microbiol. Rev. 26, 239–256. doi: 10.1111/j.1574-6976.2002.tb00613.x
Levin, D. E. (2011). Regulation of cell wall biogenesis in Saccharomyces
Klis, F. M., Boorsma, A., and De Groot, P. W. J. (2006). Cell wall construction
cerevisiae: the cell wall integrity signaling pathway. Genetics 189, 1145–1175.
in Saccharomyces cerevisiae. Yeast 23, 185–202. doi: 10.1002/yea.1349
doi: 10.1534/genetics.111.128264
Kock, C., Arlt, H., Ungermann, C., and Heinisch, J. J. (2016). Yeast cell wall
Li, B., Xie, C. Y., Yang, B. X., Gou, M., Xia, Z. Y., Sun, Z. Y., et al. (2020).
integrity sensors form specific plasma membrane microdomains important for
The response mechanisms of industrial Saccharomyces cerevisiae to acetic acid
signalling. Cell. Microbiol. 18, 1251–1267. doi: 10.1111/cmi.12635
and formic acid during mixed glucose and xylose fermentation. Proc. Biochem.
Kock, C., Dufrêne, Y. F., and Heinisch, J. J. (2015). Up against the wall: 91, 319–329. doi: 10.1016/j.procbio.2020.01.002
is yeast cell wall integrity ensured by mechanosensing in plasma membrane
Li, S., Dean, S., Li, Z., Horecka, J., Deschenes, R. J., and Fassler, J. S. (2002).
microdomains? Appl. Environ. Microbiol. 81, 806–811. doi: 10.1128/AEM.03273-
The eukaryotic two-component histidine kinase Sln1p regulates OCH1 via the
14
transcription factor, Skn7p. Mol. Biol. Cell 13, 412–424. doi: 10.1091/mbc.01-09-
Kolhe, N., Kulkarni, A., Zinjarde, S., and Acharya, C. (2021). Transcriptome 0434
response of the tropical marine yeast Yarrowia lipolytica on exposure to uranium.
Li, Y., Li, Y., Li, R., Liu, L., Miao, Y., Weng, P., et al. (2021). Metabolic changes
Curr. Microbiol. 78, 2033–2043. doi: 10.1007/s00284-021-02459-z
of Issatchenkia orientalis under acetic acid stress by transcriptome profile using
Kolhe, N., Zinjarde, S., and Acharya, C. (2020). Impact of uranium exposure RNA-sequencing. Int. Microbiol. doi: 10.1007/s10123-021-00217-6 [Epub ahead
on marine yeast, Yarrowia lipolytica: insights into the yeast strategies to withstand of print].
uranium stress. J. Hazard. Mater. 381:121226. doi: 10.1016/j.jhazmat.2019.121226
Lide, D. R. (2003). “Dissociation constants of organic acids and bases,” in CRC
Kollár, R., Petráková, E., Ashwell, G., Robbins, P. W., and Cabib, E. (1995). Handbook of Chemistry and Physics, ed. D. R. Lide (Boca. Raton, FL: CRC Press),
Architecture of the yeast cell wall. the linkage between chitin and β(1 → 3)-glucan. 8–46.
J. Biol. Chem. 270, 1170–1178. doi: 10.1074/jbc.270.3.1170
Liszkowska, W., and Berlowska, J. (2021). Yeast fermentation at low
Kollár, R., Reinhold, B. B., Petráková, E., Yeh, H. J. C., Ashwell, G., Drgonová, J., temperatures: adaptation to changing environmental conditions and formation of
et al. (1997). Architecture of the yeast cell wall. beta(1–>6)-glucan interconnects volatile compounds. Molecules 26:1035. doi: 10.3390/molecules26041035
mannoprotein, beta(1–>)3-glucan, and chitin. J. Biol. Chem. 272, 17762–17775.
Liu, G., Bi, X., Tao, C., Fei, Y., Gao, S., Liang, J., et al. (2021). Comparative
doi: 10.1074/jbc.272.28.17762
transcriptomics analysis of Zygosaccharomyces mellis under high-glucose stress.
Kong, M., Li, X., Li, T., Zhao, X., Jin, M., Zhou, X., et al. (2021). Overexpressing Food Sci. Hum. Wellness 10, 54–62. doi: 10.1016/j.fshw.2020.05.006
CCW12 in Saccharomyces cerevisiae enables highly efficient ethanol production
Liu, G., Tao, C., Zhu, B., Bai, W., Zhang, L., Wang, Z., et al. (2016). Identification
from lignocellulose hydrolysates. Bioresour. Technol. 337:125487. doi: 10.1016/j.
of Zygosaccharomyces mellis strains in stored honey and their stress tolerance. Food
biortech.2021.125487
Sci. Biotechnol. 25, 1645–1650. doi: 10.1007/s10068-016-0253-x
Krause, E. L., Villa-García, M. J., Henry, S. A., and Walker, L. P. (2007).
Determining the effects of inositol supplementation and the opi1 mutation on Longo, V., Ždralević, M., Guaragnella, N., Giannattasio, S., Zolla, L.,
ethanol tolerance of Saccharomyces cerevisiae. Ind. Biotechnol. (New Rochelle. N. and Timperio, A. M. (2015). Proteome and metabolome profiling of
Y). 3, 260–268. doi: 10.1089/ind.2007.3.260 wild-type and YCA1-knock-out yeast cells during acetic acid-induced
programmed cell death. J. Proteomics 128, 173–188. doi: 10.1016/j.jprot.2015.
Kuanyshev, N., Ami, D., Signori, L., Porro, D., Morrissey, J. P., and 08.003
Branduardi, P. (2016). Assessing physio-macromolecular effects of lactic acid on
Zygosaccharomyces bailii cells during microaerobic fermentation. FEMS Yeast Res. Lucena, R. M., Dolz-Edo, L., Brul, S., de Morais, M. A., and Smits, G. (2020).
16:fow058. doi: 10.1093/femsyr/fow058 Extreme low cytosolic pH is a signal for cell survival in acid stressed yeast. Genes
11:56. doi: 10.3390/genes11060656
Kubiak-Szymendera, M., Skupien-Rabian, B., Jankowska, U., and Celiñska,
E. (2022). Hyperosmolarity adversely impacts recombinant protein synthesis Madhavan, A., Jose, A. A., Binod, P., Sindhu, R., Sukumaran, R. K., Pandey,
by Yarrowia lipolytica-molecular background revealed by quantitative A., et al. (2017). Synthetic biology and metabolic engineering approaches and its
proteomics. Appl. Microbiol. Biotechnol. 106, 349–367. doi: 10.1007/s00253-021- impact on non-conventional yeast and biofuel production. Front. Energy Res. 5:8.
11731-y doi: 10.3389/fenrg.2017.00008
Kubota, S., Takeo, I., Kume, K., Kanai, M., Shitamukai, A., Mizunuma, M., et al. Madrid, M., Vázquez-Marín, B., Franco, A., Soto, T., Vicente-Soler, J., Gacto,
(2004). Effect of ethanol on cell growth of budding yeast: genes that are important M., et al. (2016). Multiple crosstalk between TOR and the cell integrity
for cell growth in the presence of ethanol. Biosci. Biotechnol. Biochem. 68, 968–972. MAPK signaling pathway in fission yeast. Sci. Rep. 6, 1–17. doi: 10.1038/srep
doi: 10.1271/bbb.68.968 37515
Kuranda, M. J., and Robbins, P. W. (1991). Chitinase is required for cell Mager, W. H., and Siderius, M. (2002). Novel insights into the osmotic stress
separation during growth of Saccharomyces cerevisiae. J. Biol. Chem. 266, 19758– response of yeast. FEMS Yeast Res. 2, 251–257. doi: 10.1016/S1567-1356(02)
19767. 00116-2
Magnelli, P., Cipollo, J. F., and Abeijon, C. (2002). A refined method for Murata, Y., Homma, T., Kitagawa, E., Momose, Y., Sato, M. S., Odani, M., et al.
the determination of Saccharomyces cerevisiae cell wall composition and beta- (2006). Genome-wide expression analysis of yeast response during exposure to 4
1,6-glucan fine structure. Anal. Biochem. 301, 136–150. doi: 10.1006/abio.2001. degrees C. Extremophiles 10, 117–128. doi: 10.1007/s00792-005-0480-1
5473
Nandy, S. K., and Srivastava, R. K. (2018). A review on sustainable yeast
Martín, H., Rodríguez-Pachón, J. M., Ruiz, C., Nombela, C., and Molina, M. biotechnological processes and applications. Microbiol. Res. 207, 83–90. doi: 10.
(2000). Regulatory mechanisms for modulation of signaling through the cell 1016/j.micres.2017.11.013
integrity Slt2-mediated pathway in Saccharomyces cerevisiae. J. Biol. Chem. 275,
Narayanan, V., Sànchez i Nogué, V., van Niel, E. W. J., and Gorwa-Grauslund,
1511–1519. doi: 10.1074/jbc.275.2.1511
M. F. (2016). Adaptation to low pH and lignocellulosic inhibitors resulting in
Martins, L. C., Monteiro, C. C., Semedo, P. M., and Sá-Correia, I. (2020). ethanolic fermentation and growth of Saccharomyces cerevisiae. AMB Exp. 6:8.
Valorisation of pectin-rich agro-industrial residues by yeasts: potential and doi: 10.1186/s13568-016-0234-8
challenges. Appl. Microbiol. Biotechnol. 104, 6527–6547. doi: 10.1007/s00253-020- Navarro-Tapia, E., Querol, A., and Pérez-Torrado, R. (2018). Membrane
10697-7 fluidification by ethanol stress activates unfolded protein response in yeasts.
Mattanovich, D., Gasser, B., Hohenblum, H., and Sauer, M. (2004). Stress in Microb. Biotechnol. 11, 465–475. doi: 10.1111/1751-7915.13032
recombinant protein producing yeasts. Jounal 113, 121–135. doi: 10.1016/j.jbiotec. Nguyen, T. H., Fleet, G. H., and Rogers, P. L. (1998). Composition of the
2004.04.035 cell walls of several yeast species. Appl. Microbiol. Biotechnol. 50, 206–212. doi:
Mazur, P., Morin, N., Baginsky, W., El-Sherbeini, M., Clemas, J. A., Nielsen, J. B., 10.1007/s002530051278
et al. (1995). Differential expression and function of two homologous subunits Nielsen, J. (2019). Yeast systems biology: model organism and cell factory.
of yeast 1,3-beta-D-glucan synthase. Mol. Cell. Biol. 15, 5671–5681. doi: 10.1128/ Biotechnol. J. 14, e1800421. doi: 10.1002/biot.201800421
MCB.15.10.5671
Noti, O., Vaudano, E., Pessione, E., and Garcia-Moruno, E. (2015). Short-
Mensonides, F. I. C., Brul, S., Klis, F. M., Hellingwerf, K. J., and De Mattos, term response of different Saccharomyces cerevisiae strains to hyperosmotic stress
M. J. T. (2005). Activation of the protein kinase C1 pathway upon continuous caused by inoculation in grape must: RT-qPCR study and metabolite analysis. Food
heat stress in Saccharomyces cerevisiae is triggered by an intracellular increase Microbiol. 52, 49–58. doi: 10.1016/j.fm.2015.06.011
in osmolarity due to trehalose accumulation. Appl. Environ. Microbiol. 71, 4531–
4538. doi: 10.1128/AEM.71.8.4531-4538.2005 Ogawa, Y., Nitta, A., Uchiyama, H., Imamura, T., Shimoi, H., and Ito, K. (2000).
Tolerance mechanism of the ethanol-tolerant mutant of sake yeast. J. Biosci.
Miao, Y., Xiong, G., Li, R., Wu, Z., Zhang, X., and Weng, P. (2018). Bioeng. 90, 313–320. doi: 10.1016/S1389-1723(00)80087-0
Transcriptome profiling of Issatchenkia orientalis under ethanol stress. AMB Exp.
8:39. doi: 10.1186/s13568-018-0568-5 Ohsawa, S., Yurimoto, H., and Sakai, Y. (2017). Novel function of Wsc proteins
as a methanol-sensing machinery in the yeast Pichia pastoris. Mol. Microbiol. 104,
Mira, N. P., Becker, J. D., and Sá-Correia, I. (2010a). Genomic expression 349–363. doi: 10.1111/mmi.13631
program involving the haa1p-regulon in Saccharomyces cerevisiae response to
acetic acid. Omi. A J. Integr. Biol. 14, 587–601. doi: 10.1089/omi.2010.0048 Orlean, P. (2012). Architecture and biosynthesis of the Saccharomyces cerevisiae
cell wall. Genetics 192, 775–818. doi: 10.1534/genetics.112.144485
Mira, N. P., Palma, M., Guerreiro, J. F., and Sá-Correia, I. (2010b). Genome-wide
identification of Saccharomyces cerevisiae genes required for tolerance to acetic Ortiz-Merino, R. A., Kuanyshev, N., Byrne, K. P., Varela, J. A., Morrissey,
acid. Microb. Cell Fact. 9:79. doi: 10.1186/1475-2859-9-79 J. P., Porro, D., et al. (2018). Transcriptional response to lactic acid stress in the
hybrid yeast Zygosaccharomyces parabailii. Appl. Environ. Microbiol. 84, 1–17.
Mira, N. P., Teixeira, M. C., and Sá-Correia, I. (2010c). Adaptive response and doi: 10.1128/AEM.02294-17
tolerance to weak acids in Saccharomyces cerevisiae: a genome-wide view. OMICS
14, 525–540. doi: 10.1089/omi.2010.0072 Osmond, B. C., Specht, C. A., and Robbins, P. W. (1999). Chitin synthase III:
synthetic lethal mutants and “stress related” chitin synthesis that bypasses the
Mira, N. P., Henriques, S. F., Keller, G., Teixeira, M. C., Matos, R. G., Arraiano, CSD3/CHS6 localization pathway. Proc. Natl. Acad. Sci. U.S.A. 96, 11206–11210.
C. M., et al. (2011). Identification of a DNA-binding site for the transcription factor doi: 10.1073/pnas.96.20.11206
Haa1, required for Saccharomyces cerevisiae response to acetic acid stress. Nucleic
Acids Res. 39, 6896–6907. doi: 10.1093/nar/gkr228 Pacheco, A., Talaia, G., Sá-Pessoa, J., Bessa, D., Gonçalves, M. J., Moreira, R.,
et al. (2012). Lactic acid production in Saccharomyces cerevisiae is modulated by
Mira, N. P., Lourenço, A. B., Fernandes, A. R., Becker, J. D., and Sá-Correia, expression of the monocarboxylate transporters Jen1 and Ady2. FEMS Yeast Res.
I. (2009). The RIM101 pathway has a role in Saccharomyces cerevisiae adaptive 12, 375–381. doi: 10.1111/j.1567-1364.2012.00790.x
response and resistance to propionic acid and other weak acids. FEMS Yeast Res.
9, 202–216. doi: 10.1111/j.1567-1364.2008.00473.x Palma, M., Dias, P. J., Roque, F., de, C., Luzia, L., Guerreiro, J. F., et al. (2017).
The Zygosaccharomyces bailii transcription factor Haa1 is required for acetic
Mo, W., Wang, M., Zhan, R., Yu, Y., He, Y., and Lu, H. (2019). Kluyveromyces acid and copper stress responses suggesting subfunctionalization of the ancestral
marxianus developing ethanol tolerance during adaptive evolution with significant bifunctional protein Haa1/Cup2. BMC Geno. 18:75. doi: 10.1186/s12864-016-
improvements of multiple pathways. Biotechnol. Biofuels 12, 1–15. doi: 10.1186/ 3443-2
s13068-019-1393-z
Palma, M., Guerreiro, J. F., and Sá-Correia, I. (2018). Adaptive response and
Mollapour, M., Shepherd, A., and Piper, P. W. (2009). Presence of the tolerance to acetic acid in Saccharomyces cerevisiae and Zygosaccharomyces bailii:
Fps1p aquaglyceroporin channel is essential for Hog1p activation, but suppresses a physiological genomics perspective. Front. Microbiol. 9:274. doi: 10.3389/fmicb.
Slt2(Mpk1)p activation, with acetic acid stress of yeast. Microbiology 155, 3304– 2018.00274
3311. doi: 10.1099/mic.0.030502-0
Palma, M., and Sá-Correia, I. (2019). “Physiological genomics of the highly
Moradas-Ferreira, P., and Costa, V. (2000). Adaptive response of the yeast weak-acid-tolerant food spoilage yeasts of Zygosaccharomyces bailii sensu lato,”
Saccharomyces cerevisiae to reactive oxygen species: defences, damage and death. in Progress in Molecular and Subcellular Biology, ed. I. Sá-Correia (Cham: Springer
Redox Rep. 5, 277–285. doi: 10.1179/135100000101535816 International Publishing), doi: 10.1007/978-3-030-13035-0_4
Morano, K. A., Grant, C. M., and Moye-Rowley, W. S. (2012). The response Panahi, H., Dehhaghi, M., Kinder, J. E., and Ezeji, T. C. (2019). A review on
to heat shock and oxidative stress in Saccharomyces cerevisiae. Genetics 190, green liquid fuels for the transportation sector: a prospect of microbial solutions
1157–1195. doi: 10.1534/genetics.111.128033 to climate change. Biofuel Res. J. 6, 995–1024. doi: 10.18331/BRJ2019.6.3.2
Morimoto, Y., and Tani, M. (2015). Synthesis of mannosylinositol Pandey, P., Zaman, K., Prokai, L., and Shulaev, V. (2020). Comparative
phosphorylceramides is involved in maintenance of cell integrity of yeast proteomics analysis reveals unique early signaling response of Saccharomyces
Saccharomyces cerevisiae. Mol. Microbiol. 95, 706–722. doi: 10.1111/mmi.12896 cerevisiae to oxidants with different mechanism of action. Int. J. Mol. Sci. 22:167.
Morris, G. J., Winters, L., Coulson, G. E., and Clarke, K. J. (1983). Effect doi: 10.3390/ijms22010167
of osmotic stress on the ultrastructure and viability of the yeast Saccharomyces Parapouli, M., Vasileiadis, A., Afendra, A. S., and Hatziloukas, E. (2020).
cervisiae. J. Gen. Microbiol. 132, 2023–2034. doi: 10.1099/00221287-132-7- Saccharomyces cerevisiae and its industrial applications. AIMS Microbiol. 6, 1–31.
2023 doi: 10.3934/microbiol.2020001
Mota, M. N., Martins, L. C., and Sá-Correia, I. (2021). The identification Patel, P. K., and Free, S. J. (2019). The genetics and biochemistry of cell wall
of genetic determinants of methanol tolerance in yeast suggests differences in structure and synthesis in Neurospora crassa, a model filamentous fungus. Front.
methanol and ethanol toxicity mechanisms and candidates for improved methanol Microbiol. 10:1–18. doi: 10.3389/fmicb.2019.02294
tolerance engineering. J. Fungi 7, 1–28. doi: 10.3390/jof7020090
Petkova, M. I., Pujol-Carrion, N., Arroyo, J., García-Cantalejo, J., and De La
Mukherjee, V., Radecka, D., Aerts, G., Verstrepen, K. J., Lievens, B., and Torre-Ruiz, M. A. (2010). Mtl1 is required to activate general stress response
Thevelein, J. M. (2017). Phenotypic landscape of non-conventional yeast species through TOR1 and RAS2 inhibition under conditions of glucose starvation
for different stress tolerance traits desirable in bioethanol fermentation. Biotechnol. and oxidative stress. J. Biol. Chem. 285, 19521–19531. doi: 10.1074/jbc.M109.
Biofuels 10, 1–19. doi: 10.1186/s13068-017-0899-5 085282
Pillet, F., Lemonier, S., Schiavone, M., Formosa, C., Martin-yken, H., Francois, and global biofuel research. Int. J. Biol. Macromol. 193, 2304–2319. doi: 10.1016/j.
J. M., et al. (2014). Uncovering by atomic force microscopy of an original circular ijbiomac.2021.11.063
structure at the yeast cell surface in response to heat shock. BMC Biol. 12:1–11.
Sakihama, Y., Hasunuma, T., and Kondo, A. (2015). Improved ethanol
doi: 10.1186/1741-7007-12-6
production from xylose in the presence of acetic acid by the overexpression of
Piper, P., Mahé, Y., Thompson, S., Pandjaitan, R., Holyoak, C., Egner, R., et al. the HAA1 gene in Saccharomyces cerevisiae. J. Biosci. Bioeng. 119, 297–302. doi:
(1998). The pdr12 ABC transporter is required for the development of weak 10.1016/j.jbiosc.2014.09.004
organic acid resistance in yeast. EMBO J. 17, 4257–4265. doi: 10.1093/emboj/17.
Saldaña, C., Villava, C., Ramírez-Villarreal, J., Morales-Tlalpan, V., Campos-
15.4257
Guillen, J., Chávez-Servín, J., et al. (2021). Rapid and reversible cell volume
Popolo, L., Gilardelli, D., Bonfante, P., and Vai, M. (1997). Increase in chitin changes in response to osmotic stress in yeast. Brazilian J. Microbiol. 52, 895–903.
as an essential response to defects in assembly of cell wall polymers in the ggp11 doi: 10.1007/s42770-021-00427-0
mutant of Saccharomyces cerevisiae. J. Bacteriol. 179, 463–469. doi: 10.1128/jb.179.
Samakkarn, W., Ratanakhanokchai, K., and Soontorngun, N. (2021).
2.463-469.1997
Reprogramming of the ethanol stress response in Saccharomyces cerevisiae
Porro, D., Bianchi, M. M., Brambilla, L., Menghini, R., Bolzani, D., Carrera, by the transcription factor Znf1 and its effect on the biosynthesis of glycerol
V., et al. (1999). Replacement of a metabolic pathway for large-scale production and ethanol. Appl. Environ. Microbiol. 87:e0058821. doi: 10.1128/AEM.
of lactic acid from engineered yeasts. Appl. Environ. Microbiol. 65, 4211–4215. 00588-21
doi: 10.1128/aem.65.9.4211-4215.1999
Sanz, A. B., García, R., Rodríguez-Peña, J. M., and Arroyo, J. (2018). The CWI
Pøibylová, L., Montigny, J., and Sychrova, H. (2007). Osmoresistant yeast pathway: regulation of the transcriptional adaptive response to cell wall stress in
Zygosaccharomyces rouxii:the two most studied wild-type strains (ATCC 2623 and yeast. J. Fungi 4:1. doi: 10.3390/jof4010001
ATCC 42981) differ in osmotolerance and glycerol metabolism. Yeast 24, 171–180.
Sauer, M., Porro, D., Mattanovich, D., and Branduardi, P. (2010). 16 years
Ragni, E., Fontaine, T., Gissi, C., Latgè, J. P., and Popolo, L. (2007). The gas research on lactic acid production with yeast–ready for the market? Biotechnol.
family of proteins of Saccharomyces cerevisiae: characterization and evolutionary Genet. Eng. Rev. 27, 229–256. doi: 10.1080/02648725.2010.10648152
analysis. Yeast 24, 297–308. doi: 10.1002/yea.1473
Schade, B., Jansen, G., Whiteway, M., Entian, K. D., and Thomas, D. Y. (2004).
Raj, T., Chandrasekhar, K., Naresh Kumar, A., Rajesh Banu, J., Yoon, J.-J., Cold adaptation in budding yeast. Mol. Biol. Cell 15, 5492–5502. doi: 10.1091/mbc.
Kant Bhatia, S., et al. (2022). Recent advances in commercial biorefineries e04-03-0167
for lignocellulosic ethanol production: current status, challenges and future
perspectives. Bioresour. Technol. 344:126292. doi: 10.1016/j.biortech.2021. Schiavone, M., Formosa-Dague, C., Elsztein, C., Teste, M. A., Martin-Yken, H.,
126292 De Morais, M. A., et al. (2016). Evidence for a role for the plasma membrane
in the nanomechanical properties of the cell wall as revealed by an atomic force
Rajavel, M., Philip, B., Buehrer, B. M., Errede, B., and Levin, D. E. (1999). Mid2 microscopy study of the response of Saccharomyces cerevisiae to ethanol stress.
is a putative sensor for cell integrity signaling in Saccharomyces cerevisiae. Mol. Appl. Environ. Microbiol. 82, 4789–4801. doi: 10.1128/AEM.01213-16
Cell. Biol. 19, 3969–3976. doi: 10.1128/mcb.19.6.3969
Schiavone, M., Vax, A., Formosa, C., Martin-Yken, H., Dague, E., and François,
Ram, A. F. J., Kapteyn, J. C., Montijn, R. C., Caro, L. H. P., Douwes, J. E., J. M. (2014). A combined chemical and enzymatic method to determine
Baginsky, W., et al. (1998). Loss of the plasma membrane-bound protein gas1p quantitatively the polysaccharide components in the cell wall of yeasts. FEMS Yeast
in Saccharomyces cerevisiae results in the release of β1,3-glucan into the medium Res. 14, 933–947. doi: 10.1111/1567-1364.12182
and induces a compensation mechanism to ensure cell wall integrity. J. Bacteriol.
180, 1418–1424. doi: 10.1128/jb.180.6.1418-1424.1998 Schöppner, P., Lutz, A. P., Lutterbach, B. J., Brückner, S., Essen, L.-O., and
Mösch, H.-U. (2022). Structure of the yeast cell wall integrity sensor Wsc1 reveals
Ramírez-Cota, G. Y., López-Villegas, E. O., Jiménez-Aparicio, A. R., and an essential role of surface-exposed aromatic clusters. J. Fungi 8:379. doi: 10.3390/
Hernández-Sánchez, H. (2021). Modeling the ethanol tolerance of the probiotic jof8040379
yeast Saccharomyces cerevisiae var. boulardii CNCM I-745 for its possible use in
a functional beer. Probiot. Antimicrob. Proteins 13, 187–194. doi: 10.1007/s12602- Seike, T., Narazaki, Y., Kaneko, Y., Shimizu, H., and Matsuda, F. (2021).
020-09680-5 Random transfer of Ogataea polymorpha genes into Saccharomyces cerevisiae
reveals a complex background of heat tolerance. J. Fungi 7:302. doi: 10.3390/
Ribeiro, R. A., Godinho, C. P., Vitorino, M. V., Robalo, T. T., Fernandes, F., jof7040302
Rodrigues, M. S., et al. (2022). Crosstalk between yeast cell plasma membrane
ergosterol content and cell wall stiffness under acetic acid stress involving Pdr18. Sekova, V. Y., Dergacheva, D. I., Isakova, E. P., Gessler, N. N., Tereshina,
J. Fungi. 8:103. doi: 10.3390/jof8020103 V. M., and Deryabina, Y. I. (2019). Soluble sugar and lipid readjustments in
the Yarrowia lipolytica yeast at various temperatures and pH. Metabolites 9:307.
Ribeiro, R. A., Vitorino, M. V., Godinho, C. P., Bourbon-Melo, N., Robalo, T. T., doi: 10.3390/metabo9120307
Fernandes, F., et al. (2021). Yeast adaptive response to acetic acid stress involves
structural alterations and increased stiffness of the cell wall. Sci. Rep. 11:12652. Serra-Cardona, A., Canadell, D., and Ariño, J. (2015). Coordinate responses
doi: 10.1038/s41598-021-92069-3 to alkaline pH stress in budding yeast. Microb. Cell 2, 182–196. doi: 10.15698/
mic2015.06.205
Rodríguez-Peña, J. M., García, R., Nombela, C., and Arroyo, J. (2010). The
high-osmolarity glycerol (HOG) and cell wall integrity (CWI) signalling pathways Serrano, R., Martín, H., Casamayor, A., and Ariño, J. (2006). Signaling alkaline
interplay: a yeast dialogue between MAPK routes. Yeast 27, 495–502. doi: 10.1002/ pH stress in the yeast Saccharomyces cerevisiae through the Wsc1 cell surface
yea sensor and the Slt2 MAPK pathway. J. Biol. Chem. 281, 39785–39795. doi: 10.1074/
jbc.M604497200
Roemer, T., Paravicini, G., Payton, M. A., and Bussey, H. (1994).
Characterization of the yeast (1–>6)-beta-glucan biosynthetic components, Sha, W., Martins, A. M., Laubenbacher, R., Mendes, P., and Shulaev, V. (2013).
Kre6p and Skn1p, and genetic interactions between the PKC1 pathway and The genome-wide early temporal response of saccharomyces cerevisiae to oxidative
extracellular matrix assembly. J. Cell Biol. 127, 567–579. doi: 10.1083/jcb.127.2.567 stress induced by cumene hydroperoxide. PLoS One 8:1–15. doi: 10.1371/journal.
pone.0074939
Roncero, C. (2002). The genetic complexity of chitin synthesis in fungi. Curr.
Genet. 41, 367–378. doi: 10.1007/s00294-002-0318-7 Simões, T., Mira, N., Fernandes, A., and Sá-correia, I. (2006). The SPI1
gene, encoding a glycosylphosphatidylinositol-anchored cell wall protein, plays a
Rossignol, T., Dulau, L., Julien, A., and Blondin, B. (2003). Genome-wide prominent role in the development of yeast resistance to lipophilic weak-acid food
monitoring of wine yeast gene expression during alcoholic fermentation. Yeast 20, preservatives. Appl. Environ. Microbiol. 18, 7168–7165. doi: 10.1128/AEM.01476-
1369–1385. doi: 10.1002/yea.1046 06
Runguphan, W., and Keasling, J. D. (2014). Metabolic engineering of Simpson, W. J., and Hammond, J. R. M. (1989). The response of brewing yeasts
Saccharomyces cerevisiae for production of fatty acid-derived biofuels and to acid washing. J. Inst. Brew. 95, 347–354.
chemicals. Metab. Eng. 21, 103–113. doi: 10.1016/j.ymben.2013.07.003 Skoneczny, M. (2018). Stress Response Mechanisms in Fungi: Theoretical and
Rzechonek, D. A., Day, A. M., Quinn, J., and Miroñczuk, A. M. (2018). Influence Practical Aspects. Cham: Springer, doi: 10.1007/978-3-030-00683-9
of ylHog1 MAPK kinase on Yarrowia lipolytica stress response and erythritol Songdech, P., Ruchala, J., Semkiv, M. V., Jensen, L. T., Sibirny, A.,
production. Sci. Rep. 8, 1–12. doi: 10.1038/s41598-018-33168-6 Ratanakhanokchai, K., et al. (2020). Overexpression of transcription factor ZNF1
Sahara, T., Goda, T., and Ohgiya, S. (2002). Comprehensive expression analysis of glycolysis improves bioethanol productivity under high glucose concentration
of time-dependent genetic responses in yeast cells to low temperature. J. Biol. and enhances acetic acid tolerance of Saccharomyces cerevisiae. Biotechnol. J. 15,
Chem. 277, 50015–50021. doi: 10.1074/jbc.M209258200 1–10. doi: 10.1002/biot.201900492
Saini, S., and Sharma, K. K. (2021). Fungal lignocellulolytic enzymes and Sousa-Lopes, A., Antunes, F., Cyrne, L., and Marinho, H. S. (2004). Decreased
lignocellulose: a critical review on their contribution to multiproduct biorefinery cellular permeability to H2 O2 protects Saccharomyces cerevisiae cells in stationary
phase against oxidative stress. FEBS Lett. 578, 152–156. doi: 10.1016/j.febslet.2004. kinase required for cell integrity. Eukaryot. Cell 6, 744–752. doi: 10.1128/EC.
10.090 00009-07
Staleva, L., Hall, A., and Orlow, S. J. (2004). Oxidative stress activates FUS1 and Udom, N., Chansongkrow, P., Charoensawan, V., and Auesukaree, C. (2019).
RLM1 transcription in the yeast Saccharomyces cerevisiae in an oxidant-dependent Coordination of the cell wall integrity and highosmolarity glycerol pathways in
manner. Mol. Biol. Cell 15, 5574–5582. doi: 10.1091/mbc.E04-02-0142 response to ethanol stress in Saccharomyces cerevisiae. Appl. Environ. Microbiol.
85, e551. doi: 10.1128/AEM.00551-19
Stanley, D., Bandara, A., Fraser, S., Chambers, P. J., and Stanley, G. A. (2010).
The ethanol stress response and ethanol tolerance of Saccharomyces cerevisiae. Ullah, A., Chandrasekaran, G., Brul, S., and Smits, G. J. (2013). Yeast adaptation
J. Appl. Microbiol. 109, 13–24. doi: 10.1111/j.1365-2672.2009.04657.x to weak acids prevents futile energy expenditure. Front. Microbiol. 4:142. doi:
10.3389/fmicb.2013.00142
Stewart, G. G. (2017). “The structure and function of the yeast cell wall, plasma
membrane and periplasm,” in Brewing and Distilling Yeasts, ed. G. G. Stewart Usmani, Z., Sharma, M., Awasthi, A. K., Lukk, T., Tuohy, M. G., Gong, L., et al.
(Cham: Springer), 55–75. doi: 10.1007/978-3-319-69126-8_5 (2021). Lignocellulosic biorefineries: the current state of challenges and strategies
Stovicek, V., Dato, L., Almqvist, H., Schöpping, M., Chekina, K., Pedersen, L. E., for efficient commercialization. Renew. Sustain. Energy Rev. 148:111258. doi: 10.
et al. (2022). Rational and evolutionary engineering of Saccharomyces cerevisiae 1016/j.rser.2021.111258
for production of dicarboxylic acids from lignocellulosic biomass and exploring Valdivieso, M. H., Ferrario, L., Vai, M., Duran, A., and Popolo, L. (2000).
genetic mechanisms of the yeast tolerance to the biomass hydrolysate. Biotechnol. Chitin synthesis in a gas1 mutant of Saccharomyces cerevisiae. J. Bacteriol. 182,
Biofuels Bioprod. 15:22. doi: 10.1186/s13068-022-02121-1 4752–4757. doi: 10.1128/JB.182.17.4752-4757.2000
Sugiyama, M., Akase, S. P., Nakanishi, R., Horie, H., Kaneko, Y., and Harashima, Valiante, V., Macheleidt, J., Föge, M., and Brakhage, A. A. (2015). The Aspergillus
S. (2014). Nuclear localization of Haa1, which is linked to its phosphorylation fumigatus cell wall integrity signalling pathway: drug target, compensatory
status, mediates lactic acid tolerance in Saccharomyces cerevisiae. Appl. Environ. pathways and virulence. Front. Microbiol. 6:1–12. doi: 10.3389/fmicb.2015.
Microbiol. 80, 3488–3495. doi: 10.1128/AEM.04241-13 00325
Suhr, K. I., and Nielsen, P. V. (2004). Effect of weak acid preservatives on growth van Leeuwe, T. M., Arentshorst, M., Punt, P. J., and Ram, A. F. J. (2020).
of bakery product spoilage fungi at different water activities and pH values. Int. J. Interrogation of the cell wall integrity pathway in Aspergillus niger identifies a
Food Microbiol. 95, 67–78. doi: 10.1016/j.ijfoodmicro.2004.02.004 putative negative regulator of transcription involved in chitin deposition. Gene
Suvarna, S., Dsouza, J., Ragavan, M. L., and Das, N. (2018). Potential probiotic 5:100028. doi: 10.1016/j.gene.2020.100028
characterization and effect of encapsulation of probiotic yeast strains on survival Van Maris, A. J. A., Winkler, A. A., Porro, D., Van Dijken, J. P., and Pronk, J. T.
in simulated gastrointestinal tract condition. Food Sci. Biotechnol. 27, 745–753. (2004). Homofermentative lactate production cannot sustain anaerobic growth of
doi: 10.1007/s10068-018-0310-8 engineered Saccharomyces cerevisiae: possible consequence of energy-dependent
Suzuki, T., Sugiyama, M., Wakazono, K., Kaneko, Y., and Harashima, S. (2012). lactate export. Appl. Environ. Microbiol. 70, 2898–2905. doi: 10.1128/AEM.70.5.
Lactic-acid stress causes vacuolar fragmentation and impairs intracellular amino- 2898-2905.2004
acid homeostasis in Saccharomyces cerevisiae. J. Biosci. Bioeng. 113, 421–430. doi: van Uden, N. (1985). Ethanol toxicity and ethanol tolerance in yeasts. Annu.
10.1016/j.jbiosc.2011.11.010 Reports Ferment. Proc. 8, 11–58.
Swinnen, S., Henriques, S. F., Shrestha, R., Ho, P. W., Correia, I. S., and van Voorst, F., Houghton-Larson, J., Jønson, L., Kielland-Brandt, M. C., and
Nevoigt, E. (2017). Improvement of yeast tolerance to acetic acid through Haa1 Brandt, A. (2006). Genome-wide identification of genes required for growth of
transcription factor engineering: towards the underlying mechanisms. Microb. Cell Saccharomyces cerevisiae under ethanol stress. Yeast 23, 351–359. doi: 10.1002/
Fact. 16, 1–15. doi: 10.1186/s12934-016-0621-5 yea.1359
Tanaka, K., Ishii, Y., Ogawa, J., and Shima, J. (2012). Enhancement of acetic Varol, D., Purutçuoǧlu, V., and Yılmaz, R. (2018). Transcriptomic analysis of
acid tolerance in Saccharomyces cerevisiae by overexpression of the Haa1 gene, the heat stress response for a commercial baker’s yeast Saccharomyces cerevisiae.
encoding a transcriptional activator. Appl. Environ. Microbiol. 78, 8161–8163. Genes Geno. 40, 137–150. doi: 10.1007/s13258-017-0616-6
doi: 10.1128/AEM.02356-12
Vaz, E., Lee, J., and Levin, D. (2020). Crosstalk between S. cerevisiae SAPKs
Tanaka, S., and Tani, M. (2018). Mannosylinositol phosphorylceramides and Hog1 and Mpk1 is mediated by glycerol accumulation. Fungal Biol. 124, 361–367.
ergosterol coodinately maintain cell wall integrity in the yeast Saccharomyces doi: 10.1016/j.funbio.2019.10.002
cerevisiae. FEBS J. 285, 2405–2427. doi: 10.1111/febs.14509
Vélez-Segarra, V., González-Crespo, S., Santiago-Cartagena, E., Vázquez-
Taymaz-Nikerel, H., Cankorur-Cetinkaya, A., and Kirdar, B. (2016). Genome- Quiñones, L. E., Martínez-Matías, N., Otero, Y., et al. (2020). Protein interactions
wide transcriptional response of Saccharomyces cerevisiae to stress-induced of the mechanosensory proteins Wsc2 and Wsc3 for stress resistance in
perturbations. Front. Bioeng. Biotechnol. 4:17. doi: 10.3389/fbioe.2016.00017 Saccharomyces cerevisiae. G3 Gen. Geno. Genet. 10, 3121–3135. doi: 10.1534/g3.
Teixeira, M. C., Raposo, L. R., Mira, N. P., Lourenço, A. B., and Sá-Correia, 120.401468
I. (2009). Genome-wide identification of Saccharomyces cerevisiae genes required Verghese, J., Abrams, J., Wang, Y., and Morano, K. A. (2012). Biology of the heat
for maximal tolerance to ethanol. Appl. Environ. Microbiol. 75, 5761–5772. doi: shock response and protein chaperones: budding yeast (Saccharomyces cerevisiae)
10.1128/AEM.00845-09 as a model system. Microbiol. Mol. Biol. Rev. 76, 115–158. doi: 10.1128/MMBR.
Teixeira, M. C., Raposo, L. R., Palma, M., and Sá-Correia, I. (2010). 05018-11
Identification of genes required for maximal tolerance to high-glucose Verna, J., Lodder, A., Lee, K., Vagts, A., and Ballester, R. (1997). A family of
concentrations, as those present in industrial alcoholic fermentation media, genes required for maintenance of cell wall integrity and for the stress response
through a chemogenomics approach. OMICS 14, 201–210. doi: 10.1089/omi.2009. in Saccharomyces cerevisiae. Proc. Natl. Acad. Sci. U.S.A. 94, 13804–13809. doi:
0149 10.1073/pnas.94.25.13804
Tenreiro, S., Nunes, P. A., Viegas, C. A., Neves, M. S., Teixeira, M. C., Cabral, Viladevall, L., Serrano, R., Ruiz, A., Domenech, G., Giraldo, J., Barceló, A.,
M. G., et al. (2002). AQR1 gene (ORF YNL065w) encodes a plasma membrane et al. (2004). Characterization of the calcium-mediated response to alkaline stress
transporter of the major facilitator superfamily that confers resistance to short- in Saccharomyces cerevisiae. J. Biol. Chem. 279, 43614–43624. doi: 10.1074/jbc.
chain monocarboxylic acids and quinidine in Saccharomyces cerevisiae. Biochem. M403606200
Biophys. Res. Commun. 292, 741–748. doi: 10.1006/bbrc.2002.6703
Thomé, P. E. (2007). Cell wall involvement in the glycerol response to Vilella, F., Herrero, E., Torres, J., and De La Torre-Ruiz, M. A. (2005). Pkc1 and
high osmolarity in the halotolerant yeast Debaryomyces hansenii. antonie van the upstream elements of the cell integrity pathway in Saccharomyces cerevisiae,
leeuwenhoek. Int. J. Gen. Mol. Microbiol. 91, 229–235. doi: 10.1007/s10482-006- Rom2 and Mtl1, are required for cellular responses to oxidative stress. J. Biol.
9112-8 Chem. 280, 9149–9159. doi: 10.1074/jbc.M411062200
Thorwall, S., Schwartz, C., Chartron, J. W., and Wheeldon, I. (2020). Wada, K., Fujii, T., Akita, H., and Matsushika, A. (2020). IoGAS1, a GPI-
Stress-tolerant non-conventional microbes enable next-generation chemical anchored protein derived from Issatchenkia orientalis, confers tolerance of
biosynthesis. Nat. Chem. Biol. 16, 113–121. doi: 10.1038/s41589-019-0452-x Saccharomyces cerevisiae to multiple acids. Appl. Biochem. Biotechnol. 190, 1349–
1359. doi: 10.1007/s12010-019-03187-8
Torres, L., Martín, H., García-Saez, M. I., Arroyo, J., Molina, M., Sánchez,
M., et al. (1991). A protein kinase gene complements the lytic phenotype of Walker, G. M., and Basso, T. O. (2020). Mitigating stress in
Saccharomyces cerevisiae lyt2 mutants. Mol. Microbiol. 11, 2845–2854. doi: 10. industrial yeasts. Fungal Biol. 124, 387–397. doi: 10.1016/j.funbio.2019.
1111/j.1365-2958.1991.tb01993.x 10.010
Truman, A. W., Millson, S. H., Nuttall, J. M., Mollapour, M., Prodromou, C., Walker, G., and Stewart, G. (2016). Saccharomyces cerevisiae in the
and Piper, P. W. (2007). In the yeast heat shock response, Hsf1-directed induction production of fermented beverages. Beverages 2:30. doi: 10.3390/beverages20
of Hsp90 facilitates the activation of the Slt2 (Mpk1) mitogen-activated protein 40030
Wang, G., Kell, D. B., and Borodina, I. (2021). Harnessing the yeast Yen, H. W., Chen, T. J., Pan, W. C., and Wu, H. J. (2010). Effects of neutralizing
Saccharomyces cerevisiae for the production of fungal secondary metabolites. agents on lactic acid production by Rhizopus oryzae using sweet potato starch.
Essays Biochem. 65, 277–291. doi: 10.1042/EBC20200137 World J. Microbiol. Biotechnol. 26, 437–441. doi: 10.1007/s11274-009-0186-0
Wei, X., Chi, Z., Liu, G.-L., Hu, Z., and Chi, Z.-M. (2021). The genome-wide Zeng, L., Huang, J., Feng, P., Zhao, X., Si, Z., Long, X., et al. (2022).
mutation shows the importance of cell wall integrity in growth of the psychrophilic Transcriptomic analysis of formic acid stress response in Saccharomyces
yeast metschnikowia australis W7-5 at different temperatures. Microb. Ecol. 81, cerevisiae. World J. Microbiol. Biotechnol. 38:34. doi: 10.1007/s11274-021-
52–66. doi: 10.1007/s00248-020-01577-8 03222-z
Wu, H., Zheng, X., Araki, Y., Sahara, H., Takagi, H., and Shimoi, H. (2006). Zhang, C., Ma, Y., Miao, H., Tang, X., Xu, B., Wu, Q., et al. (2020).
Global gene expression analysis of yeast cells during sake brewing. Appl. Environ. Transcriptomic analysis of Pichia pastoris (Komagataella phaffii) GS115
Microbiol. 72, 7353–7358. doi: 10.1128/AEM.01097-06 during heterologous protein production using a high-cell-density fed-
batch cultivation strategy. Front. Microbiol. 11:463. doi: 10.3389/fmicb.2020.
Xu, Y., Yang, H., Brennan, C. S., Coldea, T. E., and Zhao, H. (2020). Cellular
00463
mechanism for the improvement of multiple stress tolerance in brewer’s yeast
by potassium ion supplementation. Int. J. Food Sci. Technol. 55, 2419–2427. doi: Zhang, Q., and Xu, Y. (2021). UDP-glucose pyrophosphorylase ugp1 is required
10.1111/ijfs.14491 in heat stress response of Saccharomyces cerevisiae. J. Bio. Mater. Bioenergy 15,
188–193. doi: 10.1016/j.bbrc.2015.10.090
Yamaguchi, M., Namiki, Y., Okada, H., Mori, Y., Furukawa, H., and Wang, J.
(2011). Structome of Saccharomyces cerevisiae determined by freeze-substitution Zhao, C., Jung, U. S., Garrett-Engele, P., Roe, T., Cyert, M. S., and Levin,
and serial ultrathin-sectioning electron microscopy. J. Electron Microsci. 60, 321– D. E. (1998). Temperature-induced expression of yeast FKS2 is under the dual
335. doi: 10.1093/jmicro/dfr052 control of protein kinase C and calcineurin. Mol. Cell. Biol. 18, 1013–1022. doi:
10.1128/MCB.18.2.1013
Yamakawa, C. K., Qin, F., and Mussatto, S. I. (2018). Advances and
opportunities in biomass conversion technologies and biorefineries for the Zhao, X. Q., Xue, C., Ge, X. M., Yuan, W. J., Wang, J. Y., and
development of a bio-based economy. Bio. Bioenergy 119, 54–60. doi: 10.1016/j. Bai, F. W. (2009). Impact of zinc supplementation on the improvement
jbiotec.2018.07.041 of ethanol tolerance and yield of self-flocculating yeast in continuous
ethanol fermentation. J. Biotechnol. 139, 55–60. doi: 10.1016/j.jbiotec.2008.
Yang, Z., Huang, J., Geng, J., Nair, U., and Klionsky, J. (2006). Atg22 recycles
08.013
amino acids to link the degradative and recycling functions of autophagy. Mol.
Biol. Cell 17, 5094–5104. doi: 10.1091/mbc.E06 Zhong, W., Yang, M., Hao, X., Sharshar, M. M., Wang, Q., and Xing,
J. (2021). Improvement of D-lactic acid production at low pH through
Yasokawa, D., Murata, S., Iwahashi, Y., Kitagawa, E., Nakagawa, R., Hashido, expressing acid-resistant gene IoGAS1 in engineered Saccharomyces
T., et al. (2010). Toxicity of methanol and formaldehyde towards Saccharomyces cerevisiae. J. Chem. Technol. Biotechnol. 96, 732–742. doi: 10.1002/jctb.
cerevisiae as assessed by DNA microarray analysis. Appl. Biochem. Biotechnol. 160, 6587
1685–1698. doi: 10.1007/s12010-009-8684-y
Zlotnik, H., Pilar Fernandez, M., Bowers, B., and Cabib, E. (1984).
Yazawa, H., Iwahashi, H., and Uemura, H. (2007). Disruption of URA7 and Saccharomyces cerevisiae mannoproteins form an external cell wall layer that
GAL6 improves the ethanol tolerance and fermentation capacity of Saccharomyces determines wall porosity. J. Bacteriol. 159, 1018–1026. doi: 10.1128/jb.159.3.1018-
cerevisiae. Yeast 24, 551–560. doi: 10.1002/yea.1492 1026.1984